License: CC BY-SA 4.0
arXiv:2401.11534v1 [astro-ph.GA] 21 Jan 2024

Comparing dark matter and MOND hyphotheses from the distribution function of A, F, early-G stars in the solar neighbourhood

M. A. Syaifudin,11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPT M. I. Arifyanto,2,323{}^{2,3}start_FLOATSUPERSCRIPT 2 , 3 end_FLOATSUPERSCRIPT H. R. T. Wulandari2,323{}^{2,3}start_FLOATSUPERSCRIPT 2 , 3 end_FLOATSUPERSCRIPT, F. A. M. Mulki2,323{}^{2,3}start_FLOATSUPERSCRIPT 2 , 3 end_FLOATSUPERSCRIPT
11{}^{1}start_FLOATSUPERSCRIPT 1 end_FLOATSUPERSCRIPTAstronomy Master Program, Bandung Institute of Technology, Jl. Ganesa No. 10, 40135, Indonesia,
22{}^{2}start_FLOATSUPERSCRIPT 2 end_FLOATSUPERSCRIPTDepartment of Astronomy, Bandung Institute of Technology, Jl. Ganesa No. 10, 40135, Indonesia,
33{}^{3}start_FLOATSUPERSCRIPT 3 end_FLOATSUPERSCRIPTBosscha Observatory, Bandung Institute of Technology, Jl. Peneropongan Bintang No. 1, 40391, Indonesia
(January 21, 2024)
Abstract

Dark matter is hypothetical matter believed to address the missing mass problem in galaxies. However, alternative theories, such as Modified Newtonian Dynamics (MOND), have been notably successful in explaining the missing mass problem in various astrophysical systems. The vertical distribution function of stars in the solar neighbourhood serves as a proxy to constrain galactic dynamics in accordance to its contents. We employ both the vertical positional and velocity distribution of stars in cylindrical coordinates with a radius of 150 pc and a half-height of 200 pc from the galactic plane. Our tracers consist of main-sequence A, F, and early-G stars from the GAIA, RAVE, APOGEE, GALAH, and LAMOST catalogues. We attempt to solve the missing mass in the solar neighbourhood, interpreting it as either dark matter or MOND. Subsequently, we compare both hypotheses newtonian gravity with dark matter and MOND, using the Bayes factor (BF) to determine which one is more favoured by the data. We found that the inferred dark matter in the solar neighbourhood is in range of (0.010.07)Mpc3similar-toabsent0.010.07subscriptMdirect-productsuperscriptpc3\sim(0.01\textup{--}0.07)\,\textrm{M}_{\odot}\,\textrm{pc}^{-3}∼ ( 0.01 – 0.07 ) M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. We also determine that the MOND hypothesis’s acceleration parameter a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is (1.26±0.13)×1010ms2plus-or-minus1.260.13superscript1010msuperscripts2(1.26\pm 0.13)\times 10^{-10}\,\textrm{m}\,\textrm{s}^{-2}( 1.26 ± 0.13 ) × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT m s start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT for simple interpolating function. The average of bayes factor for all tracers between the two hypotheses is logBF0.1similar-toBF0.1\log\textrm{BF}\sim 0.1roman_log BF ∼ 0.1, meaning no strong evidence in favour of either the dark matter or MOND hypotheses.

Keywords: Dark Matter – Gravitation – Galaxy: kinematics and dynamics

1   INTRODUCTION

The study of dark matter constitutes a crucial topic in modern astrophysics. The discovery of flat rotation curves at large radii in spiral galaxies offered a glimpse into the existence of dark matter (Rubin et al. 1980). The local density of dark matter, typically observed a few hundred parsecs from the Sun, holds significant importance in constraining the nature of dark matter within the Milky Way. This value then imparts information about the shape of the Milky Way’s dark matter halo in the midplane. Furthermore, it serves as a constraint in galaxy formation models and cosmology (Read 2014), and even extends its importance to alternative theories of gravity (Milgrom 2001; Nipoti et al. 2007).

One of the first documented explorations of the missing mass problem was conducted by Oort (1932), who examined the vertical force on stars in the vicinity of the Sun. He observed that the vertical force exceeded the expected value based on visible matter alone. A similar problem was also identified by Zwicky (1933) through the observation of the mass-to-light ratio in the Coma cluster. Zwicky found that the mass-to-light ratio derived from dynamical means significantly exceeded that derived from the luminous matter alone.

The mainstream explanation for the missing mass problem is the existence of non-baryonic dark matter. This hypothetical matter interacts solely through gravity with baryonic matter, resulting in an additional gravitational force among the stars. The adoption of dark matter particles in various astrophysical systems has proven highly successful in numerous areas. Examples include explaining the rotation curve of spiral galaxies (Rubin et al. 1980), ensuring the stability of galactic discs (Ostriker & Peebles 1973), addressing mass deficiencies in ultrafaint dwarf galaxies (Lin & Ishak 2016), and contributing to our understanding of cosmological evolutions (Davis 2000).

The local dark matter density is typically estimated by measuring the vertical force on stars near the Sun. This is commonly accomplished by measuring the vertical velocity, a practice that had been studied since Oort (1932). Read (2014) provides an extensive review of the methods used to measure local dark matter. The estimated value of the local dark matter density usually hovers around 0.01similar-toabsent0.01\sim 0.01∼ 0.01 Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT, although some studies show higher values, reaching up to 0.1similar-toabsent0.1\sim 0.1∼ 0.1 Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT (Bahcall 1984a). Fig. 1 displays the values of local dark matter density from various studies.

Refer to caption
Figure 1: The estimated local dark matter density from a few studies in the last decade, including our own. The points are from Moni Bidin et al. (2012; MB12), Garbari et al. (2011; G12), Zhang et al. (2013; Z13), Bienaymè et al. (2014; B14), Xia et al. (2016; X16), Sivertsson et al. (2018; S18), and Buch et al. (2019; B19) for type A, F, and early G stars.

Alternative studies suggest that the apparent effect of missing mass on the vertical (and radial) force may not be attributed to dark matter itself but rather to modifications in the gravitational force law. This is typically achieved by adjusting the Newtonian gravitational force law, introducing slight changes to the effect of gravity at large distances. One classic example of such a modification is the Modified Newtonian Dyanimcs (MOND) theory, proposed by Milgrom (1983). Subsequently, extensive studies have been conducted to replace dark matter with modified gravity theories, yielding varying degrees of success (Famaey & McGaugh 2012).

The MOND theory has proven highly successful in elucidating the dynamics of stars in spiral galaxies (Milgrom 1994). This theory mimics the effects attributed to dark matter, but notably, it does not necessitate the presence of additional matter. Nevertheless, there are some differing predictions between the MOND theory and the dark matter theory. Lisanti et al. (2019) attempt to exploit these differences, particularly in the ratio of the vertical force to the radial force in the Milky Way. Their findings reveal that the MOND theory predicts a distinct ratio compared to the dark matter theory, providing a means to test which theory aligns more closely with the observed data.

Another method employed to test the hypotheses of dark matter and MOND involves comparing the centripetal acceleration of wide binary stars with their acceleration calculated using classical Newtonian gravity. Hernandez (2023) discovered anomalous behavior in the low acceleration regime, favoring MOND. Chae (2023) even identified evidence at a 10σ10𝜎10\sigma10 italic_σ level in favor of MOND. Conversely, Pittordis & Sutherland (2023) found that standard gravity is more preferred. A more rigorous Bayesian statistical analysis conducted by Banik et al. (2023) revealed results consistent with Newtonian dynamics, ruling out MOND with a confidence level of 16σ16𝜎16\sigma16 italic_σ. This complex array of findings indicates the need for further studies to rigorously test these two competing hypotheses.

In this paper, we will explore the influence of matter distribution (with or whitout dark matter) on generating gravitational potential and shaping the vertical distribution of stars near the Sun. We assume that the stars are in dynamical equilibrium and employ the collisionless Boltzmann equation to model the distribution of stars (Holmberg & Flynn 2004). The vertical density profile of tracers and their vertical velocity dispersion are measured. Subsequently, we attempt to fit the vertical distribution using mass models that assume isothermal components in dynamical equilibrium. This analysis is conducted for both dark matter and MOND hypotheses. The comparison of both best-fittings is then performed by evaluating their Bayes factor.

The paper is organized as follows: In Section 2, we provide a detailed description of the data collection and the subsequent data cleaning process. Section 3 outlines the mass model, theoretical background, and the two hypotheses. Section 4 explains the methodology employed for data fitting using Markov Chain Monte Carlo (MCMC). It includes an explanation of the likelihood function, the prior, and the numerical method used for solving the gravitational potential and calculating the Bayes factor. Section 5 presents a the analysis of the results, discussing the dark matter density from the dark matter hypothesis, the acceleration parameter from the MOND hypothesis, and a comparison of the two hypotheses using Bayes factors. Finally, Section 6 summarizes and concludes the findings of our study. Additionally, Section A contains the derivation of the likelihood function for inferring number density.

2   DATA

The data utilised in this study consist of the three-dimensional positions and kinematics of stars around the Sun. These data are compiled from Gaia DR3, RAVE 6, APOGEE-2 (SDSS), GALAH DR3, and LAMOST DR7 catalogues. They are collected either by downloading the bulk data directly from their repositories or queried using astroquery through virtual observatory services.

The Gaia catalogue serves as the primary source, providing three-dimensional positions (α,δ,ϖ𝛼𝛿italic-ϖ\alpha,\delta,\varpiitalic_α , italic_δ , italic_ϖ) and kinematics (μα,μδ,vrsubscript𝜇𝛼subscript𝜇𝛿subscript𝑣𝑟\mu_{\alpha},\mu_{\delta},v_{r}italic_μ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , italic_μ start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT) of stars. Other catalogues are employed to fill in missing radial velocity data from Gaia DR3. Table 1 displays the distribution of radial velocity data from each catalogue. It is evident that the contribution of radial velocity data from catalogues other than Gaia is not significant, rendering the subsequent analysis unaffected by data collected solely from Gaia DR3.

Table 1: Combined contribution of radial velocity data from Gaia, RAVE, APOGEE, GALAH, and LAMOST catalogues.
Catalogue #
GALAH 78,055
APOGEE 370,523
RAVE 27,826
LAMOST 442,140
Gaia 23,191,519

Additional information from 2MASS photometry is also utilised to obtain the colours J𝐽Jitalic_J and Kssubscript𝐾𝑠K_{s}italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT of the stars. The colour JKs𝐽subscript𝐾𝑠J-K_{s}italic_J - italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is used for selecting the main sequence cut, following a methodology similar to that employed in Bovy’s work (2017a). Infrared photometry is chosen because it is less susceptible to the effects of interstellar extinction. Hence, for this study, there is no need to correct for the extinction effect. The 2MASS All-sky Point Source Catalogue (PSC) encompasses photometry data for 470,992,970 stars in the near-infrared wavelength (J𝐽Jitalic_J, H𝐻Hitalic_H, and Kssubscript𝐾𝑠K_{s}italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT bands).

2.1 Sample Selection

Firstly, all stars from Gaia DR3 are selected across all directions (all sky). Subsequently, we crossmatch these stars with the 2MASS PSC to acquire the colours J𝐽Jitalic_J and Kssubscript𝐾𝑠K_{s}italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. This crossmatching process utilises the 2MASS-Gaia joined catalogue, a compilation previously conducted by the Gaia team (Marrese, P. M. et al. 2019) and accessible from the Gaia archive111https://gea.esac.esa.int/archive/. Both catalogues are outer joined, combining entries whether there are matching stars from 2MASS to Gaia or not. The resulting catalogue, named Gaia+2MASS, encompasses astrometry, kinematics, and photometry data for 470,992,970 stars.

The remaining catalogues are obtained by downloading the data from their respective archives. Fortunately, for each catalogue, there exist a pre-existing crossmatch with Gaia DR3 conducted by the Gaia team. These crossmatches are established by simply matching the IDs from the Gaia+2MASS catalogue with the corresponding join table catalogue. Specifically, there are 450,978 crossmatched stars from RAVE, 573,733 stars from APOGEE, 588,061 stars from GALAH, and 2,730,472 stars from LAMOST. It is required for the crossmatched catalogue to include radial velocity data for the stars.

Further filtering is conducted to exclude stars with poor-quality photometry, astrometry, and outliers. Adopting the criteria outlined in Kim & Lépine’s work (2021), the following selection criteria are applied to the stars:

  • Parallax ϖ>0italic-ϖ0\varpi>0italic_ϖ > 0: Only stars with a positive parallax are selected to avoid the complexities associated with distance inference for stars with negative parallax.

  • Parallax error σϖ/ϖ<0.15subscript𝜎italic-ϖitalic-ϖ0.15\sigma_{\varpi}/\varpi<0.15italic_σ start_POSTSUBSCRIPT italic_ϖ end_POSTSUBSCRIPT / italic_ϖ < 0.15: Stars with good-quality parallax data, characterized by a relative error less than 15 per cent, are chosen.

  • 3.0<GBPGRP<6.03.0subscript𝐺BPsubscript𝐺RP6.0-3.0<G_{\textup{BP}}-G_{\textup{RP}}<6.0- 3.0 < italic_G start_POSTSUBSCRIPT BP end_POSTSUBSCRIPT - italic_G start_POSTSUBSCRIPT RP end_POSTSUBSCRIPT < 6.0 and 3.0<G<21.03.0𝐺21.03.0<G<21.03.0 < italic_G < 21.0: This criterion aim to eliminate outliers in the colour-magnitude diagram.

  • σ(FBP)/FBP<0.10𝜎subscript𝐹BPsubscript𝐹BP0.10\sigma\left(F_{\textup{BP}}\right)/F_{\textup{BP}}<0.10italic_σ ( italic_F start_POSTSUBSCRIPT BP end_POSTSUBSCRIPT ) / italic_F start_POSTSUBSCRIPT BP end_POSTSUBSCRIPT < 0.10 and σ(FRP)/FRP<0.10𝜎subscript𝐹RPsubscript𝐹RP0.10\sigma\left(F_{\textup{RP}}\right)/F_{\textup{RP}}<0.10italic_σ ( italic_F start_POSTSUBSCRIPT RP end_POSTSUBSCRIPT ) / italic_F start_POSTSUBSCRIPT RP end_POSTSUBSCRIPT < 0.10: Only stars with good-quality photometry in GBPsubscript𝐺BPG_{\textup{BP}}italic_G start_POSTSUBSCRIPT BP end_POSTSUBSCRIPT and GRPsubscript𝐺RPG_{\textup{RP}}italic_G start_POSTSUBSCRIPT RP end_POSTSUBSCRIPT bands, characterized by relative errors less than 10 per cent, are selected.

