Constraining Intermediate-Mass Black Holes From The Stellar Disc of Sgra

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

MNRAS 000, 000–000 (0000) Preprint 26 May 2023 Compiled using MNRAS LATEX style file v3.

Constraining intermediate-mass black holes from the


stellar disc of SgrA*

Jean-Baptiste
1
Fouvry1 , Marı́a José Bustamante-Rosell2 , Aaron Zimmerman3
CNRS and Sorbonne Université, UMR 7095, Institut d’Astrophysique de Paris, 98 bis Boulevard Arago, F-75014 Paris, France
2 Department of Astronomy and Astrophysics, University of California, Santa Cruz, CA 95064, USA
3 Weinberg Institute, University of Texas at Austin, Austin, TX 78712, USA

26 May 2023
arXiv:2305.15998v1 [astro-ph.GA] 25 May 2023

ABSTRACT
Stars evolving around a supermassive black hole see their orbital orientations diffuse
efficiently, a process called “vector resonant relaxation”. In particular, stars within
the same disc, i.e. neighbors in orientations, will slowly diffuse away from one another
through this stochastic process. We use jointly (i) detailed kinetic predictions for the
efficiency of this dilution and (ii) the recent observation of a stellar disc around SgrA*,
the supermassive black hole at the centre of the Milky-Way, to constrain SgrA*’s unob-
served stellar cluster. Notably, we investigate quantitatively the impact of a population
of intermediate mass black holes on the survivability of the stellar disc.
Key words: Diffusion - Gravitation - Galaxies: kinematics and dynamics - Galaxies:
nuclei

1 INTRODUCTION Magnan et al. 2022; Máthé et al. 2023) possibly leading to


formation of discs of IMBHs (Szölgyén & Kocsis 2018); (iv)
Recent outstanding observations keep providing us with new the rapid (resonant) dynamical friction imposed by a stellar
information on the dense cluster orbiting around SgrA*, the disc onto an IMBH (Ginat et al. 2022; Levin 2022).
supermassive black hole (BH) at the centre of the Milky- Similarly, the potential fluctuations generated by the
Way. These include a thorough census of stellar popula- old background stellar cluster and its putative IMBHs can
tions (Ghez et al. 2008; Gillessen et al. 2017) in particular drive the spontaneous dilution of the presently observed
highlighting the presence of a clockwise stellar disc (Levin & discs of young S-stars. As such, we argue that the present
Beloborodov 2003; Paumard et al. 2006; Bartko et al. 2009; survival of the stellar disc requires that VRR fluctuations
Lu et al. 2009; Yelda et al. 2014; von Fellenberg et al. 2022), are not driven too strongly throughout the disc’s life. In the
along with the relativistic precession of the star S2 (GRAV- present work, we build upon Giral Martı́nez et al. (2020),
ITY Collaboration et al. 2020). This wealth of information hereafter G20, and use quantitatively the process of “neigh-
regarding the dynamical status of SgrA*’s nuclear stellar bor separation” by VRR to compare VRR predictions to
cluster (Neumayer et al. 2020) is expected to offer new in- the observed S-stars. Our aim is to leverage VRR to con-
sights on the possible presence of intermediate mass black strain the stellar content of SgrA*’s cluster, in particular its
holes (IMBHs) in these dense regions (see, e.g., Portegies possible IMBHs.
Zwart & McMillan 2002). The paper is organised as follows. In §2, we recall the
The dynamics of stars orbiting a supermassive BH in- key elements of VRR. In §3, we detail our inference method.
volves an intricate hierarchy of dynamical processes (Rauch In §4, we apply this method to SgrA* to (loosely) constrain
& Tremaine 1996; Alexander 2017). In the present work, the properties of the likely population of IMBHs orbiting
we focus on the process of vector resonant relaxation therein. Finally, we conclude in §5. The calculations and
(VRR) (Kocsis & Tremaine 2015), i.e. the mechanism numerical details presented in the main text are kept to a
through which coherent torques between orbits lead to an minimum and are spread through Appendices.
efficient diffusion of the stellar orbital orientations. The an-
alytical and numerical studies of VRR have recently seen a
surge of new developments to understand in particular (i)
the warping of stellar discs (Kocsis & Tremaine 2011); (ii)
2 VECTOR RESONANT RELAXATION
the enhancement of binary merger rates embedded in galac-
tic nuclei (Hamers et al. 2018); (iii) the non-trivial equilib- We consider a set of N ≫ 1 stars orbiting around a super-
rium distribution of orientations (Roupas et al. 2017; Takács massive BH of mass M• . We call this background of stars the
& Kocsis 2018; Touma et al. 2019; Gruzinov et al. 2020; “bath”. They represent the unobserved populations (old low

