A publishing partnership

Inferred Evidence for Dark Matter Kinematic Substructure with SDSS–Gaia

, , and

Published 2019 March 15 © 2019. The American Astronomical Society. All rights reserved.
, , Citation Lina Necib et al 2019 ApJ 874 3 DOI 10.3847/1538-4357/ab095b

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/874/1/3

Abstract

We use the distribution of accreted stars in Sloan Digital Sky Survey–Gaia DR2 to demonstrate that a nontrivial fraction of the dark matter halo within galactocentric radii of 7.5–10 kpc and $| z| \gt 2.5\,\mathrm{kpc}$ is in substructure and thus may not be in equilibrium. Using a mixture likelihood analysis, we separate the contributions of an old, isotropic stellar halo and a younger anisotropic population. The latter dominates and is uniform within the region studied. It can be explained as the tidal debris of a disrupted massive satellite on a highly radial orbit and is consistent with mounting evidence from recent studies. Simulations that track the tidal debris from such mergers find that the dark matter traces the kinematics of its stellar counterpart. If so, our results indicate that a component of the nearby dark matter halo that is sourced by luminous satellites is in kinematic substructure referred to as debris flow. These results challenge the Standard Halo Model, which is discrepant with the distribution recovered from the stellar data, and have important ramifications for the interpretation of direct detection experiments.

Export citation and abstract BibTeX RIS

1. Introduction

The motions of nearby stars have proven invaluable in the study of the local dark matter (DM) distribution, starting from the work of Kapteyn (1922), Jeans (1922), and Oort (1932). As the mapping of stellar velocities has dramatically improved with surveys such as Hipparcos (ESA 1997), the Sloan Digital Sky Survey (SDSS; Ahn et al. 2012), and the RAdial Velocity Experiment (Kunder et al. 2017), the measurement of the local DM density has greatly improved (Read 2014). In contrast, the DM velocity distribution has remained relatively unexplored. In this work, we use data from the second Gaia data release (Gaia Collaboration et al. 2016, 2018), cross-matched with SDSS, to characterize the local population of accreted stars and infer properties of the DM velocity distribution.

In the ΛCDM paradigm, the Milky Way halo is built from the mergers of smaller satellite galaxies (White & Rees 1978). As these galaxies fall into the Milky Way, they experience strong tidal forces that gradually tear them apart. In the initial stages of disruption, the DM in the outskirts of the satellite is stripped away, but as the galaxy is slowly eaten down, its more tightly bound stars and DM are also removed. This tidal debris litters the Milky Way, a fossil remnant of the Galaxy's accretion history.

DM that accreted early has time to virialize with the host galaxy. We can thus imagine that it forms an isotropic isothermal halo (Ostriker et al. 1974; Bahcall & Soneira 1980; Caldwell & Ostriker 1981) with a mass distribution that is consistent with a flat rotation curve (Burton & Gordon 1978; Blitz 1979; Clemens 1985; Knapp et al. 1985; Fich et al. 1989; Pont et al. 1994). Self-consistently, the local DM velocities follow a Maxwell–Boltzmann distribution (Drukier et al. 1986; Freese et al. 1988). This scenario is often referred to as the Standard Halo Model (SHM).

More recent mergers, however, can leave residual structures in the DM phase-space distribution. For example, the debris from the most recent mergers may be in cold phase-space structures called tidal streams, which are coherent in position and velocity space (Diemand et al. 2008; Vogelsberger et al. 2009; Zemp et al. 2009; Kuhlen et al. 2010; Elahi et al. 2011; Maciejewski et al. 2011; Vogelsberger & White 2011). With time, the tidal debris becomes increasingly more mixed. In this process, any velocity features typically persist longer than spatial ones (Helmi & White 1999). The resulting kinematic substructure, referred to as debris flow, may consist of many overlapping streams, shells, or plumes from the debris of one or more satellites that have made several orbits before dissolving (Lisanti & Spergel 2012; Kuhlen et al. 2012).

Numerical simulations are an invaluable tool in exploring the range of DM distributions possible in Milky Way–like galaxies. They have demonstrated, for example, that the solar neighborhood could have been built from a wide variety of accretion histories (Diemand et al. 2007; Wang et al. 2011). DM-only simulations find fairly marked deviations from the Maxwell–Boltzmann assumption, especially on the high-end tail of the velocity distribution (Fairbairn & Schwetz 2009; March-Russell et al. 2009; Vogelsberger et al. 2009; Kuhlen et al. 2010; Mao et al. 2013). Simulations that also model the baryonic physics typically exhibit closer alignment with the Maxwell–Boltzmann distribution, although significant scatter is observed between them (Ling et al. 2010; Pillepich et al. 2014; Bozorgnia et al. 2016; Kelso et al. 2016; Sloane et al. 2016; Bozorgnia & Bertone 2017).

Ultimately, we would like to determine the local DM distribution from observations. One possibility is to use the motion of stars to constrain the local gravitational potential (or density) and to subsequently infer the velocity distribution using the Jeans theorem. A variety of proposals of this nature have been made (Hansen & Moore 2006; Chaudhury et al. 2010; Lisanti et al. 2011; Catena & Ullio 2012; Bhattacharjee et al. 2013; Bozorgnia et al. 2013; Fornasa & Green 2014; Mandal et al. 2018), but they typically rely on the assumption that the DM is isotropic and/or in equilibrium, either of which may be violated depending on the Milky Way's accretion history.

An alternative proposal is to identify populations of stars that share the same kinematics as the DM. These are stars that were also accreted onto the Milky Way from merging satellites. As such, they typically have distinctive kinematic and chemical properties compared to the population that is born in the Galaxy. Using accreted stars as direct kinematic tracers for the DM is beneficial because it makes no assumption about steady state. The potential downside to this approach is that it does not account for DM that originates from nonluminous satellites or that was accreted diffusely.

Numerical simulations have demonstrated excellent correspondence between the DM and accreted stars. Herzog-Arbeitman et al. (2018a) recently showed that the oldest and most metal-poor stars in the halo trace the velocities of the virialized DM using the Eris hydrodynamic simulation. Stellar substructure in the form of debris flow also traces similar kinematic features in the DM distribution, as was demonstrated using the Via Lactea simulation (Lisanti et al. 2015). As this substructure arises from more recent accretion events than the virialized component, it may be associated with more metal-rich stars. More recently, the correspondence between the stellar and DM velocities has also been explicitly demonstrated (for the virialized component and debris flow at redshift z = 0) in two additional simulated halos in the Fire suite (Necib et al. 2018).

These arguments motivate a close study of the local accreted stellar population as a means of characterizing the DM distribution. Recent observations have begun to change our understanding of the stellar halo, disfavoring the viewpoint that a large fraction was born in situ from stars that were kicked up from the Galactic disk (Helmi 2008). In contrast, evidence has been building for a two-component model that consists of an isotropic population from old accretion events, and an anisotropic population from a more recent—and quite significant—merger (Belokurov et al. 2018a; Deason et al. 2018; Helmi et al. 2018; Lancaster et al. 2018; Myeong et al. 2018a, 2018b). In this picture, the majority of the local stellar halo originates from this one merger, which also deposits DM. This work presents the first modeling of the velocity and metallicities of the stars accreted in this event, providing clues about the corresponding DM debris as well.

The second Gaia data release (DR2) presents a unique opportunity to study the kinematics of this accreted population. Cross-matching Gaia DR2 with SDSS yields metallicities for 193,162 of its stars. This data set allows us to characterize the velocity distribution of 21,443 stars with metallicities down to [Fe/H] ∼ −3, galactocentric radii of 7.5–10 kpc, and $2.5\,\mathrm{kpc}\lt | z| \lt 10\,\mathrm{kpc}$. We build a model that successfully describes the observations over the full range of metallicities and velocities. As we will argue, the properties of the local stellar distribution suggest that a fraction of the local DM is in substructure called debris flow, challenging the SHM.