  • ruwe <1.4absent1.4<1.4< 1.4: Only stars with good astrometry, defined by a renormalized unit weight error (ruwe) less than 1.4, are chosen.

  • 1.0+0.015(GBPGRP)2<E<1.3+0.06(GBPGRP)21.00.015superscriptsubscript𝐺BPsubscript𝐺RP2𝐸1.30.06superscriptsubscript𝐺BPsubscript𝐺RP21.0+0.015\left(G_{\textup{BP}}-G_{\textup{RP}}\right)^{2}<E<1.3+0.06\left(G_{% \textup{BP}}-G_{\textup{RP}}\right)^{2}1.0 + 0.015 ( italic_G start_POSTSUBSCRIPT BP end_POSTSUBSCRIPT - italic_G start_POSTSUBSCRIPT RP end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT < italic_E < 1.3 + 0.06 ( italic_G start_POSTSUBSCRIPT BP end_POSTSUBSCRIPT - italic_G start_POSTSUBSCRIPT RP end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT: This criterion is applied to select stars with a good colour excess E𝐸Eitalic_E.

Additional filtering is carried out by incorporating the stars’ infrared photometry from the 2MASS PSC to obtain the highest-quality infrared data. Following Bovy’s methodology (2017a), the criteria for selecting the best infrared data are:

  • 2<J<13.52𝐽13.52<J<13.52 < italic_J < 13.5: Establishing a lower bound to eliminate outliers and an upper bound to ensure greater than 99 per cent completeness.

  • ph_qual = A_A: Selecting high-quality photometry based on the photometric quality flag.

  • rd_flg = 1 or 3: Verifying data quality using the read flag, specifically selecting stars with flag values of 1 or 3.

  • use_src =1absent1=1= 1: Choosing stars that satisfy the criteria for good 2MASS data.

All the aforementioned criteria are applied to the Gaia+2MASS catalogue, yielding 29,144,488 stars with Gaia ID, referred to as the pre-combined catalogue. Additionally, 50,250,399 stars are identified with 2MASS infrared photometry only, serving as the reference dataset.

In the final step, the pre-combined catalogue is crossmatched with the remaining catalogues using the Gaia DR3 source ID. This crossmatched catalogue is denoted as the combined catalogue, encompassing numerous radial velocity data from various catalogues. The radial velocity data employed in this study corresponds to the radial velocity with the smallest error from the respective original catalogue. Stars lacking radial velocity data are excluded from the combined catalogue, resulting in a dataset comprising 24,110,063 stars with complete astrometry (α,δ,ϖ𝛼𝛿italic-ϖ\alpha,\delta,\varpiitalic_α , italic_δ , italic_ϖ), kinematics (μα,μδ,vrsubscript𝜇𝛼subscript𝜇𝛿subscript𝑣𝑟\mu_{\alpha},\mu_{\delta},v_{r}italic_μ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , italic_μ start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT), and photometry (J,Ks𝐽subscript𝐾𝑠J,K_{s}italic_J , italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) data.

2.2 Midplane Selection and Main Sequence Cut

Inferences regarding the stellar dynamics in the Milky Way due to matter distribution are drawn by analyzing the positions and motions of appropriate stellar populations, referred to as tracers. For this study, we specifically focus on stars in the solar neighborhood, defining them as stars within a cylindrical volume in galactice coordinate with a vertical distance of |z|<200𝑧200|z|<200| italic_z | < 200 pc from the midplane and a radius of 150150150150 pc. Main sequence stars are chosen and categorized into multiple groups based on their colour JKs𝐽subscript𝐾𝑠J-K_{s}italic_J - italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Following the methodology used by Bovy (2017a), we employ the mean dwarf stellar locus from Pecaut & Mamajek (2013) and introduce a reasonable thickness to the absolute magnitude MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT. Utilising a similar approach, we define the main sequence by tracing the main branch of the Hertzsprung-Russell diagram (HRD) of nearby stars, establishing upper and lower boundaries in absolute magnitude as displayed by the shaded area in Fig. 2.

Refer to caption
Figure 2: The Hertzsprung-Russell (HR) diagram depicts stars with stellar type A to early-G in the solar neighborhood, with the shaded area denoting the main sequence cut.

The stellar populations under consideration in this study must have achieved dynamical equilibrium within the Galactic disc in recent times. Following the recommendation of Buch et al. (2019), we opt for stars characterized by lower scale height and a shorter equilibration timescale, specifically focusing on younger A stars up to early G stars. Consequently, we choose the interval between A0 stars and G3 stars and divide the group into 15 subgroups, as detailed in Table 2. The color division for each group is carefully selected to ensure a minimum of 1000 stars in the midplane, guaranteeing robust statistical analyses for each subgroup. The spectral type of each group is determined from the color JKs𝐽subscript𝐾𝑠J-K_{s}italic_J - italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT using the relation provided by Pecaut & Mamajek (2013)222We use version 2022.04.16, https://www.pas.rochester.edu/ emamajek/EEM_dwarf_UBVIJHK_colours_Teff.txt.

Table 2: Tracer groups of stars utilised in the study, categorized by colour. The midplane column indicates the number of stars within |z|<50𝑧50|z|<50| italic_z | < 50 pc.
JKs𝐽subscript𝐾𝑠J-K_{s}italic_J - italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT Spectral Type |z|<200𝑧200|z|<200| italic_z | < 200 pc Midplane
[0.019[-0.019[ - 0.019, 0.124]0.124]0.124 ] A0 - A8 2422 1006
[0.124[0.124[ 0.124, 0.188]0.188]0.188 ] A9 - F1 2719 1014
[0.188[0.188[ 0.188, 0.219]0.219]0.219 ] F2 - F4 2867 1027
[0.219[0.219[ 0.219, 0.241]0.241]0.241 ] F5 2919 1014
[0.241[0.241[ 0.241, 0.259]0.259]0.259 ] F6 2970 1033
[0.259[0.259[ 0.259, 0.275]0.275]0.275 ] F7 3291 1049
[0.275[0.275[ 0.275, 0.288]0.288]0.288 ] F8 3165 1032
[0.288[0.288[ 0.288, 0.300]0.300]0.300 ] F9 3315 1038
[0.300[0.300[ 0.300, 0.312]0.312]0.312 ] F9.5 3523 1060
[0.312[0.312[ 0.312, 0.323]0.323]0.323 ] F9.5 3546 1001
[0.323[0.323[ 0.323, 0.333]0.333]0.333 ] G0 3493 1007
[0.333[0.333[ 0.333, 0.343]0.343]0.343 ] G1 3580 1028
[0.343[0.343[ 0.343, 0.353]0.353]0.353 ] G1 3803 1067
[0.353[0.353[ 0.353, 0.362]0.362]0.362 ] G2 3517 1044
[0.362[0.362[ 0.362, 0.376]0.376]0.376 ] G2 - G3 3460 1529

2.3 Selection Functions

Stellar information from the combined catalogue is subject to incompleteness arising from various factors, potentially introducing bias to the inferred density distribution. To mitigate this issue, it is crucial to correct the measured density distribution by accounting for the selection function of the stars. Following the approach outlined by Bovy (2017a), we estimate the selection function of the stars in the combined catalogue by comparing them to the assumed complete sample of stars from the reference catalogue (refer to subsection 2.1).

The selection function is defined as the probability of observing a star with specific properties in the catalogue or survey. Let S(J,JKs)𝑆𝐽𝐽subscript𝐾𝑠S(J,J-K_{s})italic_S ( italic_J , italic_J - italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) represent the ratio between the number of stars in the combined catalogue and the number of stars in the reference catalogue with color JKs𝐽subscript𝐾𝑠J-K_{s}italic_J - italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and absolute magnitude MJsubscript𝑀𝐽M_{J}italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT. The effective selection function is then defined as:

𝔖(x,y,z)=ρCMD(MJ,c|x,y,z)S(MJ,c)dcdMJ,𝔖𝑥𝑦𝑧subscript𝜌CMDsubscript𝑀𝐽conditional𝑐𝑥𝑦𝑧𝑆subscript𝑀𝐽𝑐d𝑐dsubscript𝑀𝐽\displaystyle\mathfrak{S}(x,y,z)=\int\rho_{\textup{CMD}}(M_{J},c|x,y,z)S(M_{J}% ,c)\textup{d}c\textup{d}M_{J},fraktur_S ( italic_x , italic_y , italic_z ) = ∫ italic_ρ start_POSTSUBSCRIPT CMD end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT , italic_c | italic_x , italic_y , italic_z ) italic_S ( italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT , italic_c ) d italic_c d italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT , (1)

where c=JKs𝑐𝐽subscript𝐾𝑠c=J-K_{s}italic_c = italic_J - italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, and ρCMD(MJ,JKs|x,y,z)subscript𝜌CMDsubscript𝑀𝐽𝐽conditionalsubscript𝐾𝑠𝑥𝑦𝑧\rho_{\textup{CMD}}(M_{J},J-K_{s}|x,y,z)italic_ρ start_POSTSUBSCRIPT CMD end_POSTSUBSCRIPT ( italic_M start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT , italic_J - italic_K start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT | italic_x , italic_y , italic_z ) represents the density distribution of stars in the color-magnitude diagram (CMD) at spatial position (x,y,z)𝑥𝑦𝑧(x,y,z)( italic_x , italic_y , italic_z ). Equation (1) can be seen as a conversion from color-magnitude space completeness to position space completeness. This equation is then carried to calculate the effective volume completeness, providing an estimate of the completeness for a given region ΠΠ\Piroman_Π, defined as:

Ξ(Π)=Π𝔖(x,y,z)d3xΠd3x,ΞΠsubscriptΠ𝔖𝑥𝑦𝑧superscriptd3𝑥subscriptΠsuperscriptd3𝑥\displaystyle\Xi(\Pi)=\frac{\int_{\Pi}\mathfrak{S}(x,y,z)\textup{d}^{3}x}{\int% _{\Pi}\textup{d}^{3}x},roman_Ξ ( roman_Π ) = divide start_ARG ∫ start_POSTSUBSCRIPT roman_Π end_POSTSUBSCRIPT fraktur_S ( italic_x , italic_y , italic_z ) d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_x end_ARG start_ARG ∫ start_POSTSUBSCRIPT roman_Π end_POSTSUBSCRIPT d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_x end_ARG , (2)

where in this case ΠΠ\Piroman_Π represents a cylindrical volume with a radius of 150150150150 pc and varying height z𝑧zitalic_z with some thickness ΔzΔ𝑧\Delta zroman_Δ italic_z.

For this study, we make the assumption that the density distribution in the Color-Magnitude Diagram (CMD) is uniform. The values can thus be inferred from the Hertzsprung-Russell (HR) diagram of stars in the solar neighborhood, as illustrated in Fig. 2.

The effective volume completeness is employed to adjust the density distribution of stars inferred from the combined catalogue to the (supposedly) true density distribution of stars in the solar neighborhood. This correction involves dividing the density distribution of stars in the combined catalogue by the effective volume completeness, resulting in what is termed the corrected density distribution. Figure 3 displays samples from four groups of stars alongside their corresponding effective volume completeness. Notably, the effective volume completeness remains relatively constant for higher values of |z|𝑧|z|| italic_z |, indicating that correction is essentially unnecessary for those regions.

Refer to caption
Figure 3: The effective volume completeness function for stars in the solar neighborhood. This graph illustrates the behavior for four selected groups of stars, with other groups exhibiting similar trends.

3   MODELS

3.1 Distribution Function

In a self-gravitating system such as a galaxy, the distribution function (DF) of a star represents the probability density function of finding a star with a given position and velocity (or momentum), denoted as f(𝒙,𝒗)𝑓𝒙𝒗f(\bm{x},\bm{v})italic_f ( bold_italic_x , bold_italic_v ). In a collisionless system and under the condition of dynamical equilibrium, the DF adheres to the collisionless Boltzmann equation (CBE),