© 0000 The Authors


2 J.-B. Fouvry, M. J. Bustamante-Rosell, A. Zimmerman
of G20. Ultimately, §B provides us with a prediction for
the time evolution of the probability distribution function
(PDF) of the angular separation
t 7→ P (ϕ | t), (2)
as a function of time. We emphasise that the statistics of
P (ϕ | t) depends on (i) the orbital distribution of the back-
ground bath, n(K); (ii) the orbital parameters of the two
test stars, (K1 , K2 ); (iii) their initial angular separation, ϕ0 .

3 INFERENCE METHODS
Since VRR drives a stochastic dynamics, constraints on the
properties of SgrA*’s stellar background, i.e. n(K), can only
be obtained in a statistical sense. To do so, we use maximum
likelihood estimations. We detail our approach in §D and
Figure 1. Illustration of the VRR dilution of a disc of 7 stars over
reproduce here only the key assumptions.
3 Myr when embedded within the Top-Heavy model from §C3.
Here, each point represents the instantaneous orbital orientation, Our goal is to constrain the parameters, θ, of SgrA*’s
L,
b of a star on the unit sphere. background bath, given some observed angular separa-
tions between disc stars. For example, parameters can be
the collection θ = {n(K), ϕ0 , T⋆ }, with T⋆ the ages of the
mass stars, IMBHs, etc.). They are responsible for fluctua- observed stars. For a given pair of test stars (i, j), we
tions in the gravitational potential, and hence drive VRR. have at our disposal their (conserved) orbital parameters,
Provided that one considers dynamics on timescales longer {Ki , Kj }, along with their current angular separation, ϕij .
than the Keplerian motion induced by the central BH (e.g., For each θ, we compute the associated predicted PDF,
∼ 16 yr for S2) and the in-plane precession driven by rela- ϕij 7→ P (ϕij | θ, Ki , Kj ), according to which ϕij could have
tivistic corrections (e.g., ∼ 3×104 yr for S2), one can orbit- been drawn. The likelihood of the parameters θ is then
average the dynamics over each star’s mean anomaly and proportional to P (ϕij | θ, Ki , Kj ). This is the gist of our
pericentre phase. Stars are formally replaced by massive an- method.
nuli, see, e.g., fig. 1 in G20. These are characterised by To improve upon this approach, we include a few addi-
K = (m, a, e), (1) tional effects by (i) marginalising over the dispersion of the
predicted log-normal distributions which we use to model
with m the individual mass, a the semi-major axis, and e the likelihood (§D1); (ii) accounting jointly for all the dis-
the eccentricity. The normpof the angular momentum of a tinct pairs of test stars of a given disc (§D2); (iii) incor-
given annulus is L(K) = m GM• a(1−e2 ) and is conserved porating measurement errors in the observed S-stars (§D3).
during the VRR dynamics. The remaining dynamical quan- In practice, all these effects are included by bootstrapping
tity is the instantaneous orbital orientation, denoted with over independent samples. It is essential to assess the perfor-
the unit vector L.b We assume that the bath is on average mance of this approach by applying it to tailored numerical
isotropic. Its distribution of orbital parameters is charac- simulations of increasing complexity, see §C for details. This
terised by a distribution function (DF), n(K), normalised to is what we now detail.
b n(K) = N . Our conventions are spelled out in §A.
R
dK dL First, we consider a simple one-population background
In addition to this bath, we consider a population of test bath (§C2). In Fig. 2, we report our inferred likelihoods from
stars. These correspond to the observed S-stars forming the these fiducial simulations, as one varies respectively (m, γ),
clockwise stellar disc. We neglect the mass of these test stars: the individual mass and power-law index of the bath, or
they only probe the potential fluctuations in the system but (ϕ0 , T⋆ ), the initial angular separation and age of the test
do not contribute to it. Ultimately, these stars’ orientations stars. The likelihood here corresponds to synthetic obser-
are driven away from one another via VRR, i.e. the disc vations mimicking our actual analysis of the S-stars around
dissolves as illustrated in Fig. 1. This is the observational SgrA*. In fact, we display the expectation value of the likeli-
signature that we leverage in order to constrain the stellar hood over multiple realisations of the observations, in order
content of the background bath, i.e. n(K). to provide a stringent test of the model, as detailed in §C2.
Given that the disc stars are massless, investigating the Although Fig. 2 suffers from a slight bias, our method is
dilution of the disc is formally equivalent to investigating the able to recover the parameters of the simulated bath. This
dilution of pairs of test stars (see §E in G20). Let us consider is reassuring.
two disc stars of orbital parameters K1 and K2 . We assume We now turn our interest to a astrophysically more rele-
that at the time of their birth, these stars’ respective angu- vant astrophysical system, namely the two-population (stars
lar momentum vectors were separated by some small angle and IMBHs) Top-Heavy model from Generozov & Madigan
ϕ0 . As a result of VRR, these two stars are slowly stirred (2020) (§C3), in which we inject discs of 7 test stars mimick-
away from one another. This is the process of “neighbor sep- ing the observed S-stars. In Fig. 3, we illustrate our expected
aration” driving the stochastic evolution of ϕ(t). This was likelihood as one varies respectively (m• , γ• ) the individual
the process investigated in detail in G20. mass and power-index of the IMBH population or (ϕ0 , T⋆ )
In §B, we reproduce and improve upon the main results the initial angular size of the disc and their (common) stellar

