23
Draft version November 13, 2018 Preprint typeset using L A T E X style emulateapj v. 12/16/11 ACTION-BASED DYNAMICAL MODELLING FOR THE MILKY WAY DISK Wilma H. Trick 1,2 , Jo Bovy 3 , and Hans-Walter Rix 1 Draft version November 13, 2018 ABSTRACT We present RoadMapping, a full-likelihood dynamical modelling machinery that aims to recover the Milky Way’s (MW) gravitational potential from large samples of stars in the Galactic disk. RoadMap- ping models the observed positions and velocities of stars with a parametrized, three-integral distri- bution function (DF) in a parametrized axisymmetric potential. We investigate through differential test cases with idealized mock data how the breakdown of model assumptions and data properties affect constraints on the potential and DF. Our key results are: (i) If the MW’s true potential is not included in the assumed model potential family, we can—in the axisymmetric case—still find a robust estimate for the potential, with only . 10% difference in surface density within |z|≤ 1.1 kpc inside the observed volume. (ii) Modest systematic differences between the true and model DF are inconse- quential. E.g, when binning stars to define sub-populations with simple DFs, binning errors do not affect the modelling as long as the DF parameters of neighbouring bins differ by < 20%. In addition, RoadMapping ensures unbiased potential estimates for either (iii) small misjudgements of the spatial selection function (i.e., . 15% at the survey volume’s edge), (iv) if distances are known to within 10%, or (v) if proper motion uncertainties are known within 10% or are smaller than δμ . 1 mas yr -1 . Challenges are the rapidly increasing computational costs for large sample sizes. Overall, RoadMap- ping is well suited to making precise new measurements of the MW’s potential with data from the upcoming Gaia releases. Keywords: Galaxy: disk — Galaxy: fundamental parameters — Galaxy: kinematics and dynamics — Galaxy: structure 1. INTRODUCTION Through dynamical modelling we can infer the Milky Way’s (MW) gravitational potential from stellar motions (Binney & Tremaine 2008; Binney 2011; Rix & Bovy 2013). Observational information on the 6D phase-space coordinates of stars is currently growing at a rapid pace, and will be taken to a whole new level in quantity and precision by the upcoming data from the Gaia mission (Perryman et al. 2001). Yet, rigorous and practical mod- elling tools that turn position-velocity data of individual stars into constraints both on the gravitational potential and on the distribution function (DF) of stellar orbits are scarce (Rix & Bovy 2013). The Galactic gravitational potential is fundamental for understanding the MW’s dark matter and baryonic structure (Famaey 2012; Rix & Bovy 2013; Strigari 2013; Read 2014) and the stellar-population-dependent orbit DF is a basic constraint on the Galaxy’s formation his- tory (Binney 2013; Sanders & Binney 2015). There is a variety of practical approaches to dy- namical modelling of discrete collisionless tracers, such as the stars in the MW, e.g., Jeans modelling (Kui- jken & Gilmore 1989; Bovy & Tremaine 2012; Gar- bari et al. 2012; Zhang et al. 2013; B¨ udenbender et al. 2015), action-based DF modelling (with parametric DFs: Bovy & Rix 2013; Piffl et al. 2014; Sanders & Binney 2015; Das & Binney 2016; with marginalization over non-parametric DFs: Magorrian 2014), torus modelling 1 Max-Planck-Institut f¨ ur Astronomie, K¨ onigstuhl 17, D-69117 Heidelberg, Germany 2 Correspondence should be addressed to [email protected]. 3 Department of Astronomy and Astrophysics, University of Toronto, 50 St. George Street, Toronto, ON, M5S 3H4, Canada (McMillan & Binney 2008; McMillan & Binney 2012; McMillan & Binney 2013), or made-to-measure mod- elling (Syer & Tremaine 1996; de Lorenzi et al. 2007; Hunt & Kawata 2014). Most of them—explicitly or implicitly—describe the stellar distribution through a DF. Not all of them avoid binning to exploit the full discrete information content of the data. Recently, Binney (2012b) and Bovy & Rix (2013) pro- posed constraining the MW’s gravitational potential by combining parametrized axisymmetric potential models with DFs that are simple analytic functions of the three orbital actions to model discrete data. Bovy & Rix (2013) (BR13 hereafter) put this in prac- tice by implementing a rigorous modelling approach for so-called mono-abundance populations (MAPs), i.e, sub- sets of stars with similar [Fe/H] and [α/Fe] within the Galactic disk, which seem to follow simple DFs (Bovy et al. 2012a,b,c). Given an assumed (axisymmetric) model for the Galactic potential and action-based DF (Binney 2010; Binney & McMillan 2011; Ting et al. 2013) they calculated the likelihood of the observed (~x,~v) for each MAP, using SEGUE G-dwarf stars (Yanny et al. 2009). They also accounted for the complex, but known selection function of the kinematic tracers (Bovy et al. 2012c). For each MAP the modelling resulted in an in- dependent estimate of the same gravitational potential. Taken as an ensemble, they constrained the disk surface density over a wide range of radii (4 - 9 kpc), and powerfully constrained the disk mass scale length and the stellar-disk-to-dark-matter ratio at the Solar radius. BR13 made however a number of quite severe and ide- alizing assumptions about the potential, the DF and the knowledge of observational effects. These idealizations arXiv:1605.08601v1 [astro-ph.GA] 27 May 2016

RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

  • Upload
    others

  • View
    9

  • Download
    0

Embed Size (px)

Citation preview

Page 1: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Draft version November 13, 2018Preprint typeset using LATEX style emulateapj v. 12/16/11

ACTION-BASED DYNAMICAL MODELLING FOR THE MILKY WAY DISK

Wilma H. Trick1,2, Jo Bovy3, and Hans-Walter Rix1

Draft version November 13, 2018

ABSTRACT

We present RoadMapping, a full-likelihood dynamical modelling machinery that aims to recover theMilky Way’s (MW) gravitational potential from large samples of stars in the Galactic disk. RoadMap-ping models the observed positions and velocities of stars with a parametrized, three-integral distri-bution function (DF) in a parametrized axisymmetric potential. We investigate through differentialtest cases with idealized mock data how the breakdown of model assumptions and data propertiesaffect constraints on the potential and DF. Our key results are: (i) If the MW’s true potential is notincluded in the assumed model potential family, we can—in the axisymmetric case—still find a robustestimate for the potential, with only . 10% difference in surface density within |z| ≤ 1.1 kpc insidethe observed volume. (ii) Modest systematic differences between the true and model DF are inconse-quential. E.g, when binning stars to define sub-populations with simple DFs, binning errors do notaffect the modelling as long as the DF parameters of neighbouring bins differ by < 20%. In addition,RoadMapping ensures unbiased potential estimates for either (iii) small misjudgements of the spatialselection function (i.e., . 15% at the survey volume’s edge), (iv) if distances are known to within 10%,or (v) if proper motion uncertainties are known within 10% or are smaller than δµ . 1 mas yr−1.Challenges are the rapidly increasing computational costs for large sample sizes. Overall, RoadMap-ping is well suited to making precise new measurements of the MW’s potential with data from theupcoming Gaia releases.Keywords: Galaxy: disk — Galaxy: fundamental parameters — Galaxy: kinematics and dynamics —

Galaxy: structure

1. INTRODUCTION

Through dynamical modelling we can infer the MilkyWay’s (MW) gravitational potential from stellar motions(Binney & Tremaine 2008; Binney 2011; Rix & Bovy2013). Observational information on the 6D phase-spacecoordinates of stars is currently growing at a rapid pace,and will be taken to a whole new level in quantity andprecision by the upcoming data from the Gaia mission(Perryman et al. 2001). Yet, rigorous and practical mod-elling tools that turn position-velocity data of individualstars into constraints both on the gravitational potentialand on the distribution function (DF) of stellar orbitsare scarce (Rix & Bovy 2013).

The Galactic gravitational potential is fundamentalfor understanding the MW’s dark matter and baryonicstructure (Famaey 2012; Rix & Bovy 2013; Strigari 2013;Read 2014) and the stellar-population-dependent orbitDF is a basic constraint on the Galaxy’s formation his-tory (Binney 2013; Sanders & Binney 2015).

There is a variety of practical approaches to dy-namical modelling of discrete collisionless tracers, suchas the stars in the MW, e.g., Jeans modelling (Kui-jken & Gilmore 1989; Bovy & Tremaine 2012; Gar-bari et al. 2012; Zhang et al. 2013; Budenbender et al.2015), action-based DF modelling (with parametric DFs:Bovy & Rix 2013; Piffl et al. 2014; Sanders & Binney2015; Das & Binney 2016; with marginalization overnon-parametric DFs: Magorrian 2014), torus modelling

1 Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117Heidelberg, Germany

2 Correspondence should be addressed to [email protected] Department of Astronomy and Astrophysics, University of

Toronto, 50 St. George Street, Toronto, ON, M5S 3H4, Canada

(McMillan & Binney 2008; McMillan & Binney 2012;McMillan & Binney 2013), or made-to-measure mod-elling (Syer & Tremaine 1996; de Lorenzi et al. 2007;Hunt & Kawata 2014). Most of them—explicitly orimplicitly—describe the stellar distribution through aDF. Not all of them avoid binning to exploit the fulldiscrete information content of the data.

Recently, Binney (2012b) and Bovy & Rix (2013) pro-posed constraining the MW’s gravitational potential bycombining parametrized axisymmetric potential modelswith DFs that are simple analytic functions of the threeorbital actions to model discrete data.

Bovy & Rix (2013) (BR13 hereafter) put this in prac-tice by implementing a rigorous modelling approach forso-called mono-abundance populations (MAPs), i.e, sub-sets of stars with similar [Fe/H] and [α/Fe] within theGalactic disk, which seem to follow simple DFs (Bovyet al. 2012a,b,c). Given an assumed (axisymmetric)model for the Galactic potential and action-based DF(Binney 2010; Binney & McMillan 2011; Ting et al. 2013)they calculated the likelihood of the observed (~x,~v) foreach MAP, using SEGUE G-dwarf stars (Yanny et al.2009). They also accounted for the complex, but knownselection function of the kinematic tracers (Bovy et al.2012c). For each MAP the modelling resulted in an in-dependent estimate of the same gravitational potential.Taken as an ensemble, they constrained the disk surfacedensity over a wide range of radii (∼ 4 − 9 kpc), andpowerfully constrained the disk mass scale length andthe stellar-disk-to-dark-matter ratio at the Solar radius.

BR13 made however a number of quite severe and ide-alizing assumptions about the potential, the DF and theknowledge of observational effects. These idealizations

arX

iv:1

605.

0860

1v1

[as

tro-

ph.G

A]

27

May

201

6

Page 2: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

2 Trick et al.

could plausibly translate into systematic errors on theinferred potential, well above the formal error bars ofthe upcoming surveys with their wealth and quality ofdata.

In this work we present RoadMapping (“Recoveryof the Orbit Action Distribution of Mono-AbundancePopulations and Potential INference for our Galaxy”)—an improved, refined, flexible, robust and well-tested ver-sion of the original dynamical modelling machinery byBR13. Our goal is to explore which of the assump-tions BR13 made and which other aspects of data, modeland machinery limit RoadMapping’s recovery of the truegravitational potential.

We investigate the following aspects of the RoadMap-ping machinery that become especially important for alarge number of stars: (i) Numerical inaccuracies mustnot be an important source of systematics (Section 2.6).(ii) As parameter estimates become much more precise,we need more flexibility in the potential and DF modeland efficient strategies to find the best fit parameters.The improvements made in RoadMapping as comparedto the machinery used in BR13 are presented in Section2.8. (iii) RoadMapping should be an unbiased estimator(Section 3.1).

We also explore how different aspects of the observa-tional experiment design impact the parameter recovery:(i) We consider the importance of the survey volumegeometry, size, shape and position within the MW toconstrain the potential (Section 3.2). (ii) We ask whathappens if our knowledge of the sample selection func-tion is imperfect, and potentially biased (Section 3.3).(iii) We investigate how to best account for individual,and possibly misjudged, measurement uncertainties (Sec-tion 3.4). (iv) We determine which of several stellar sub-populations is best for constraining the potential (Section3.7).

One of the strongest assumptions is restricting the dy-namical modelling to a certain family of parametrizedfunctions for the gravitational potential and the DF. Weinvestigate how well we can hope to recover the true po-tential, when our models do not encompass the true DF(Section 3.5) and potential (Section 3.6).

For all of the above aspects we show some plausible andillustrative examples on the basis of investigating mockdata. The mock data is generated from galaxy modelsoutlined in Sections 2.1-2.4 following the procedure inAppendices A-B and analysed according to the descrip-tion of the RoadMapping machinery in Sections 2.5-2.8.Section 3 compiles our results on the investigated mod-elling aspects. In particular, our key results about thesystematics introduced by using wrong DF or potentialmodels are presented in the Sections 3.5 and 3.6. Section4 finally summarizes and discusses our findings.

2. DYNAMICAL MODELLING

In this section we summarize the basic elements ofRoadMapping, the dynamical modelling machinery pre-sented in this work, which in many respects follows BR13and makes extensive use of the galpy Python package forgalactic dynamics4 (Bovy 2015).

4 galpy is an open-source code that is being developed on http://github.com/jobovy/galpy. The latest documentation can befound at http://galpy.readthedocs.org/en/latest/.

2.1. Coordinate system

Our modelling takes place in the Galactocentric rest-frame with cylindrical coordinates x ≡ (R,φ, z) and cor-responding velocity components v ≡ (vR, vT , vz). If thestellar phase-space data is given in observed heliocen-tric coordinates, position x ≡ (RA,Dec,m−M) in rightascension RA, declination Dec and distance modulus(m−M), and velocity v ≡ (µRA · cos(Dec), µDec, vlos) asproper motions and line-of-sight velocity, the data (x, v)has to be converted into the Galactocentric rest-framecoordinates (x,v) using the Sun’s position and velocity.We assume for the Sun

(R, φ, z) = (8 kpc, 0, 0 kpc)

(vR, vT, vz) = (0, 230, 0) km s−1.

2.2. Actions

Stellar orbits in (axisymmetric) gravitational poten-tials are best described and fully specified by the threeactions J ≡ (JR, Jz, Jφ = Lz), defined as

Ji ≡1

∮orbit

pi dxi, (1)

which is evaluated along the orbit with position x(t) andmomentum p(t) in a given potential Φ. Actions have sev-eral convenient properties which make them excellent or-bit labels and ideal as arguments for orbit DFs: Actionsare integrals of motion; actions have an intuitive physi-cal meaning as they quantify the amount of oscillation ofthe orbit in each coordinate direction; actions—togetherwith a set of angle coordinates θ—form canonical con-jugate phase-space coordinates, i.e., the Jacobian deter-minant |∂(J ,θ)/∂(x,v)| = 1 (with Cartesian x and v).The angles θ(t) ∝ t evolve linearly in time and specifythe position of the star along the orbit. (For a full intro-duction to angle-action variables see Binney & Tremaine2008, §3.5.)

Action calculation from a star’s phase-space coordi-

nates, (x,v)Φ−→ J , is typically very computationally ex-

pensive. Only for some special, separable potentials doesEquation (1) simplify significantly. The triaxial Stackelpotentials (de Zeeuw 1985) are the most general poten-tials, that allow exact action calculations using a singlequadrature. Some flattened axisymmetric Stackel poten-tials are quite similar to our Galaxy’s potential (Bin-ney & Tremaine 2008, §3.5.3; Batsleer & Dejonghe 1994;Famaey & Dejonghe 2003). The spherical isochrone po-tential (Henon 1959; Binney & Tremaine 2008, §3.5.3) isthe most general special case for which the action calcu-lation is analytic without any integration. In all otherpotentials actions have to be numerically estimated; seeSanders & Binney (2016) for a recent review of actionestimation methods. According to Sanders & Binney(2016) the best compromise of speed and accuracy forthe Galactic disk is the Stackel fudge by Binney (2012a)for axisymmetric potentials. In addition we use actioninterpolation grids (Binney 2012a; Bovy 2015) to speedup the calculation. The latter is one of the improvementsemployed by RoadMapping, which was not used in BR13.