(𝒙f)𝒗(𝒙Φ)(𝒗f)subscript𝒙𝑓𝒗subscript𝒙Φsubscript𝒗𝑓\displaystyle(\nabla_{\bm{x}}f)\cdot\bm{v}-(\nabla_{\bm{x}}\Phi)\cdot(\nabla_{% \bm{v}}f)( ∇ start_POSTSUBSCRIPT bold_italic_x end_POSTSUBSCRIPT italic_f ) ⋅ bold_italic_v - ( ∇ start_POSTSUBSCRIPT bold_italic_x end_POSTSUBSCRIPT roman_Φ ) ⋅ ( ∇ start_POSTSUBSCRIPT bold_italic_v end_POSTSUBSCRIPT italic_f ) =0absent0\displaystyle=0= 0
(𝒙f)𝒗subscript𝒙𝑓𝒗\displaystyle(\nabla_{\bm{x}}f)\cdot\bm{v}( ∇ start_POSTSUBSCRIPT bold_italic_x end_POSTSUBSCRIPT italic_f ) ⋅ bold_italic_v =(𝒙Φ)(𝒗f),absentsubscript𝒙Φsubscript𝒗𝑓\displaystyle=(\nabla_{\bm{x}}\Phi)\cdot(\nabla_{\bm{v}}f),= ( ∇ start_POSTSUBSCRIPT bold_italic_x end_POSTSUBSCRIPT roman_Φ ) ⋅ ( ∇ start_POSTSUBSCRIPT bold_italic_v end_POSTSUBSCRIPT italic_f ) , (3)

where ΦΦ\Phiroman_Φ is the gravitational potential of the system in equilibrium. When considering the DF of a star in close proximity to the midplane, it can be approximated by separating the contributions into the vertical and horizontal directions,

f(𝒙,𝒗)𝑓𝒙𝒗\displaystyle f(\bm{x},\bm{v})italic_f ( bold_italic_x , bold_italic_v ) =Z(z,w)T(r,vr,ϕ,vϕ),absent𝑍𝑧𝑤𝑇𝑟subscript𝑣𝑟italic-ϕsubscript𝑣italic-ϕ\displaystyle=Z(z,w)T(r,v_{r},\phi,v_{\phi}),= italic_Z ( italic_z , italic_w ) italic_T ( italic_r , italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT , italic_ϕ , italic_v start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT ) , (4)

where Z(z,w)𝑍𝑧𝑤Z(z,w)italic_Z ( italic_z , italic_w ) is the DF in the vertical direction, and T(r,vr,ϕ,vϕ)𝑇𝑟subscript𝑣𝑟italic-ϕsubscript𝑣italic-ϕT(r,v_{r},\phi,v_{\phi})italic_T ( italic_r , italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT , italic_ϕ , italic_v start_POSTSUBSCRIPT italic_ϕ end_POSTSUBSCRIPT ) is the DF in the horizontal direction. Equation (3) can be expressed as,

wZzΦzZw𝑤𝑍𝑧Φ𝑧𝑍𝑤\displaystyle w\frac{\partial Z}{\partial z}-\frac{\partial\Phi}{\partial z}% \frac{\partial Z}{\partial w}italic_w divide start_ARG ∂ italic_Z end_ARG start_ARG ∂ italic_z end_ARG - divide start_ARG ∂ roman_Φ end_ARG start_ARG ∂ italic_z end_ARG divide start_ARG ∂ italic_Z end_ARG start_ARG ∂ italic_w end_ARG =0.absent0\displaystyle=0.= 0 . (5)

The differential equation (5) has a general solution Z(z,w)=Z(Ez)Z(w2/2Φ(z))𝑍𝑧𝑤𝑍subscript𝐸𝑧𝑍superscript𝑤22Φ𝑧Z(z,w)=Z(E_{z})\equiv Z(w^{2}/2\Phi(z))italic_Z ( italic_z , italic_w ) = italic_Z ( italic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ) ≡ italic_Z ( italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 roman_Φ ( italic_z ) ), where Ezsubscript𝐸𝑧E_{z}italic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT is the energy in the vertical direction. Thus, the DF in the vertical direction is a function of energy only. Assuming the star has a vertical velocity dispersion f0(w)subscript𝑓0𝑤f_{0}(w)italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_w ) in the midplane, the DF in the vertical direction can then be expressed as

Z(z,w)𝑍𝑧𝑤\displaystyle Z(z,w)italic_Z ( italic_z , italic_w ) =Z(Ez)absent𝑍subscript𝐸𝑧\displaystyle=Z(E_{z})= italic_Z ( italic_E start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT )
=Z(w2/2+Φ(z))absent𝑍superscript𝑤22Φ𝑧\displaystyle=Z(w^{2}/2+\Phi(z))= italic_Z ( italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 + roman_Φ ( italic_z ) )
=Z(w02/2+Φ(z0))absent𝑍superscriptsubscript𝑤022Φsubscript𝑧0\displaystyle=Z(w_{0}^{2}/2+\Phi(z_{0}))= italic_Z ( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 + roman_Φ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) )
f0(w0)proportional-toabsentsubscript𝑓0subscript𝑤0\displaystyle\propto f_{0}(w_{0})∝ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT )
f0(w2+2[Φ(z)Φ(z0)]),proportional-toabsentsubscript𝑓0superscript𝑤22delimited-[]Φ𝑧Φsubscript𝑧0\displaystyle\propto f_{0}\left(\sqrt{w^{2}+2[\Phi(z)-\Phi(z_{0})]}\right),∝ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( square-root start_ARG italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 [ roman_Φ ( italic_z ) - roman_Φ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ] end_ARG ) , (6)

where, z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT represent the equivalent midplane position and velocity of the star that has the same energy as the star at (z,w)𝑧𝑤(z,w)( italic_z , italic_w ).

If we assume the vertical velocity dispersion f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT follows a Gaussian distribution, then equation (6) can be solved analytically to obtain the DF of spatial vertical direction.

Fz(z)subscript𝐹𝑧𝑧\displaystyle F_{z}(z)italic_F start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ( italic_z ) =Z(z,w)dwabsentsuperscriptsubscript𝑍𝑧𝑤d𝑤\displaystyle=\int_{-\infty}^{\infty}Z(z,w)\text{d}w= ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_Z ( italic_z , italic_w ) d italic_w
=Aexp(Φ(z)σw2),absent𝐴Φ𝑧superscriptsubscript𝜎𝑤2\displaystyle=A\exp\left(-\frac{\Phi(z)}{\sigma_{w}^{2}}\right),= italic_A roman_exp ( - divide start_ARG roman_Φ ( italic_z ) end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (7)
ν(z)absent𝜈𝑧\displaystyle\Rightarrow\nu(z)⇒ italic_ν ( italic_z ) =ν0exp(Φ(z)σw2),absentsubscript𝜈0Φ𝑧superscriptsubscript𝜎𝑤2\displaystyle=\nu_{0}\exp\left(-\frac{\Phi(z)}{\sigma_{w}^{2}}\right),= italic_ν start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_exp ( - divide start_ARG roman_Φ ( italic_z ) end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (8)

where A𝐴Aitalic_A is a normalization constant, σwsubscript𝜎𝑤\sigma_{w}italic_σ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT is the vertical velocity dispersion, and Φ(z0)Φsubscript𝑧0\Phi(z_{0})roman_Φ ( italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is chosen to be zero. If we only consider N𝑁Nitalic_N identical stars, then the vertical distribution (7) is proportional to the expected number density ν(z)𝜈𝑧\nu(z)italic_ν ( italic_z ) of the stars, leading to (8).

The vertical velocity distribution can be calculated by integrating the DF in the spatial vertical direction,

Fw(w)subscript𝐹𝑤𝑤\displaystyle F_{w}(w)italic_F start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT ( italic_w ) =Z(z,w)dzabsentsuperscriptsubscript𝑍𝑧𝑤d𝑧\displaystyle=\int_{-\infty}^{\infty}Z(z,w)\text{d}z= ∫ start_POSTSUBSCRIPT - ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_Z ( italic_z , italic_w ) d italic_z
=a*f0(w)absentsuperscript𝑎subscript𝑓0𝑤\displaystyle=a^{*}f_{0}(w)= italic_a start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_w )
ν(w)absent𝜈𝑤\displaystyle\Rightarrow\nu(w)⇒ italic_ν ( italic_w ) =aexp[(ww0)22σw2],absent𝑎superscript𝑤subscript𝑤022superscriptsubscript𝜎𝑤2\displaystyle=a\exp\left[-\frac{\left(w-w_{0}\right)^{2}}{2\sigma_{w}^{2}}% \right],= italic_a roman_exp [ - divide start_ARG ( italic_w - italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ] , (9)

where a𝑎aitalic_a is a normalization constant. The vertical velocity distribution (9) is a Gaussian distribution with mean w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and standard deviation σwsubscript𝜎𝑤\sigma_{w}italic_σ start_POSTSUBSCRIPT italic_w end_POSTSUBSCRIPT.

3.2 Gravitational Potential

The gravitational potential of a system can be calculated from the mass density distribution using Poisson’s equation. In a cylindrical volume around the Sun, Poisson’s equation can be expressed as

2Φsuperscript2Φ\displaystyle\nabla^{2}\Phi∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Φ =2Φz2+1rr(rΦr)=4πGρabsentsuperscript2Φsuperscript𝑧21𝑟𝑟𝑟Φ𝑟4𝜋𝐺𝜌\displaystyle=\frac{\partial^{2}\Phi}{\partial z^{2}}+\frac{1}{r}\frac{% \partial}{\partial r}\left(r\frac{\partial\Phi}{\partial r}\right)=4\pi G\rho= divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Φ end_ARG start_ARG ∂ italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG italic_r end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ italic_r end_ARG ( italic_r divide start_ARG ∂ roman_Φ end_ARG start_ARG ∂ italic_r end_ARG ) = 4 italic_π italic_G italic_ρ
2Φz2superscript2Φsuperscript𝑧2\displaystyle\frac{\partial^{2}\Phi}{\partial z^{2}}divide start_ARG ∂ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Φ end_ARG start_ARG ∂ italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG =4πGρeff,absent4𝜋𝐺subscript𝜌eff\displaystyle=4\pi G\rho_{\textup{eff}},= 4 italic_π italic_G italic_ρ start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT , (10)

where the second term of equation (10) is referred to as the rotation curve term \mathcal{R}caligraphic_R, which is effectively constant for a small volume around the Sun. The rotation curve term can be calculated as

\displaystyle\mathcal{R}caligraphic_R =14πG1r(rΦr)absent14𝜋𝐺1𝑟𝑟Φ𝑟\displaystyle=\frac{1}{4\pi G}\frac{1}{r}\frac{\partial}{\partial}\left(r\frac% {\partial\Phi}{\partial r}\right)= divide start_ARG 1 end_ARG start_ARG 4 italic_π italic_G end_ARG divide start_ARG 1 end_ARG start_ARG italic_r end_ARG divide start_ARG ∂ end_ARG start_ARG ∂ end_ARG ( italic_r divide start_ARG ∂ roman_Φ end_ARG start_ARG ∂ italic_r end_ARG )
=A2B22πGabsentsuperscript𝐴2superscript𝐵22𝜋𝐺\displaystyle=\frac{A^{2}-B^{2}}{2\pi G}= divide start_ARG italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_π italic_G end_ARG (11)

where A𝐴Aitalic_A and B𝐵Bitalic_B represent the Oort constants. The values of A𝐴Aitalic_A and B𝐵Bitalic_B are 15.3±0.4plus-or-minus15.30.415.3\pm 0.415.3 ± 0.4 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT kpc11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT and 11.9±0.4plus-or-minus11.90.4-11.9\pm 0.4- 11.9 ± 0.4 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT kpc11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, respectively (Bovy 2017b).

For the volume studied in this work, the rotation curve term is approximately constant and the gravitational potential ΦΦ\Phiroman_Φ is only dependent on z𝑧zitalic_z. Therefore, the Poisson’s equation (10) can be written as a second-order ordinary differential equation with respect to z𝑧zitalic_z,

d2Φ(z)dz2superscriptd2Φ𝑧dsuperscript𝑧2\displaystyle\frac{\textup{d}^{2}\Phi(z)}{\textup{d}z^{2}}divide start_ARG d start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Φ ( italic_z ) end_ARG start_ARG d italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG =4πGρeff(Φ,z),absent4𝜋𝐺subscript𝜌effΦ𝑧\displaystyle=4\pi G\rho_{\textup{eff}}(\Phi,z),= 4 italic_π italic_G italic_ρ start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT ( roman_Φ , italic_z ) , (12)

where ρeff(Φ,z)subscript𝜌effΦ𝑧\rho_{\textup{eff}}(\Phi,z)italic_ρ start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT ( roman_Φ , italic_z ) could be a function of ΦΦ\Phiroman_Φ and z𝑧zitalic_z. Section 3.4 will discuss the mass model used in this work in detail.

3.3 Modified Poisson Equation

In Milgrom’s work 1983, he proposed a modification to Newtonian dynamics by introducing a new constant a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, referred to as the acceleration parameter. This constant plays a crucial role in scenarios where an object experiences a low acceleration field. One approach to incorporating the acceleration parameter into Newtonian dynamics is by ensuring that the circular speed of an object in a low acceleration regime remains constant:

Centripetal acc.=Vc2r1r.Centripetal acc.superscriptsubscript𝑉𝑐2𝑟proportional-to1𝑟\displaystyle\text{Centripetal acc.}=\frac{V_{c}^{2}}{r}\propto\frac{1}{r}.Centripetal acc. = divide start_ARG italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_r end_ARG ∝ divide start_ARG 1 end_ARG start_ARG italic_r end_ARG . (13)

Such a modification can be achieved by adjusting the gravitating acceleration to follow

g=gNa0,𝑔subscript𝑔𝑁subscript𝑎0\displaystyle g=\sqrt{g_{N}a_{0}},italic_g = square-root start_ARG italic_g start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG , (14)

where gNsubscript𝑔𝑁g_{N}italic_g start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT represents the classical Newtonian gravitational acceleration.

Concerning this phenomenology, MOND introduces a different form of Poisson’s equation, as formulated by Bekenstein & Milgrom (1984) for AQUAL theory:

4πGρ=[μ(|Φ|a0)Φ],4𝜋𝐺𝜌delimited-[]𝜇Φsubscript𝑎0Φ\displaystyle 4\pi G\rho=\nabla\cdot\left[\mu\left(\frac{|\nabla\Phi|}{a_{0}}% \right)\Phi\right],4 italic_π italic_G italic_ρ = ∇ ⋅ [ italic_μ ( divide start_ARG | ∇ roman_Φ | end_ARG start_ARG italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ) roman_Φ ] , (15)

where a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT represents the MOND acceleration parameter constant, and μ𝜇\muitalic_μ is the MOND μ𝜇\muitalic_μ-interpolating function. The μ𝜇\muitalic_μ-interpolating function is defined as

μ(x)𝜇𝑥\displaystyle\mu(x)italic_μ ( italic_x ) ={1if x1xif x1.absentcases1much-greater-thanif 𝑥1𝑥much-less-thanif 𝑥1\displaystyle=\begin{cases}1&\text{if }x\gg 1\\ x&\text{if }x\ll 1\end{cases}.= { start_ROW start_CELL 1 end_CELL start_CELL if italic_x ≫ 1 end_CELL end_ROW start_ROW start_CELL italic_x end_CELL start_CELL if italic_x ≪ 1 end_CELL end_ROW . (16)

Equation (16) ensures that the modified Poisson’s equation reduces to the Newtonian Poisson’s equation when the acceleration is much larger than the MOND acceleration parameter a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. In small volumes, where the total acceleration |Φ|Φ|\nabla\Phi|| ∇ roman_Φ | is approximately constant, the modified Poisson’s equation (15) can be simplified to:

2Φ=4πGρeffμ0,superscript2Φ4𝜋𝐺subscript𝜌effsubscript𝜇0\displaystyle\nabla^{2}\Phi=\frac{4\pi G\rho_{\text{eff}}}{\mu_{0}},∇ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Φ = divide start_ARG 4 italic_π italic_G italic_ρ start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG , (17)

where ρeffsubscript𝜌eff\rho_{\text{eff}}italic_ρ start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT is the effective mass density (ρ4πG𝜌4𝜋𝐺\rho-4\pi G\mathcal{R}italic_ρ - 4 italic_π italic_G caligraphic_R), and μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the value of the interpolating function at the Sun’s position.