MNRAS 000, 000–000 (0000)


Constraining IMBHs from SgrA*’s disc 3
2.0 1.0
68%

M<• a /(M<• a + M<★a )


68%

0
1.8 95% 0.8 95%
99% 99%
1.6 0.6 Top-Heavy
γ

0
1.4 0.4

0
1.2 0.2
1.0 0.0
0.0 0.5 1.0 1.5 2.0 0 20 40 60 80 100
m m• [M⊙ ]
40
68% Figure 4. Likelihood estimated from the S-stars’ resolved or-
95%
30 99% bits. We assume that SgrA*’s background cluster follows the
T★ 20 two-population Top-Heavy model from §C3, and respectively
vary (m• , M• (< a0 )/(M• (< a0 )+M⋆ (< a0 ))) with the total en-
10 closed mass, M• (< a0 )+M⋆ (< a0 ), and all the other parameters
fixed. The black dot corresponds to the Top-Heavy model (Gen-
0 erozov & Madigan 2020). See §D for details.
0 1 2 3 4 5 6
ϕ0[°]

Figure 2. Likelihood for the fiducial one-population simulations


(§C2) as one varies (m, γ) (top) or (ϕ0 , T⋆ ) (bottom). The black
dot corresponds to the bath effectively simulated and the colored

m• /m★
lines to the confidence levels. See §D1 for details.