Characterizing the DM velocity distribution is critical for interpreting results of direct detection experiments, which search for the recoils of nuclei from a collision with a DM particle (see Jungman et al. 1996; Freese et al. 2013, for reviews). The rate of such scattering interactions depends on the DM speed upon collision. Indeed, for certain DM speeds and/or models the velocity distribution can make the difference between observing a signal or seeing nothing at all. The characterization of the DM velocity distribution function is one of the largest sources of uncertainty in the interpretation of such experimental results (Del Nobile 2014; Green 2017). If a subset of the local DM is indeed in substructure, it could potentially change the landscape of exclusion limits on DM masses and scattering cross sections. We, however, caution that the analysis performed in this paper only holds for $| z| \gt 2.5\,\mathrm{kpc}$. While we extrapolate the results to the Sun's position for illustration, a dedicated study is needed to confirm the behavior closer to the midplane.

Our paper is organized as follows. Section 2 describes the data preparation and likelihood procedure used in the study. Section 3 presents the results of the SDSS–Gaia DR2 analysis and describes the characteristics of the disk, halo, and substructure stars that we identify. Section 4 discusses the implications of this stellar substructure for DM detection. Figures 8 and 9 summarize the impact on experimental limits for spin-independent interactions. We conclude in Section 5. The Appendix includes supplementary material that further substantiates the results presented in the main text. Interpolations of the heliocentric velocity distribution, which can be used to calculate DM scattering rates, are provided at the following github repository: https://linoush.github.io/DM_Velocity_Distribution/.

2. Analysis Methodology

2.1. Data Selection

In this analysis, we use the SDSS DR9 (Ahn et al. 2012) data set, cross-matched to Gaia DR2 (Gaia Collaboration et al. 2018). The cross-match was performed inside the Whole Sky Database (WSDB), an archive providing SQL access to catalogs from all major wide-area surveys. In particular, we utilized Q3C, spatial indexing, and cross-matching plug-in (Koposov & Bartunov 2006) to select the nearest SDSS neighbor for each Gaia source within a 1'' aperture, while taking into account the proper motion of all objects and the time difference between the observations. The combination of the two data sets allows us to take advantage of the large number of halo stars in the SDSS spectroscopic data set, while simultaneously using the unprecedented accuracies of the proper motions provided by the Gaia survey. From the SDSS catalog, we select main-sequence (MS) stars, which are considered standard candles in this context, that satisfy $| b| \gt 10^\circ $, Ag < 0.5 mag, σRV < 50 km s−1, ${\rm{S}}/{\rm{N}}\gt 10,3.5\,\lt \mathrm{log}(g)\lt 5$, 0.2 < gr < 0.8, 0.2 < g − i < 4, $4500\,{\rm{K}}\,\lt {T}_{\mathrm{eff}}\lt 8000\,{\rm{K}}$, and 15 < r < 19.5. All stellar magnitudes are dereddened using the maps of Schlegel et al. (1998). Admittedly, the MS stars are not true standard candles, but accurate distances can be estimated using the so-called photometric parallax method, which was recently tested against Gaia parallaxes (see Deason et al. 2018). While other possible choices exist, e.g., (i) a mix of all stellar populations with Gaia parallaxes or (ii) red giants with photometric parallaxes and (iii) blue horizontal branch stars, the MS sample has a combination of advantageous properties making it a better choice for the analysis presented here. More precisely, the MS sample is not limited by Gaia's parallax error to a relatively short distance range and is not dominated by the disk stars. The photometric parallax uncertainties for the MS stars do not vary as strongly with metallicity as, e.g., those for red giants (whose distance error can easily reach 100% for the metal-poor giants compared to the typical 10%–20% for the MS stars). Finally, MS stars are much more numerous compared to the blue horizontal branch or red giant stars.

Distances to these predominantly MS stars are calculated using Equations (A2), (A3), and (A7) of Ivezic et al. (2008), with distance errors within the region of interest of ∼2%, or of the order of ∼100 pc. Recently, the validity of this photometric parallax calibration was verified by Deason et al. (2018) for a subset of stars with accurate Gaia DR2 parallaxes. We have also confirmed that the fractional errors on the distance as derived from the parallaxes are consistently larger than those derived using the photometric parallax. The celestial coordinates, heliocentric distances, proper motions, and radial velocities are then used to calculate stellar velocity components in spherical polar coordinates. We marginalize over the local standard of rest value assuming that it is described by a Gaussian with a center at 238 km s−1 and a dispersion of 9 km s−1 (Schönrich 2012). The components of the solar peculiar motion are those presented in Coşkunoǧlu et al. (2011). The measured parameters' uncertainties—including covariances in the proper motion components—are then propagated using Monte Carlo sampling to obtain estimates of the uncertainties in the spherical velocity components (see Belokurov et al. 2018b, for details). For the analysis described in the sections below, each star's Monte Carlo samples are modeled with a Gaussian distribution to obtain the full covariance matrix.

There are no strong reasons to believe that the SDSS spectroscopic sample of MS stars used in our analysis suffers from any appreciable kinematic bias (Belokurov et al. 2018b). However, as the sample uses a mixture of SDSS target categories, a moderate metallicity bias toward more metal-poor stars is expected (see Yanny et al. 2009, for details). We also repeated our benchmark analysis selecting only F/G stars, which are known to have no biases in either metallicity or kinematics. This reduces the sample size substantially, increasing the uncertainties on the recovered fit parameters. However, the overall results—most importantly the fractional contribution of the stellar components—remain essentially unchanged, within uncertainties.

Figure 1 shows the spatial distribution of stellar counts in SDSS–Gaia DR2, as a function of galactocentric radius, r, and vertical distance from the Galactic plane, z.

Figure 1.

Figure 1. Spatial distribution of the SDSS–Gaia DR2 sample in terms of galactocentric radius, r, and vertical distance from the Galactic plane, z. We characterized the disk, halo, and substructure populations in regions within the dashed aqua box, which spans r ∈ [7.5, 10.0] kpc and $| z| \gt 2.5\,\mathrm{kpc}$.

Standard image High-resolution image

2.2. Model Motivation

Figure 2 shows the distribution of the spherical velocity components as a function of stellar metallicity5 for the SDSS–Gaia DR2 subsample within r ∈ [7.5, 8.5] kpc and $| z| \gt 2.5\,\mathrm{kpc}$. For Figure 2 (and subsequent figures), we sample from the posterior distributions of the mean and covariance matrices to construct the velocity components vr, vθ, vϕ for the best-fit model.

Figure 2.

Figure 2. Chemodynamic distribution of stars in the SDSS–Gaia DR2 sample within r ∈ [7.5, 8.5] kpc and $| z| \gt 2.5\,\mathrm{kpc}$. The panels show how the distributions vary in iron abundance [Fe/H] and the spherical galactocentric radial coordinates vr (left), vθ (middle), and vϕ (right). The disk population is pronounced at [Fe/H] ∼ −0.8, and a nearly isotropic halo population is apparent at [Fe/H] ≲ −1.8. A highly radial population at [Fe/H] ∼ −1.4 constitutes a large fraction of the sample and is an example of kinematic substructure. The 95% contours of the distributions recovered from the likelihood analysis are also shown (see Section 2.3); the disk, halo, and substructure best fits are shown in green, pink, and blue, respectively.