Both Poisson equations (12) and (17) are similar, with a minor difference represented by a factor of μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. This value can be interpreted as the additional boost of the gravitational potential due to the MOND effect. In the dark matter hypothesis, this additional boost is generated by additional mass density from nonluminous dark matter. Essentially, the MOND effect can be mimicked by adding extra mass density to the system. The two equations can result in the exact same gravitational potential if the additional missing mass follows the baryonic mass distribution. Fortunately, most dark matter models predict that the dark matter distribution does not precisely follow the baryonic mass distribution (Navarro et al. 1997). As a result, the two equations should yield different gravitational potentials in theory. The extent of the difference between the two equations is determined by the value of μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. If μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is close to 1, it becomes extremely challenging to distinguish between the hypotheses of dark matter and MOND.

The exact function of the μ𝜇\muitalic_μ-interpolating function is not precisely known. However, for many studies focusing on the low acceleration regime, the specific form of the interpolating function is often not critical. In this study, where the acceleration is not low enough to be in the low acceleration regime, the exact function of the interpolating function becomes crucial for inferring the value of the acceleration parameter a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We consider two common interpolating functions: the simple and the standard. Table 3 provides the exact functions for these two interpolating functions.

Table 3: The formulas for the interpolating function μ(x)𝜇𝑥\mu(x)italic_μ ( italic_x ).
Interpolating Function μ(x)𝜇𝑥\mu(x)italic_μ ( italic_x )
Simple x1+x𝑥1𝑥\displaystyle\frac{x}{1+x}divide start_ARG italic_x end_ARG start_ARG 1 + italic_x end_ARG
Standard x1+x2𝑥1superscript𝑥2\displaystyle\frac{x}{\sqrt{1+x^{2}}}divide start_ARG italic_x end_ARG start_ARG square-root start_ARG 1 + italic_x start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG

3.4 Mass Model

The model employs in this study encompasses analytical functions for the vertical distribution of stars (8) and the vertical velocity distribution of stars (9). This model is derived by solving the Poisson equation (10 or 17) with the provided mass density distribution. The mass density distribution for this study is the sum of Nbsubscript𝑁bN_{\textrm{b}}italic_N start_POSTSUBSCRIPT b end_POSTSUBSCRIPT baryon components and a constant dark matter (DM) halo contribution in the midplane. The baryon mass density ρbsubscript𝜌b\rho_{\textrm{b}}italic_ρ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT is described by the Bahcall model, which includes isothermal gas, stars, and stellar remnants (Bahcall 1984b, 1984c, 1984a).

ρb(z)subscript𝜌b𝑧\displaystyle\rho_{\textrm{b}}(z)italic_ρ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT ( italic_z ) =i=1Nbρi(0)exp(Φ(z)σz,i2),absentsuperscriptsubscript𝑖1subscript𝑁𝑏subscript𝜌𝑖0Φ𝑧superscriptsubscript𝜎𝑧𝑖2\displaystyle=\sum_{i=1}^{N_{b}}\rho_{i}(0)\exp\left(-\frac{\Phi(z)}{\sigma_{z% ,i}^{2}}\right),= ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) roman_exp ( - divide start_ARG roman_Φ ( italic_z ) end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_z , italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (18)

where ρi(0)subscript𝜌𝑖0\rho_{i}(0)italic_ρ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( 0 ) represents the midplane mass density of the i𝑖iitalic_i-th component at the midplane, and σz,isubscript𝜎𝑧𝑖\sigma_{z,i}italic_σ start_POSTSUBSCRIPT italic_z , italic_i end_POSTSUBSCRIPT denotes the vertical velocity dispersion of the i𝑖iitalic_i-th component. Table 4 provides the parameters of the Bahcall model utilized in this study.

Table 4: Baryonic mass components employed in this study. The table is extracted from table 2 of Buch et al. (2019).
Component ρi,0[M/pc2]subscript𝜌𝑖0delimited-[]subscriptMdirect-productsuperscriptpc2\rho_{i,0}\left[\textrm{M}_{\odot}/\textup{pc}^{2}\right]italic_ρ start_POSTSUBSCRIPT italic_i , 0 end_POSTSUBSCRIPT [ M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT / pc start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] σz,i[km/s]subscript𝜎𝑧𝑖delimited-[]km/s\sigma_{z,i}\left[\textup{km/s}\right]italic_σ start_POSTSUBSCRIPT italic_z , italic_i end_POSTSUBSCRIPT [ km/s ]
Molecular gas (H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT) 0.0104±0.00312plus-or-minus0.01040.003120.0104\pm 0.003120.0104 ± 0.00312 3.7±0.2plus-or-minus3.70.23.7\pm 0.23.7 ± 0.2
Cold atomic gas (HI(1)I1\,\text{\scriptsize I}(1)I ( 1 )) 0.0277±0.00554plus-or-minus0.02770.005540.0277\pm 0.005540.0277 ± 0.00554 7.1±0.5plus-or-minus7.10.57.1\pm 0.57.1 ± 0.5
Warm atomic gas (HI(2)I2\,\text{\scriptsize I}(2)I ( 2 )) 0.0073±0.00070plus-or-minus0.00730.000700.0073\pm 0.000700.0073 ± 0.00070 22.1±2.4plus-or-minus22.12.422.1\pm 2.422.1 ± 2.4
Hot ionized gas (H II) 0.0005±0.00003plus-or-minus0.00050.000030.0005\pm 0.000030.0005 ± 0.00003 39.0±4.0plus-or-minus39.04.039.0\pm 4.039.0 ± 4.0
Giant stars 0.0006±0.00006plus-or-minus0.00060.000060.0006\pm 0.000060.0006 ± 0.00006 15.5±1.6plus-or-minus15.51.615.5\pm 1.615.5 ± 1.6
MV<3subscript𝑀𝑉3M_{V}<3italic_M start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT < 3 0.0018±0.00018plus-or-minus0.00180.000180.0018\pm 0.000180.0018 ± 0.00018 7.5±2.0plus-or-minus7.52.07.5\pm 2.07.5 ± 2.0
3<MV<43subscript𝑀𝑉43<M_{V}<43 < italic_M start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT < 4 0.0018±0.00018plus-or-minus0.00180.000180.0018\pm 0.000180.0018 ± 0.00018 12.0±2.4plus-or-minus12.02.412.0\pm 2.412.0 ± 2.4
4<MV<54subscript𝑀𝑉54<M_{V}<54 < italic_M start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT < 5 0.0029±0.00029plus-or-minus0.00290.000290.0029\pm 0.000290.0029 ± 0.00029 18.0±1.8plus-or-minus18.01.818.0\pm 1.818.0 ± 1.8
5<MV<85subscript𝑀𝑉85<M_{V}<85 < italic_M start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT < 8 0.0072±0.00072plus-or-minus0.00720.000720.0072\pm 0.000720.0072 ± 0.00072 18.5±1.9plus-or-minus18.51.918.5\pm 1.918.5 ± 1.9
MV>8subscript𝑀𝑉8M_{V}>8italic_M start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT > 8 (M dwarfs) 0.0216±0.00280plus-or-minus0.02160.002800.0216\pm 0.002800.0216 ± 0.00280 18.5±4.0plus-or-minus18.54.018.5\pm 4.018.5 ± 4.0
White dwarfs 0.0056±0.00100plus-or-minus0.00560.001000.0056\pm 0.001000.0056 ± 0.00100 20.0±5.0plus-or-minus20.05.020.0\pm 5.020.0 ± 5.0
Brown dwarfs 0.0015±0.00050plus-or-minus0.00150.000500.0015\pm 0.000500.0015 ± 0.00050 20.0±5.0plus-or-minus20.05.020.0\pm 5.020.0 ± 5.0

The DM halo contribution is assumed to be constant in the vicinity of the Sun, so the total mass density distribution is given by

ρ(z)𝜌𝑧\displaystyle\rho(z)italic_ρ ( italic_z ) =ρb(z)+ρDM.absentsubscript𝜌b𝑧subscript𝜌DM\displaystyle=\rho_{\textrm{b}}(z)+\rho_{\textrm{\scriptsize DM}}.= italic_ρ start_POSTSUBSCRIPT b end_POSTSUBSCRIPT ( italic_z ) + italic_ρ start_POSTSUBSCRIPT DM end_POSTSUBSCRIPT . (19)

Based on the mass density distribution, the gravitational potential is calculated by solving the Poisson equation (10) or (17). This potential is then used to compute the vertical distribution of stars (8) and the vertical velocity distribution of stars (9).

Solving the Poisson equation with this mass model is not trivial because it is not analytically solvable or separable. Numerical methods are needed to solve the Poisson equation. Fundamentally, from the Poisson equation, given mass model ρ𝜌\rhoitalic_ρ, there is a unique gravitational potential ΦΦ\Phiroman_Φ.

In this study, we explore three hypotheses: the baseline hypothesis involves Classic Newtonian gravity without Dark Matter (NO), while the other two hypotheses consider Newtonian gravity with dark matter (DM) and Modified Newtonian Dynamics (MOND). The DM model incorporates the DM halo contribution into the mass density distribution, while the MOND model relies solely on baryonic mass contributions. Despite the exclusion of the DM halo contribution in the MOND model’s equation (19), it introduces the MOND effect into the Poisson equation (17). Equation (17) can be perceived as the traditional Poisson equation with baryonic mass density, scaled by a factor of 1/μ01subscript𝜇01/\mu_{0}1 / italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT times the baryonic matter. Essentially, for our specific study, the MOND model is equivalent to a scenario without a DM model using classical Newtonian gravity, albeit with adjusted mass density.

4   METHODS

4.1 Density Inference

The most straightforward method for inferring the density distribution of stars from their spatial distribution is to count the number of stars within a specified volume. While this approach is rudimentary, its simplicity is a powerful tool, particularly when dealing with a large number of stars. However, a drawback of this method is the omission of uncertainties in star positions, which may be significant in certain cases.

Previous attempts, as demonstrated by Buch et al. (2019); Xia et al. (2016), Sivertsson (2018), among others, have aimed to infer density uncertainty by employing Poisson error or simple standard deviation within a given volume. While this approach is superior to having no uncertainty estimation, it often assumes that the distribution follows a Gaussian (or even log-Gaussian) shape, which is not always accurate. For a large number of stars, the Poisson distribution can be well approximated by a Gaussian distribution. However, when dealing with a small number of stars, assuming a Gaussian distribution might lead to inaccuracies. In such cases, the distribution should be considered back to Poisson distribution.

While the Poisson error in a given volume accounts for uncertainty, it does not consider the individual star’s position uncertainty. For datasets with high-quality parallax data, individual star position uncertainties may be negligible. However, in our study, we aim to develop a method applicable to any type of data, including those with poor quality and significant uncertainty. Appendix A outlines the methodology for inferring the density distribution of objects from their ‘position’. Using this approach, we can infer the density distribution of stars from their spatial distribution, taking into account the uncertainty in their positions. In other words, for a given distance z𝑧zitalic_z from the midplane, there is a probability density function P(ν,z)𝑃𝜈𝑧P(\nu,z)italic_P ( italic_ν , italic_z ) representing the likelihood of the density being ν𝜈\nuitalic_ν.

We employ the technique proposed by Bailer-Jones (2015) to obtain the probability distribution function of the distance r𝑟ritalic_r with a constant volume density prior:

p(r|ϖ,σϖ)r2exp[12σϖ2(ϖ1r)2].proportional-to𝑝conditional𝑟italic-ϖsubscript𝜎italic-ϖsuperscript𝑟212superscriptsubscript𝜎italic-ϖ2superscriptitalic-ϖ1𝑟2\displaystyle p(r|\varpi,\sigma_{\varpi})\propto r^{2}\exp\left[-\frac{1}{2% \sigma_{\varpi}^{2}}\left(\varpi-\frac{1}{r}\right)^{2}\right].italic_p ( italic_r | italic_ϖ , italic_σ start_POSTSUBSCRIPT italic_ϖ end_POSTSUBSCRIPT ) ∝ italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_exp [ - divide start_ARG 1 end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT italic_ϖ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( italic_ϖ - divide start_ARG 1 end_ARG start_ARG italic_r end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] . (20)

Subsequently, we convert this distribution to the vertical distance z𝑧zitalic_z,

p(z|ϖ,σϖ,b)z2exp[12σϖ2(ϖsinbz)2],proportional-to𝑝conditional𝑧italic-ϖsubscript𝜎italic-ϖ𝑏superscript𝑧212superscriptsubscript𝜎italic-ϖ2superscriptitalic-ϖ𝑏𝑧2\displaystyle p(z|\varpi,\sigma_{\varpi},b)\propto z^{2}\exp\left[-\frac{1}{2% \sigma_{\varpi}^{2}}\left(\varpi-\frac{\sin b}{z}\right)^{2}\right],italic_p ( italic_z | italic_ϖ , italic_σ start_POSTSUBSCRIPT italic_ϖ end_POSTSUBSCRIPT , italic_b ) ∝ italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_exp [ - divide start_ARG 1 end_ARG start_ARG 2 italic_σ start_POSTSUBSCRIPT italic_ϖ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( italic_ϖ - divide start_ARG roman_sin italic_b end_ARG start_ARG italic_z end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] , (21)

where (ϖ,σϖ)italic-ϖsubscript𝜎italic-ϖ(\varpi,\sigma_{\varpi})( italic_ϖ , italic_σ start_POSTSUBSCRIPT italic_ϖ end_POSTSUBSCRIPT ) represent the parallax and its error, z𝑧zitalic_z denotes the vertical distance from the Galactic midplane, and b𝑏bitalic_b is the Galactic latitude.