2.0
68%
1.8 95%
99% m★ [M⊙ ]
γ•
1.6
1.4 Figure 5. Same as in Fig. 4, this time jointly varying
1.2 (m⋆ , m• /m⋆ ) keeping all other parameters fixed. The dashed line
corresponds to m• = 100M⊙ above which the heavy objects are
1.0 usually considered as IMBHs. The hashed contours correspond to
0 20 40 60 80 100
m• the 68% level lines constrained in Tep et al. (2021) by considering
3 eccentricity relaxation.
68%
95%
99%
T★ [Myr]

2
the stars to T⋆ = 7.1 Myr (see §A). We marginalise our like-
lihood over the observational uncertainties in the S-star’s
1
orbital parameters from Gillessen et al. (2017), and over the
initial orientation of the stellar orbits within a narrow dis-
0
0 1 2 3 4 5 6 tribution around ϕ0 = 3◦ , following Eq. (C2).
ϕ0[°] In Fig. 4, we plot this likelihood, varying the indi-
vidual mass m• as well as the total IMBH mass fraction
Figure 3. Same as in Fig. 2, but for the two-population simula- M• (< a0 )/(M• (< a0 )+M⋆ (< a0 )), keeping the mass enclosed
tions from §C3. See §D2 for details.
within a fiducial scale radius a0 , M• (< a0 )+M⋆ (< a0 ), fixed.
Interestingly, we find that the Top-Heavy model from Gen-
ages. We average over many realisations of the observations erozov & Madigan (2020) is incompatible with SgrA*’s ob-
(§D2). We note that the reconstruction is not ideal and suf- served stellar disc. The survival of the disc through the VRR
fers from some bias. One should keep these limitations in dynamics requires indeed a quieter bath, i.e. a smaller IMBH
mind when applying VRR neighbor separation to SgrA*’s fraction.
clockwise disc. Finally, in Fig. 5, we let the individual masses of the
background cluster, m⋆ and m• , vary with the natural
constraint m• ≥ m⋆ , while fixing the total enclosed masses
M⋆ (< a0 ) and M• (< a0 ). In that figure, we find that the sur-
4 RESULTS FOR SGRA*
vival of SgrA*’s disc requires a drastically smaller individ-
We now apply our approach to the clockwise disc of S-stars ual mass for the background population of heavy particles.
(see §A). The realm of possible models for SgrA*’s back- Assuming that all the other parameters of the background
ground is extremely large. For the sake of simplicity, we only bath are fixed, the current presence of SgrA*’s disc seems
consider baths similar to the Top-Heavy model from Gen- in tension with the presence of IMBHs, as illustrated by the
erozov & Madigan (2020) (§C3). These clusters are there- dotted line in that plot.
fore characterised by two background populations (stars and A similar parameter exploration has recently been pre-
IMBHs) each with their own individual mass and power- sented in the bottom right panel of fig. 3 of Tep et al. (2021).
law index, an initial angular separation for the disc, ϕ0 (see There, the authors used the eccentricity relaxation of the S-
Eq. C2), and a common stellar age, T⋆ , for the S-stars. For stars to constrain SgrA*’s stellar content, assuming that the
simplicity, we arbitrarily impose ϕ0 = 3◦ , and fix the age of S-stars were initially born from an initial quasi- circular disc.

MNRAS 000, 000–000 (0000)