Page 3: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 3

Table 1Axisymmetric gravitational potential models used throughout this work. The potential parameters are fixed

for the mock data creation at the values given in this table, which we susequently aim to recover withRoadMapping. The parameters of DHB-Pot and KKS-Pot were chosen to resemble the MW14-Pot (see Figure 1).

We use vcirc(R) = 230 km s−1 as the circular velocity at the Sun for all potentials in this work.

name potential model parameters pΦ action calculation

Iso-Pot isochrone potential(a) b 0.9 kpc analytic and exact(Henon 1959) (Binney & Tremaine 2008, §3.5.2)

KKS-Pot 2-component ∆ 0.3 exactKuzmin-Kutuzov-

(ac

)Disk

20 using interpolation

Stackel potential(b):(ac

)Halo

1.07 on action griddisk and halo k 0.28 (Binney 2012a; Bovy 2015)(Batsleer & Dejonghe 1994)

DHB-Pot Disk+Halo+Bulge potential(c): adisk 3 kpc approximateMiyamoto-Nagai disk, bdisk 0.28 kpc (fixed) using Stackel fudgeNFW halo, fhalo 0.35/0.95 (Binney 2012a)Hernquist bulge ahalo 16 kpc (fixed) and interpolation on action grid(same as MW14-Pot, fbulge 0.05/1.0 (fixed)except of bulge) abulge 0.6 kpc (fixed)

MW14-Pot MW-like potential(d): approximateMiyamoto-Nagai disk, using Stackel fudgeNFW halo,cut-off power-law bulge(Bovy 2015)

(a) The free parameter of the spherical Iso-Pot is the isochrone scale length b.(b) The coordinate system of each of the two Stackel-potential components of the KKS-Pot is R2/(τi,p+αp)+z2/(τi,p+γp) = 1 with p ∈ Disk,Halo and τi,p ∈ λp, νp. Both components have the same focal distance ∆ ≡

√γp − αp, to

ensure that the superposition itself is a Stackel potential. The axis ratio of the coordinate surfaces (a/c)p :=√αp/γp

describes the flatness of each component. k is the relative contribution of the disk mass to the total mass.(c) The parameters of the DHB-Pot are the Miyamoto-Nagai disk scale length adisk and height bdisk, the NFW haloscale length ahalo and its relative contribution to v2circ(R) with respect to the total disk+halo contribution, fhalo,and the Hernquist bulge scale length abulge and its contribution to the total v2circ(R), fbulge. We keep all exceptvcirc(R), adisk and fhalo fixed to their true values in the analysis.(d) The MWPotential2014 by Bovy (2015) (see their Table 1) has vcirc(R) = 220 km s−1. We use however vcirc(R) =230 km s−1.

−10 −5 0 5 10R [kpc]

−5

0

5

z [k

pc]

Iso−Pot

−10 −5 0 5 10R [kpc]

KKS−Pot

−10 −5 0 5 10R [kpc]

DHB−Pot

−10 −5 0 5 10R [kpc]

MW14−Pot

SunGalacticplane

Figure 1. Density distribution of the four reference galaxy potentials in Table 1. These potentials are used throughout this work to createand model mock data with RoadMapping.

2.3. Potential models

For the gravitational potential in our modelling weassume a family of parametrized models. We use: AMW-like potential with disk, halo and bulge (DHB-Pot);the spherical isochrone potential (Iso-Pot); and the 2-component Kuzmin-Kutuzov Stackel potential (Batsleer& Dejonghe 1994; KKS-Pot), which also displays a diskand halo structure. Table 1 summarizes all reference po-tentials used in this work together with their free param-eters pΦ. The true circular velocity at the Sun was chosento be vcirc(R) = 230 km s−1 for all potential models.The Iso-Pot allows both accurate and particularly fastaction calculations; we use it therefore for tests requiringa large number of analyses. The KKS-Pot and DHB-Potwere chosen for their more realistic galaxy shape and be-cause their closed-form expression for Φ(R, z) makes the

computation of forces and densities fast and easy. TheKKS-Pot allows also for exact action calculations, whilethe DHB-Pot has physically more intuitive potential pa-rameters, but requires the Stackel fudge to estimate ac-tions. The parameter values of KKS-Pot and DHB-Pot inTable 1 were chosen to resemble the MW potential fromBovy (2015) (MW14-Pot). The density distribution of allthese potentials is illustrated in Figure 1.

2.4. Stellar distribution functions

A stellar distribution function DF(x,v) can be con-sidered as the probability of a star to be found at (x,v).Using instead orbit DFs in terms of (J ,θ) has the advan-tage that the distribution of stars in θ is uniform and theorbit DF reduces effectively to a function of the actionsJ only. As |∂(J ,θ)/∂(x,v)| = 1, the function DF(J)

Page 4: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

4 Trick et al.

Table 2Reference parameters for the qDF in Equations (2)-(7), used to create 6Dphase-space mock data sets for stellar populations of different kinematic

temperature. (The parameter X is explained in Section 2.4.)

name qDF parameters pDF

hR [kpc] σR,0 [km s−1] σz,0 [km s−1] hσ,R [kpc] hσ,z [kpc]

hot 2 55 66 8 7cool 3.5 42 32 8 7cooler 3 27.5 33 8 7colder 2 +X% 55−X% 66−X% 8 7warmer 3.5−X% 42 +X% 32 +X% 8 7

can still be thought of as a probability in (x,v).The action-based quasi-isothermal distribution func-

tion (qDF) by Binney (2010) and Binney & McMillan(2011) is a simple DF which we will employ as a specificexample throughout this work to describe individual stel-lar sub-populations. This is motivated by the findingsof Bovy et al. (2012a,b,c) and Ting et al. (2013) on thesimple phase-space structure of stellar MAPs and BR13’ssuccessful application. The qDF has the form

qDF(J | pDF)

= fσR (JR, Lz | pDF)× fσz (Jz, Lz | pDF) (2)

with some free parameters pDF and

fσR (JR, Lz | pDF) =n× Ω

πσ2R(Rg)κ

exp

(− κJRσ2R(Rg)

)× [1 + tanh (Lz/L0)] (3)

fσz (Jz, Lz | pDF) =ν

2πσ2z(Rg)

exp

(− νJzσ2z(Rg)

)(4)

(Binney & McMillan 2011). Here Rg, Ω, κ and ν arefunctions of Lz and denote respectively the guiding-center radius, circular, radial/epicycle and vertical fre-quency of the near-circular orbit with angular momen-tum Lz in a given potential. The term [1 + tanh (Lz/L0)]suppresses counter-rotation for orbits in the disk withLz < L0 (with L0 ∼ 10 km s−1 kpc).

Following BR13, we choose the functional forms

n(Rg | pDF)∝ exp

(−RghR

)(5)

σR(Rg | pDF) =σR,0 × exp

(−Rg −R

hσ,R

)(6)

σz(Rg | pDF) =σz,0 × exp

(−Rg −R

hσ,z

), (7)

which indirectly set the stellar number density and radialand vertical velocity dispersion profiles. The qDF hastherefore a set of five free parameters pDF: the densityscale length of the tracers hR, the radial and vertical ve-locity dispersion at the Solar position R, σR,0 and σz,0,and the scale lengths hσ,R and hσ,z, that describe theradial decrease of the velocity dispersion. RoadMappingallows to fit any number of DF parameters simultane-ously, while BR13 kept σR,0, hσ,R fixed. Throughoutthis work we make use of a few example stellar pop-ulations whose qDF parameters are given in Table 2:Most tests use the hot and cool qDFs, which correspondto kinematically hot and cool populations, respectively.

The warmer (cooler and colder) qDFs in Table 2 werechosen to have the same anisotropy σR,0/σz,0 as the cool(hot) qDF, with X being a free parameter describing thetemperature difference. Hotter populations have shortertracer scale lengths (Bovy et al. 2012c) and the veloc-ity dispersion scale lengths were fixed according to Bovyet al. (2012b).

One indispensable step in our dynamical modellingtechnique (Section 2.5-2.6), as well as in creating mockdata (Appendix A), is to calculate the (axisymmetric)spatial tracer density ρDF(x | pΦ, pDF) for a given DFand potential. Analogously to BR13,

ρDF(R, |z| | pΦ, pDF)

=

∫ ∞−∞

DF(J [R, z,v | pΦ] | pDF) d3v

≈∫ nσσR(R|pDF)

−nσσR(R|pDF)

∫ nσσz(R|pDF)

−nσσz(R|pDF)

∫ 1.5vcirc(R)

0

DF(J [R, z,v | pΦ] | pDF) dvT dvz dvR, (8)

where σR(R | pDF) and σz(R | pDF) are given by Equa-tions (6) and (7).5 Each integral is evaluated using aNv-th order Gauss-Legendre quadrature. For a given pΦ

and pDF we explicitly calculate the density on Nx ×Nxregular grid points in the (R, z) plane and interpolateln ρDF in between using bivariate spline interpolation.The grid is chosen to cover the extent of the observations(for |z| ≥ 0, because the model is symmetric in z by con-struction). The total number of actions to be calculatedto set up the density interpolation grid is N2

x×N3v , which

is one of the factors limiting the computation speed. Tocomplement the work by BR13, we will specifically workout in Section 2.6 and Figure 2 how large Nx, Nv and nσhave to be chosen to get the density with a sufficientlyhigh numerical accuracy.

2.5. Data likelihood

As data D we consider here the positions and veloc-ities of a sub-population of stars within a given surveyselection function SF(x),

D ≡ xi,vi | (star i in given sub-population)

∧ (SF(xi) > 0).For simplicity we assume in most tests of this study

contiguous, spherical SFs centred on the Sun, which are

5 The integration ranges over the velocities are motivated byFigure A2 and nσ should be chosen as nσ ∼ 5 (see Figure 2). Theintegration range [0, 1.5vcirc(R)] over vT is in general sufficient,only for observation volumes with larger mean stellar vT this upperlimit needs to be increased.

Page 5: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 5

functions of x only and which we motivate in AppendixB. The maximum radius of this spherical observed vol-ume is denoted by rmax.

We fit a model potential and DF (here: the qDF) whichare specified by a number of fixed and free model param-eters,

pM ≡ pDF, pΦ.

The orbit of the i-th star in a potential with pΦ is labelledby the actions J i ≡ J [xi,vi | pΦ] and the DF evaluatedfor the i-th star is then DF(J i | pM ) ≡ DF(J [xi,vi |pΦ] | pDF).

The likelihood of the data given the model is, followingBR13 and McMillan & Binney (2013),

L (D | pM )

≡N∗∏i

p(xi,vi | pM )

=

N∗∏i

DF(J i | pM ) · SF(xi)∫DF(J | pM ) · SF(x) d3xd3v

∝N∗∏i

DF(J i | pM )∫ρDF(R, |z| | pM ) · SF(x) d3x

, (9)

where N∗ is the number of stars in D, and in the last step

we used Equation (8).6∏N∗i SF(xi) is independent of

pM , so we treat it as unimportant proportionality factor.We find the best fitting pM by maximizing the posteriorprobability distribution pdf(pM | D), which is, accordingto Bayes’ theorem

pdf(pM | D) ∝ L (D | pM ) · p(pM ),

where p(pM ) is some prior probability distribution on themodel parameters. We assume flat priors in both pΦ and

pDF := lnhR, lnσR,0, lnσz,0, lnhσ,R, lnhσ,z (10)

(see Section 2.4) throughout this work. Then pdf andlikelihood are proportional to each other and differ onlyin units.

In this case, where we use uninformative priors, amaximum-likelihood estimation procedure (e.g., via theexpectation-maximization (EM) algorithm and param-eter uncertainty estimates from the Fisher informationmatrix) would lead to the same result as the Bayesianinferential procedure described in this work (see Section2.8). We expect however that in due course increasinglyinformative priors will become available (like, e.g., ro-tation curve measurements from maser sources by Reidet al. (2009)) and Bayesian inference is therefore the pre-ferred framework.

2.6. Likelihood normalisation

The normalisation in Equation (9) is a measure for thetotal number of tracers inside the survey volume,

Mtot ≡∫ρDF(R, |z| | pM ) · SF(x) d3x. (11)

6 Because |∂(J ,θ)/∂(x,v)| = 1, the integration over phase-spacein the normalisation term can be performed either over (J ,θ) orCartesian (x,v).

In the case of an axisymmetric Galaxy model andSF(x) = 1 within the observation volume (as in mosttests in this work), the normalisation is essentially a two-dimensional integral in the (R, z) plane over ρDF withfinite integration limits. We evaluate the integrals usingGauss-Legendre quadratures of order 40. The integralover the azimuthal direction can be solved analytically.

It turns out that a sufficiently accurate evaluation ofthe likelihood is computationally expensive, even for onlyone set of model parameters. This expense is dominatedby the number of action calculations required, which inturn depends on N∗ and the numerical accuracy of thetracer density interpolation grid with N2

x×N3v grid points

in Equation (8) needed for the likelihood normalisation inEquation (11). The accuracy of the normalisation has tobe chosen high enough, such that the resulting numericalerror

δMtot ≡Mtot,approx(Nx, Nv, nσ)−Mtot

Mtot(12)

does not dominate the numerically calculated log-likelihood, i.e.,

ln Lapprox(D | pM )

=

N∗∑i

ln DF(Ji | pM )−N∗ ln(Mtot)

−N∗ ln(1 + δMtot), (13)

with

ln(1 + δMtot) ≤1

N∗, (14)

and therefore δMtot . 1/N∗. Otherwise numerical inac-curacies could lead to systematic biases in the potentialand DF recovery. For data sets as large as N∗ = 20, 000stars, which in the age of Gaia could very well be thecase, one needs a numerical accuracy of 0.005% in thenormalisation. We made sure that this is satisfied for allanalyses in this work. Figure 2 demonstrates how the nu-merical accuracy for analyses with the DHB-Pot dependson the spatial and velocity resolution of the grid and thatthe accuracy we use, Nx = 16, Nv = 24 and nσ = 5, issufficient.7 It has to be noted however, that the opti-mal values for Nx, Nv and nσ depend not only on N∗,but also on the kinematic temperature of the population(and to a certain degree even on the choice of potential8)and it has to be checked on a case-by-case basis what theoptimal accuracy is.

McMillan & Binney (2013), who use a similar mod-elling approach and likelihood normalisation, argued thatthe required accuracy for the normalisation scales aslog10 (1 + δMtot) ≤ 1/N∗ ⇒ δMtot . 2.3/N∗, which issatisfied for our tests as well. They evaluate the inte-grals in the normalisation via Monte-Carlo integration

7 The accuracy used in this work’s analyses is slightly higherthan in BR13, where N∗ was only a few ∼ 100.

8 In Figure 17 we will show a comparison for the qDF param-eters of two very similar mock data distributions in two differentpotentials, the MW14-Pot and a best fit potential of the parametricform of the KKS-Pot. As some of the qDF parameters in both po-tentials are very different, and even more different from the actualphysical scale lengths and velocity dispersions, an optimal nσ hasto be estimated first for a given potential model before running theRoadMapping analysis.