Standard image High-resolution image

Several features in Figure 2 are apparent by visual inspection. The first is the disk population, which is centered at [Fe/H] ∼ −0.8, vr,θ ∼ 0 km s−1, and vϕ ∼ 130 km s−1. The second is a population with [Fe/H] ∼ −1.4 with large radial anisotropy. The third is a population that extends down to [Fe/H] ≲ −1.8 with nearly isotropic velocities.6

In this work, we will refer to the second population as "substructure" and the third population as the "halo." We are envisioning that both originate from the disruption of accreted satellites in the Milky Way. What we refer to as the "halo" is intended to encapsulate the tidal debris from the oldest mergers, which will typically be the most metal-poor and fully well mixed in phase space. What we call "substructure" constitutes tidal debris that is not fully phase mixed; such a component may exhibit interesting features in spatial and/or velocity coordinates, such as streams or debris flow. The prefix "sub-" suggests that this population is less dominant than the halo population; we adopt this terminology, as it is standard in the DM literature, but make no assumptions on its relative dominance in our study.

Evidence has been building for a multicomponent inner stellar halo that is dominated by the tidal debris of one massive merger (Deason et al. 2015; Fiorentino et al. 2015; Belokurov et al. 2018a; Helmi et al. 2018; Myeong et al. 2018a). The large radial anisotropy of the stars with [Fe/H] ≳ −1.7 in Figure 2 was first identified using the SDSS–Gaia DR1 sample (Belokurov et al. 2018b).7 This work noted that the "sausage"–like feature in the data appears to be non-Gaussian and estimated its contribution to be ∼66% of the nondisk population over the full SDSS footprint. They found that the radial anisotropy of the sample drops markedly at [Fe/H] ≲ −1.7, suggesting that a separate isotropic and metal-poor population is also present. It is unlikely that the radial and isotropic populations originated in the Milky Way, as their iron abundances are in line with those observed in Milky Way dwarf spheroidal galaxies (Venn et al. 2004) and their velocities are distinct from disk stars.

Recent work using local MS stars from SDSS–Gaia DR2, as well as a separate sample of more distant blue horizontal branch stars, demonstrated that the orbits of the most highly eccentric stars share a common apocenter at r ∼ 20 kpc (Deason et al. 2018). This radius is coincident with the observed break in the Milky Way's stellar density distribution (Deason et al. 2013), suggesting that the radial stellar population is the tidal debris of a recent and large merger that dominates the inner halo. Simulations have shown that the density of stars from a particular merger may exhibit a break at the average apocenter of the stars (Deason et al. 2013).

If the radial substructure is indeed associated with a large merger, one might expect that globular clusters were also stripped from the satellite progenitor as it was disrupted. Indeed, a number of globular clusters on highly radial orbits were recently identified that may be associated with the large merger(s) causing the sausage-like feature in the SDSS–Gaia data (Myeong et al. 2018b). The number of these clusters suggests a total progenitor mass of ∼1010 M; their tracks in age–metallicity space bound the maximum infall redshift to less than ∼3.

These new results have direct implications for the local DM distribution. Previous work demonstrated that the virialized DM component is traced by the most metal-poor stars in the Milky Way—this corresponds to the population that we refer to as the halo here (Herzog-Arbeitman et al. 2018a; Necib et al. 2018). The virialized DM is well tracked by stars with metallicities [Fe/H] ≲ −3 in the Eris simulation (Herzog-Arbeitman et al. 2018a) and for [Fe/H] ≲ −2 in the Fire simulation (Necib et al. 2018). The relevant metallicity range depends on the time when the earliest mergers occurred, which depends on a galaxy's detailed merger history. Additionally, the kinematic substructure observed in the data is highly reminiscent of debris flow (Kuhlen et al. 2012; Lisanti & Spergel 2012). Indeed, a study of the stellar halo in Via Lactea (where star particles were painted onto the most bound DM particles in subhalos) found precisely the same kind of radial substructure becoming apparent in the SDSS–Gaia data, and that the kinematics of the accreted stars correlate with that of the DM debris (Lisanti et al. 2015). The stellar–DM correlation for debris flow has since been verified using the Fire simulations (Necib et al. 2018). Therefore, if we want to infer the kinematic properties of the local DM, we will need to model the velocities of the halo and substructure populations.

2.3. Likelihood Procedure

To isolate the accreted stellar population, we can place a hard upper cutoff on the metallicity of the sample. The downside to this conservative approach is that it ignores the high-metallicity tail of the accreted stellar distribution that overlaps with disk stars. For this reason, we use a mixture model analysis to statistically identify the individual populations of accreted stars over the full metallicity range of the sample.

Each star, labeled by the index i, is associated with a set of observable quantities such as its velocity and metallicity, ${O}_{i}=\left({{\boldsymbol{v}}}_{i},{[\mathrm{Fe}/{\rm{H}}]}_{i}\right)$, as well as the variance for each. We assign each star a flag j = d, h, or s that designates whether it belongs to the disk, halo, or substructure population, respectively. The likelihood of observing Oi for a disk star is

Equation (1)

where θ is the set of free parameters and ${ \mathcal N }$ denotes the normal distribution. The set θ includes the velocity distribution mean ${{\boldsymbol{\mu }}}^{d}$ and covariance matrix ${{\boldsymbol{\Sigma }}}_{i}^{d}$, as well as the metallicity distribution mean ${\mu }_{[\mathrm{Fe}/{\rm{H}}]}^{d}$ and dispersion ${\sigma }_{[\mathrm{Fe}/{\rm{H}}],i}^{d}$. The covariance matrix depends on the individual velocity dispersions σr,θ,ϕ and the correlation coefficients ρ, ρ, and ρθϕ. Note that the velocity covariance matrix and the metallicity dispersion vary between stars because the observed covariance depends on the true value and the measurement error—specifically, ${{\boldsymbol{\Sigma }}}_{\mathrm{obs}}={{\boldsymbol{\Sigma }}}_{\mathrm{true}}+{{\boldsymbol{\Sigma }}}_{\mathrm{err}}$. There are 11 parameters associated with this model. The likelihood for a halo star is also given by Equation (1), except with $d\to h$, and thus comes with an additional 11 parameters.

Modeling the substructure population is more challenging, as initial evidence suggests that its radial velocities are non-Gaussian (Belokurov et al. 2018b). Therefore, we assume that the velocities are a sum of two multivariate normal distributions with equivalent parameters, except for equal and opposite mean in vr:

Equation (2)

where ${{\boldsymbol{\mu }}}^{\tilde{s}}={\left(-{\mu }_{r},{\mu }_{\theta },{\mu }_{\phi }\right)}^{s}$. This model can vary over unimodal and bimodal distributions in vr. For example, when ${\mu }_{r}^{s}\to 0$, Equation (2) approaches a single Gaussian distribution peaked at zero. In the limit where ${\mu }_{r}^{s}\gg {\sigma }_{r}^{s}$ and ${\mu }_{r}^{s}\ne 0$, the radial lobes are very pronounced. If, in contrast, ${\mu }_{r}^{s}\ll {\sigma }_{r}^{s}$, then the overlap between the two lobes increases and the radial velocity distribution is more box-like. We assume that the radial lobes, if present, are symmetric about vr = 0 km s−1, as this would be expected if the tidal debris originates from a satellite as it moves toward (vr < 0) and then away from (vr > 0) the Galactic center.8

The likelihood for the complete set of N stars is

Equation (3)

where the braces around Oi indicate the full list of N values. Qj is the probability that the star belongs to the jth population; these represent two additional parameters in the model, as Qh = 1 − Qd − Qs.