4 J.-B. Fouvry, M. J. Bustamante-Rosell, A. Zimmerman
In Fig. 5, we reproduce the bottom-right panel of fig. 3 of Tep the semi-major axes and eccentricities of each annuli are
et al. (2021). It is very interesting to note that both works, supposed to be conserved. This amounts to neglecting the
although they are using different dynamical processes (scalar possible contributions of scalar resonant relaxation and non-
vs. vector resonant relaxation) offer similar constraints on resonant relaxation (Rauch & Tremaine 1996). Although
SgrA*’s background cluster. They jointly constrain the like- slower, these additional processes could still marginally im-
lihood of IMBHs orbiting SgrA*. Naturally, one should keep pact the disc’s dilution.
in mind all the limitations of the present exploration, in par- Finally, we expect that future observations, such as
ticular the small number of parameters effectively probed. ELT (Pott et al. 2018; Davies et al. 2018) and TMT (Do
Future works will focus on exploring larger parameter spaces et al. 2019), will soon provide the community with a wealth
by sampling from the full likelihood. of resolved orbits around SgrA*, along with their stellar
ages. This additional statistics will prove paramount to
tighten the present constraints on SgrA*’s stellar content.
In particular, the same data will allow one to constrain both
5 CONCLUSIONS the efficiency of eccentricity relaxation (Tep et al. 2021) as
We illustrated how a statistical characterisation of VRR well as orientation relaxation (this work). Ultimately, these
used jointly with the recent observation of SgrA*’s clockwise two dynamical windows should offer complementary con-
stellar disc allows one to place constraints on the content straints on the statistics of the possible IMBHs orbiting
of SgrA*’s underlying stellar cluster, in particular IMBHs. around SgrA*.
More precisely, to be compatible with the disc’s being still
present today, VRR fluctuations cannot be too large, i.e.
cannot be too efficient at driving the spontaneous dilution Data Distribution
of the disc. Keeping in mind the small range of parameters The code and the data underlying this article is available
explored, we showed that the survival of SgrA*’s disc tends through reasonable request to the authors.
to reduce the likelihood for heavy background particles, like
IMBHs, to orbit around SgrA*.
Naturally, this paper is only a first step toward fully
ACKNOWLEDGEMENTS
leveraging VRR to constrain SgrA*’s stellar content: a more
thorough exploration of parameter space is a logical next JBF is partially supported by grant Segal ANR-19-CE31-
step. We now conclude with a few possible additional venues 0017 of the French Agence Nationale de la Recherche, and
for future works. by the Idex Sorbonne Université. AZ was supported by
The statistical description of neighbor separation by NSF Grants PHY-1912578 and PHY-2207594. JBF warmly
VRR (§B) relies extensively on the assumption of an thanks F. Leclercq for various suggestions.
isotropic background bath. This is key to simplify the an-
alytical calculations However, this does not account for
any possible anisotropic clustering in orientations, such as
APPENDIX A: STELLAR CUSPS IN SGRA*
IMBHs discs (Szölgyén & Kocsis 2018) and the associated
resonant dynamical friction (see, e.g., Szölgyén et al. 2021). In this Appendix, we specify our convention for the de-
Indeed, a rotating background stellar cluster is expected to scription of the background bath. Following Gillessen et al.
source (resonant) dynamical friction, possibly leading to an (2017), we fix SgrA*’s mass to M• = 4.28×106 M⊙ . We take
efficient disruption of the disc (Levin 2022). Going beyond SgrA*’s distance from the Earth to be d = 8.178 kpc (GRAV-
isotropy, e.g., axisymmetric distributions, will be the focus ITY Collaboration et al. 2019). This is used to obtain the
of future works. S-stars’ semi-major axes from table 3 in Gillessen et al.
We also assumed the limit of test particles for the disc (2017). Following fig. 12 of Gillessen et al. (2017), we as-
stars, i.e. we neglected the self-gravity among them. Contri- sume that the disc is composed of a total of seven stars,
butions from this additional coupling were already investi- namely (S66, S67, S83, S87, S91, S96, S97). Their ages are
gated numerically in Kocsis & Tremaine (2011), see figs. 6 inferred from the main-sequence ages measured in table 2
and 7 therein. It showed that self-gravity increased the co- of Habibi et al. (2017), albeit for different S-stars. In prac-
herence of the disc, hence delaying its dilution: the more tice, we assume that the disc stars’ ages are identical and im-
massive the disc, the longer its survival. Incorporating this pose T⋆ = 7.1 Myr, the average stellar age of the S-stars with
additional effect is no easy analytical undertaking. both resolved orbits and constrained main-sequence stellar
We neglected the effect of SgrA*’s spin, which via ages. This large uncertainty on the stars’ ages is one the
the Lense–Thirring precession (see, e.g., Merritt 2013) can main limitations of the results from the main text.
also drive a spontaneous dilution of the disc (Levin & Be- A given bath is composed of different stellar popula-
loborodov 2003; Fragione & Loeb 2022). We point out that tions within which all the orbits share the same individ-
these effects could still be somewhat small here because the ual mass. In pratice, we make two important assumptions:
disc stars are quite far from SgrA*, adisc ≳ 50 mpc. In ad- (i) each population follows a power-law distributin in semi-
dition, the appropriate timescale for dilution is the one of major axes; (ii) each population follows a thermal distribu-
phase mixing, i.e. the orbital planes of the disc stars must tion in eccentricities. As a consequence, a given population
precess respectively from one another. It would be worth- is characterised by three numbers, namely (i) γ, the slope
while to include this additional effect. of the power-law profile in semi-major axis; (ii) m, the in-
VRR relies on a double orbit-average over both the fast dividual mass of the particles; (iii) M (< a0 ), the total mass
Keplerian motion and the in-plane precession. As a result, physically enclosed within a radius a0 , a radius of reference.