Page 6: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

6 Trick et al.

10-5

10-4

10-3

10-2

10-1

100

hot population

norm

alis

ation error

|δMtot| [

%]

Nv = 24, nσ = 5

δMtot ≲ 1/N ∗N ∗=20,000

Nx = 16, nσ = 5 Nx = 16, Nv = 4.8 nσ

0.0 0.2 0.4 0.6 0.8spatial resolution rmax/Nx

10-5

10-4

10-3

10-2

10-1

100

cold population

norm

alis

ation error

|δMtot| [

%]

rmax of obs.volume:

200pc1kpc2kpc3kpc4kpc

0.1 0.2 0.3 0.4 0.5 0.6velocity resolution nσ /Nv

3 4 5 6 7velocity range nσ

Figure 2. Relative error of the likelihood normalisation, δMtot in Equation (12), depending on the accuracy of the grid-based densitycalculation in Equation (8) (and surrounding text) in five spherical observation volumes with different radius rmax and depending on thekinematic temperature of the population in the DHB-Pot. (Test 1 in Table 3 summarizes the model parameters.) The tracer density inEquation (8) is calculated on Nx × Nx spatial grid points in R ∈ [R ± rmax] and |z| ∈ [0, rmax]. The integration over the velocitiesis performed with Gauss-Legendre quadratures of order Nv within an integration range of ±nσ times the dispersion σR(R) and σz(R)(and [0, 1.5vcirc] in vT ). (We vary Nx, Nv and nσ separately and keep the other two fixed at the values indicated above each panel.) Wecalculate the “true” normalisation Mtot in Equation (12) with high accuracy as Mtot ≡Mtot,approx(Nx = 32, Nv = 68, nσ = 7). The blackstars indicate the accuracy used in analyses with the DHB-Pot, Tests 5 and 7: It is better than 0.005% (dotted line), which is requiredfor N∗ = 20, 000 stars. We find that the spatial resolution of the grid is important and depends on the kinematic temperature of thepopulation, as cooler populations have a steeper density gradient in z-direction, which has to be sampled sufficiently.

with ∼ 109 sample points in action space. Our approachuses a tracer density interpolation grid for which the res-olution needs to be optimized by hand, but it has theadvantage that it then only requires the calculation ofN2x ×N3

v ∼ 4 · 106 − 107 actions per normalisation.

2.7. Measurement uncertainties

Measurement uncertainties of the data have to beincorporated in the likelihood. We assume Gaussianuncertainties in the observable space y ≡ (x, v) =(RA,Dec, (m − M), µRA · cos(Dec), µDec, vlos), i.e., thei-th star’s observed yi is drawn from the normal distri-

bution N [yi′, δyi] ≡

∏6i N [yi,k

′, δyi,k] =∏6i exp−(yk −

yi,k′)2/(2δy2

i,k)/√

2πδy2i,k, with yi

′ being the star’s true

phase-space position, δyi its uncertainty, and yk the k-thcoordinate component of y. Stars follow the DF(J [y′ |pΦ] | pDF) (≡ DF(y′) ≡ for short) convolved with themeasurement uncertainties N [0, δyi]. The selection func-tion SF(y) acts on the space of (uncertainty affected)observables. Then the probability of one star becomes

p(yi | pΦ, pDF, δyi)

≡SF(yi) ·

∫DF(y′) ·N [yi, δyi] d6y′∫ (

DF(y′) ·∫

SF(y) ·N [y′, δyi] d6y)

d6y′. (15)

In the case of uncertainties in distance and/or (RA,Dec)the evaluation of this is computational very expensive—especially if the stars have heteroscedastic δyi, which isthe case for realistic data sets, and the normalisationneeds to be calculated for each star separately. In prac-tice we compute the convolution using Monte Carlo (MC)

integration with Nsamples samples,

papprox(yi | pΦ, pDF, δyi)

≈ SF(xi)

Mtot· 1

Nsamples

Nsamples∑n

DF(xi,v[y′i,n]) (16)

with

y′i,n ∼ N [yi, δyi].

In addition, this approximation assumes that the star’sposition xi is perfectly measured. As the SF is alsovelocity independent, this simplifies the normalisationdrastically to Mtot in Equation (11). Measurementuncertainties in RA and Dec are often negligible any-way. The uncertainties in the Galactocentric velocitiesvi = (vR,i, vT,i, vz,i) depend not only on δµ and δvlos

but also on the distance and its uncertainty, which wedo not neglect when drawing MC samples y′i,n from thefull uncertainty distribution N [yi, δyi].

An analogous but one-dimensional treatment of mea-surement uncertainties in only vz was already appliedby BR13. Similar approaches ignoring measurement un-certainties in the likelihood normalisation and using MCsampling of the error ellipses were also used by McMillan& Binney (2013) and Das & Binney (2016). In Section3.4, Figure 9 (Test 6.2 in Table 3), we will investigatethe breakdown of our approximation for non-negligibledistance uncertainties.

Figure 3 demonstrates that in the absence of positionuncertainties the Nsamples needed for the convolution in-

Page 7: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 7

0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3

δvmax/σ

101

102

103

Nsamples

Nsamples ∝ (δvmax)2

Iso−Pot parameters:vcirc(R⊙)Iso−Pot b

bias due to in-sufficient sampling:

0.25 SE

0.5 SE

1.0 SE

2.0 SE

Figure 3. Number of MC samples Nsamples needed for the nu-merical convolution of the model probability with the measurementuncertainties in Equation (16), given the maximum velocity uncer-tainty δvmax within the stellar sample with respect to the sample’skinematic temperature σ. Insufficient sampling introduces system-atic biases in the parameter recovery; the size of the bias (in unitsof the standard error (SE) on the parameter estimate) is indicatedin the legend. The relation found here, Nsamples ∝ δv2

max, wasdistilled from analyses (with different Nsamples) of mock data sets

with different proper motion uncertainties δµ ∈ [2, 5] mas yr−1 (seeTest 2 in Table 3). As the reference for the converged convolutionintegral, we used Nsamples = 800 and 1200 for δµ ≤ 3 mas yr−1

and δµ > 3 mas yr−1, respectively (see also left panels in Figure9). We plot δvmax in units of the sample temperature, which wequantify by σ ≡ (σR,0 + σz,0)/2 (see Table 2 for the hot qDF).

tegral to converge depends as

Nsamples ∝ (δv)2

on the uncertainties in the (1D) velocities. Figure 3 isbased on analyses of mock data sets with different propermotion uncertainties δµ (see Test 2 in Table 3 for all pa-rameters). The proper motion uncertainty δµ translatesto heteroscedastic velocity uncertainties according to

δv[km s−1] ≡ 4.74047 · r[kpc] · δµ[mas yr−1],

with r being the distance star—Sun. Stars with largerδv require more Nsamples for the integral over their mea-surement uncertainties to converge; Figure 3 thereforeshows how the Nsamples—needed for the pdf of the wholedata set to be converged—depends on the largest velocityerror δvmax ≡ δv(rmax) within the data set.

These mock data sets contained each N∗ = 10, 000stars. We found that for N∗ = 5, 000 the requiredNsamples to reach a given accuracy becomes smaller forvcirc(R), but remains similar for b. The former is consis-tent with our expectation that we need higher accuracyand therefore more Nsamples for larger data sets. Thelatter seems to be a special property of the Iso-Pot (seealso the discussion in Section 3.3).

2.8. Fitting procedure

To search the (pΦ, pDF) parameter space for the max-imum of the pdf in Equation (9), we go beyond the sin-gle fixed grid search by BR13 and employ an efficient

two-step procedure: Nested-grid search and Monte-CarloMarkov Chain (MCMC).

The first step employs a nested-grid search to find theapproximate peak and width of the pdf in the high-dimensional pM space with a low number of likelihoodevaluations:

• Initialization. ForNp free model parameters pM westart with a sufficiently large grid with 3Np regularpoints.

• Evaluation. We evaluate the pdf at each grid-pointsimilar to BR13 (their Figure 9): An outer loopiterates over the potential parameters pΦ and pre-calculates all N∗ ×Nsamples +N2

x ×N3v actions re-

quired for the likelihood calculation (see Equations(8), (9) and (16)). Then an inner loop evaluatesEquation (9) (or (16)) for all DF parameters pDF

in the given potential.

• Iteration. For each of the model parameters pMwe marginalize the pdf . A Gaussian is fitted to themarginalized pdf and the peak ± 4σ become theboundaries of the next grid with 3Np grid points.The grid might be still too coarse or badly posi-tioned to fit Gaussians. In that case we either zoominto the grid point with the highest probability orshift the current range to find new approximategrid boundaries. We proceed with iteratively eval-uating the pdf on finer and finer grids, until wehave found a reliable 4σ fit range in each of thepM dimensions. The central grid point is then veryclose to the best fit pM , and the grid range is ofthe order of the pdf width.

• The fiducial qDF. To save time by pre-calculatingactions, they have to be independent of the choiceof pDF. However, the normalisation in Equation(11) requires actions on a N2

x×N3v grid and the grid

ranges in velocity space do depend on the currentpDF (see Equation (8)). To relax this, we followBR13 and use a fixed set of qDF parameters (thefiducial qDF ) to set the velocity grid boundaries inEquation (8) globally for a given pΦ. Choosing afiducial qDF that is very different from the true DFcan however lead to large biases in the pM recov-ery. BR13 did not account for that. RoadMappingavoids this as follows: To get successively closer tothe optimal fiducial qDF—with the (yet unknown)best fit pDF—we use in each iteration step of thenested-grid search the central grid point of the cur-rent pM grid as the fiducial qDF’s pDF. As thenested-grid search approaches the best fit values,the fiducial qDF approaches its optimum as well.

• Computational expense. Overall the computationspeed of this nested-grid approach is dominated(in descending order of importance) by a) the com-plexity of potential and action calculation, b) theN∗ × Nsamples + N2

x × N3v actions required to be

calculated per pΦ, c) the number of potential pa-rameters and d) the number of DF parameters.

The second step samples the shape of the pdf usingMCMC. Formally, calculating the pdf on a fine grid like

Page 8: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

8 Trick et al.

228

229

230

231

vcirc(R⊙) [km s−1 ]

0.8

0.9

1.0

b [k

pc]

0.8

0.9

1.0

b [kpc]

−1.45

−1.40

−1.35

−1.30

ln(h

R/8

kpc)

−1.

45

−1.

40

−1.

35

−1.

30ln(hR /8kpc)

−1.40

−1.38

−1.36

ln(σ

R,0/2

20

km s−1

)

−1.

40

−1.

38

−1.

36ln(σR,0/220km s−1 )

−1.22

−1.20

−1.18

−1.16

ln(σ

z,0/

22

0km

s−1

)

−1.

22

−1.

20

−1.

18

−1.

16ln(σz,0/220km s−1 )

−0.05

0.00

0.05

0.10

ln(h

σ,R

/8kp

c)

−0.

05

0.00

0.05

0.10

ln(hσ,R/8kpc)

228

229

230

231

vcirc(R⊙) [km s−1 ]

−0.20

−0.15

−0.10

ln(h

σ,z/8

kpc)

0.8

0.9

1.0

b [kpc]−1.

45

−1.

40

−1.

35

−1.

30

ln(hR /8kpc)−1.

40

−1.

38

−1.

36

ln(σR,0/220km s−1 )−1.

22

−1.

20

−1.

18

−1.

16

ln(σz,0/220km s−1 )−0.

050.

000.

050.

10

ln(hσ,R/8kpc)−0.

20

−0.

15

−0.

10

ln(hσ,z/8kpc)

−0.

20

−0.

15

−0.

10ln(hσ,z/8kpc)

Figure 4. The pdf in the parameter space pM = pΦ, pDF for one example mock data set (see Test 3.1 in Table 3). Blue indicates the pdffor the potential parameters pΦ, green the qDF parameters pDF. The true parameters are marked by dotted lines. The dark, medium andbright contours in the 2D distributions represent 1σ, 2σ and 3σ confidence regions, respectively. The parameters are weakly to moderatelycovariant, but their level of covariance depends on the actual choice of the mock data’s pM . The pdf here was sampled using MCMC. Thedashed lines in the 1D distributions are Gaussian fits to the histogram of MCMC samples. This demonstrates that for such a large numberof stars, the pdf approaches the shape of a multi-variate Gaussian that also projects into Gaussians when considering the marginalized pdffor all the individual pM , as expected for a maximum likelihood estimator.

BR13 (e.g., with K = 11 grid points in each dimen-sion) would provide the same information. However thenumber of expensive pdf evaluations scales as KNp . Fora high-dimensional pM (Np > 4), a MCMC approachmight sample the pdf much faster: We use emcee byForeman-Mackey et al. (2013) and release the walkersvery close to the best fit pM found by the nested-gridsearch, which assures fast convergence in much less thanKNp pdf evaluations. We also use the best fit pM of thegrid-search as fiducial qDF for the whole MCMC. In do-ing so, the normalisation varies smoothly with differentpM and is less sensitive to the accuracy in Equation (8).

3. RESULTS

We are now in a position to examine the limitations ofaction-based modelling posed in the introduction usingour RoadMapping machinery. We explore: (i) whetherthe parameter estimates are unbiased, (ii) the role of thesurvey volume, (iii) imperfect selection functions, (iv)measurement uncertainties, and what happens if the true(v) DF or (vi) potential are not included in the space ofmodels.

We will rely on mock data as input to explore the lim-

itations of the modelling. The mock data is generateddirectly from the fiducial potential and DF models intro-duced in Sections 2.3 and 2.4, following the proceduredescribed in Appendix A. With the exception of the testsuite on measurement uncertainties in Section 3.4, we as-sume that phase-space uncertainties are negligible. Alltests are also summarized in Table 3.

We do not explore the breakdown of the assumptionthat the system is axisymmetric and in steady state northe impact of resonances, which is not possible in thecurrent setup using mock data drawn from axisymmetricgalaxy models. We plan however to investigate this in afuture paper, where we will apply RoadMapping to N-body simulations of disk galaxies.

3.1. Model parameter estimates in the limit of largedata sets

The MAPs in BR13 contained between 100 and 800objects, which implied broad pdfs for the model parame-ters pM . Several consequences arise in the limit of muchlarger samples, say N∗ = 20, 000: (i) As outlined in Sec-tion 2.6 and investigated in Figure 2 (Test 1 in Table 3),higher numerical accuracy is needed due to the likelihood

Page 9: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 9

1.5 2.0 2.5 3.0 3.5 4.0

Standard Error (SE)[% of true value]

−4

−3

−2

−1

0

1

2

3

4

bia

s [S

E]

= (

best

fit

- t

rue v

alu

e)

/ SE

rmax of obs. volume (color)

4kpc

3kpc

2kpc

1kpc

200pc

# analyses

Iso−

Pot: b

8 10 12 14 16 18 20 22

SE [%]

−4

−3

−2

−1

0

1

2

3

4

bia

s [S

E]

Stellar population (symbol)

hot cool

#

qD

F: ln(h

σ,z /8

kpc)

Gaussian fit to histogram

Normal distribution

Figure 5. Lack of bias in the parameter estimates. Maximumlikelihood estimators converge to the true parameter values forlarge numbers of data points and have a Gaussian spread—if themodel assumptions are fulfilled. To test that these conditions aresatisfied for RoadMapping, we create 320 mock data sets, whichcome from two different stellar populations and five spherical ob-servation volumes (see legends). (All model parameters are sum-marized in Table 3 as Test 3.2.) Bias and relative standard error(SE) are derived from the marginalized pdf for two model param-eters (isochrone scale length b in the first row and qDF parameterhσ,z in the second row). The second column displays a histogramof the 320 bias offsets. As it closely follows a normal distribution,our modelling method is therefore well-behaved and unbiased. Theblack dots denote the pdf expectation value for the 32 analyses be-longing to the same pM .

normalisation requirement δMtot . 1/N∗ (see Equation(14)), which drives the computing time. (ii) The pdfs ofthe pM become Gaussian, with a pdf width (i.e., the stan-dard error (SE) on the parameter estimate) that scalesas 1/

√N∗. The former is demonstrated in Figure 4 (Test

3.1 in Table 3) and we also verified that the latter is true.(iii) Any bias in the pdf expectation value has to be con-siderably less than the SE. Figure 5 (Test 3.2 in Table 3)illustrates that RoadMapping behaves like an unbiasedmaximum likelihood estimator: The average parameterestimates from many mock data sets are very close to theinput pM , and the distribution of the actual parameterestimates are Gaussian around it.