We use emcee (Foreman-Mackey et al. 2013) to find the posterior distributions of all 35 free parameters. In particular, we use 250 walkers, with 5000 steps, and a burn-in period of 10,000 steps. The priors for the separate parameters are provided in Table 1. We perform the mixture analysis in separate regions within the dashed aqua box of Figure 1, which spans the range r ∈ [7.5, 10.0] kpc and $| z| \gt 2.5\,\mathrm{kpc}$. We find that the fit is well behaved in this radial span, as gauged primarily by its ability to reproduce the expected properties of the baryonic disk. Below $| z| \sim 2.5\,\mathrm{kpc}$, we find a persistent systematic bias in the fitting procedure that results from modeling the azimuthal disk velocities as a single Gaussian (Schönrich & Binney 2012), so we do not present those results here.

Table 1.  Parameters and Associated Prior Types/Ranges for the Disk, Halo, and Substructure Populations

Parameter Type Priors
    Disk Halo Substructure
μr Linear [−70, 70] [−70, 70] [0, 250]
μθ Linear [−70, 70] [−70, 70] [−70, 70]
μϕ Linear [0, 300] [−70, 70] [−70, 70]
σr,θ,ϕ Linear [0, 200] [0, 200] [0, 200]
ρ,,θϕ Linear [−1, 1] [−1, 1] [−1, 1]
${\mu }_{[\mathrm{Fe}/{\rm{H}}]}$ Linear [−1.5, 0.5] [−3, −1] [−3, −1]
${\sigma }_{[\mathrm{Fe}/{\rm{H}}]}$ Linear [0, 2] [0, 2] [0, 2]
Q Linear [0, 1] [0, 1]

Download table as:  ASCIITypeset image

3. The Stellar Distribution

The 95% contours in metallicity–velocity space that are recovered from the analysis are overlaid on the separate panels of Figure 2. These results apply specifically to the SDSS–Gaia DR2 sample in the region where r ∈ [7.5, 8.5] kpc and $| z| \gt 2.5\,\mathrm{kpc}$. Clearly, the best-fit distributions do an excellent job of picking out the population clusters that we identified by eye at the offset. Note that the priors are, for the most part, uninformative. They make no assumption about the relative metallicities of the halo and substructure populations, or the means and dispersions of their velocity distributions. The choice of priors gives the analysis enough freedom to explore both unimodal and bimodal distributions for the radial velocity of the substructure.

Figure 3 shows the one-dimensional vr, vθ, vϕ, and [Fe/H] best-fit distributions for the same region. The disk, halo, and substructure distributions are indicated by the solid green, dashed red, and dotted blue lines, respectively, while the data are represented by the gray histograms. In the Appendix, we also provide the residuals between the model and data as Figure 11, as well as the corresponding corner plots as Figures 1214. One of the most surprising results is the degree to which the halo distribution is subdominant relative to the substructure. To better understand this, let us review in some detail the behavior of each population separately and characterize its evolution as we vary the lower bound on $| z| $ in the range from 2.5 to 4.0 kpc.

Figure 3.

Figure 3. Best-fit distributions of the spherical galactocentric velocity components and metallicity for the SDSS–Gaia DR2 sample in the region within r ∈ [7.5, 8.5] kpc and $| z| \gt 2.5\,\mathrm{kpc}$ (clockwise from top left: vr, vθ, [Fe/H], and vϕ). These distributions are found from sampling the posteriors of the model parameters. In each panel, the disk, halo, and substructure distributions are shown as solid green, dashed red, and dotted blue lines, respectively. The data are represented by gray histograms. The corresponding corner plots are shown in the Appendix as Figures 1214.

Standard image High-resolution image

From Figure 3, it is clear that the best-fit disk population has a median metallicity of [Fe/H] = −0.779 and ${\mu }_{\phi }={136.3}_{-1.8}^{+1.9}$ km s−1, consistent with the data. Figure 4 demonstrates how the disk's velocity and metallicity distribution vary away from the midplane. The [Fe/H] and vθ distributions remain essentially constant as one moves from $| z| \gt 2.5$ to >4 kpc. The vr distribution broadens slightly and the median vϕ shifts to lower values, as expected from asymmetric drift (Bond et al. 2010).

Figure 4.

Figure 4. Best-fit distributions for the disk population in the region with r ∈ [7.5, 8.5] kpc (clockwise from top left: vr, vθ, [Fe/H], and vϕ). We vary the distance from the midplane from $| z| \gt 2.5$ to 4.0 kpc.

Standard image High-resolution image

Figure 5 shows the corresponding distributions for the halo, which remain constant over the full z-range explored here. This population is clearly very metal-poor with a median [Fe/H] = −1.82. Its velocity distribution is nearly isotropic as ${\sigma }_{r}={136.1}_{-3.6}^{+3.6}$ km s−1, ${\sigma }_{\theta }={112.5}_{-3.8}^{+4.1}$ km s−1, and ${\sigma }_{\phi }\,={139.1}_{-5.2}^{+5.5}$ km s−1. The radial and azimuthal means are nonzero, with ${\mu }_{r}={10.0}_{+4.9}^{-4.6}$ km s−1 and ${\mu }_{\phi }={24.9}_{+4.6}^{-5.6}$ km s−1. All three correlation coefficients are small: $({\rho }_{r\theta },{\rho }_{r\phi },{\rho }_{\theta \phi })\,=(-{0.03}_{-0.03}^{+0.03},-{0.08}_{-0.03}^{+0.03},{0.06}_{-0.02}^{+0.01})$.

Figure 5.

Figure 5. Same as Figure 4, except for the halo population. The chemodynamic properties of the halo are invariant as one moves away from the Galactic midplane.

Standard image High-resolution image

The halo is subdominant to the substructure, which is distinctive in both chemical abundance and kinematics. As shown in Figure 6, the median metallicity of the substructure remains constant at [Fe/H] = −1.39 over all z-values. This population is more metal-rich, on average, than the halo, but more metal-poor than the disk. The radial velocity lobes are centered at ${\mu }_{r}=\pm {147.6}_{-6.4}^{+7.2}$ km s−1 with ${\sigma }_{r}={113.6}_{-3.0}^{+3.1}$ km s−1. There is no evidence for rotation in the polar direction (${\mu }_{\theta }=-{2.8}_{-1.6}^{+1.5}$ km s−1 and ${\sigma }_{\theta }={65.2}_{-1.2}^{+1.1}$ km s−1); however, there is a larger offset in the azimuthal direction, with ${\mu }_{\phi }={27.9}_{-2.9}^{+2.8}$ km s−1 and ${\sigma }_{\phi }={61.9}_{-2.9}^{+2.6}$ km s−1. The correlations ρ and ρθϕ are consistent with zero, while ${\rho }_{r\theta }={0.18}_{-0.03}^{+0.03}$.

Figure 6.

Figure 6. Same as Figure 4, except for the substructure population. The radial distribution has characteristic lobes at ±148 km s−1 that are likely related to tidal debris that is stripped as a merging satellite moves toward/away from the Galactic center on its orbit. The distributions remain constant over the entire z-range.

Standard image High-resolution image

Figure 7 shows the fractional contribution of the disk, halo, and substructure stars in the data set. We see that the disk contribution reduces from 40% at $| z| \gt 2.5\,\mathrm{kpc}$ down to 25% at $| z| \gt 4\,\mathrm{kpc}$. The halo contribution increases mildly in this range, as might be expected. However, the relative fraction of the substructure to the nondisk stellar population (right panel) is constant at ∼60%. We remind the reader that these fractional contributions pertain only to the data set and that metallicity biases can potentially affect the extrapolation to the Galaxy. However, the fact that the results are unchanged (within uncertainties) when the analysis is repeated on the subset of F/G stars, which exhibit minimal bias, gives us confidence in the results presented here. Figure 15 shows a similar behavior as a function of Galactocentric distance, where the disk contribution decreases as expected as a function of r.

Figure 7.