A similar approach is used to determine the probability distribution function of vertical velocity. The formula used to calculate w𝑤witalic_w from its proper motion and radial velocity is given by:

w(μb,b,vr,r)=κμbrcosb+vrsinb,𝑤subscript𝜇𝑏𝑏subscript𝑣𝑟𝑟𝜅subscript𝜇𝑏𝑟𝑏subscript𝑣𝑟𝑏\displaystyle w(\mu_{b},b,v_{r},r)=\kappa\mu_{b}r\cos b+v_{r}\sin b,italic_w ( italic_μ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , italic_b , italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT , italic_r ) = italic_κ italic_μ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT italic_r roman_cos italic_b + italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT roman_sin italic_b , (22)

where κ=4.74𝜅4.74\kappa=4.74italic_κ = 4.74 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT pc11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT arcsec yr11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT, μbsubscript𝜇𝑏\mu_{b}italic_μ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT is the proper motion in the galactic latitude direction, and vrsubscript𝑣𝑟v_{r}italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT is the radial velocity.

We generate 100,000 samples of (ϖ,μb,vritalic-ϖsubscript𝜇𝑏subscript𝑣𝑟\varpi,\mu_{b},v_{r}italic_ϖ , italic_μ start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , italic_v start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT) for each star, then calculate their z𝑧zitalic_z and w𝑤witalic_w values. Subsequently, we bin both z𝑧zitalic_z and w𝑤witalic_w for all star groups and calculate their weights W𝑊Witalic_W according to the methods outlined in Appendix A. Fig 4 shows one such inference for the tracer of the first group (A0-A8).

4.2 Likelihood Function

The likelihood function is divided into two parts: the likelihood function of the vertical number density z𝑧zitalic_z and the likelihood function of the vertical velocity dispersion w𝑤witalic_w. Given a set of parameters ζ𝜁\zetaitalic_ζ that describe the vertical distribution of stars, the likelihood function of finding N(zi|ζ)𝑁conditionalsubscript𝑧𝑖𝜁N(z_{i}|\zeta)italic_N ( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ζ ) stars in the i𝑖iitalic_i-th bin is expressed as

(N(zi|ζ))𝑁conditionalsubscript𝑧𝑖𝜁\displaystyle\mathcal{L}(N(z_{i}|\zeta))caligraphic_L ( italic_N ( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ζ ) ) =n=0|z=ziW(n)Poisson(n|N(zi|ζ)),\displaystyle=\sum_{n=0|z=z_{i}}^{\infty}W(n)\textup{Poisson}\left(n|N(z_{i}|% \zeta)\right),= ∑ start_POSTSUBSCRIPT italic_n = 0 | italic_z = italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_W ( italic_n ) Poisson ( italic_n | italic_N ( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ζ ) ) , (23)

where W(n)𝑊𝑛W(n)italic_W ( italic_n ) represents the weight of finding n𝑛nitalic_n stars in the i𝑖iitalic_i-th bin, calculated from the data collected in Section 2. The value of N(zi|ζ)𝑁conditionalsubscript𝑧𝑖𝜁N(z_{i}|\zeta)italic_N ( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ζ ) is calculated using (8). The term Poisson(n|N)Poissonconditional𝑛𝑁\textup{Poisson}(n|N)Poisson ( italic_n | italic_N ) is the probability density function of the Poisson distribution, representing the likelihood of finding n𝑛nitalic_n stars given the expected number of stars N𝑁Nitalic_N. For a more detailed derivation, refer to Appendix A.

The same function is used to calculate the likelihood function of finding N(wi|ξ)𝑁conditionalsubscript𝑤𝑖𝜉N(w_{i}|\xi)italic_N ( italic_w start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ξ ) stars in the i𝑖iitalic_i-th bin of the velocity distribution, where ξ𝜉\xiitalic_ξ is a set of parameters defining the vertical velocity distribution of the stars. Therefore,

(N(wi|ξ))𝑁conditionalsubscript𝑤𝑖𝜉\displaystyle\mathcal{L}(N(w_{i}|\xi))caligraphic_L ( italic_N ( italic_w start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ξ ) ) =n=0|z=ziW(n)Poisson(n|N(wi|ξ)).\displaystyle=\sum_{n=0|z=z_{i}}^{\infty}W(n)\textup{Poisson}\left(n|N(w_{i}|% \xi)\right).= ∑ start_POSTSUBSCRIPT italic_n = 0 | italic_z = italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_W ( italic_n ) Poisson ( italic_n | italic_N ( italic_w start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_ξ ) ) . (24)

The total likelihood function is the product of the likelihood function for the vertical number density and the likelihood function for the vertical velocity dispersion. The complete set of free parameters θ𝜃\thetaitalic_θ encompasses both the combined dynamics parameters ζ𝜁\zetaitalic_ζ and the kinematics parameters ξ𝜉\xiitalic_ξ.

4.3 Priors

The prior distribution of the parameters ζ𝜁\zetaitalic_ζ and ξ𝜉\xiitalic_ξ is summarized in Table 6 and Table 5, respectively. We assume that for each group, there are actually two populations of stars: one with a broader velocity dispersion and another with a narrower velocity dispersion. Most likely, these two populations correspond to the thin disc and the thick disc and/or halo. The thin disc is assumed to be the dominant population, so its contribution is assumed to be higher than that of the contaminant component. This assumption is based on the fact that the thin disc is the dominant population in the solar neighborhood. Adopting this approach, we observe that the fitting curve is improved compared to using only one population.

Only two types of prior distributions are used: uniform distribution and normal distribution. In Table 6 and Table 5, the “Constraints” column is defined as lower and upper bounds [a,b]𝑎𝑏[a,b][ italic_a , italic_b ] if the distribution is uniform, and mean and standard deviation [μ,σ]𝜇𝜎[\mu,\sigma][ italic_μ , italic_σ ] if the distribution is normal.

We also impose restrictions on the joint prior distribution of the parameters σw,1subscript𝜎𝑤1\sigma_{w,1}italic_σ start_POSTSUBSCRIPT italic_w , 1 end_POSTSUBSCRIPT and σw,2subscript𝜎𝑤2\sigma_{w,2}italic_σ start_POSTSUBSCRIPT italic_w , 2 end_POSTSUBSCRIPT to be σw,1<σw,2subscript𝜎𝑤1subscript𝜎𝑤2\sigma_{w,1}<\sigma_{w,2}italic_σ start_POSTSUBSCRIPT italic_w , 1 end_POSTSUBSCRIPT < italic_σ start_POSTSUBSCRIPT italic_w , 2 end_POSTSUBSCRIPT and on the parameters a1subscript𝑎1a_{1}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and a2subscript𝑎2a_{2}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT to be a1>a2subscript𝑎1subscript𝑎2a_{1}>a_{2}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. Parameter a𝑎aitalic_a is the sum of a1subscript𝑎1a_{1}italic_a start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and a2subscript𝑎2a_{2}italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. These restrictions are implemented based on the assumption that the thin disc is the dominant population and that the thin disc has a narrower velocity dispersion than the contaminant population.

Table 5: Summary of kinematic priors, with units indicated in parentheses corresponding to the argument of logarithm.
Parameter Distribution Contraints Unit
w0subscript𝑤0w_{0}italic_w start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Uniform [15,0]150[-15,0][ - 15 , 0 ] km/s
logσw,isubscript𝜎𝑤𝑖\log\sigma_{w,i}roman_log italic_σ start_POSTSUBSCRIPT italic_w , italic_i end_POSTSUBSCRIPT Uniform [ln3,ln50]350[\ln 3,\ln 50][ roman_ln 3 , roman_ln 50 ] [km/s]
loga𝑎\log aroman_log italic_a Normal [lnmax(N(wi)),2]max𝑁subscript𝑤𝑖2[\ln\textup{max}(N(w_{i})),2][ roman_ln max ( italic_N ( italic_w start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ) , 2 ] -
loga2subscript𝑎2\log a_{2}roman_log italic_a start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Uniform [lnmax(N(wi))5,lnmax(N(wi))]max𝑁subscript𝑤𝑖5max𝑁subscript𝑤𝑖[\ln\textup{max}(N(w_{i}))-5,\ln\textup{max}(N(w_{i}))][ roman_ln max ( italic_N ( italic_w start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ) - 5 , roman_ln max ( italic_N ( italic_w start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ) ] -
Table 6: Summary of dynamic priors. The constraints for baryon parameters are referenced to Table 4.
Parameter Distribution Contraints Unit
ρb,isubscript𝜌b𝑖\rho_{\textrm{b},i}italic_ρ start_POSTSUBSCRIPT b , italic_i end_POSTSUBSCRIPT Normal Table 4 Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT
σz,isubscript𝜎𝑧𝑖\sigma_{z,i}italic_σ start_POSTSUBSCRIPT italic_z , italic_i end_POSTSUBSCRIPT Normal Table 4 km s11{}^{-1}start_FLOATSUPERSCRIPT - 1 end_FLOATSUPERSCRIPT
ρDMsubscript𝜌DM\rho_{\textup{\tiny DM}}italic_ρ start_POSTSUBSCRIPT DM end_POSTSUBSCRIPT Uniform [-0.05, 0.15] Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT
lnμ0subscript𝜇0\ln\mu_{0}roman_ln italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Uniform [ln0.1,ln2.00.12.0\ln 0.1,\ln 2.0roman_ln 0.1 , roman_ln 2.0] -
lnν0subscript𝜈0\ln\nu_{0}roman_ln italic_ν start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Normal [lnmax(N(zi))max𝑁subscript𝑧𝑖\ln\textup{max}(\vec{N}(z_{i}))roman_ln max ( over→ start_ARG italic_N end_ARG ( italic_z start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ), 3] -
\mathcal{R}caligraphic_R Normal [3.4, 0.6]×103absentsuperscript103\times 10^{-3}× 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT
z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT Uniform [-150, 150] pc

4.4 MCMC

We make use of the Markov Chain Monte Carlo (MCMC) method to sample the posterior distribution of the parameters θ𝜃\thetaitalic_θ. The MCMC simulations are conducted using hammer-and-sample333https://crates.io/crates/hammer-and-sample by Adam Reichold, a Rust implementation of the affine invariant ensemble sampler introduced by Goodman & Weare (2010), which is a reimplementation of emcee Python library by Foreman-Mackey et al. (2013).

For each group, the MCMC is run for 50,000 steps with 10 ×\times× ndim walkers, where ndim =33absent33=33= 33 for DM or MOND model, and ndim =32absent32=32= 32 for classic NO model. The first 4,000 iterations are discarded as burn-in, and the samples are thinned by a factor of 20.

The MCMC program is implemented in the Rust programming language to enhance performance and reliability. With 33 free parameters for each group in the DM model, spanning 15 groups based on infrared color, and considering 2 gravitational models, a substantial amount of computation is involved. Although the MCMC program is easy to implement, optimizing it can be quite challenging. Rust demonstrates superior performance, outperforming even well-written Python code. While Rust may not be as user-friendly as Python, and its ecosystem is not as mature in Science and Engineering, the significant performance gains justify the effort. Despite this, the program is also ported to Python for the sake of easier data analysis and visualization.

4.5 Solving the Poisson Equation

The solution to the Poisson equation is achieved using the Dormand-Prince method of order 5(4) with dense output of order 4 (Dopri5). This method is implemented in the Rust crate ode_solvers444https://crates.io/crates/ode_solvers by Sylvain Renevey. The choice of this method is driven by its speed, ease of use, and direct integration into Rust. The output from the integration seamlessly interfaces with the rest of the program in Rust, eliminating the need for data structure conversion or cross-language interfaces. The native code compilation enhances the program’s performance blazingly fast.

The general idea of the method involves providing the initial conditions Φ(0)=0Φ00\Phi(0)=0roman_Φ ( 0 ) = 0 and zΦ(0)=0subscript𝑧Φ00\partial_{z}\Phi(0)=0∂ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT roman_Φ ( 0 ) = 0, along with a set of parameters θ𝜃\thetaitalic_θ. The method then performs the integration of the Poisson equation at discrete values of z𝑧zitalic_z, yielding the values of Φ(z)Φ𝑧\Phi(z)roman_Φ ( italic_z ) and zΦ(z)subscript𝑧Φ𝑧\partial_{z}\Phi(z)∂ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT roman_Φ ( italic_z ) at those points. The resulting integration is subsequently interpolated using a simple linear interpolation method to obtain the value of Φ(z)Φ𝑧\Phi(z)roman_Φ ( italic_z ) at any arbitrary z𝑧zitalic_z. This calculated Φ(z)Φ𝑧\Phi(z)roman_Φ ( italic_z ) value is then used to determine the vertical number density of stars using (8).

4.6 Bayes Factor Estimation

The computation of the Bayes factor from the MCMC chain involves employing the approach presented in Robert & Wraith (2009). The harmonic means method is utilized to estimate the evidence term for each MCMC simulation, given by

EN(i=1Nϕ(θi)(θi)p0(θi))1,𝐸𝑁superscriptsuperscriptsubscript𝑖1𝑁italic-ϕsubscript𝜃𝑖subscript𝜃𝑖subscript𝑝0subscript𝜃𝑖1\displaystyle E\approx N\left(\sum_{i=1}^{N}\frac{\phi(\theta_{i})}{\mathcal{L% }(\theta_{i})p_{0}(\theta_{i})}\right)^{-1},italic_E ≈ italic_N ( ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT divide start_ARG italic_ϕ ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG start_ARG caligraphic_L ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) end_ARG ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (25)

where N𝑁Nitalic_N denotes the number of samples obtained from the MCMC simulations, and ϕ(θi)italic-ϕsubscript𝜃𝑖\phi(\theta_{i})italic_ϕ ( italic_θ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) is any arbitrary probability distribution function over θ𝜃\thetaitalic_θ. While a simple harmonic means method estimates the evidence term by selecting ϕ=p0italic-ϕsubscript𝑝0\phi=p_{0}italic_ϕ = italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, it can be risky due to the instability of the approximation (Newton & Raftery 1994). In our case, we implement the uniform distribution ϕ(θ)italic-ϕ𝜃\phi(\theta)italic_ϕ ( italic_θ ) with boundaries around the peak of the posterior distribution p(θ)𝑝𝜃p(\theta)italic_p ( italic_θ ), following the strategy outlined in Robert & Wraith (2009), which stabilizes the approximation. It is crucial to note that for this method to be effective, the proper prior distribution function p0subscript𝑝0p_{0}italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT must be used. The Bayes factor is defined as the ratio between the evidence terms of different models, B12=E1/E2subscript𝐵12subscript𝐸1subscript𝐸2B_{12}=E_{1}/E_{2}italic_B start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_E start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. We calculate the approximation 10 times, then take the mean and the standard deviation as the estimated value.