3.2. The role of the survey volume geometry

To explore the role of the survey volume at given sam-ple size, we devise two suites of mock data sets.

The first suite draws mock data for two different po-tentials (Iso-Pot and DHB-Pot) and four volume wedges(see Appendix B) with different extent and at differ-ent positions within the Galaxy, illustrated in the up-per panel of Figure 6. Otherwise the data sets are gen-erated from the same pM (see Test 4 in Table 3). Toisolate the role of the survey volume geometry and po-sition, the mock data sets all have the same number ofstars (N∗ = 20, 000), and are drawn from identical totalsurvey volumes (4.5 kpc3, achieved by adjusting the an-

3 4 5 6 7 8 9 10 11R [kpc]

−2

−1

0

1

2

3

z [k

pc]

Galactic plane

Sun

Extent in φ direction such thatall wedges have the same volume.

Galactic plane

Sun

Extent in φ direction such thatall wedges have the same volume.

Galactic plane

Sun

Extent in φ direction such thatall wedges have the same volume.

Galactic plane

Sun

Extent in φ direction such thatall wedges have the same volume.

Shape & position of observed volume

1 2 3 4 5Standard Error (SE)[% of true value]

−3−2−1

0123

bia

s [S

E]

adisk

2 4 6 8 10SE [%]

fhalo

0.0 0.2 0.4 0.6 0.8 1.0−3−2−1

0123

bia

s [S

E]

= (

best

fit

- t

rue v

alu

e)

/ SE

vcirc(R⊙)

1 2 3 4 5

b

Iso−P

ot

DHB−P

ot

Figure 6. Bias vs. standard error in recovering the potentialparameters for mock data sets drawn from four different wedge-shaped test observation volumes within the Galaxy (illustratedin the upper panel; the corresponding analyses are colour-coded)and two different potentials (Iso-Pot and DHB-Pot from Table 1;see also Test 4 in Table 3 for all model parameters used). Stan-dard error and offset were determined from a Gaussian fit to themarginalized pdf . The angular extent of each wedge-shaped obser-vation volume was adapted such that all have a volume of 4.5 kpc3,even though their extent in (R, z) is different. (The recovery ofthe free potential parameter vcirc(R) in the different wedges isvery similar for both potentials and therefore only shown for theIso-Pot). Minor expected differences can be seen (e.g., vcirc(R)and tracer density scale lengths requiring larger radial extent), butoverall there is no clear trend that an observation volume aroundthe Sun, above the disk or at smaller Galactocentric radii shouldgive remarkably better constraints on the potential than the othervolumes.

gular width of the wedges). We investigate these ratherunrealistic survey volumes to test (i) if there are regionsin the Galaxy where stars are on intrinsically more in-formative orbits and (ii) if spatial cuts applied to thesurvey volume (e.g., to avoid regions of large dust ex-tinction or measurement uncertainties) would thereforestrongly affect the precision of the potential constraints.To make this effect—if it exists—noticeable, we choosesome extreme, but illustrative examples.

The results are shown in Figure 6: The wedges allhave the same volume and all give results of similarprecision. There are some minor and expected differ-ences, e.g., vcirc(R) and radial scale lengths (b and adisk)are slightly better recovered for large radial extent and

Page 10: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

10 Trick et al.

the halo fraction at the Sun, fhalo, for volumes centeredaround R. In the case of an axisymmetric model galaxy,the extent in φ direction is not expected to matter. Over-all radial extent and vertical extent seem to be equallyimportant to constrain the potential. Figure 6 impliestherefore that volume offsets or spatial cuts of the sur-vey volume in the radial or vertical direction have at mosta modest impact—even in case of the very large samplesize at hand.

The second suite of mock data sets was already in-troduced in Section 3.1 (see also Test 3.2 in Table 3),where mock data sets were drawn from five spherical vol-umes around the Sun with different rmax, for two differ-ent stellar populations. The results of this second suiteare shown in Figure 5 and exemplify the effect of the sizeof the survey volume.

Figure 5 demonstrates that, given a choice of pDF, alarger volume always results in tighter constraints. Thereis no obvious trend that a hotter or cooler populationwill always give better results; it depends on the surveyvolume and the model parameter in question.

While it appears that the argument for significant ra-dial and vertical extent is generic, we have not done afull exploration of all combinations of pM and volumes.

That in reality different regions in the Galaxy havedifferent stellar number densities and different measure-ment uncertainties, should therefore be the major fac-tor to drive the precision of the potential recovery whenchoosing a survey volume.

3.3. Impact of misjudging the selection function of thedata set

The survey SF (see also Appendix B) is sometimes notperfectly determined. While the pattern of the surveyarea on the sky may be complex, it is usually preciselyknown. It is the uncertainties (e.g., in completeness) inthe line-of-sight directions that is most prone to system-atic misassessment. We therefore focus here on complete-ness misjudgements (with a simplified angular pattern)and investigate how much this could affect the recoveryof the potential. We do this by creating mock data inthe DHB-Pot within a spherical survey volume with ra-dius rmax around the Sun (see Test 5 in Table 3) and aspatially varying completeness function

completeness(r) ≡ 1− εrr

rmax, (17)

which drops linearly with distance r from the Sun. Thecompleteness function can be understood as the prob-ability of a star at distance r to be detected (see alsoEquation B1). In the RoadMapping analysis on the otherhand, we assume constant completeness (εr = 0). Theincompleteness parameter εr of the mock data quantifiestherefore by how much we misjudge the SF. This mocktest captures the relevant case of stars being less likelyto be observed (than assumed) the further away they are(e.g., due to unknown dust obscuration).

Figure 7 demonstrates that the potential recovery withRoadMapping is quite robust against somewhat wrongassumptions about the completeness of the data, i.e.,εr . 0.15 for the hot and εr . 0.2 for the cool pop-ulation. The cool population is more robust, because itis less affected by the SF misjudgement at high |z| thanthe hot population. Our simple model SF affects stars

226

228

230

232

v circ(R⊙)

[km

s−1

]

hot population cool population

2.8

3.0

3.2

3.4

3.6

adisk [

kpc]

0.0 0.2 0.4 0.6

incompleteness ǫr

−1.4

−1.2

−1.0

−0.8

−0.6

ln(h

R/8

kpc)

True parameter values

DHB−Pot potential

hot qDF

cool qDF

0.0 0.2 0.4 0.6

incompleteness ǫr

Figure 7. Impact of misjudging the completeness of the data onthe parameter recovery with RoadMapping. Each mock data setwas created with a different incompleteness parameter εr (shownon the x-axis, see Equation (17)). (The model parameters aregiven as Test 5 in Table 3.) The analysis however assumed thatall data sets had constant completeness within the survey volume(εr = 0). The violins show the full shape of the projected pdfsfor each model parameter, and the solid lines their true values.The RoadMapping method seems to be robust against modest de-viations between the true and the assumed data incompleteness.(The potential parameter fhalo and the other qDF parameters arerecovered to a comparable accuracy and are therefore not shownhere.)

at large and small radii in equal proportion. As long asthe misjudgement is small, the tracer scale length pa-rameter hR can still be reliable recovered, and with itthe potential.

We have also investigated several test suites using theIso-Pot. The recovery of vcirc(R) and the qDF pa-rameters at different εr is qualitatively and quantita-tively similar to Figure 7 for the DHB-Pot. The isochronescale length b however is recovered independently of εr—probably because rotation curve measurements in theplane alone, which are not affected by the SF cuts, givereliable constraints on b. When not including tangen-tial velocity measurements in the analysis (which is doneby marginalizing the likelihood in Equation (9) over vT ),the parameters are well recovered only for εr . 0.15 andεr . 0.2 for the hot and cool population respectively. Asthis is in concordance with our findings for the DHB-Pot,this result seems to be valid for different choices of po-tentials.

For spatial completeness functions varying with thedistance from the plane |z| only, the Iso-Pot potentialrecovery is similarly robust as long as vT measurementsare included.

3.4. Measurement uncertainties and their effect on theparameter recovery

Measurement uncertainties in proper motions and dis-tance dominate over uncertainties in position on the sky(RA, Dec) and line-of-sight velocity, which can be more

Page 11: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 11

0 1 2 3 4 5 6

δµ [mas yr−1 ]

0.10

0.15

0.20

0.25

0.30

0.35v c

irc(R⊙): SE [%]

0 1 2 3 4 5 6

δµ [mas yr−1 ]

3

4

5

6

7

8

Iso−P

ot b: SE [%]

hot, δµ>0

cool, δµ>0

hot, no δv

cool, no δv

Figure 8. Effect of proper motion uncertainties δµ on the pre-cision of potential parameter recovery for two stellar populationsof different kinematic temperature (see Test 6.1 in Table 3 for allmodel parameters). The relative standard error (SE) derived fromthe marginalized pdf for each model parameter was determined forprecise data sets without measurement uncertainties (solid lines,with dotted lines indicating the error) and for data sets affectedby different proper motion uncertainties δµ and δvlos = 2 km s−1

(data points with error bars), but no uncertainties in position. Theerrors come from taking the mean over several data sets.

−3−2−1

0123

bia

s [S

E]

=(b

est

fit

- t

rue v

alu

e)

/ SE

vcirc(R⊙)

−3−2−1

0123

bia

s [S

E]

Iso−Pot bUncertainties inproper motion δµ:

1 mas yr−1

2 mas yr−1

3 mas yr−1

4 mas yr−1

5 mas yr−1

0.20 0.25 0.30 0.35Standard Error (SE)[% of true value]

−202468

1012

bia

s [S

E]

3 4 5 6 7SE [%]

−6

−4

−2

0

2

bia

s [S

E]

Uncertainties inproper motion δµ &distance modulus δ(m−M):

1 mas yr−1 , 0.07 mag

2 mas yr−1 , 0.2 mag

3 mas yr−1 , 0.3 mag

4 mas yr−1 , 0.4 mag

5 mas yr−1 , 0.5 mag

Figure 9. Potential parameter recovery using the approxima-tion for the model probability convolved with measurement uncer-tainties in Equation (16). We show pdf offset and relative width(i.e., standard error SE) for potential parameters recovered frommock data sets (which were created according to Test 6.2 in Ta-ble 3). The data sets in the upper panels are affected only byproper motion uncertainties δµ (and δvlos = 2 mas yr−1), whilethe data sets in the lower panels also have distance (modulus)uncertainties δ(m − M), as indicated in the legend. For datasets with δµ ≤ 3 mas yr−1 Equation (16) was evaluated withNsamples = 800, for δµ > 3 mas yr−1 we used Nsamples = 1200.In absence of distance uncertainties Equation (16) gives unbiasedresults. For δ(m −M) > 0.2 mag (i.e., δr/r > 0.1; for r ∼ 3 kpc)however biases of several σ are introduced, as Equation (16) is onlyan approximation for the true likelihood in this case.

accurately determined.The range of proper motion uncertainties we will in-

vestigate in this section, 1 − 5 mas yr−1, is the approx-imate measurement accuracy that can be achieved bycombining catalogues from ground-based surveys like theSloan Digital Sky Survey (SDSS; Abazajian et al. 2003),the USNO-B catalogue (Monet et al. 2003), 2MASS(Skrutskie et al. 2006) and the Pan-STARRS1 photo-metric catalogue (PS1; Kaiser et al. 2010).9 Space-

9 Combining observations from the SDSS Data Release 1 withthe USNO-B catalogue based on the Palomar Observatory Sky Sur-vey’s (POSS) photographic plates from the 1950s lead to propermotion measurements precise to δµ ∼ 3 or 5 mas yr−1 depend-ing on magnitude r < 18 or r < 20 respectively (Munn et al.

228

230

232

234

236

238

v circ(R⊙)

[km

s−1

]

Iso−Pot parameters

1.0 1.5 2.0 2.5 3.0

true δµ [mas yr−1 ]

0.8

1.0

1.2

1.4

1.6

b [k

pc]

−2.0

−1.8

−1.6

−1.4

−1.2

ln(σ

z,0/

22

0km

s−1

)

qDF parameters

qDF of population &underestimation of δµhot & 10%

hot & 50%

cool & 10%

cool & 50%

1.0 1.5 2.0 2.5 3.0

true δµ [mas yr−1 ]

−0.2

−0.1

0.0

0.1

0.2

ln(h

σ,z/8

kpc)

true parameter values

Figure 10. Effect of a systematic underestimation of proper mo-tion uncertainties δµ on the recovery of the model parameters.(The true model parameters used to create the mock data aresummarized as Test 6.3 in Table 3, four of them are indicatedas black dotted lines in this figure.) The mock data was perturbedaccording to proper motion uncertainties δµ = δµDec = δµRA asindicated on the x-axis. In the RoadMapping analysis (see likeli-hood in Equation (16)) however, we underestimated the true δµby 10% (circles, solid lines) and 50% (triangles, dashed lines). Thesymbols denote the best fit parameters with 1σ error bars of sev-eral mock data sets. The lines connect the mean of correspondingdata realisations to guide the eye.

based surveys can do even better: The Hipparcos (ESA1997) and Tycho-2 (Høg et al. 2000) catalogues achieveδµ ∼ 2.5 mas yr−1 (and even δµ . 1 mas yr−1 for allstars with V < 12), which will be soon superseded byGaia with only δµ ∼ 0.3 mas yr−1 at its faint end atmagnitude G ∼ 20 (de Bruijne et al. 2014).

We first investigate the impact of (perfectly known)proper motion uncertainties on the precision of the po-tential parameter recovery (see Test 6.1 in Table 3). Fig-ure 8 demonstrates that for data sets with δµ as high as5 mas yr−1 the precision degrades by a factor of no morethan ∼ 2 as compared to a data set without measurementuncertainties. The precision gets monotonically betterfor smaller δµ, being larger only by a factor of ∼ 1.15 atδµ = 1 mas yr−1. With relative standard errors on therecovered parameters of only a few percent at most for10,000 stars, this means we still get quite precise con-straints on the potential, as long as we know the propermotion uncertainties perfectly.

We also note that in this case the relative and absolutedifference in recovered precision between the precise andthe uncertainty-affected data sets does not seem to de-pend strongly on the kinematic temperature of the stellarpopulation.

Secondly, we investigate the impact of additional mea-surement uncertainties in distance (modulus). In ab-sence of distance uncertainties the uncertainty-convolvedmodel probability given in Equation (16) is unbiased (see

2004, 2008; Gould & Kollmeier 2004). The same accuracy canbe achieved when using the four years of measurements by thePS1 only. By careful calibration of USNO-B and 2MASS withPS1, Sesar et al. (2015) even got proper motions as accurate asδµ ∼ 1.5 mas yr−1 for r . 18. The Large Synoptic SurveyTelescope (LSST, Ivezic et al. 2008) planned for 2021 might evenachieve δµ . 1 mas yr−1 during its 10 years of scanning the sky(Ivezic et al. 2008).