Figure 7. Left: fractional contribution of the disk (green), halo (red), and substructure (blue) populations in the data set, for the region r ∈ [7.5, 8.5] kpc and $| z| \gt {z}_{\mathrm{cut}}$. Percentiles of 2.1, 50, and 97.9 are shown here. Right: fraction of the substructure relative to all nondisk stars in the data set.

Standard image High-resolution image

We made the corresponding versions of Figures 47 for $| z| \gt 2.5\,\mathrm{kpc}$ and varying r in five equally sized bins from 7.5 to 10.0 kpc as in Figures 1618. These figures are provided in the Appendix. We observe no significant change in the velocity and metallicity distribution of the halo in this range. The substructure distributions are also stable, except that the radial lobes are further apart and more pronounced closer to the center of the Galaxy, as shown in Figure 18. The mean of the radial distribution is at ${\mu }_{r}={140.5}_{-7.7}^{+8.4}$ km s−1 with ${\sigma }_{r}={114.5}_{-3.9}^{+4.1}$ km s−1 for r ∈ [7.5, 8.0] kpc, while it drops to ${\mu }_{r}={115.0}_{-3.1}^{+3.1}$ km s−1 with ${\sigma }_{r}={104.2}_{-2.6}^{+2.7}$ km s−1 for r ∈ [9.5, 10] kpc. This may be related to features of the orbit that change with r.

The trends we observe are consistent with the interpretation that the substructure originates from the merger of a fairly massive satellite on a highly radial orbit (Belokurov et al. 2018a; Deason et al. 2018; Myeong et al. 2018a, 2018b). We distinguish two individual lobes in the radial velocity distribution, consistent with tidal debris that is preferentially stripped as the satellite moves toward/away from the Galactic center. The small, but nonzero, azimuthal rotation may also be linked to the properties of the orbit. We note that our analysis cannot distinguish between one or more mergers. The latter situation seems unlikely, as the metallicity of the substructure remains constant over the entire spatial range probed, which suggests a single progenitor. If multiple mergers were at cause, the satellites would have to have similar masses and orbital properties, which seems fine-tuned.

It is challenging to directly compare our best-fit values to previous studies of the stellar halo, because we break down the sample into a Gaussian and non-Gaussian component at the likelihood level. As a result, the halo velocity distribution published in other works would be the weighted sum of our halo and substructure populations. However, the general trends we observe are roughly consistent with previous results. For example, previous studies found that halo stars with intermediate metallicities ([Fe/H] ∼ −1.4) are radially anisotropic (Smith et al. 2009; Bond et al. 2010; Belokurov et al. 2018b), while stars on the more metal-poor end of the spectrum become more isotropic (Carollo et al. 2007, 2010; Herzog-Arbeitman et al. 2018b; Belokurov et al. 2018b).

4. The DM Distribution

We now discuss how our results concerning the accreted stellar population translate to the DM distribution. In particular, we will argue that the presence of an anisotropic population of stars suggests that a component of the DM halo may have also originated from a more recent merger and may consequently not have reached steady state. Section 4.1 outlines a simple model framework for this two-component DM scenario, and Section 4.2 describes the consequences for direct detection experiments.

4.1. Two-component DM Model

Let us assume that the DM in the region of study can be divided into two populations. For the moment, we remain agnostic to the origin of these two populations, but we will return to this shortly. In this scenario, the total DM phase-space distribution satisfies

Equation (4)

where fi is the normalized velocity distribution and ξi is the relative fraction of the ith population. Note that ξ1 + ξ2 = 1. Starting from the distribution function in Equation (4), it is possible to derive the Jeans equations in spherical coordinates. To simplify the derivation and highlight the most important conceptual points, we make several assumptions about these two populations.

First, we assume that population 1 is isotropic and spherically symmetric and that its velocity components are uncorrelated. Most importantly, we take it to be in steady state, which allows us to ignore partial derivatives of f1(v) with respect to time. When we state that a population is in equilibrium, we mean specifically that it is in steady state with ∂fi/∂t = 0.

As a point of contrast, we will assume that population 2 has not reached steady state—or, at the very least, that we cannot confirm whether it has. Additionally, we will assume that its spatial density is spherically symmetric, that its velocity components are uncorrelated and have vanishing mean at present-day, and that the mean velocities and dispersions are spatially invariant in the region of interest.

The radial Jeans equation for this two-component model is

Equation (5)

where νi is the number density, μr,i is the radial velocity mean, σr,i, σθ,i, σϕ,i are the velocity dispersions for the ith population, and Φ is the gravitational potential.

Using Equation (5), one can recover the classical argument laid out in Drukier et al. (1986) that motivates a Maxwell–Boltzmann velocity distribution for DM. The observation of a flat rotation curve near the solar position suggests a logarithmic potential for the Milky Way halo of the form ${\rm{\Phi }}(r)\,={v}_{c}^{2}\mathrm{ln}(r)+\ \mathrm{constant}$, where vc is the circular velocity. If we assume that all of the DM is in steady state (ξ1 = 1, ξ2 = 0) and that its number density is described as a falling power law, ν1 ∝ rb, then Equation (5) predicts a tight link between the value of the power-law index b, the potential Φ, and the phase-space distribution. In this limit, the radial Jeans equation becomes

Equation (6)

Taking the local circular velocity to be vc ∼ 235 km s−1 and a dispersion of σr,1 ∼ 160 km s−1 yields a power-law slope of b ∼ 2. A density distribution of the form ${\nu }_{1}\sim {r}^{-2}$ is isothermal, and one can use Poisson's equation to show that the associated velocity distribution is Maxwell–Boltzmann.

However, notice from Equation (5) that the simple derivation of the isothermal model breaks down in the presence of a second DM population. In particular, the tight predictive link between the potential Φ and the number density ν1 no longer holds when ${\xi }_{2}\ne 0$. This is true regardless of whether the second DM population is in steady state or not. However, the challenge is exacerbated for the latter case, as we then need to quantify ∂μr,2/∂t.

This two-component model for the DM distribution10 is motivated in light of recent observations of the nearby accreted stellar distributions. As already discussed, the observation of a metal-poor and isotropic stellar halo is likely associated with tidal debris from the oldest luminous mergers that built up the Milky Way, while the anisotropic component at intermediate metallicities is due to tidal debris from a more recent merger. As DM would have also been stripped from these accreted satellites, it is reasonable to assume that it should also be separated into at least two populations as well. It is of course possible that the DM can be divided into more than two populations. This would be relevant if a significant fraction of the DM originated from nonluminous satellites or diffuse accretion, neither of which should be correlated with stellar distributions.

For now, we solely focus on the subset of DM from luminous mergers, as that is where the results of this work have the most consequence. In this case, the recent observations from Gaia suggest that we should be moving to a two-component model for the DM, as per Equation (4). Population 1 would correspond to the DM from the oldest mergers (the "halo" component), and population 2 would correspond to the DM from the recent merger (the "substructure" component).

Numerical simulations have demonstrated that DM accreted from the oldest luminous mergers is well traced by metal-poor stars—this corresponds to the halo population (Herzog-Arbeitman et al. 2018a; Necib et al. 2018). Additionally, DM in debris flow is well traced by intermediate-metallicity stars (Lisanti et al. 2015; Necib et al. 2018). This suggests that one can use the stellar velocity distribution for the halo and substructure populations, recovered in Section 3, to model the individual DM populations.

As the halo population is the most metal-poor and is also isotropic, we expect that it consists of tidal debris from the earliest mergers that have reached stead state today. The metallicity range and velocity profiles that we recover share similar features to the distribution of stars accreted at redshifts z > 3 in a study of two separate Milky Way–like halos in the suite of Fire hydrodynamic simulations (Necib et al. 2018). We believe that this population of tidal material is accreted while the proto-galaxy is still forming, and is fully relaxed today.