5   RESULTS AND DISCUSSION

After running the MCMC for each group and model, the results are in the form of a 3-dimensional array with respect to the number of walkers, number of parameters, and number of steps. These results are then flattened (combining all walkers) to obtain the full samples of the posterior distribution functions θ𝜃\thetaitalic_θ. The dark matter density ρDMsubscript𝜌DM\rho_{\textrm{\tiny DM}}italic_ρ start_POSTSUBSCRIPT DM end_POSTSUBSCRIPT and boost interpolating function value μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are extracted by constructing their histograms. Appropriate functions are used to obtain their central values and credible intervals.

5.1 Dark Matter Density

The full set of estimated parameters θ𝜃\thetaitalic_θ is best-fitted to the inferred density profile using MCMC (Fig. 4). This reveals that the contribution of dark matter in the halo is approximately on the order of 0.01M,pc30.01subscriptMdirect-productsuperscriptpc30.01\,\textrm{M}_{\odot},\textrm{pc}^{-3}0.01 M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT , pc start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. The results exhibit fluctuations and are not single-valued, as illustrated in Fig. 5. The figure also displays the inferred total baryonic matter density profile, which aligns with the initial values provided in Table 4. This indicates that the fitting process accurately captures the contribution of the known baryonic matter.

Refer to caption
Figure 4: Density inference (error-bars) and best-fitting curve (shaded area) to the vertical profile density and vertical velocity dispersion for one of the tracer groups (A0-A8). All uncertainties are presented in 68, 90, and 95 percent credible intervals.

In broad terms, the inferred dark matter densities from A-type and G-type stars exhibit agreement, hovering around 0.010.03similar-toabsent0.010.03\sim 0.01-0.03∼ 0.01 - 0.03 Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT, while the value for F-type stars is notably higher, reaching up to 0.07similar-toabsent0.07\sim 0.07∼ 0.07 Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT. The combined dark matter density for all groups is (3.710.36+0.39)×102subscriptsuperscript3.710.390.36superscript102(3.71^{+0.39}_{-0.36})\times 10^{-2}( 3.71 start_POSTSUPERSCRIPT + 0.39 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.36 end_POSTSUBSCRIPT ) × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT, while if F-type stars are excluded, the combined infered dark matter density becomes (1.830.52+0.53)×102subscriptsuperscript1.830.530.52superscript102(1.83^{+0.53}_{-0.52})\times 10^{-2}( 1.83 start_POSTSUPERSCRIPT + 0.53 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.52 end_POSTSUBSCRIPT ) × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT Mdirect-product{}_{\odot}start_FLOATSUBSCRIPT ⊙ end_FLOATSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT. Theoretically, the dark matter density should be consistent across all spectral types, as the dark matter gravitational potential contribution is expected to be independent of the spectral type of stars. The implication of this result could stem from either the unsuitability of the dark matter model employed in this study or the violation of some assumptions made in the analysis. Let us discuss the possible causes of this result.

Firstly, it’s important to note that the dark matter model employed in this study is relatively simple, assuming a constant density up to 200 pc above and below the galactic plane. However, simulations of Milky Way-like galaxies, as indicated by Garbari et al. (2011), suggest that the dark matter density remains approximately constant up to 1 kpc above and below the galactic plane, even after 4 Gyrs of evolution. While the simulation may not perfectly reflect the real Galaxy, it provides valuable insights into the nature of dark matter distribution. Consequently, it seems unlikely that refining the dark matter density model would significantly alter the obtained results.

Secondly, among the various assumptions made in this study, two stand out as potential contributors to the observed peculiar results. The first assumption involves the baryon mass distribution, following the Bahcall model, which consists of several isothermal components. The isothermality assumption implies that the velocity dispersion remains constant up to 200 pc. Fortunately, simulations of Milky Way-like galaxies, as suggested by Garbari et al. (2011), indicate that the velocity dispersion is approximately constant up to 1 kpc above and below the galactic plane. Thus, this assumption is likely valid. The second assumption that raises suspicion is the assumption of dynamical equilibrium, implying that the stars under consideration maintain the same distribution function after years of interaction and mixing. Achieving dynamical equilibrium requires that the stars are old enough to interact and mix but not so old that they have been evaporated from the galaxy. Furthermore, dynamical equilibrium can only be sustained if there are no significant perturbations to the system, such as a merger with other galaxies (Malhan et al. 2022). This might be the primary cause of the fluctuating results in the inferred dark matter density for different groups of stars within distinct color ranges.

The lack of convergence in the results for dark matter density strongly suggests that the sampled data is not in dynamical equilibrium. This finding aligns with the results reported by Buch et al. (2019), indicating that the dark matter density differs for various spectral types of stars. To address this issue, a modification to the collisionless Boltzmann equation may be necessary to account for the effects of disequilibrium. One potential approach involves introducing external wave-like perturbations to the distribution function, as proposed by (Banik et al. 2017). This presents an intriguing avenue for future research.

Refer to caption
Figure 5: Inferred dark matter density for each group of stars. The vertical dashed line separates different spectral types, while the horizontal dashed line represents the initial total baryonic matter density from Table 4, with the gray diagonal shaded area indicating the uncertainty. The red error bar plot represents the inferred baryonic matter density, and the blue error bar plot depicts the inferred dark matter density from the fitting process. All uncertainties are presented in 68, 90, and 95 per cent credible intervals.

5.2 Acceleration Parameters

Fitting the data without including dark matter contribution reveals that the inferred baryonic matter density consistently exceeds the initial value from Table 4. This suggests the presence of another matter component, such as dark matter, or the possibility that the additional matter is an effect of modified gravity.

Equation (17) shows the relationship between baryonic matter and the additional μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT boost from MOND theory. Our analysis yields a value of μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT around 0.64±0.03similar-toabsentplus-or-minus0.640.03\sim 0.64\pm 0.03∼ 0.64 ± 0.03 for the Sun’s location in the galaxy, without delving into the specifics of MOND theory.

A similar method to test MOND theory was conducted by Lisanti et al. (2019), revealing that the value of the ν𝜈\nuitalic_ν-interpolating function is 1.440.08+0.13subscriptsuperscript1.440.130.081.44^{+0.13}_{-0.08}1.44 start_POSTSUPERSCRIPT + 0.13 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.08 end_POSTSUBSCRIPT. Our equivalent ν𝜈\nuitalic_ν-interpolating function value is 1.56±0.06plus-or-minus1.560.061.56\pm 0.061.56 ± 0.06, which falls within the uncertainty range. This suggests that our analysis results are consistent with their findings.

MOND theory places particular emphasis on determining the value of the acceleration parameter a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Converting the value of μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT into acceleration parameters a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT requires a specific interpolating function. We use two interpolating functions: the simple one and the standard one, as outlined in Table 3. The resulting value of a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for the simple interpolating function is approximately (1.260.13+0.14)×1010subscriptsuperscript1.260.140.13superscript1010(1.26^{+0.14}_{-0.13})\times 10^{-10}( 1.26 start_POSTSUPERSCRIPT + 0.14 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.13 end_POSTSUBSCRIPT ) × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT m s22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT, while for the standard interpolating function, it is around (2.690.17+0.19)×1010subscriptsuperscript2.690.190.17superscript1010(2.69^{+0.19}_{-0.17})\times 10^{-10}( 2.69 start_POSTSUPERSCRIPT + 0.19 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.17 end_POSTSUBSCRIPT ) × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPTm s22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT. Fig. 6 shows the probability density function for simple and standard interpolating function. An intriguing observation is that the values of a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for these two interpolating functions differ significantly. Upon closer inspection, it is noted that the uncertainties of both results do not overlap, suggesting the potential to distinguish between the two models.

Previous studies by Gentile et al. (2011) and Begeman et al. (1991) found that the value of a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is approximately 1.2×10101.2superscript10101.2\times 10^{-10}1.2 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT m s22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT by fitting the rotation curve of galaxies. Using the methodology of our approach, constraining the gravitational potential by proxy of the local distribution function, we found a similar value. If MOND is valid, the simple interpolating function is more likely to be true to achieve the same result from a global galactic-scale approach.

Refer to caption
Figure 6: The acceleration parameter probability density derived from the infered μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. The gray curve in inset plot are the infered μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for each groups and dark bold curve is the combined values for all groups, around 0.640.640.640.64.

Adopting the standard interpolating function, the result of a01.2×1010similar-tosubscript𝑎01.2superscript1010a_{0}\sim 1.2\times 10^{-10}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ 1.2 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT m s22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT is consistent with other studies employing MOND on Galactic scales. This finding indicates that even in the solar neighborhood, the prospect to observe the MOND effect is significant and compelling. Further comprehensive investigations on this scale could shed light on the true nature of MOND and Dark Matter.

5.3 Comparisons

Bayes factors (BF) serve as a tool for comparing two or more models based on the observed data. The Bayes factor represents the integrated posterior probability of a model across all potential parameter values, often referred to as the “evidence” term in Bayesian statistics. A higher Bayes factor indicates that a model is more suitable for explaining the observed data.

The three hypotheses being compared are Mno=subscript𝑀noabsentM_{\textrm{no}}=italic_M start_POSTSUBSCRIPT no end_POSTSUBSCRIPT = (Newtonian Dynamics) with no dark matter, Mdm=subscript𝑀dmabsentM_{\textrm{dm}}=italic_M start_POSTSUBSCRIPT dm end_POSTSUBSCRIPT = Dark Matter, and Mmond=subscript𝑀mondabsentM_{\textrm{mond}}=italic_M start_POSTSUBSCRIPT mond end_POSTSUBSCRIPT = MOND. Fig. 7 presents the complete results. The mean log\logroman_log BF values with respect to Mno=subscript𝑀noabsentM_{\textrm{no}}=italic_M start_POSTSUBSCRIPT no end_POSTSUBSCRIPT = for all groups of stars are 1.801.801.801.80 (dm-no) and 1.701.701.701.70 (mond-no), suggesting that any additional matter from the baryonic dark matter hypothesis (no) is strongly unlikely. This trend is more pronounced for stars of type F, where the value spikes to 6666. Apart from these spikes, the log\logroman_log BF values are close to zero, indicating that the three hypotheses are equally likely to be true. The only strong conclusion that can be drawn is Newtonian dynamics without including non-baryonic dark matter is not feasible.

Compare to the Bayes factor between the Dark Matter and MOND hypotheses; the two are neck and neck. The log\logroman_log BF fluctuates around zero, with a mean of only 0.10 in favor of dark matter. This result suggests that both hypotheses are equally plausible in explaining the data we have. From our analysis, we cannot definitively say which hypothesis is more favored.

One way to better discriminate between the two hypotheses is by incorporating more data from stars at high altitudes (|z|>200𝑧200|z|>200| italic_z | > 200 pc). This approach aims to reveal potential differences in the predictions of the distribution function between the two hypotheses. However, a challenge arises for large altitudes as the assumption that the distribution function is separable becomes invalid. Consequently, solving the full form of the distribution function is necessary (see (3)). This poses a complex and challenging problem to address.

Other studies, such as the one conducted by Lisanti et al. (2019), have employed statistical methods to compare the Dark Matter and MOND hypotheses, specifically by calculating the Bayesian Information Criterion (BIC). Their findings strongly favor the Dark Matter hypothesis based on the difference in BIC values (ΔBICΔBIC\Delta\textup{BIC}roman_Δ BIC), which contrasts with our results. In our analysis, the Bayes factors for the Dark Matter hypothesis are only marginally higher than those for the MOND hypothesis. However, it is important to note that our results do not suggest strong evidence against the MOND hypothesis, with mean log\logroman_log BF values of 0.10.10.10.1. According to Kass & Raftery (1995), a log Bayes factor of less than 0.5 is not even worth more than a bare mention. More data and robust methods are needed to provide conclusive answers.

Refer to caption
Figure 7: Bayes factors for each group of stars for the three hypotheses. (Top) Bayes factors of Dark Matter and MOND hypotheses with respect to NO hypothesis. The bayes factor for MOND-NO points are slightly shifted to the left for clarity.(Bottom) Bayes factors comparing the Dark Matter hypothesis to the MOND hypothesis.

6   CONCLUSION

In this study, we have developed a method to estimate the dark matter density in the solar neighbourhood by leveraging the vertical distribution of stars obtained from Gaia, RAVE, APOGEE, GALAH, and LAMOST catalogues. The data collected from these various catalogues are dominated by Gaia DR3. In principle, the analysis done in this study is unaffected by the addition of other catalogues.

Our approach utilises both the vertical height distribution and vertical velocity distribution of stars as the vertical distribution function. In turn, the distribution function is directly correlated with the Galactic gravitational potential through CBE, which dictates the mass model for our hypotheses. By proxy of the distribution function, the dark matter density and the acceleration parameter a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is inferred.

Our analysis reveals a dark matter density in the solar neighbourhood of approximately (0.010.03) M0.010.03subscript Mdirect-product(0.01\textrm{--}0.03)\textrm{ M}_{\odot}( 0.01 – 0.03 ) M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT pc33{}^{-3}start_FLOATSUPERSCRIPT - 3 end_FLOATSUPERSCRIPT, and an acceleration parameter a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of 1.260.13+0.14×1010subscriptsuperscript1.260.140.13superscript10101.26^{+0.14}_{-0.13}\times 10^{-10}1.26 start_POSTSUPERSCRIPT + 0.14 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT - 0.13 end_POSTSUBSCRIPT × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT m s22{}^{-2}start_FLOATSUPERSCRIPT - 2 end_FLOATSUPERSCRIPT for a simple interpolating function. These fluctuations could be the results of disequilibria of tracers being used, suggesting challenges to improve the approach for more robust results in future studies.