Page 12: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

12 Trick et al.

upper left panel in Figure 9). When including distance(modulus) uncertainties, Equation (16) is just an approx-imation for the true likelihood; the systematic bias thusintroduced in the parameter recovery gets larger with thesize of δ(m−M), as demonstrated in Figure 9, lower pan-els (see also Test 6.2 in Table 3). We find however thatin case of δ(m−M) . 0.2 mag (if also δµ . 2 mas yr−1

and a maximum distance of rmax = 3 kpc, see Test 6.2in Table 3) the potential parameters can still be recov-ered within 2σ. This corresponds to a relative distanceuncertainty of ∼ 10%. The overall precision of the po-tential recovery is also not degraded much by introducingdistance uncertainties of less than 10%.

How does this compare with the distance uncertain-ties expected for Gaia? For a typical red clump giantstar with MI ∼ 0 mag and (V − I) ∼ 1 mag at a dis-tance of r = 3 kpc we estimate (using the magnitudetransformation by Jordi et al. (2010) and the uncertaintyparametrization by de Bruijne et al. (2014)) a parallaxuncertainty of δπ ∼ 11 µas, which is consistent with adistance uncertainty of less than 5%, and a proper mo-tion uncertainty of δµ ∼ 6 µas yr−1, which is negligible.For the case that the modelling is not restricted to giantstars only, a quick investigation by Rene Andrae (privatecommunication) of stars at r = 3 kpc±5 pc from the GaiaUniverse model snapshot catalogue (GUMS; Robin et al.2012) revealed that a magnitude cut at G ∼ 15 mag inthe overall Gaia data set should keep all distance uncer-tainties within 3 kpc below ∼ 10%, while also preservingGaia’s simple SF.

We therefore found that in case we perfectly knowthe measurement uncertainties (and the distance uncer-tainty is negligible or of the order of the uncertaintiesexpected from Gaia within ∼ 3 kpc), the convolution ofthe model probability with the measurement uncertain-ties gives precise and accurate constraints on the modelparameters—even if the measurement uncertainty itselfis quite large.

Lastly, Figure 10 investigates the effect of a systematicunderestimation of the true proper motion uncertaintiesδµ by 10% and 50% (see also Test 6.3 in Table 3). Wefind that this causes a bias in the parameter recovery thatgrows seemingly linear with δµ. For an underestimationof only 10% however, the bias becomes . 2σ for 10,000stars—even for δµ ∼ 3 mas yr−1.

The size of the bias also depends on the kinematic tem-perature of the stellar population and the model param-eter considered (see Figure 10). The qDF parameters arefor example better recovered by hotter populations. Thisis, because the relative difference between the true σi(R)(with i ∈ R, z) and measured σi(R) (which comes fromthe deconvolution with an underestimated velocity un-certainty) is smaller for hotter populations.

3.5. The impact of deviations of the data from theidealized distribution function

Our modelling approach assumes that each stellar pop-ulation follows a simple DF; here we use the qDF. In thissection we explore what happens if this idealization doesnot hold. We investigate this issue by creating mock datasets that are drawn from two distinct qDFs of different

Exam

ple

1:

#st

ars

Mixed stellar populationsmock data set(mix of 2 pop.)

best fit(with single qDF)

Y1% of hot qDF +Y2% of cooler qDF

Y1 ,Y2=0,100

Y1 ,Y2=30,70

Y1 ,Y2=70,30

Y1 ,Y2=100,0

50 100

150

200

250

300

vT [km s−1 ]

Exam

ple

2a:

#st

ars

−20

0

−10

0 010

020

030

0

vz [km s−1 ]

50% of hot qDF +50% of qDFbeing X% colder

X=70

X=40

X=20

X=0

Figure 11. Distribution of mock data vT and vz created by mix-ing stars drawn from two different qDFs (solid lines), and the dis-tribution predicted by the best fit of a single qDF and potentialto the data (dotted lines). (The model parameters used to createthe mock data are given in Table 3 as Test 7, Example 1 & 2a,with the qDF parameters referred to in the legend given in Ta-ble 2.) The corresponding single qDF best fit curves were derivedfrom the best fit parameters found in Figures 13 and 14. (Thedata sets are colour-coded in the same way as the correspondinganalyses in Figures 13 and 14.) We use the mixtures of two qDFsto demonstrate how RoadMapping behaves for data sets followingDFs with shapes slightly differing from a single qDF. For large de-viations it might already become visible from directly comparingthe mock data and best fit distribution, that a single qDF is a badassumption for the stars’ true DF.

6 7 8 9 10

R [kpc]

−2

−1

0

1

2

z [k

pc]

−10

0

10

(D−M

)/√ M

Example 2a:50% of hot qDF +

50% of qDF being X=60% colder

−10 0 10

(D−M)/√M

−4

−2

0

2

4

resi

dual si

gnific

ance

(D−M

)/√ M

Figure 12. Residual significance (D −M)/√M of one example

mock data set D and its best fit single qDF model M in the (R, z)plane. The mock data set D was created by mixing a hot and aX% colder population in equal proportion (see also Table 3, Test7, Example 2a, with X = 60%). The best fit distribution M wasderived analogously to the ones shown for the velocity componentsin Figure 11. This is an extreme example where the best fit singleqDF is not a good fit anymore (see Figure 14, Example 2a, X =60%), but it illustrates how we constructed mock data distributionswith radial and vertical density profiles differing from a single qDFby mixing two different qDF populations for the Test suite 7.

Page 13: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 13

228

230

232

234

v circ(R⊙)

[km

s−1

]

Mixed stellar populationsExample 1

3.0

3.5

4.0

adisk [

kpc]

all hot 50% hot50% cooler

all cooler

amount of stars drawn from each qDF

−1.5

−1.0

−0.5

0.0

ln(h

R/8

kpc)

True parameter valuesof mix with different qDFs

DHB−Pot potential

hot qDF

cooler qDF

Figure 13. The dependence of the parameter recovery on degreeof pollution and temperature of the stellar population. We mix(i.e., “pollute”) varying amounts of stars from a hot stellar popu-lation with stars from a very different cooler population (see Table2), as indicated on the x-axis. (All model parameters used to cre-ate the mock data are given as Test 7, Example 1, in Table 3.)The composite polluted mock data set follows a true DF that hasa slightly different shape than the qDF. We then analyse it usingRoadMapping and fit a single qDF only. The violins represent themarginalized pdfs for the best fit model parameters. Some mockdata sets are shown in Figure 11, first row, in the same coloursas the violins here. We find that a hot population is much lessaffected by pollution with stars from a cooler population than viceversa. (The potential parameter fhalo is recovered to a similar oreven slightly better accuracy than adisk at each given mixing rateand is therefore not shown here.)

temperature10 (see Table 2 and Test 7 in Table 3) in theDHB-Pot, and analyse the composite mock data set by fit-ting a single qDF to it. The velocity distribution of somemock data sets and their best fit qDFs are illustrated inFigure 11, and Figure 12 shows the tracer density residu-als between data and best fit in the (R, z) plane. Figures13 and 14 compare the input and best fit parameters.In Example 1 we choose qDFs of widely different tem-perature and vary their relative fraction of stars in thecomposite mock data set (Figure 13); in Example 2 wealways mix mock data stars from two different qDFs inequal proportion, but vary by how much the qDFs’ tem-peratures differ (Figure 14).

The first set of tests mimics a DF that has wider wingsor a sharper core in velocity space than a qDF (see Fig-ure 11) and slightly different radial and vertical tracerdensity profiles (similar to Figure 12). The second testcould be understood as mixing neighbouring MAPs inthe [α/Fe]-vs.-[Fe/H] plane due to large bin sizes or abun-dance measurement errors (cf. BR13).

We consider the impact of the DF deviations on the

10 Following the observational evidence, our mock data popula-tions with cooler qDFs also have longer tracer scale lengths.

228

230

232

234

v circ(R⊙)

[km

s−1

]

Mixed stellar populationsExample 2a Example 2b

3.0

3.5

4.0

adisk [

kpc]

0% 20% 40% 60%X

difference in qDF parameters

−1.5

−1.0

−0.5

0.0

ln(h

R/8

kpc)

True parameter valuesof 50/50 mixture

hot qDF

X% colder qDF

DHB−Pot potential

0% 20% 40% 60%X

True parameter valuesof 50/50 mixture

cool qDF

X% warmer qDF

DHB−Pot potential

Figure 14. The dependence of the parameter recovery on thedifference in qDF parameters of a 50/50 mixture of two stellarpopulations and their temperature. The two qDFs from which thestars in each mock data set were drawn are indicated in the legend,with the qDF parameters σR,0, σz,0 and hR differing by X% (seealso Table 2 and Section 2.4), as indicated on the x-axis. (Themodel parameters used for the mock data creation are given asTest 7, Example 2a & b, in Table 3.) Each composite mock dataset is fitted with a single qDF and the marginalized pdfs are shownas violins. Some mock data sets of Example 2a and their best fitdistributions are shown in Figure 11, last row (colour-coded anal-ogous to the violins here), and Figure 12 shows the correspondingresiduals in the (R, z) plane. By mixing populations with varyingdifference in their qDF parameters, we model the effect of finite binsize or abundance errors when sorting stars into different MAPs inthe [α/Fe]-vs.-[Fe/H] plane and assuming they follow single qDFs(cf. BR13). We find that the bin sizes should be chosen such thatthe difference in qDF parameters between neighbouring MAPs isless than 20%. (The potential parameter fhalo is recovered to asimilar or even slightly better accuracy than adisk at each given Xand is therefore not shown here.)

recovery of the potential and the qDF parameters sepa-rately.

We find from Example 1 that the potential parameterscan be more robustly recovered, if a mock data popula-tion is polluted by a modest fraction (. 30%) of starsdrawn from a much cooler qDF, as opposed to the samepollution of stars from a hotter qDF. When consideringthe case of a 50/50 mix of contributions from differentqDFs in Example 2, there is a systematic, but mostlysmall, bias in recovering the potential parameters, mono-tonically increasing with the qDF parameter difference.In particular for fractional differences in the qDF param-eters of . 20% the systematics are insignificant even forsample sizes of N∗ = 20, 000, as used in the mock data.

Overall, the circular velocity at the Sun is very reliablyrecovered to within 2% in all these tests. But the best fitvcirc(R) is not always unbiased at the implied precision.

The recovery of the effective qDF parameters, in lightof non-qDF mock data, is quite intuitive (in Figures13 and 14 we therefore show only hR): the effective

Page 14: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

14 Trick et al.

qDF temperature lies between the two temperatures fromwhich the mixed DF of the mock data was drawn; in allcases the scale lengths of the velocity dispersion fall-off,hσ,R and hσ,z, are shorter than the true scale lengths,because the stars drawn form the hotter qDF dominateat small radii, while stars from the cooler qDF (with itslonger tracer scale length) dominate at large radii; the re-covered tracer scale lengths, hR, vary smoothly betweenthe input values of the two qDFs that entered the mixof mock data. The latter is also demonstrated in Figure12: The radial tracer density profile of the mock datais steeper than a single qDF in the mid-plane and moreshallow at higher |z|; overall the best fit hR lies thereforein between.

We note that in the cases where the systematic bias inthe potential parameter recovery becomes several σ large,a direct comparison of the true mock data set and best fitdistribution (see Figure 11) can sometimes already revealthat the assumed DF is not a good model for the data.

We performed the same tests also for the sphericalIso-Pot instead of the galaxy-like DHB-Pot and for amuch higher sampling of the mixing rate and qDF dif-ference X. The results are qualitatively and quantita-tively very similar and therefore independent of the exactchoice of potential.

Overall, we find that the potential inference is quiterobust to modest deviations of the data from the assumedDF.

3.6. The implications of a gravitational potential notfrom the space of model potentials

We now explore what happens when the mock datawere drawn from one axisymmetric potential family, hereMW14-Pot, and is then modelled considering potentialsfrom another axisymmetric family, here KKS-Pot (see Ta-ble 1 and Figure 1). In the analysis we assume the circu-lar velocity at the Sun to be fixed and known and onlyfit the parametric potential form.11

We analyse a mock data set from a hot and cool stellarpopulation each (see Test 8 in Table 3) with high numer-ical accuracy. The distributions generated from the bestfit parameters reproduce the data in configuration spacevery well (see Figure 15 for the spatial distribution andthe circles in Figure 17 for the velocity distribution).

The comparison between true and best fit potentialsare shown in Figure 16. We find that the potential re-covered by RoadMapping is in good agreement with thetrue potential inside of the observed volume of mock trac-ers. Outside of it we can make predictions at least to acertain extent. Especially the potential forces, to whichthe stellar orbits are sensitive, are recovered and tightlyconstrained. This robust recovery of the radial and ver-tical forces leads to small errors on the estimated cir-cular velocity curve (. 5%) and surface density within|z| = 1.1 kpc (. 10%), respectively. We get the bestresults for the local density, the surface density and disk-to-halo ratio between R ∼ 4 kpc and R ∼ 8 kpc, i.e.,where most of the tracer stars used in the analysis arelocated (see Figure 15).

11 We made sure that vcirc(R) can be very well recovered whenincluded in the fit of a cool population. The model assumptionthat vcirc(R) is known does therefore not affect the discussionqualitatively.

4 6 8 10 12 14

R [kpc]

# s

tars

Wrong potential model

mock datain MW14−Potbest fitwith KKS−Pot

−4 −

2 0 2 4

z [kpc]

stellar pop.hot

cool

Figure 15. Comparison of the spatial distribution of mock datain R and z created in the MW14-Pot potential and with two differentstellar populations (see Test 8 in Table 3 for all mock data modelparameters), and the best fit distribution recovered by fitting thefamily of KKS-Pot potentials to the data. The best fit potentials areshown in Figure 16 and the corresponding best fit qDF parametersin Figure 17. The data is very well recovered, even though thefitted potential family did not incorporate the true potential.

The local density distribution is in general less reliablyconstrained than the forces, but we still capture the es-sentials. Exceptions are the inner regions R . 3 kpc,where the KKS-Pot model is missing a bulge by con-struction, and the local radial density profile, which issomewhat misjudged by the KKS-Pot model. The coolpopulation, where most stars are confined to regions closeto the mid-plane, recovers the flatness of the disk betterthan the hot population, but overall the best fit disk isslightly less dense in the mid-plane than the true disk.While it is in general possible to generate very flatteneddensity distributions from Stackel potentials, it might bedifficult to simultaneously have a roundish halo and torequire that both Stackel components have the same fo-cal distance (see Table 1).

The disk-to-halo surface density fraction within |z| =1.1 kpc is not tightly constrained (& 20%), but recoveredwithin the errors inside of the survey volume. Using awrong potential model does therefore not necessarily leadto biases in local dark matter measurements.

Overplotted in Figure 16 is also the KKS-Pot with theparameters from Table 1, which were fixed based on a(by-eye) fit directly to the force field (within rmax =4 kpc from the Sun) and rotation curve of the MW14-Pot.The potential found with the RoadMapping analysis isan even better fit. This demonstrates that RoadMappingfitting infers a potential that in its actual properties re-sembles the input potential for the mock data in regionsof large tracer density as closely as possible, given thedifferences in functional forms.