The substructure population arises from a more recent merger, with estimated accretion redshift in the range of z ∼ 1–3 (Belokurov et al. 2018a; Myeong et al. 2018b). Necib et al. (2018) showed that the tidal debris from satellites accreted at z ≲ 3 leaves distinctive features in velocity space that stand out from the behavior of older tidal debris, which is fully relaxed. Further study is required to determine whether this new substructure has reached steady state.

4.2. Experimental Implications

In this section, we derive the heliocentric speed distribution for the halo and substructure populations, which we assume to trace the DM, and use it to calculate the scattering rate of a DM particle off a nuclear target. Ideally, we need the stellar velocity distribution at the solar position, but this is also the region where the disk contribution dominates. Any mismodeling of the disk may therefore strongly bias the fit results in this regime. For this reason, we restrict ourselves to $| z| \gt 2.5\,\mathrm{kpc}$. The results from Section 3 give us confidence that the halo and substructure distributions are likely invariant in z and that we can extrapolate them into the plane; however, this should be verified explicitly. Previous work (Belokurov et al. 2018b) also finds evidence for the radial substructure down to $| z| =1\,\mathrm{kpc}$.

The best-fit speed distribution in the heliocentric frame is shown in the left panel of Figure 8, for heliocentric distances of d < 4 kpc and $| z| \gt 2.5\,\mathrm{kpc}$. To obtain this distribution, we draw values of vr, vθ, and vϕ from the full posterior distribution in the Galactic frame. Assuming that the stars are spatially uniform, we then transform to the heliocentric frame using the local rest-frame velocities v⊙,pec = (U, V, W) = (8.50, 13.38, 6.49) km s−1 (Coşkunoǧlu et al. 2011) and local circular velocity vc = 235 km s−1. The substructure and halo components are plotted separately (blue dotted and red dashed lines, respectively). An important subtlety arises when summing these two contributions, as we need to know the relative amount of DM that is contributed by the accreted satellites in each population—i.e., the fractions ξ1,2 in Equation (4). For now, we make the simplifying assumption that the satellites have comparable mass-to-light ratios. In this case, the total contribution is shown as the solid black line. Necib et al. (2018) provide a more detailed prescription to estimate the relative DM fractions associated with each population.

Figure 8.

Figure 8. Left: best-fit speed distribution for the halo (red dashed line) and substructure (blue dotted line) components. The solid black line represents the total contribution. These results are based on fits to the SDSS–Gaia DR2 data within heliocentric distances of d < 4 kpc and $| z| \gt 2.5\,\mathrm{kpc}$. For comparison, we show the SHM (gray dashed line), defined in Equation (7). The empirical distribution does not include contributions from DM accreted from nonluminous satellites or diffusely. Right: 95% background-free C.L. limits on the DM–nucleon scattering cross section, σχn, for spin-independent interactions as a function of DM mass, mχ, assuming a xenon target with an exposure of 1 kton × year exposure and a 4.9 keVnr energy threshold. These limits are illustrative and do not account for experimental energy efficiencies near threshold (Aprile et al. 2018). Note that the "Total" distribution assumes that the relative ratio of DM to stars contributed by accreted satellites in the halo and substructure populations is comparable. This assumption is further addressed in Necib et al. (2018).

Standard image High-resolution image

For comparison, we also plot the SHM as the gray dashed line. The SHM is the speed distribution associated with the Maxwell–Boltzmann distribution (after integrating over the angular coordinates) and is defined as

Equation (7)

The dispersion of the SHM is closest to that of the halo posterior, albeit slightly higher; the SHM is isotropic with σ ∼ 156 km s−1, while the best fits for the halo are $({\sigma }_{r},\,{\sigma }_{\theta },\,{\sigma }_{\phi })\,=\left({140.3}_{-4.9}^{+4.2},{114.2}_{-1.8}^{+3.3},{125.9}_{-3.4}^{+4.1}\right)$ km s−1. The primary discrepancy with the SHM arises from the substructure population. When this component is included, the total speed distribution is discrepant with the SHM. In this case, the polar and azimuthal velocities of the substructure are Gaussian with means $({\mu }_{\theta },{\mu }_{\phi })=(-{3.1}_{-0.9}^{+0.9},{35.5}_{-1.8}^{+1.8})$ km s−1 and dispersions $({\sigma }_{\theta },{\sigma }_{\phi })=({57.7}_{-0.8}^{+0.7},{61.2}_{-1.5}^{+1.5})$ km s−1, but the radial distribution has peaks at $\pm {117.7}_{-2.1}^{+1.8}$ km s−1, with a dispersion ${\sigma }_{r}={108.2}_{-1.3}^{+1.2}$ km s−1 for each. The substructure component arises from a more recent merger, as underlined by the fact that it is more metal-rich and highly radial. Because it is not necessarily in steady state, there is no reason why the SHM should model it well.

The empirical distribution clearly underestimates the fraction of high-speed DM particles, as compared to the SHM, which affects DM models where the minimum scattering speed, vmin, needed to create a nuclear recoil of energy Enr is high. For elastic scattering, the minimum speed depends on both the DM particle properties and the experiment as follows:

Equation (8)

where mN is the nuclear mass and μ is the DM–nucleus reduced mass. If the scattering is inelastic, then the minimum speed is even larger. The differential scattering rate, per unit detector mass, for the most common operators is then

Equation (9)

where mχ is the DM particle mass and ρχ its local density, σ(q) is an effective scattering cross section that depends on the momentum transfer q, and the mean inverse speed is defined as

Equation (10)

where $\tilde{f}(v)$ is the heliocentric velocity distribution.

The right panel of Figure 8 shows the corresponding limits on the DM mass and DM–nucleon scattering cross section, σχn, assuming the simplest spin-independent operator. For this example, we assume a xenon target, energy threshold of 4.9 keVnr, and exposure of 1 kton × year. The 95% one-sided Poisson C.L. limit (three events) obtained using the velocity distribution inferred from SDSS–Gaia DR2 is shown in solid black and compared to the SHM in dashed gray. The substructure component drives the sensitivity at all masses, while the halo contribution is subdominant but becomes more important at lower masses. In both cases, the exclusion is significantly weakened for mχ ≲ 30 GeV relative to that obtained using the SHM. For mχ ≳ 100 GeV, the black and gray dashed lines approach each other because ${v}_{\min }\to 0$ in Equation (10).

The overall effect of the empirical velocity distribution on the scattering limit depends on the details of the nuclear target, experimental threshold, and DM mass—all parameters that feed into the minimum scattering speed defined in Equation (8). A more model- and experiment-independent way of understanding these effects is to study the dependence of the time-averaged inverse speed, $\langle g({v}_{\min })\rangle $, as a function of the minimum speed, as this term captures the dependence of the scattering rate on the DM velocities. The left panel of Figure 9 plots this quantity for the empirical speed distribution obtained in this work (solid black) and the SHM (dashed gray). The scattering rate for the empirical distribution is reduced relative to that for the SHM at vmin ≳ 300 km s−1; it is enhanced for lower minimum speeds. The scattering rate is completely suppressed for vmin ≳ 550 km s−1, whereas the SHM continues to contribute events above this point.

Figure 9.