Bayesian analysis, through the calculation of Bayes factors, presents strong evidence against classical Newtonian gravity without additional missing mass. However, when comparing the two competing hypotheses in explaining the missing mass, we found a marginally higher likelihood for the dark matter hypothesis compared to the Modified Newtonian Dynamics (MOND) hypothesis, with logBF=0.1BF0.1\log\textrm{BF}=0.1roman_log BF = 0.1. It is important to note, however, that the Bayes factors provide no substantial support for the dark matter hypothesis, falling short of conclusive evidence to reject the MOND hypothesis outright.

Acknowledgements

We thanked the Faculty of Mathematics and Natural Sciences of Bandung Institute of Technology for financial support via the PPMI 2022 funding program.

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

This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation.

Funding for Rave has been provided by: the Leibniz Institute for Astrophysics Potsdam (AIP); the Australian Astronomical Observatory; the Australian National University; the Australian Research Council; the French National Research Agency; the German Research Foundation (SPP 1177 and SFB 881); the European Research Council (ERC-StG 240271 Galactica); the Istituto Nazionale di Astrofisica at Padova; the Johns Hopkins University; the National Science Foundation of the USA (AST-0908326); the W. M. Keck foundation; the Macquarie University; the Netherlands Research School for Astronomy; the Natural Sciences and Engineering Research Council of Canada; the Slovenian Research Agency; the Swiss National Science Foundation; the Science & Technology Facilities Council of the UK; Opticon; Strasbourg Observatory; the Universities of Basel, Groningen, Heidelberg and Sydney. This work has made use of data from the European Space Agency (ESA) mission Gaia (https://www.cosmos.esa.int/gaia), processed by the Gaia Data Processing and Analysis Consortium (DPAC, https://www.cosmos.esa.int/web/gaia/dpac/consortium). Funding for the DPAC has been provided by national institutions, in particular the institutions participating in the Gaia Multilateral Agreement.

Guoshoujing Telescope (the Large Sky Area Multi-Object Fiber Spectroscopic Telescope LAMOST) is a National Major Scientific Project built by the Chinese Academy of Sciences. Funding for the project has been provided by the National Development and Reform Commission. LAMOST is operated and managed by the National Astronomical Observatories, Chinese Academy of Sciences.

This work made use of the Third Data Release of the GALAH Survey (Buder et al. 2021). The GALAH Survey is based on data acquired through the Australian Astronomical Observatory, under programs: A/2013B/13 (The GALAH pilot survey); A/2014A/25, A/2015A/19, A2017A/18 (The GALAH survey phase 1); A2018A/18 (Open clusters with HERMES); A2019A/1 (Hierarchical star formation in Ori OB1); A2019A/15 (The GALAH survey phase 2); A/2015B/19, A/2016A/22, A/2016B/10, A/2017B/16, A/2018B/15 (The HERMES-TESS program); and A/2015A/3, A/2015B/1, A/2015B/19, A/2016A/22, A/2016B/12, A/2017A/14 (The HERMES K2-follow-up program). We acknowledge the traditional owners of the land on which the AAT stands, the Gamilaraay people, and pay our respects to elders past and present. This paper includes data that has been provided by AAO Data Central (datacentral.org.au).

Funding for the Sloan Digital Sky Survey V has been provided by the Alfred P. Sloan Foundation, the Heising-Simons Foundation, the National Science Foundation, and the Participating Institutions. SDSS acknowledges support and resources from the Center for High-Performance Computing at the University of Utah. The SDSS web site is www.sdss.org.

SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions of the SDSS Collaboration, including the Carnegie Institution for Science, Chilean National Time Allocation Committee (CNTAC) ratified researchers, the Gotham Participation Group, Harvard University, Heidelberg University, The Johns Hopkins University, L’Ecole polytechnique fédérale de Lausanne (EPFL), Leibniz-Institut für Astrophysik Potsdam (AIP), Max-Planck-Institut für Astronomie (MPIA Heidelberg), Max-Planck-Institut für Extraterrestrische Physik (MPE), Nanjing University, National Astronomical Observatories of China (NAOC), New Mexico State University, The Ohio State University, Pennsylvania State University, Smithsonian Astrophysical Observatory, Space Telescope Science Institute (STScI), the Stellar Astrophysics Participation Group, Universidad Nacional Autónoma de México, University of Arizona, University of Colorado Boulder, University of Illinois at Urbana-Champaign, University of Toronto, University of Utah, University of Virginia, Yale University, and Yunnan University.

Data Availability

The original raw data are available from publicly availabe source at their respective repositories. The reduced data underlying this article will be shared upon reasonable request to the corresponding author.

References

  • Rubin et al. [1980] V. C. Rubin, Jr. Ford, W. K., and N. Thonnard. Rotational properties of 21 SC galaxies with a large range of luminosities and radii, from NGC 4605 (R=4kpc) to UGC 2885 (R=122kpc). ApJ, 238:471–487, June 1980. doi: 10.1086/158003.
  • Read [2014] J. I. Read. The local dark matter density. J. Phys. G: Nucl. Phys, 41(6):063101, June 2014. doi: 10.1088/0954-3899/41/6/063101.
  • Milgrom [2001] Mordehai Milgrom. The shape of ‘dark matter’ haloes of disc galaxies according to mond. MNRAS, 326(4):1261–1264, 2001. doi: 10.1111/j.1365-2966.2001.04653.x.
  • Nipoti et al. [2007] Carlo Nipoti, Pasquale Londrillo, HongSheng Zhao, and Luca Ciotti. Vertical dynamics of disc galaxies in modified Newtonian dynamics. MNRAS, 379(2):597–604, August 2007. doi: 10.1111/j.1365-2966.2007.11835.x.
  • Oort [1932] J. H. Oort. The force exerted by the stellar system in the direction perpendicular to the galactic plane and some related problems. Bull. Astron. Inst. Neth., 6:249, August 1932.
  • Zwicky [1933] F. Zwicky. Die Rotverschiebung von extragalaktischen Nebeln. Helv. Phys. Acta, 6:110–127, January 1933.
  • Ostriker and Peebles [1973] J. P. Ostriker and P. J. E. Peebles. A numerical study of the stability of flattened galaxies: or, can cold galaxies survive? ApJ, 186:467–480, December 1973. doi: 10.1086/152513.
  • Lin and Ishak [2016] Weikang Lin and Mustapha Ishak. Ultra faint dwarf galaxies: an arena for testing dark matter versus modified gravity. JCAP, 2016(10):025, October 2016. doi: 10.1088/1475-7516/2016/10/025.
  • Davis [2000] M. Davis. The cosmological matter density. Phys. Rep., 333:147–165, August 2000. doi: 10.1016/S0370-1573(00)00020-X.
  • Bahcall [1984a] J. N. Bahcall. K giants and the total amount of matter near the sun. ApJ, 287:926–944, December 1984a. doi: 10.1086/162750.
  • Moni Bidin et al. [2012] C. Moni Bidin, G. Carraro, R. A. Méndez, and R. Smith. Kinematical and Chemical Vertical Structure of the Galactic Thick Disk. II. A Lack of Dark Matter in the Solar Neighborhood. ApJ, 751(1):30, May 2012. doi: 10.1088/0004-637X/751/1/30.
  • Garbari et al. [2011] Silvia Garbari, Justin I Read, and George Lake. Limits on the local dark matter density. MNRAS, 416(3):2318–2340, 2011. doi: 10.1051/epjconf/20121901008.
  • Zhang et al. [2013] Lan Zhang, Hans-Walter Rix, Glenn van de Ven, Jo Bovy, Chao Liu, and Gang Zhao. The gravitational potential near the sun from segue k-dwarf kinematics. ApJ, 772(2):108, August 2013. doi: 10.1088/0004-637X/772/2/108.
  • Bienaymé et al. [2014] O. Bienaymé, B. Famaey, A. Siebert, K. C. Freeman, B. K. Gibson, G. Gilmore, E. K. Grebel, J. Bland-Hawthorn, G. Kordopatis, U. Munari, J. F. Navarro, Q. Parker, W. Reid, G. M. Seabroke, A. Siviero, M. Steinmetz, F. Watson, R. F. G. Wyse, and T. Zwitter. Weighing the local dark matter with RAVE red clump stars. A& A, 571:A92, November 2014. doi: 10.1051/0004-6361/201424478.
  • Xia et al. [2016] Qiran Xia, Chao Liu, Shude Mao, Yingyi Song, Lan Zhang, RJ Long, Yong Zhang, Yonghui Hou, Yuefei Wang, and Yue Wu. Determining the local dark matter density with lamost data. MNRAS, 458(4):3839–3850, 2016. doi: 10.1093/mnras/stw565.
  • Sivertsson et al. [2018] Sofia Sivertsson, Hamish Silverwood, Justin I Read, Gianfranco Bertone, and Pascal Steger. The localdark matter density from sdss-segue g-dwarfs. MNRAS, 478(2):1677–1693, 2018. doi: 10.1093/mnras/sty977.
  • Buch et al. [2019] Jatan Buch, Shing Chau John Leung, and JiJi Fan. Using gaia dr2 to constrain local dark matter density and thin dark disk. JCAP, 2019(04):026, 2019. doi: 10.1088/1475-7516/2019/04/026.
  • Milgrom [1983] M. Milgrom. A modification of the Newtonian dynamics as a possible alternative to the hidden mass hypothesis. ApJ, 270:365–370, July 1983. doi: 10.1086/161130.
  • Famaey and McGaugh [2012] Benoit Famaey and Stacy McGaugh. Modified Newtonian Dynamics (MOND): Observational Phenomenology and Relativistic Extensions. Living rev. relativ., 15:10, 2012. doi: 10.12942/lrr-2012-10.
  • Milgrom [1994] M. Milgrom. Dynamics with a nonstandard inertia-acceleration relation: An alternative to dark matter in galactic systems. Ann. Phys., 229(2):384–415, 1994. ISSN 0003-4916. doi: https://doi.org/10.1006/aphy.1994.1012. URL https://www.sciencedirect.com/science/article/pii/S0003491684710128.
  • Lisanti et al. [2019] Mariangela Lisanti, Matthew Moschella, Nadav Joseph Outmezguine, and Oren Slone. Testing dark matter and modifications to gravity using local milky way observables. Phys. Rev. D, 100(8):083009, 2019. doi: 10.1103/PhysRevD.100.083009.
  • Hernandez [2023] X. Hernandez. Internal kinematics of Gaia DR3 wide binaries: anomalous behaviour in the low acceleration regime. MNRAS, 525(1):1401–1415, October 2023. doi: 10.1093/mnras/stad2306.
  • Chae [2023] Kyu-Hyun Chae. Breakdown of the Newton-Einstein Standard Gravity at Low Acceleration in Internal Dynamics of Wide Binary Stars. ApJ, 952(2):128, August 2023. doi: 10.3847/1538-4357/ace101.
  • Pittordis and Sutherland [2023] Charalambos Pittordis and Will Sutherland. Wide Binaries from GAIA EDR3: preference for GR over MOND? OJAp, 6:4, February 2023. doi: 10.21105/astro.2205.02846.
  • Banik et al. [2023] Indranil Banik, Charalambos Pittordis, Will Sutherland, Benoit Famaey, Rodrigo Ibata, Steffen Mieske, and Hongsheng Zhao. Strong constraints on the gravitational law from Gaia DR3 wide binaries. MNRAS, November 2023. doi: 10.1093/mnras/stad3393.
  • Holmberg and Flynn [2004] Johan Holmberg and Chris Flynn. The local surface density of disc matter mapped by Hipparcos. MNRAS, 352(2):440–446, August 2004. doi: 10.1111/j.1365-2966.2004.07931.x.
  • Bovy [2017a] Jo Bovy. Stellar inventory of the solar neighbourhood using gaia dr1. MNRAS, 470(2):1360–1387, 2017a. doi: 10.1093/mnras/stx1277.
  • Marrese, P. M. et al. [2019] Marrese, P. M., Marinoni, S., Fabrizio, M., and Altavilla, G. Gaia data release 2 - cross-match with external catalogues: algorithms and results. A&A, 621:A144, 2019. doi: 10.1051/0004-6361/201834142. URL https://doi.org/10.1051/0004-6361/201834142.
  • Kim and Lépine [2021] Bokyoung Kim and Sebastien Lépine. Stars in the local galactic thick disc and halo in Gaia EDR3: a catalogue of half a million local main-sequence stars with photometric metallicities. MNRAS, 510(3):4308–4329, 12 2021. ISSN 0035-8711. doi: 10.1093/mnras/stab3671.
  • Pecaut and Mamajek [2013] Mark J Pecaut and Eric E Mamajek. Intrinsic colors, temperatures, and bolometric corrections of pre-main-sequence stars. ApJS, 208(1):9, 2013. doi: 10.1088/0067-0049/208/1/9.
  • Bovy [2017b] Jo Bovy. Galactic rotation in gaia dr1. MNRAS, 468(1):L63–L67, 2017b. doi: 10.1093/mnrasl/slx027.
  • Bekenstein and Milgrom [1984] J. Bekenstein and M. Milgrom. Does the missing mass problem signal the breakdown of Newtonian gravity? ApJ, 286:7–14, November 1984. doi: 10.1086/162570.
  • Navarro et al. [1997] Julio F. Navarro, Carlos S. Frenk, and Simon D. M. White. A Universal Density Profile from Hierarchical Clustering. ApJ, 490(2):493–508, December 1997. doi: 10.1086/304888.
  • Bahcall [1984b] J. N. Bahcall. The distribution of stars perpendicular to galactic disk. ApJ, 276:156–168, January 1984b. doi: 10.1086/161600.
  • Bahcall [1984c] J. N. Bahcall. Self-consistent determinations of the total amount of matter near the sun. ApJ, 276:169–181, January 1984c. doi: 10.1086/161601.
  • Bailer-Jones [2015] Coryn A. L. Bailer-Jones. Estimating Distances from Parallaxes. PASP, 127(956):994, October 2015. doi: 10.1086/683116.
  • Goodman and Weare [2010] Jonathan Goodman and Jonathan Weare. Ensemble samplers with affine invariance. Commun. appl. math. comput. sci., 5(1):65–80, 2010. doi: 10.2140/camcos.2010.5.65.
  • Foreman-Mackey et al. [2013] Daniel Foreman-Mackey, David W Hogg, Dustin Lang, and Jonathan Goodman. emcee: the mcmc hammer. PASP, 125(925):306, 2013. doi: 10.1086/670067.
  • Robert and Wraith [2009] C. P. Robert and D. Wraith. Computational methods for Bayesian model choice. In Paul M. Goggans and Chun-Yong Chan, editors, Bayesian Inference and Maximum Entropy Methods in Science and Engineering: The 29th International Workshop on Bayesian Inference and Maximum Entropy Methods in Science and Engineering, volume 1193 of American Institute of Physics Conference Series, pages 251–262, December 2009. doi: 10.1063/1.3275622.
  • Newton and Raftery [1994] Michael A. Newton and Adrian E. Raftery. Approximate bayesian inference with the weighted likelihood bootstrap. J. R. Stat. Soc., B (Methodol.), 56(1):3–48, 1994. ISSN 00359246. URL http://www.jstor.org/stable/2346025.
  • Malhan et al. [2022] Khyati Malhan, Rodrigo A Ibata, Sanjib Sharma, Benoit Famaey, Michele Bellazzini, Raymond G Carlberg, Richard D’souza, Zhen Yuan, Nicolas F Martin, and Guillaume F Thomas. The global dynamical atlas of the milky way mergers: Constraints from gaia edr3–based orbits of globular clusters, stellar streams, and satellite galaxies. ApJ, 926(2):107, 2022.
  • Banik et al. [2017] Nilanjan Banik, Lawrence M. Widrow, and Scott Dodelson. Galactoseismology and the local density of dark matter. MNRAS, 464(4):3775–3783, February 2017. doi: 10.1093/mnras/stw2603.
  • Gentile et al. [2011] Gianfranco Gentile, B Famaey, and WJG de Blok. Things about mond. A&J, 527:A76, 2011. doi: 10.1051/0004-6361/201015283.
  • Begeman et al. [1991] K. G. Begeman, A. H. Broeils, and R. H. Sanders. Extended rotation curves of spiral galaxies : dark haloes and modified dynamics. MNRAS, 249:523, April 1991. doi: 10.1093/mnras/249.3.523.
  • Kass and Raftery [1995] Robert E. Kass and Adrian E. Raftery. Bayes factors. J. Am. Stat. Assoc., 90(430):773–795, 1995. doi: 10.1080/01621459.1995.10476572. URL https://www.tandfonline.com/doi/abs/10.1080/01621459.1995.10476572.