Figure 17 compares the true qDF parameters with thebest fit qDF parameters belonging to the best fit po-tentials from Figure 16, and we also overplot the ac-tual physical scale lengths and velocity dispersion as es-timated directly from the mock data. While we recoverhR, σR,0 and hσ,R within the errors, we misjudge theparameters of the vertical velocity dispersion (σ0,z andespecially hσ,z), even though the actual mock data dis-tribution is well reproduced. This discrepancy could beconnected to the KKS-Pot not being able to reproducethe flatness of the disk. Also, σz and σR in Equations(6)-(7) are scaling profiles for the qDF (cf. BR13) andhow close they are to the actual velocity profile dependson the choice of potential; that is, the physical veloc-ity dispersion is well recovered, even if the qDF velocitydispersion parameters are not. Figure 17 stresses once

Page 15: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 15

2 4 6 8 10 12 14R [kpc]

−6

−4

−2

0

2

4

6

z [k

pc]

Radial force contours, FR (R,z)

2 4 6 8 10 12 14R [kpc]

−6

−4

−2

0

2

4

6

z [k

pc]

Vertical force contours, Fz (R,z)

2 4 6 8 10 12 14R [kpc]

−6

−4

−2

0

2

4

6

z [k

pc]

Density contours, ρΦ (R,z)

0 2 4 6 8 10 12 14 1610-3

10-2

10-1

100

101

ρΦ(R

,z=0)

[M⊙

pc−

3]

Radial local density at z=0

−4 −2 0 2 4

10-2

10-1

ρΦ(R

=8k

pc,z)

[M

⊙ pc−

3]

Vertical local density at R=8kpc

0 2 4 6 8 10 12 14 16R [kpc]

−50

0

50

diff. [

%]

−4 −2 0 2 4z [kpc]

−50

0

50

diff. [

%]

2 4 6 8 10 12 1480

100

120

140

160

180

200

220

240

v circ(R)

[km

s−1

]

vcirc(8 kpc) = 230 km s−1

fixed by construction

Circular velocity

2 4 6 8 10 12 14R [kpc]

−5

0

5

diff. [

%]

MW14−Pot, true potentialof mock dataobserved volume

best fit KKS−Pot shaped potentialfrom hot population

best fit KKS−Pot shaped potentialfrom cool population

KKS−Pot from Table 1(as reference)

2 4 6 8 10 12 14101

102

103

Σ(R

,|z|≤

1.1k

pc)

[M

⊙ pc−

2]

Surface density

2 4 6 8 10 12 140

2

4

6

8

10

12

14

Σdisk/Σ

halo(R

,|z|≤

1.1k

pc)

Disk-to-halo ratio

2 4 6 8 10 12 14R [kpc]

−15

0

15

diff. [

%]

2 4 6 8 10 12 14R [kpc]

−20

0

20

diff. [

%]

Figure 16. Recovery of the gravitational potential if the assumed potential model family (KKS-Pot with fixed vcirc(R)) and the truepotential of the (mock data) stars (MW14-Pot in Table 1) have slightly different parametric forms. In addition to contours of equal densityρΦ, radial and vertical force FR and Fz in the (R, z) plane (left comlumn), we show local density profiles ρΦ(R, z = 0) and ρΦ(R = 8 kpc, z),

as well as the circular velocity curve vcirc(R), the total surface density profile within |z| ≤ 1.1 kpc, Σ(R) ≡∫ 1.1kpc−1.1kpc ρΦ(R, z) dz, and the

ratio of the disk and halo contributions to the total surface density, Σdisk(R)/Σhalo(R). We compare the true potential (black lines) with100 sample potentials (red and blue lines) drawn from the pdf found with MCMC for a hot (red) and a cool (blue) stellar population andalso display the relative difference in % of the true value. (All mock data model parameters are given as Test 8 in Table 3.) Overall, thetrue potential is well recovered—especially in regions where most of the observed stars are located.

more that the actual parameter values of action-basedDFs have always to be considered together with the po-tential in which they were derived. This is of importancein studies that use a fiducial potential to fit action-basedDFs to stellar data, like, e.g., Sanders & Binney (2015)and Das & Binney (2016).

3.7. The influence of the stellar population’s kinematictemperature

Overall, we found that it does not make a big differenceif we use hot or cool stellar populations in our modelling.

How precise and reliable model parameters can be re-covered does to a certain extent depend on the kinematictemperature of the data, as well as on the model parame-

Page 16: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

16 Trick et al.

hot

cool

−1.4

−1.2

−1.0

−0.8

ln(h

R/8

kpc)

hot

cool

−1.7

−1.6

−1.5

−1.4

−1.3

ln(σ

R,0/2

20

km s−1

)

hot

cool

−2.2

−2.0

−1.8

−1.6

−1.4

−1.2

ln(σ

z,0/2

20

km s−1

)

hot

cool

0.0

0.1

0.2

0.3

0.4

ln(h

σ,R

/8kp

c)

hot

cool

−0.4

−0.2

0.0

0.2

ln(h

σ,z/8

kpc)

best fit qDFin best fit KKS−Pot

true qDF of hot pop.±5%, in true MW14−Pottrue qDF of cool pop.±5%, in true MW14−Pot

true physical scale lengthsand velocity dispersionestimated from mock data

physical scale lengthsand velocity dispersionof best fit distribution

Figure 17. Recovery of the qDF parameters for the case where the true and assumed potential deviate from each other (see Test 8in Table 3). The thick red (blue) lines represent the true qDF parameters of the hot (cool) qDF in Table 2 used to create the mockdata, surrounded by a 5% error region. The grey violins are the marginalized pdfs for the qDF parameters found simultaneously with thepotential constraints shown in Figure 16. We compare the qDF parameters with the actual physical scale lenghts and velocity dispersionat the Sun estimated from the mock data and the best fit distribution by fitting exponential functions to the data. Firstly this showsthat—apart from some small deviations in the velocity dispersion scale lengths—the velocity distribution of the mock data is very wellreproduced by the best fit. Secondly this demonstrates how the qDF parameters in different potentials do not necessarily agree with eachother or with the actual physical velocity distribution.

ter in question and on the observation volume. But thereis no easy rule of thumb, what combination would givethe best results (see Figure 5). There are two exceptions.

First, the circular velocity at the Sun, vcirc(R), isalways best recovered with cooler populations (see Fig-ures 8, 10, 13, 14 and for the recovery of vcirc(R) atR 6= R see Figure 16), because more stars are on near-circular orbits (see Figure A1). Cooler populations arealso less sensitive to misjudgements of (spatial) selectionfunctions at large |z| (see Figure 7). There is howeverthe caveat, that cool populations are more susceptible tonon-axisymmetric streaming motions in the disk.

Second, hotter populations seem to be less sensitiveto misjudgements of proper motion measurement uncer-tainties (see Figure 10) and pollution with stars from acooler population (see Figures 13 and 14), because oftheir higher intrinsic velocity dispersion (see Figure A2).

In addition we find indications in Figure 16, that differ-ent regions within the Galaxy are probed best by popu-lations of different kinematic temperature: The hot pop-ulation gives the best constraints on the radial local andsurface density profiles at a smaller radius than the coolpopulation because of its smaller tracer scale length. Thecool population with most stars close to the mid-planerecovers the flatness of the disk more reliably.

4. SUMMARY AND DISCUSSION

Recently implementations of action DF-based mod-elling of 6D data in the Galactic disk have been put forth,in part to lay the ground-work for Gaia (BR13; McMil-lan & Binney 2013; Piffl et al. 2014; Sanders & Binney2015).

We present RoadMapping, an improved implementa-tion of the dynamical modelling machinery of BR13, torecover the MW’s gravitational potential by fitting an or-bit DF to stellar populations within the Galactic disk. Inthis work we investigated the capabilities, strengths andweaknesses of RoadMapping by testing its robustnessagainst the breakdown of some of its assumptions—forwell-defined, isolated test cases using mock data. Over-all the method works very well and is robust, even whenthere are small deviations of the model assumptions fromthe “real” Galaxy.RoadMapping applies a full likelihood analysis and is

statistically well-behaved. It goes beyond BR13 by al-lowing for a straightforward and flexible implementationof different model families for potential and DF. It also

accounts for selection effects by using full 3D selectionfunctions (given some symmetries).

Computational speed: Large data sets in the age ofGaia require increasingly accurate likelihood evaluationsand flexible models. To be able to deal with thesecomputational demands, we sped up the RoadMappingcode by combining a nested-grid approach with MCMCand by faster action calculation using the Stackel(Binney 2012a) interpolation grid by Bovy (2015). Ourapproach therefore allows us to explore the full pdf ,while similar studies (e.g., Piffl et al. 2014; Sanders& Binney 2015; Das & Binney 2016) focus more onthe model with maximum likelihood only. This makesRoadMapping also slower: Fitting three DHB-Pot andfive qDF parameters in each of the analyses in Tests 5and 7 (see Table 3) takes for example ∼ 25 − 30 hourson 25 CPUs. This is still a feasible computational effortas long as we restrict ourselves to potentials with aclosed-form expression for Φ(R, z) (as done in this work).An equivalent analysis using, e.g., a double exponentialdisk (requiring integrals over Bessel functions) wouldtake several days to weeks for N∗ = 20, 000. In any case,the application of RoadMapping to millions of stars willbe a task for supercomputers and calls for even moreimprovements and speed-up in the fitting machinery.

Properties of the data set: We could show thatRoadMapping can provide potential and DF parameterestimates that are very accurate (i.e., unbiased) and pre-cise in the limit of large datasets, as long as the modellingassumptions are fulfilled.

In case the data set is affected by substantive measure-ment uncertainties, the potential can still be recovered tohigh precision, as long as these uncertainties are perfectlyknown and distance uncertainties are negligible. Forlarge proper motion uncertainties, e.g., δµ ∼ 5 mas yr−1,the formal errors on the parameters are only twice aslarge as in the case of no measurement uncertainties.However, properly accounting for measurement uncer-tainties is computationally expensive.

For the results to be accurate within 2σ (for 10,000stars), we need to know to within 10% both the truestellar distances (at rmax ≤ 3 kpc and δµ . 2 mas yr−1)and the true proper motion uncertainties (with δµ .3 mas yr−1).

Page 17: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 17

The distance condition is an artefact of the likelihoodapproximation (Equation (16)) that RoadMapping usesto save computation time, and the reason why we willhave to restrict the RoadMapping modelling to stars withsmall distance uncertainties.

Fortunately, the measurement uncertainties of the finalGaia data release with δµ . 0.3 mas yr−1 at G . 20 magand δr/r . 5% at r ∼ 3 kpc and for G < 15 mag (seeSection 3.4 and de Bruijne et al. 2014) will be well belowthese limits and promise accurate potential constraints.Before the final Gaia data release however we might haveto restrict the modelling to suitable giant tracers withsmall uncertainties.

The main caveat of Tests 2 and 6.1-6.3 in Section 3.4(see Table 3) concerning measurement uncertainties isthe use of the Iso-Pot, which we chose computationalspeed reasons. However, Tests 5 and 7, which we run forboth DHB-Pot and Iso-Pot, gave qualitatively and quan-titatively very similar results for both potentials. Thismakes us confident that also our results about measure-ment uncertainties are independent of the actual choiceof potential.

We also found that the location of the survey volumewithin the Galaxy matters little. At given sample size alarger survey volume with large coverage in both radialand vertical direction will give the tightest constraintson the model parameters.

The potential recovery with RoadMapping seems tobe robust against minor misjudgements of the spatialdata SF, in particular to a completeness overestimationof . 15−20% at the edge of a survey volume with rmax =3 kpc.

We found indications that populations of differentscale lengths and temperature probe different regions ofthe Galaxy, because the best potential constraints areachieved where most of the stellar tracers are located.This supports the approach by BR13, who measured foreach MAP the surface mass density only at one singlebest radius to account for missing flexibility in their po-tential model.

While cooler populations probe the Galaxy rotationcurve better and hotter populations are less sensitiveto pollution, overall stellar populations of differentkinematic temperature seem to be equally well-suitedfor dynamical modelling.

Deviations from the DF assumption: RoadMap-ping assumes that stellar sub-populations can be de-scribed by simple DFs. We investigated how much themodelling would be affected if the assumed family of DFswould differ from the stars’ true DF.

In Example 1 in Section 3.5 we considered true stel-lar DFs being (i) hot with more stars with low velocitiesand less stars at small radii than assumed (reddish datasets in Figure 11 and 13), or (ii) cool with broader ve-locity dispersion wings and less stars at large radii thanassumed (bluish data sets). We find that case (i) wouldgive more reliable results for the potential parameter re-covery.

Binning of stars into MAPs in [α/Fe] and [Fe/H], asdone by BR13, could introduce systematic errors due toabundance uncertainties or too large bin sizes—alwaysassuming MAPs follow simple DF families (e.g., theqDF). In Example 2 in Section 3.5 we found that, in

the case of 20,000 stars per bin, differences of . 20% inthe qDF parameters of two neighbouring bins can stillgive quite good constraints on the potential parameters.

The relative differences in the qDF parameters σR,0and σz,0 of neighbouring MAPs in Figure 6 of BR13(which have bin sizes of [Fe/H] = 0.1 dex and ∆[α/Fe] =0.05 dex) are indeed smaller than 20%. For the hR pa-rameter however the bin sizes in Figure 6 of BR13 mightnot yet be small enough to ensure no more than 20% ofdifference in neighbouring bins.

The qDF is a specific example for a simple DF forstellar sub-populations which we used in this paper. Butit is not essential for the RoadMapping approach. Futurestudies might apply slight alternatives or completelydifferent DFs to data.

Gravitational potential beyond the parametrizedfunctions considered: In addition to the DF,RoadMapping also assumes a parametric model for thegravitational potential. We test how using a potentialof Stackel form (KKS-Pot, Batsleer & Dejonghe 1994)affects the RoadMapping analysis of mock data from adifferent potential family with halo, bulge and exponen-tial disk (MW14-Pot, Bovy 2015). The potential recoveryis quite successful: We properly reproduce the mock datadistribution in configuration space; and the best fit po-tential is—within the limits of the model—as close as itgets to the true potential, even outside of the observationvolume of the stellar tracers.

For as many as 20,000 stars constraints become alreadyso tight that it should presumably be possible to distin-guish between different parametric MW potential models(e.g., the DHB-Pot and the KKS-Pot).

Fitting parametrized potentials of Stackel form toMW data (see, e.g., Batsleer & Dejonghe 1994; Famaey& Dejonghe 2003) has the advantage of allowing actioncalculations that are accurate and fast. It does howeverlimit the space of potentials that can be investigated, asdifferent potential components are all required to havethe same focal distance. Using the Stackel fudge (Binney2012a) together with parametrized potentials made upfrom physically motivated building blocks (exponentialdisks, power-law dark matter halo etc.), as was done byBR13, seems to be the most promising approach—eventhough there remain still several challenges concerningcomputational speed to be solved.

Different modelling approaches using action-based DFs: BR13 focussed on MAPs for a numberof reasons: First, they seem to permit simple DFs (Bovyet al. 2012a,b,c), i.e., approximately qDFs (Ting et al.2013). Second, all stars must orbit in the same potential.While each MAP can yield different DF parameters, itwill also provide a (statistically) independent estimate ofthe potential. This allows for a valuable cross-checkingreference. In some sense, the RoadMapping approachfocusses on constraining the potential, treating the DFparameters as nuisance parameters. That we were ableto show in this work that RoadMapping results are quiterobust to the form of the DF not being entirely correctmotivates this approach further.

Magorrian (2014) introduced a framework whichavoids specific parametrizations of action-based DFs andmarginalizes over all possible DFs to constrain the po-

Page 18: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

18 Trick et al.

tential. While this is the proper way to treat a nuisanceDF, it appears to be computationally very challenging.

For reasons of galaxy and chemical evolution, the DFproperties are astrophysically linked between differentMAPs (Sanders & Binney 2015). In its current imple-mentation, RoadMapping treats all MAPs as indepen-dent and does not exploit such correlations. Ultimately,the goal is to do a consistent chemodynamical modelthat simultaneously fits the potential and DF(J , [X/H])(where [X/H] is [Fe/H] and other elements either refer-enced to H or Fe, i.e., [X/H] denotes the whole abun-dance space) with a full likelihood analysis. This hasnot yet been attempted with RoadMapping, because thebehaviour is quite complex.