Figure 9. Left: time-averaged inverse-speed distribution defined in Equation (10) as a function of the minimum scattering speed, vmin. The substructure and halo distributions are shown as the dotted blue and dashed red lines, respectively, and their total contribution is shown as a solid black line. The SHM expectation is the dashed gray line. The vertical green lines indicate the values of vmin near threshold for a 10 GeV DM particle scattering in the DarkSide-50 (Agnes et al. 2018), DAMA (Bernabei et al. 2018), and Xenon1T (Aprile et al. 2018) detectors. Right: expected yearly modulation amplitude between June (t0 ≈ 150 days) and December as a function of vmin. Note that the "Total" distribution assumes that the relative ratio of DM to stars contributed by accreted satellites in the halo and substructure populations is comparable. This assumption is further addressed in Necib et al. (2018).

Standard image High-resolution image

To better understand the implications of these results, let us consider the concrete example of a 10 GeV DM particle interacting in several detectors. Such a DM particle needs a minimum speed of ∼570 km s−1 to scatter a xenon nucleus at an energy of ∼5 keVnr in Xenon1T (Aprile et al. 2018) when requiring both S1 and S2 signals.11 As seen from the left panel of Figure 9, this is highly suppressed relative to the SHM expectation.12 In contrast, Xenon100 can detect recoils down to 0.7 keVnr in the S2-only analysis (Aprile et al. 2016), similarly to the DarkSide-50 low-mass analysis (Agnes et al. 2018), which can detect argon recoils down to 0.6 keVnr in energy. A 10 GeV DM particle only needs speeds of ∼130–200 km s−1 to create such a recoil, and these speeds are well supported by the empirical distribution.

The empirical velocity distribution also impacts the time dependence of a signal. The DM scattering rate should modulate annually owing to Earth's motion around the Sun (Drukier et al. 1986). The right panel of Figure 9 compares the modulation amplitude assuming the newly derived velocity distribution, as compared to the SHM. To obtain the amplitude, we transform the velocities from the Galactic to the heliocentric frame, taking into account Earth's time-dependent velocity as defined in Lee et al. (2013). We do not include the effect of gravitational focusing, which may further affect the properties of the modulation signal (Lee et al. 2014).

The modulation amplitude for the SHM exhibits the expected features: a maximum when vmin ∼ 350 km s−1 and a change in phase below ∼200 km s−1 (see Freese et al. 2013, for a review). In comparison, the modulation amplitude obtained from the SDSS–Gaia DR2 distribution is maximal closer to vmin ∼ 250 km s−1 and falls off faster toward higher vmin. This is due to the fact that the empirical velocity distribution f(v) is less broad than the SHM. Therefore, the differences in the heliocentric speed distribution over the year are typically more pronounced, but over a smaller range of vmin. The DAMA experiment, which claims an annually modulating signal, has an Na I target and a threshold energy of ∼3.3 keVnr (Bernabei et al. 2018). An observable scatter of a 10 GeV DM particle off an Na nucleus requires a minimum speed of ∼270 km s−1. We note that this falls in the region where the empirical distribution has a significant effect on the modulation amplitude and motivates a more careful study of the consistency with data.

The results shown here are specific to spin-independent interactions. However, the new velocity distribution will have an effect on other interaction operators—as the dependence of some of these operators on the DM momentum is nontrivial, the magnitude of the effects can vary from operator to operator (Fan et al. 2010; Fitzpatrick et al. 2012, 2013; Lisanti 2017). In general, any model that relies on the high-velocity tail of the velocity distribution to support the scattering rate will be affected by these results. One simple and illustrative case is that of inelastic (rather than elastic) spin-independent scattering off of nuclei. Let us return to the example of the 10 GeV DM particle scattering off a xenon target at a threshold of ∼0.7 keVnr. The minimum scattering velocity for elastic scattering is ∼210 km s−1, which is well supported by the velocity distribution as demonstrated in Figure 9. However, if one considers a case where there are nearly degenerate states separated by ∼15 keV in mass, then the inelastic scattering requires vmin ∼ 560 km s−1, which is suppressed relative to the SHM expectation.

Similarly, the velocity distribution will also be relevant for the interpretation of DM–electron scattering interactions (see Battaglieri et al. 2017, for a review) and axion experiments (Ling et al. 2004; Hoskins et al. 2016; Sloan et al. 2016; Millar et al. 2017; Vergados & Semertzidis 2017; Foster et al. 2018).

5. Conclusions

We performed a mixture model analysis on MS stars in the SDSS–Gaia DR2 catalog within the range r ∈ [7.5, 10.0] kpc and $| z| \gt 2.5\,\mathrm{kpc}$. The full chemodynamic properties of the stars (v, [Fe/H]) were used to identify the populations most likely belonging to the disk and halo, as well as any potential kinematic substructure. The velocities of the disk and halo stars were modeled as multivariate normal distributions, while the substructure component was given the freedom to scan non-Gaussian possibilities.

The recovered disk, halo, and substructure populations have median metallicities of [Fe/H] = −0.8, −1.8, and −1.4 respectively. The disk component acts as expected within the region studied here. The halo component is very metal-poor, and its velocity ellipsoid is nearly isotropic. The analysis identifies a substructure population of intermediate-metallicity stars whose radial velocities are best modeled with a non-Gaussian distribution. This population had been identified in previous work as the Gaia Sausage (Belokurov et al. 2018b). Our analysis provides the first model for its velocity and metallicity distribution and clearly distinguishes its contribution relative to the disk and metal-poor halo over the full metallicity range of the sample.

The substructure population is anisotropic in velocity, with two broad lobes centered at vr ∼ ±150 km s−1. It can be explained as tidal debris from a satellite galaxy on a highly radial orbit. The lobes are consistent with debris that is torn off as the satellite moves toward/away from the Galactic center while orbiting. The distinctive metallicity of the substructure strongly suggests that it is sourced by a single progenitor.

To illustrate these points, Figure 10 compares the orbits of a likely halo, disk, and substructure star, chosen at random—see also Deason et al. (2018). We use the gala package (Price-Whelan et al. 2017) to integrate the orbits back 1 Myr, given the star's present-day position and velocity, and assume the default Milky Way potential from Bovy (2015). In the left panel, we plot the orbit of a likely disk (green line) and halo (red line) star. The disk star is confined to the plane, as expected, while the halo star's orbit is more isotropic. We contrast this to the orbits of two likely substructure stars in the right panel, indicated by the dark-blue and light-blue lines. The orbits of these stars are highly radial and have a lower inclination angle relative to the midplane.

Figure 10.

Figure 10. Left: orbits of stars that likely belong to the disk (green) and halo (red) populations. Right: corresponding orbits for two likely substructure stars, labeled as "Subs1" and "Subs2." The main figure shows the projection in the xz plane, while the inset shows that of the xy plane. The black circle shows the location of the Sun.

Standard image High-resolution image

As the substructure is fairly hot and exhibits no obvious spatial features in the local region studied here, it likely originated from an old merger. Using the stellar mass–metallicity relation of Kirby et al. (2013) and the median metallicity of the substructure component, we estimate that its progenitor had stellar mass M* ∼ 107–108 M. We note that the Sagittarius dwarf galaxy (Ibata et al. 1994) cannot be the progenitor of the substructure, as it is a younger merger event and its orbit is less eccentric (Law & Majewski 2010; Purcell et al. 2011). In particular, the apocenter-to-pericenter ratio of Sagittarius is 5:1, whereas it is 20:1 for the substructure.

Our study of the local stellar distribution has direct relevance for DM. Numerical simulations have demonstrated that the old metal-poor halo is a good tracer of the virialized DM kinematics. Additionally, kinematic substructure such as DM debris flow has been shown to have stellar counterparts. If the accreted stellar components in the SDSS–Gaia DR2 sample (e.g., the halo and substructure populations) are adequate DM tracers, then our results imply that a fraction of the local DM is in debris flow.