Appendix A Density Inference

Let x𝑥xitalic_x be an observed value with some uncertainty sampled from a distribution 𝒯𝒯\mathcal{T}caligraphic_T, with mean μ𝜇\muitalic_μ and standard deviation σ𝜎\sigmaitalic_σ. The mean value μ𝜇\muitalic_μ can be inferred using Bayes’ theorem:

p(μ|x,σ)𝑝conditional𝜇𝑥𝜎\displaystyle p(\mu|x,\sigma)italic_p ( italic_μ | italic_x , italic_σ ) =p(x|μ,σ)p0(μ|σ)E(x,σ)absent𝑝conditional𝑥𝜇𝜎subscript𝑝0conditional𝜇𝜎𝐸𝑥𝜎\displaystyle=\frac{p(x|\mu,\sigma)p_{0}(\mu|\sigma)}{E(x,\sigma)}= divide start_ARG italic_p ( italic_x | italic_μ , italic_σ ) italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_μ | italic_σ ) end_ARG start_ARG italic_E ( italic_x , italic_σ ) end_ARG
𝒯(x|μ,σ),proportional-toabsent𝒯conditional𝑥𝜇𝜎\displaystyle\propto\mathcal{T}(x|\mu,\sigma),∝ caligraphic_T ( italic_x | italic_μ , italic_σ ) , (26)

where p0(μ|σ)1proportional-tosubscript𝑝0conditional𝜇𝜎1p_{0}(\mu|\sigma)\propto 1italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_μ | italic_σ ) ∝ 1 is assumed. If there is a set of observed values {𝒙,𝝈}𝒙𝝈\{\bm{x},\bm{\sigma}\}{ bold_italic_x , bold_italic_σ } that are independent and identically distributed (iid), the probability density function of the mean value 𝝁𝝁\bm{\mu}bold_italic_μ is

p(𝝁|𝒙,𝝈)𝑝conditional𝝁𝒙𝝈\displaystyle p(\bm{\mu}|\bm{x},\bm{\sigma})italic_p ( bold_italic_μ | bold_italic_x , bold_italic_σ ) i=1N𝒯(xi|μi,σi).proportional-toabsentsuperscriptsubscriptproduct𝑖1𝑁𝒯conditionalsubscript𝑥𝑖subscript𝜇𝑖subscript𝜎𝑖\displaystyle\propto\prod_{i=1}^{N}\mathcal{T}(x_{i}|\mu_{i},\sigma_{i}).∝ ∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT caligraphic_T ( italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) . (27)

Consider an interval x𝑥xitalic_x and x+Δx𝑥Δ𝑥x+\Delta xitalic_x + roman_Δ italic_x. Due to the noisy nature of the data, the number of observed values in that interval is a random variable. Let N𝑁Nitalic_N be the number of data points with mean μ𝜇\muitalic_μ value within the interval. The probability density function of N𝑁Nitalic_N is given by

p(N|𝒙,𝝈)𝑝conditional𝑁𝒙𝝈\displaystyle p(N|\bm{x},\bm{\sigma})italic_p ( italic_N | bold_italic_x , bold_italic_σ ) =p(N,𝝁|𝒙,𝝈)d𝝁absent𝑝𝑁conditional𝝁𝒙𝝈d𝝁\displaystyle=\int p(N,\bm{\mu}|\bm{x},\bm{\sigma})\textrm{d}\bm{\mu}= ∫ italic_p ( italic_N , bold_italic_μ | bold_italic_x , bold_italic_σ ) d bold_italic_μ
=p(N|𝝁,𝒙,𝝈)p(𝝁|𝒙,𝝈)d𝝁absent𝑝conditional𝑁𝝁𝒙𝝈𝑝conditional𝝁𝒙𝝈d𝝁\displaystyle=\int p(N|\bm{\mu},\bm{x},\bm{\sigma})p(\bm{\mu}|\bm{x},\bm{% \sigma})\textrm{d}\bm{\mu}= ∫ italic_p ( italic_N | bold_italic_μ , bold_italic_x , bold_italic_σ ) italic_p ( bold_italic_μ | bold_italic_x , bold_italic_σ ) d bold_italic_μ
=p(N|𝝁)p(𝝁|𝒙,𝝈)d𝝁absent𝑝conditional𝑁𝝁𝑝conditional𝝁𝒙𝝈d𝝁\displaystyle=\int p(N|\bm{\mu})p(\bm{\mu}|\bm{x},\bm{\sigma})\textrm{d}\bm{\mu}= ∫ italic_p ( italic_N | bold_italic_μ ) italic_p ( bold_italic_μ | bold_italic_x , bold_italic_σ ) d bold_italic_μ
=𝕀(𝝁)p(𝝁|𝒙,𝝈)d𝝁,absent𝕀𝝁𝑝conditional𝝁𝒙𝝈d𝝁\displaystyle=\int\mathbb{I}(\bm{\mu})p(\bm{\mu}|\bm{x},\bm{\sigma})\textrm{d}% \bm{\mu},= ∫ blackboard_I ( bold_italic_μ ) italic_p ( bold_italic_μ | bold_italic_x , bold_italic_σ ) d bold_italic_μ , (28)

where 𝕀(𝝁)𝕀𝝁\mathbb{I}(\bm{\mu})blackboard_I ( bold_italic_μ ) is one when the combination of 𝝁𝝁\bm{\mu}bold_italic_μ results in N𝑁Nitalic_N observed values in the interval, and zero otherwise. In reality, the random variable μ𝜇\muitalic_μ is distributed according to a Poisson distribution. Let ΛΛ\Lambdaroman_Λ be the expected number of observed values of μ𝜇\muitalic_μ in the interval; then, the probability density function of ΛΛ\Lambdaroman_Λ is given by

p(Λ|𝒙,𝝈)𝑝conditionalΛ𝒙𝝈\displaystyle p(\Lambda|\bm{x},\bm{\sigma})italic_p ( roman_Λ | bold_italic_x , bold_italic_σ ) =p(Λ,Ni|𝒙,𝝈)absent𝑝Λconditionalsubscript𝑁𝑖𝒙𝝈\displaystyle=\sum p(\Lambda,N_{i}|\bm{x},\bm{\sigma})= ∑ italic_p ( roman_Λ , italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | bold_italic_x , bold_italic_σ )
=p(Λ|Ni,𝒙,𝝈)p(Ni|𝒙,𝝈)absent𝑝conditionalΛsubscript𝑁𝑖𝒙𝝈𝑝conditionalsubscript𝑁𝑖𝒙𝝈\displaystyle=\sum p(\Lambda|N_{i},\bm{x},\bm{\sigma})p(N_{i}|\bm{x},\bm{% \sigma})= ∑ italic_p ( roman_Λ | italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , bold_italic_x , bold_italic_σ ) italic_p ( italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | bold_italic_x , bold_italic_σ )
=p(Λ|Ni)𝕀(𝝁)p(𝝁|𝒙,𝝈)d𝝁absent𝑝conditionalΛsubscript𝑁𝑖𝕀𝝁𝑝conditional𝝁𝒙𝝈d𝝁\displaystyle=\sum p(\Lambda|N_{i})\int\mathbb{I}(\bm{\mu})p(\bm{\mu}|\bm{x},% \bm{\sigma})\textrm{d}\bm{\mu}= ∑ italic_p ( roman_Λ | italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ∫ blackboard_I ( bold_italic_μ ) italic_p ( bold_italic_μ | bold_italic_x , bold_italic_σ ) d bold_italic_μ
=W(Ni)p(Λ|Ni),absent𝑊subscript𝑁𝑖𝑝conditionalΛsubscript𝑁𝑖\displaystyle=\sum W(N_{i})p(\Lambda|N_{i}),= ∑ italic_W ( italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_p ( roman_Λ | italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (29)

where p(Λ|Ni)p(Ni|Λ)proportional-to𝑝conditionalΛsubscript𝑁𝑖𝑝conditionalsubscript𝑁𝑖Λp(\Lambda|N_{i})\propto p(N_{i}|\Lambda)italic_p ( roman_Λ | italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ∝ italic_p ( italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT | roman_Λ ) is the probability density function of the Poisson distribution with mean ΛΛ\Lambdaroman_Λ given an uninformative prior, and W(Ni)𝑊subscript𝑁𝑖W(N_{i})italic_W ( italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) is the weight of finding Nisubscript𝑁𝑖N_{i}italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT observed values in the interval. The weight W(Ni)𝑊subscript𝑁𝑖W(N_{i})italic_W ( italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) is extremely hard to calculate from the integral in equation (28), so in practice we use the following approximation:

W(Ni)1Sj=1S𝕀(𝝁j),𝑊subscript𝑁𝑖1𝑆superscriptsubscript𝑗1𝑆𝕀subscript𝝁𝑗\displaystyle W(N_{i})\approx\frac{1}{S}\sum_{j=1}^{S}\mathbb{I}(\bm{\mu}_{j}),italic_W ( italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ≈ divide start_ARG 1 end_ARG start_ARG italic_S end_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_S end_POSTSUPERSCRIPT blackboard_I ( bold_italic_μ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) , (30)

where we sample 𝝁𝝁\bm{\mu}bold_italic_μ from p(𝝁|𝒙,𝝈)𝑝conditional𝝁𝒙𝝈p(\bm{\mu}|\bm{x},\bm{\sigma})italic_p ( bold_italic_μ | bold_italic_x , bold_italic_σ ) S𝑆Sitalic_S times.

Equation (29) is computationally inefficient due to the large number of possible values for Nisubscript𝑁𝑖N_{i}italic_N start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. To address this, we defined a new distribution function called the modified Poisson distribution to approximate the probability density function of ΛΛ\Lambdaroman_Λ:

Poisson*(N|Λ)=ΛNeΛΓ(N+1),superscriptPoissonconditional𝑁ΛsuperscriptΛ𝑁superscript𝑒ΛΓ𝑁1\displaystyle\textrm{Poisson}^{*}(N|\Lambda)=\frac{\Lambda^{N}e^{-\Lambda}}{% \Gamma(N+1)},Poisson start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( italic_N | roman_Λ ) = divide start_ARG roman_Λ start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT - roman_Λ end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_N + 1 ) end_ARG , (31)

where Γ(x)Γ𝑥\Gamma(x)roman_Γ ( italic_x ) is the gamma function. Our tests show that the modified Poisson distribution is a good approximation for equation (29). Thus, we express the probability density function of ΛΛ\Lambdaroman_Λ as

p(Λ|𝒙,𝝈)𝑝conditionalΛ𝒙𝝈\displaystyle p(\Lambda|\bm{x},\bm{\sigma})italic_p ( roman_Λ | bold_italic_x , bold_italic_σ ) Poisson*(Neff|Λ),absentsuperscriptPoissonconditionalsubscript𝑁effΛ\displaystyle\approx\textrm{Poisson}^{*}(N_{\textrm{eff}}|\Lambda),≈ Poisson start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ( italic_N start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT | roman_Λ ) , (32)

where Neffsubscript𝑁effN_{\textrm{eff}}italic_N start_POSTSUBSCRIPT eff end_POSTSUBSCRIPT represents the effective number of observed values in the interval. This number is not necessarily an integer. This choice minimizes the mean square error between equation (29) and equation (32), ensuring both computational efficiency and a close approximation to the true value.