Since the first application of RoadMapping by BR13there have been two similar efforts to constrain theGalactic potential and/or orbit DF for the disk:

Piffl et al. (2014) fitted both potential and a f(J) togiant stars from the RAVE survey (Steinmetz et al. 2006)and the vertical stellar number density profiles in the diskby Juric et al. (2008). They did not include any chem-ical abundances in the modelling. Instead, they used asuperposition of action-based DFs to describe the over-all stellar distribution at once: a superposition of qDFsfor cohorts in the thin disk, a single qDF for the thickdisk stars and an additional DF for the halo stars. Tak-ing proper care of the selection function requires a fulllikelihood analysis, which is computationally expensive.Piffl et al. (2014) choose to circumvent this difficulty bydirectly fitting (a) histograms of the three velocity com-ponents in eight spatial bins to the velocity distributionpredicted by the DF and (b) the vertical density profilepredicted by the DF to the profiles by Juric et al. (2008).The vertical force profile of their best fit mass modelnicely agrees with the results from BR13 for R > 6.6 kpc.The disadvantage of their approach is, that by binningthe stars spatially, a lot of information is not used.

Sanders & Binney (2015) have focussed on understand-ing the abundance-dependence of the DF, relying on afiducial potential. They developed extended distribu-tion functions (eDF), i.e., functions of both actions andmetallicity for a superposition of thin and thick disk,each consisting of several cohorts described by qDFs, aDF for the halo, a functional form of the metallicity ofthe interstellar medium at the time of birth of the stars,and a simple prescription for radial migration. They ap-plied a full likelihood analysis accounting for selectioneffects and found a best fit for the eDF in the fixed fidu-cial potential by Dehnen & Binney (1998) to the stel-lar phase-space data of the Geneva-Copenhagen Survey(Nordstrom et al. 2004; Holmberg et al. 2009), metallic-ity determinations by Casagrande et al. (2011) and thestellar density curves by Gilmore & Reid (1983). Theirbest fit predicted the velocity distribution of SEGUE G-dwarfs (Ahn et al. 2014) quite well, but had biases in themetallicity distribution, which they accounted to beinga problem with the SEGUE metallicities.

Das & Binney (2016) proceeded recently in a similarfashion to constrain an eDF for halo stars.

Future work: We know that real galaxies, includingthe MW, are not axisymmetric. Using N-body mod-els, we will explore in a subsequent paper how the re-covery of the gravitational potential with RoadMapping

will be affected when data from a non-axisymmetric diskgalaxy system with spiral arms get interpreted throughaxisymmetric models. There are several interesting sci-entific questions for which a RoadMapping investigationof galaxy simulations could be a pragmatic approach toaddress them: (i) What is the influence of spiral armsand resonances on the modelling outcome? (ii) Can werecover the potential well enough to calculate actions soaccurate that clumps in orbit space can be identified?This is important to be able to compare clumps in ac-tion space to clustering of stars in abundance space. (iii)How do results from RoadMapping, i.e., the potentialand DF, compare with results from Jeans models?

5. ACKNOWLEDGEMENTS

We thank Glenn van de Ven for suggesting the useof Kuzmin-Kutuzov Stackel potentials in this case study,the anonymous referee for her/his very helpful commentsand suggestions, James J. Binney and Payel Das (Uni-versity of Oxford) for valuable discussions, and ReneAndrae (MPIA) for an estimate of Gaia distance un-certainties. W.H.T. and H.-W.R. acknowledge fundingfrom the European Research Council under the Euro-pean Union’s Seventh Framework Programme (FP 7)ERC Grant Agreement n. [321035]. J.B. acknowledgesthe financial support from the Natural Sciences and En-gineering Research Council of Canada.

REFERENCES

Abazajian, K., Adelman-McCarthy, J. K., Agueros, M. A., et al.2003, AJ, 126, 2081

Ahn, C. P., Alexandroff, R., Allende Prieto, C., et al. 2014, ApJS,211, 17

Batsleer, P., & Dejonghe, H. 1994, A&A, 287, 43Binney, J. 2010, MNRAS, 401, 2318—. 2011, Pramana, 77, 39—. 2012a, MNRAS, 426, 1324—. 2012b, MNRAS, 426, 1328—. 2013, NewAR, 57, 29Binney, J., & McMillan, P. 2011, MNRAS, 413, 1889Binney, J., & Tremaine, S. 2008, Galactic Dynamics: Second

Edition (Princeton University Press)Bovy, J. 2015, ApJS, 216, 29Bovy, J., & Rix, H.-W. 2013, ApJ, 779, 115 (BR13)Bovy, J., Rix, H.-W., Green, G. M., Schlafly, E. F., & Finkbeiner,

D. P. 2016, ApJ, 818, 130Bovy, J., Rix, H.-W., & Hogg, D. W. 2012a, ApJ, 751, 131Bovy, J., Rix, H.-W., Hogg, D. W., et al. 2012b, ApJ, 755, 115Bovy, J., Rix, H.-W., Liu, C., et al. 2012c, ApJ, 753, 148Bovy, J., & Tremaine, S. 2012, ApJ, 756, 89Budenbender, A., van de Ven, G., & Watkins, L. L. 2015,

MNRAS, 452, 956Casagrande, L., Schonrich, R., Asplund, M., et al. 2011, A&A,

530, A138Das, P., & Binney, J. 2016, MNRAS, arXiv:1603.09332de Bruijne, J. H. J., Rygl, K. L. J., & Antoja, T. 2014, in EAS

Publications Series, Vol. 67, EAS Publications Series, 23–29de Lorenzi, F., Debattista, V. P., Gerhard, O., & Sambhus, N.

2007, MNRAS, 376, 71de Zeeuw, T. 1985, MNRAS, 216, 273Dehnen, W., & Binney, J. 1998, MNRAS, 294, 429ESA, ed. 1997, ESA Special Publication, Vol. 1200, The

HIPPARCOS and TYCHO catalogues. Astrometric andphotometric star catalogues derived from the ESAHIPPARCOS Space Astrometry Mission

Famaey, B. 2012, in SF2A-2012: Proceedings of the Annualmeeting of the French Society of Astronomy and Astrophysics,ed. S. Boissier, P. de Laverny, N. Nardetto, R. Samadi,D. Valls-Gabaud, & H. Wozniak, 15–24

Famaey, B., & Dejonghe, H. 2003, MNRAS, 340, 752

Page 19: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 19

Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J.2013, PASP, 125, 306

Garbari, S., Liu, C., Read, J. I., & Lake, G. 2012, MNRAS, 425,1445

Gilmore, G., & Reid, N. 1983, MNRAS, 202, 1025Gould, A., & Kollmeier, J. A. 2004, ApJS, 152, 103Henon, M. 1959, Annales d’Astrophysique, 22, 126Høg, E., Fabricius, C., Makarov, V. V., et al. 2000, A&A, 355, L27Holmberg, J., Nordstrom, B., & Andersen, J. 2009, A&A, 501, 941Hunt, J. A. S., & Kawata, D. 2014, MNRAS, 443, 2112Ivezic, Z., Monet, D. G., Bond, N., et al. 2008, in IAU

Symposium, Vol. 248, A Giant Step: from Milli- toMicro-arcsecond Astrometry, ed. W. J. Jin, I. Platais, &M. A. C. Perryman, 537–543

Ivezic, Z., Tyson, J. A., Abel, B., et al. 2008, ArXiv e-prints,arXiv:0805.2366

Jordi, C., Gebran, M., Carrasco, J. M., et al. 2010, A&A, 523,A48

Juric, M., Ivezic, Z., Brooks, A., et al. 2008, ApJ, 673, 864Kaiser, N., Burgett, W., Chambers, K., et al. 2010, in Proc. SPIE,

Vol. 7733, Ground-based and Airborne Telescopes III, 77330EKuijken, K., & Gilmore, G. 1989, MNRAS, 239, 605Magorrian, J. 2014, MNRAS, 437, 2230McMillan, P. J., & Binney, J. 2012, MNRAS, 419, 2251McMillan, P. J., & Binney, J. J. 2008, MNRAS, 390, 429—. 2013, MNRAS, 433, 1411Monet, D. G., Levine, S. E., Canzian, B., et al. 2003, AJ, 125, 984Munn, J. A., Monet, D. G., Levine, S. E., et al. 2004, AJ, 127,

3034

—. 2008, AJ, 136, 895Nordstrom, B., Mayor, M., Andersen, J., et al. 2004, A&A, 418,

989Perryman, M. A. C., de Boer, K. S., Gilmore, G., et al. 2001,

A&A, 369, 339Piffl, T., Binney, J., McMillan, P. J., et al. 2014, MNRAS, 445,

3133Read, J. I. 2014, Journal of Physics G Nuclear Physics, 41, 063101Reid, M. J., Menten, K. M., Zheng, X. W., et al. 2009, ApJ, 700,

137Rix, H.-W., & Bovy, J. 2013, A&A Rev., 21, 61Robin, A. C., Luri, X., Reyle, C., et al. 2012, A&A, 543, A100Sanders, J. L., & Binney, J. 2015, MNRAS, 449, 3479—. 2016, MNRAS, 457, 2107Sesar, B., Bovy, J., Bernard, E. J., et al. 2015, ApJ, 809, 59Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ, 131,

1163Steinmetz, M., Zwitter, T., Siebert, A., et al. 2006, AJ, 132, 1645Strigari, L. E. 2013, Phys. Rep., 531, 1Syer, D., & Tremaine, S. 1996, MNRAS, 282, 223Ting, Y.-S., Rix, H.-W., Bovy, J., & van de Ven, G. 2013,

MNRAS, 434, 652Yanny, B., Rockosi, C., Newberg, H. J., et al. 2009, AJ, 137, 4377Zhang, L., Rix, H.-W., van de Ven, G., et al. 2013, ApJ, 772, 108

Page 20: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

20 Trick et al.Table

3S

um

mary

of

test

suit

esin

this

work

:T

he

firs

tco

lum

nin

dic

ate

sth

ete

stsu

ite,

the

seco

nd

colu

mn

the

pote

nti

al,

DF

an

dS

Fm

od

el,

etc.

,u

sed

for

the

mock

data

crea

tion

,th

eth

ird

colu

mn

the

corr

esp

on

din

gm

od

elass

um

edin

the

RoadMapping

an

aly

sis,

an

dth

ela

stco

lum

nlist

sth

efi

gu

res

bel

on

gin

gto

the

test

suit

ean

dsu

mm

ari

zes

the

resu

lts.

Ref

eren

cep

ote

nti

als

an

dqD

Fs

are

intr

od

uce

din

Tab

les

1an

d2,

resp

ecti

vel

y.P

ara

met

ers

that

are

not

left

free

inth

ean

aly

sis,

are

alw

ays

fixed

toth

eir

tru

evalu

e.U

nle

ssst

ate

doth

erw

ise,

all

mock

data

sets

have

SF

sw

ith

com

ple

ten

ess(x

)=

1an

dn

om

easu

rem

ent

un

cert

ain

ties

,an

dw

eu

seNx

=16,N

v=

24,nσ

=5

as

nu

mer

ical

acc

ura

cyfo

rca

lcu

lati

ng

the

likel

ihood

norm

alisa

tion

(see

Sec

tion

2.6

).

Tes

tM

od

elfo

rM

ock

Data

Mod

elin

An

aly

sis

Fig

ure

s&

Res

ult

s

Tes

t1

:Potential:

DHB-Pot

-F

igu

re2

Nu

mer

ical

acc

ura

cyDF:

hot

orcool

qD

FS

uit

ab

leacc

ura

cyfo

rou

rte

sts:

inca

lcu

lati

ng

Survey

volume:

sph

ere

aro

un

dS

un

,r m

ax

=0.2,1,2,3

or

4kp

cNx

=16,N

v=

24,nσ

=5.

the

likel

ihood

Numericalaccuracy

:Nx∈

[5,3

2],Nv∈

[4,4

8],nσ∈

[3,7

]H

igh

ersp

ati

al

reso

luti

on

isn

orm

alisa

tion

requ

ired

for

coole

rp

op

ula

tion

s.T

est

2:

Potential:

Iso-Pot

Iso-Pot,

all

para

met

ers

free

Fig

ure

3N

um

eric

al

conver

gen

ceDF:

hot

qD

FqD

F,

all

para

met

ers

free

Th

enu

mb

erof

requ

ired

MC

sam

ple

sof

convolu

tion

Survey

Volume:

sph

ere

aro

un

dS

un

,r m

ax

=3

kp

c(fi

xed

&kn

ow

n)

scale

sasN

sam

ple

s∝

(δvm

ax)2

,w

ith

mea

sure

men

tUncertainties:

δR

A=δD

ec=δ(m−M

)=

0(fi

xed

&kn

ow

n)

wit

hδvm

ax

bei

ng

the

larg

estδv

un

cert

ain

ties

δvlo

s=

2km

s−1

inth

ed

ata

set.

IfN

∗is

smaller

,δµ

=2,3

,4or

5m

as

yr−

1le

ssacc

ura

cy,

i.e.

,le

ssN

sam

les,

Numericalaccuracy

:N

sam

ple

s∈

[25,1

200]

isn

eed

edto

reach

agiv

enN

∗:

10,0

00

acc

ura

cy.

Tes

t3.1

:Potential:

Iso-Pot

Iso-Pot,

all

para

met

ers

free

Fig

ure

4S

hap

eof

the

mod

elDF:

hot

qD

FqD

F,

all

para

met

ers

free

Th

epdf

bec

om

esa

mu

ltiv

ari

ate

para

met

ers’pdf

Survey

Volume:

sph

ere

aro

un

dS

un

,r m

ax

=2

kp

c(fi

xed

&kn

ow

n)

Gau

ssia

nin

the

lim

itof

for

larg

ed

ata

sets

N∗:

20,0

00

larg

ed

ata

.T

he

wid

thof

the

pdf

scale

sas

1/√N

∗.

Tes

t3.2

:Potential:

Iso-Pot

Iso-Pot,

free

para

met

er:b

Fig

ure

5P

ara

met

eres

tim

ate

sDF:

hot

orcool

qD

FqD

F,

free

para

met

ers:

RoadMapping

beh

aves

like

an

are

unb

iase

d;

lnhR,l

nσz,0,l

nhσ,z

unbia

sed

maxim

um

likel

ihood

Infl

uen

ceof

surv

eySurvey

volume:

sph

ere

aro

un

dS

un

,r m

ax

=0.2,1,2,3

or

4kp

c(fi

xed

&kn

ow

n)

esti

mato

r.L

arg

ersu

rvey

volu

mes

volu

me

size

N∗:

20,0

00

lead

toti

ghte

rco

nst

rain

ts,

even

at

the

sam

eN

∗.

Tes

t4

:Potential:

(i)Iso-Pot

or

(ii)

DHB-Pot

(i)Iso-Pot,

all

para

met

ers

free

Fig

ure

6In

flu

ence

of

(ii)

DHB-Pot,

free

para

met

ers:

Th

eex

act

posi

tion

&sh

ap

eof

the

posi

tion

&sh

ap

evcir

c(R

),a

dis

k,fhalo

surv

eyvolu

me

pla

ys

on

lya

of

surv

eyvolu

me

DF:

hot

qD

FqD

F,

all

para

met

ers

free

min

or

role

.H

avin

gb

oth

larg

eon

para

met

erre

cover

ySurvey

volume:

4d

iffer

ent

wed

ges

,se

eF

igu

re6,

up

per

pan

el(fi

xed

&kn

ow

n)

rad

ial

an

dver

tica

lex

tent

shou

ldN

∗:

20,0

00

giv

eth

eti

ghte

stco

nst

rain

ts.