We find that the halo and substructure populations do not depend on the vertical distance off the plane, at least in the region of $| z| \gt 2.5$–4 kpc. This gives us confidence in extrapolating their contributions to the solar neighborhood, which is relevant for direct detection experiments. By performing a more detailed modeling of the disk component, one could potentially extend the mixture analysis to lower z and recover the accreted stellar distribution directly in this region. We plan to pursue this in future work.

The heliocentric speed distribution that we derive from SDSS–Gaia DR2 within heliocentric distances d < 4 kpc and $| z| \gt 2.5\,\mathrm{kpc}$ is incompatible with the SHM. The inferred DM distribution has far fewer high-speed particles than expected from the SHM. If we assume that the distributions recovered in this study extrapolate down to the Galactic midplane, then this would reduce the sensitivity to DM that is not energetic enough to create observable nuclear recoils in a detector target. In this case, the limits of an experiment using a xenon target are suppressed below mDM ≲ 30 GeV for the case of spin-independent interactions. The overall size of the suppression can vary for different nuclear target masses, as well as different scattering operators. Current exclusion limits and future projections should be revisited in light of these new findings.

The SDSS–Gaia DR2 study provides the first indication that a subset of the local DM may be in kinematic substructure. Given these results, it is pressing to better quantify just how well the stars and DM track each other in simulated mergers that resemble the observations. Additionally, we need to quantify the effects of DM that is diffuse or originates from nonluminous subhalos. If either of these dominates locally, then the total DM distribution will differ from that accreted by the largest satellites. Furthermore, improved modeling is needed to better estimate the relative DM fraction contributed by the accreted satellites in the halo and substructure populations. In particular, Necib et al. (2018) present an empirical approach to estimate the relative DM contribution from each satellite, by inferring the average metallicity of the accreted stars of a merger and the ratio of DM mass to stellar mass of the merging satellite. Using such a relationship, they estimated that the ${42}_{-22}^{+26} \% $ of DM accreted from luminous satellites is in debris flow.

We would like to thank J. Bochanski, J. Bovy, G. Collin, A. Drlica-Wagner, D. Hogg, K. Hawkins, A. Kaboth, L. Lancaster, T. Li, S. McDermott, G. C. Myeong, D. Spergel, and N. Weiner for useful conversations. We also acknowledge use of the Astropy (Astropy Collaboration et al. 2013), emcee (Foreman-Mackey et al. 2013), gala (Price-Whelan et al. 2017), and IPython (Pérez & Granger 2007) software packages. This project was developed in part at the NYC Gaia DR2 Workshop in 2018 April, as well as the 2018 NYC Gaia Sprint, hosted by the Center for Computational Astrophysics of the Flatiron Institute in New York City.

L.N. is supported by the DOE under award No. DESC0011632 and the Sherman Fairchild Fellowship. M.L. is supported by the DOE under award No. DESC0007968, the Alfred P. Sloan Foundation, and the Cottrell Scholar Program through the Research Corporation for Science Advancement. V.B. is supported by the European Research Council under the European Union's Seventh Framework Programme (FP/2007-2013)/ERC grant agreement No. 308024. This paper made used of the Whole Sky Database (WSDB) created by Sergey Koposov and maintained at the Institute of Astronomy, Cambridge by Sergey Koposov, Vasily Belokurov, and Wyn Evans with financial support from the Science & Technology Facilities Council (STFC) and the European Research Council (ERC).

This work has made use of data from the European Space Agency (ESA) mission Gaia (https://www.cosmos.esa.int/gaia), processed by the Gaia Data Processing and Analysis Consortium (DPAC, https://www.cosmos.esa.int/web/gaia/dpac/consortium). Funding for the DPAC has been provided by national institutions, in particular the institutions participating in the Gaia Multilateral Agreement.

Funding for SDSS-III has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, and the U.S. Department of Energy Office of Science. The SDSS-III website is http://www.sdss3.org/.

SDSS-III is managed by the Astrophysical Research Consortium for the Participating Institutions of the SDSS-III Collaboration, including the University of Arizona, the Brazilian Participation Group, Brookhaven National Laboratory, Carnegie Mellon University, University of Florida, the French Participation Group, the German Participation Group, Harvard University, the Instituto de Astrofisica de Canarias, the Michigan State/Notre Dame/JINA Participation Group, Johns Hopkins University, Lawrence Berkeley National Laboratory, Max Planck Institute for Astrophysics, Max Planck Institute for Extraterrestrial Physics, New Mexico State University, New York University, Ohio State University, Pennsylvania State University, University of Portsmouth, Princeton University, the Spanish Participation Group, University of Tokyo, University of Utah, Vanderbilt University, University of Virginia, University of Washington, and Yale University.

Appendix:

In this Appendix, we include some additional figures to supplement the discussion in the main text. We provide residual maps that demonstrate the quality of the model fit, show how the results in the main text vary in five radial bins from r = 7.5 to 10 kpc, and provide corner plots for the halo, disk, and substructure parameters.

Figure 11.

Figure 11. Comparison of the data and best-fit model distributions in the vrvϕ plane for the SDSS–Gaia DR2 data in the region r ∈ [7.5, 8.5] kpc and $| z| \gt 2.5\,\mathrm{kpc}$. The left column shows a count map of the data (top) and model (bottom). The right column compares the two explicitly, showing the fractional (top) and total (bottom) residuals. The bottom right panel is the difference in the value of the binned histograms, where both the data and model are normalized to unity beforehand.

Standard image High-resolution image
Figure 12.

Figure 12. Corner plot for the halo model parameters in the region r ∈ [7.5, 8.5] kpc and $| z| \gt 2.5\,\mathrm{kpc}$.

Standard image High-resolution image
Figure 13.

Figure 13. Corner plot for the disk model parameters in the region r ∈ [7.5, 8.5] kpc and $| z| \gt 2.5\,\mathrm{kpc}$.

Standard image High-resolution image
Figure 14.

Figure 14. Corner plot for the substructure model parameters in the region r ∈ [7.5, 8.5] kpc and $| z| \gt 2.5\,\mathrm{kpc}$.

Standard image High-resolution image
Figure 15.

Figure 15. Same as Figure 7, except for different radial cuts.

Standard image High-resolution image
Figure 16.

Figure 16. Same as Figure 4, except for different radial cuts and fixing $| z| \gt 2.5\,\mathrm{kpc}$.

Standard image High-resolution image
Figure 17.

Figure 17. Same as Figure 5, except for different radial cuts.

Standard image High-resolution image
Figure 18.

Figure 18. Same as Figure 6, except for different radial cuts.

Standard image High-resolution image

Footnotes

  • In particular, we use the iron abundance, [Fe/H], which is defined as

    where Ni is the number density of the element i.

  • We also note a small cluster of stars at [Fe/H] ∼ −1.5 and vϕ < −200 km s−1. These stars may or may not be part of a distinct population; because they constitute only 0.5% of the total sample, we will not focus on them in this work. They are primarily flagged as "halo" stars in the likelihood study.

  • Note that we define ϕ and θ as the azimuthal and polar directions, respectively. This is the opposite of the convention used in Belokurov et al. (2018b).

  • For a visual example of the bimodal distribution, we point the reader to the top left panel of Figure 6, which we will discuss in more detail in the following section.

  • The errors on the best-fit metallicity means and dispersions quoted here are all on the order of ±0.01 dex.

  • 10 

    Specifically, the DM halo in the region of study within galactocentric radii of 7.5–10 kpc and $| z| \gt 2.5\,\mathrm{kpc}$.

  • 11 

    The S1 signal is associated with scintillation light from the initial particle collision, while the S2 signal is the scintillation light from the ionization electrons.

  • 12 

    In actuality, Xenon1T has nonzero efficiency below ∼5 keVnr, which improves its sensitivity in this range.

Please wait… references are loading.
10.3847/1538-4357/ab095b