Tes

t5

:Potential:

DHB-Pot

DHB-Pot,

free

para

met

ers:

Fig

ure

s7

Infl

uen

ceof

vcir

c(R

),a

dis

k,f

halo

For

min

or

mis

jud

gem

ents

wro

ng

ass

um

pti

on

sDF:

hot

orcool

qD

FqD

F,

all

para

met

ers

free

of

ara

dia

lly

sym

met

ric

SF

ab

ou

tth

esp

ati

al

SF

Survey

volume:

sph

ere

aro

un

dS

un

,r m

ax

=3

kp

c(fi

xed

&kn

ow

n)

(i.e

.,ε r

.0.1

5fo

rhot

an

don

para

met

erre

cover

yCompleteness:

Equ

ati

on

(17)

com

ple

ten

ess(x

)=

1,

ε r.

0.2

forcool

pop

ula

tion

s)w

ithε r∈

[0,0.7

]i.e.

,ε r

=0

the

pote

nti

al

reco

ver

yN

∗:

20,0

00

isst

ill

rob

ust

.T

est

6.1

:Potential:

Iso-Pot

Iso-Pot,

all

para

met

ers

free

Fig

ure

8E

ffec

tof

pro

per

moti

on

DF:

hot

orcool

qD

FqD

F,

all

para

met

ers

free

Ifδµ

isp

erfe

ctly

kn

ow

n,

the

un

cert

ain

ties

on

Survey

volume:

sph

ere

aro

un

dS

un

,r m

ax

=3

kp

c(fi

xed

&kn

ow

n)

pre

cisi

on

of

the

pote

nti

al

para

-p

reci

sion

of

Uncertainties:

(i)δR

A=δD

ec=δ(m−M

)=

0,

(fixed

&kn

ow

n)

met

erre

cover

yis

on

lya

pote

nti

al

reco

ver

yδvlo

s=

2km

s−1,

fact

or∼

1.1

5−

2w

ors

eδµ

=1,2,3,4

or

5m

as

yr−

1fo

rδµ∼

1−

5m

as

yr−

1

(ii)

no

mea

sure

men

tu

nce

rtain

ties

than

for

ad

ata

set

wit

hou

tN

∗:

10,0

00

pro

per

moti

on

un

cert

ain

ties

.

Page 21: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 21Table

3—

Continued

Tes

tM

od

elfo

rM

ock

Data

Mod

elin

An

aly

sis

Fig

ure

s&

Res

ult

s

Tes

t6.2

:Potential:

Iso-Pot

Iso-Pot,

all

para

met

ers

free

Fig

ure

9T

esti

ng

the

convolu

tion

DF:

hot

qD

FqD

F,

all

para

met

ers

free

Th

eap

pro

xim

ate

likel

ihood

wit

hm

easu

rem

ent

Survey

Volume:

sph

ere

aro

un

dS

un

,r m

ax

=3

kp

c(fi

xed

&kn

ow

n)

inE

qu

ati

on

(16)

isth

etr

ue

un

cert

ain

ties

inE

qu

ati

on

(16)

Uncertainties:

δR

A=δD

ec=

0,

(fixed

&kn

ow

n)

likel

ihood

inth

eab

sen

ceof

wit

h&

wit

hou

tδvlo

s=

2km

s−1,

posi

tion

erro

rs.

For

dis

tan

ced

ista

nce

un

cert

ain

ties

δµ

=1,2

,3,4

or

5m

as

yr−

1,

un

cert

ain

ties

.10%

at

(i)δ(m−M

)=

0or

r max∼

3kp

c,w

hic

his

the

case

(ii)δ(m−M

)6=

0(s

eeF

igu

re9)

for

Gaia

an

dG

.15,

the

intr

o-

N∗:

10,0

00

du

ced

bia

sis

still

less

than

.T

est

6.3

:Potential:

Iso-Pot

Iso-Pot,

all

para

met

ers

free

Fig

ure

10

Un

der

esti

mati

on

DF:

hot

orcool

qD

FqD

F,

all

para

met

ers

free

Th

ebia

sin

trod

uce

dby

un

der

-of

pro

per

moti

on

Survey

volume:

sph

ere

aro

un

dS

un

,r m

ax

=3

kp

c(fi

xed

&kn

ow

n)

esti

mati

ngδµ

gro

ws

wit

hδµ

.u

nce

rtain

ties

Uncertainties:

on

lyp

rop

erm

oti

on

un

cert

ain

ties

pro

per

moti

on

un

cert

ain

ties

Ifδµ.

3m

as

yr−

1an

dδµ

1,

2or

3m

as

yr−

110%

or

50%

un

der

esti

mate

dis

un

der

esti

mate

dby

10%

N∗:

10,0

00

the

bia

sis

on

ly∼

or

less

.T

est

7:

Potential:

DHB-Pot

DHB-Pot,

free

para

met

ers:

Fig

ure

s11,

12,

13

&14

Dev

iati

on

sof

the

vcir

c(R

),a

dis

k,f

halo

Exa

mple

1:

ass

um

edD

FDF:

mix

of

two

qD

Fs.

..si

ngle

qD

F,

all

para

met

ers

free

Hot

qD

F-l

ike

pop

ula

tion

sp

ollu

ted

from

the

(i)Exa

mple

1:

by

up

to∼

30%

of

stars

from

am

uch

stars

’tr

ue

DF

...

wit

hd

iffer

ent

mix

ing

rate

sco

ole

rp

op

ula

tion

giv

est

ill

reliab

lean

dfi

xed

qD

Fp

ara

met

ers

pote

nti

al

con

stra

ints

.T

his

isn

ot

(hot

&cooler

qD

Ffr

om

Tab

le2)

tru

efo

rth

eop

posi

teca

se.

(ii)

Exa

mple

2:

Exa

mple

2:

...

wit

h50/50

mix

ing

rate

Diff

eren

ces

of.

20%

inqD

Fan

dvary

ing

qD

Fp

ara

met

ers

(byX

%):

para

met

ers

do

not

matt

era)hot

&colder

qD

For

wh

enm

ixin

gsu

b-p

op

ula

tion

s,b)

cool

&warmer

qD

F(s

eeT

ab

le2)

e.g.,

du

eto

fin

ite

bin

nin

gin

Survey

volume:

sph

ere

aro

un

dS

un

,r m

ax

=2

kp

c(fi

xed

&kn

ow

n)

ab

un

dan

cesp

ace

and

ab

un

dan

ceN

∗:

20,0

00

erro

rs.

Tes

t8

:Potential:

MW14-Pot

KKS-Pot,

all

para

met

ers

free

,F

igu

res

15,

16

&17

Dev

iati

on

sof

the

on

lyvcir

c(R

)

=230

km

s−1

fixed

We

fin

da

good

ap

pro

xim

ati

on

ass

um

edp

ote

nti

al

mod

elDF:

hot

orcool

qD

FqD

F,

all

para

met

ers

free

for

the

tru

ep

ote

nti

al,

esp

ecia

lly

from

the

stars

’Survey

volume:

sph

ere

aro

un

dS

un

,r m

ax

=4

kp

c(fi

xed

&kn

ow

n)

wh

ere

most

stars

are

loca

ted

,tr

ue

pote

nti

al

N∗:

20,0

00

giv

enth

elim

itati

on

sof

the

wro

ng

Numericalaccuracy

:Nx

=20,N

v=

28,nσ

=5.5

ass

um

edp

ote

nti

al

mod

elfa

mily.

Page 22: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

22 Trick et al.

APPENDIX

A. MOCK DATA

The mock data in this work is generated according to the following procedure:We assume that the positions and velocities of our stellar mock sample are indeed drawn from our assumed family of

potentials (Section 2.3) and DFs (Section 2.4), with given parameters pΦ and pDF. The DF is in terms of actions, while

the transformation (xi,vi)Φ−→ J i is computationally much less expensive than its inversion. We therefore employ the

following efficient two-step method for creating mock data, which also accounts for a spatial survey selection functionSF(x) see Appendix B.

In the first step we draw stellar positions xi. We start by setting up the interpolation grid for the tracer densityρ(R, |z| | pΦ, pDF) generated according to Section 2.4.12 Next, we sample random positions (Ri, zi, φi) uniformlywithin the observable volume. Using a Monte Carlo rejection method we then shape the samples distribution to followρ(R, |z| | pΦ, pDF). To apply a non-uniform completeness function, we use the rejection method a second time. Theresulting set of positions xi follows the distribution p(x) ∝ ρDF(R, |z| | pΦ, pDF)× SF(x).

In the second step we draw velocities vi. For each of the positions (Ri, zi) we first sample velocities from a Gaussianenvelope function in velocity space which is then shaped towards DF(J [Ri, zi,v | pΦ] | pDF) using a rejection method.We now have a mock data set satisfying (xi,vi) −→ p(x,v) ∝ DF(J [x,v | pΦ] | pDF)× SF(x).

Measurement uncertainties can be added to the mock data by applying the following modifications to the aboveprocedure. We assume Gaussian uncertainties in the heliocentric phase-space coordinates x = (RA,Dec, (m−M)), v =(µRA · cos Dec, µDec, vlos) (see Section 2.1). In the case of distance and position uncertainties stars virtually scatterin and out of the observed volume. To account for this, we draw the true xi from a volume that is larger than theactual observation volume, perturb the xi according to the position uncertainties and then reject all stars that lienow outside of the observed volume. This mirrors the random scatter around the detection threshold for stars whosedistances are determined from the apparent brightness and the distance modulus. We then sample true vi (given thetrue xi) as described above and perturb them according to the velocity uncertainties.

2 4 6 8 10 12

Lz [220 km s−1 kpc]

0.2

0.4

0.6

0.8

1.0

1.2

JR

[2

20

km

s−1

kpc]

contours of constantstellar number densityin action space,containing 80%of mock data stars

contours of constantstellar number densityin action space,containing 80%of mock data stars

contours of constantstellar number densityin action space,containing 80%of mock data stars

contours of constantstellar number densityin action space,containing 80%of mock data stars

Mock Data in Action Space (2D)

2 4 6 8 10 12

Lz [220 km s−1 kpc]

0.2

0.4

0.6

0.8

1.0

1.2

Jz [

22

0 k

m s−1

kpc]

0.0 0.2 0.4 0.6 0.8 1.0 1.2

JR [220 km s−1 kpc]

0.2

0.4

0.6

0.8

1.0

1.2

Jz [

22

0 k

m s−1

kpc]

hot stellar population

cool stellar population

large volume

small volume above plane

4 6 8 10 12R [kpc]

−4

−2

0

2

4

z [k

pc]

SunGalactic

plane

20,000 stars in both volumes,angular extent: ∆φ= ±20

SunGalactic

plane

20,000 stars in both volumes,angular extent: ∆φ= ±20

SunGalactic

plane

20,000 stars in both volumes,angular extent: ∆φ= ±20

SunGalactic

plane

20,000 stars in both volumes,angular extent: ∆φ= ±20

Shape of Observed Volume

Figure A1. Distribution of mock data in action space (2D iso-density contours, enclosing 80% of the stars), depending on shapeand position of a wedge-like survey observation volume (upper leftpanel, see also Appendix B) and temperature of the stellar pop-ulation (indicated in the legend). The four mock data sets aregenerated in the KKS-Pot from Table 1 from either the hot or coolDF in Table 2. The distribution in action space visualizes howorbits with different actions reach into different regions within theGalaxy. The corresponding mock data in configuration space isshown in Figure A2.

4 5 6 7 8 9 10 11 12

R [kpc]

Mock Data Distribution (1D)... in Position

-4 -3 -2 -1 0 1 2 3 4 5

z [kpc]

-20 -15 -10 -5 0 5 10 15 20

φ [ ]

hot stellar population

cool stellar population

large volume

small volume above plane

-200 -150 -100 -50 0 50 100 150 200

vR [km s−1 ]

... in Velocity

-200 -150 -100 -50 0 50 100 150 200

vz [km s−1 ]

0 50 100 150 200 250 300 350

vT [km s−1 ]

Figure A2. Distribution of the mock data from Figure A1 in con-figuration space. The corresponding observation volumes (as indi-cated in the legend) are shown in Figure A1, upper left panel. The1D histograms illustrate that qDFs generate realistic stellar dis-tributions in Galactocentric coordinates (R, z, φ, vR, vz , vT ): Morestars are found at smaller R and |z|, and are distributed uniformlyin φ according to our assumption of axisymmetry. The distributionin radial and vertical velocities, vR and vz , is approximately Gaus-sian with the (total projected) velocity dispersion being of the orderof ∼ σR,0 and ∼ σz,0 (see Table 2). The distribution of tangentialvelocities vT is skewed because of asymmetric drift.

12 For the creation of the mock data we use Nx = 20, Nv = 40 and nσ = 5 in Equation (8).

Page 23: RoadMapping RoadMap- ping arXiv:1605.08601v1 [astro-ph.GA ... · test cases with idealized mock data how the breakdown of model assumptions and data properties a ect constraints on

Action-based Dynamical Modelling for the Milky Way Disk 23

We show examples of mock data sets (without measurement uncertainties) in action space (Figure A1) and con-figuration space (x,v) (Figure A2). The mock data generated from the qDF follow the expected distributions inconfiguration space. The distribution in action space illustrates the intuitive physical meaning of actions: The starsof the cool population have in general lower radial and vertical actions, as they are on more circular orbits. Circularorbits with JR = 0 and Jz = 0 can only be observed in the Galactic mid-plane. The different ranges of angularmomentum Lz in the two example observation volumes reflect Lz ∼ R× vcirc and the volumes’ different radial extent.The volume at larger z contains stars with higher Jz. An orbit with Lz or Lz(R) can only reach into a volumeat ∼ R, if it is more eccentric and has therefore larger JR. This together with the effect of asymmetric drift explainsthe asymmetric distribution of JR vs. Lz in Figure A1.

B. SELECTION FUNCTIONS

Any survey’s selection function (SF) can be understood as defining an effective sample sub-volume in the spaceof observables, e.g., position on the sky (limited by the pointing of the survey), distance from the Sun (limited bybrightness and detector sensitivity), colors and metallicity of the stars (limited by survey mode and targeting). The SFcan therefore be thought of as having both spatial small scale structure (due to pencil beam pointing, dust obscuration,etc.) and some overall spatial characteristics (e.g., mean height above the plane and mean Galactocentric radius of thestars). The treatment of realistic and complex SFs was already demonstrated in BR13 (who used the pencil-beam SFof the SEGUE survey (Bovy et al. 2012c)) and Bovy et al. (2016) (who investigated the effect of dust extinction). Inthis work we aim to make a generic and basic exploration of search volume shapes and, as shown by Bovy et al. (2016),this should be possible without explicitly considering spatial SF substructure. Inspired by the contiguous nature ofthe Gaia SF, which is basically only limited by a magnitude cut, and the fact that this magnitude cut would—inthe absence of small scale structure—translate to a sharp distance cut for standard candle tracer populations like redclump stars, we therefore use in our modelling a simple spatial SF of spherical shape with radius rmax around the Sun,

SF(x) ≡

completeness(x) if |x− x| ≤ rmax,

0 otherwise,(B1)

We set 0 ≤ completeness(x) ≤ 1 everywhere inside the observed volume, so it can be understood as a position-dependent detection probability for a star at x. Unless explicitly stated otherwise, we simplify to completeness(x) = 1.Additionally, we use in Figure 6 (Test 4) and Figures A1-A2 for illustrative purposes some rather unrealistic surveyvolumes which are angular segments of a cylindrical annulus (wedge), i.e., the volume with R ∈ [Rmin, Rmax], φ ∈[φmin, φmax], z ∈ [zmin, zmax] within the model Galaxy.