MNRAS 000, 000–000 (0000)


Constraining IMBHs from SgrA*’s disc 5
cos ϕ(t) conditions of the two test stars. Following Eq. (21) of G20,
1.0
these moments evolve according to1
0.8
Pℓα (cos ϕ(t)) = Dℓα CℓΩα (t) CℓDα (t).


(B2)
0.6
In Eq. (B2), Dℓα = ⟨Pℓα (cos ϕ(0))⟩ describes the statis-
0.4
tics of the angular separation at t = 0. Following Eq. (D9)
0.2 of G20, the function CℓΩα (t) describes the separation driven
0.0 t by the differences in the two particles orbital parameters,
0 20 40 60 80 100 i.e. driven by K1 ̸= K2 . It reads
 
Figure B1. Typical random walks of pairwise angular separa- X
tions, cos ϕ(t), as observed in the simple setup from §C2. CℓΩα (t) = exp − 12 Aℓα Bℓ Iℓ− (K1 , K2 , t) , (B3)

Similarly, following Eq. (D10) of G202 , the function CℓDα (t)


More precisely, for a given population the number of captures the separation sourced by the differences in the test
stars per unit semi-major axis, a, is given by particles’ initial orientations. It reads
 2−γ
Dℓα−Dℓγ
 X 
N0 a 2
N (a) = (3−γ) , (A1) CℓDα (t) = exp − EℓLα ℓℓγ Iℓ+ (K1 , K2 , t) .
a0 a0 (2ℓα +1)Dℓα
ℓ,ℓγ
where we introduced N0 = g(γ) N (< a0 ) with (B4)
√ In Eqs. (B3) and (B4), we defined the integrals
g(γ) = 2−γ π Γ(1+γ)/Γ(γ − 12 ), (A2) Z
2
with Γ(x) the gamma function and N (< a0 ) = M (< a0 )/m Iℓ− (K1 , K2 , t) = dK n(K) Jℓ [K1 , K]−Jℓ [K2 , K]
the number of stars physically enclosed within the radius
2Tc2 (K) p 
a0 . We assume that the distribution of eccentricities is ther- × χ Aℓ /2 (t/Tc (K)) ,
Aℓ
mal, neglecting effects associated with the SgrA*’s loss cone. Z
+
Eccentricities follow then the PDF Iℓ (K1 , K2 , t) = dK n(K) Jℓ [K1 , K2 ] Jℓ [K2 , K]
f (e) = 2e, (A3) 2Tc2 (K) p 
× χ Aℓ /2 (t/Tc (K)) . (B5)
Aℓ
R
which satisfies the normalisation def (e) = 1. Following the
convention from §2, the DF of a given background popula- We also introduced the coefficients
tion is therefore 1
Aℓ = ℓ(ℓ+1); Bℓ = 8π
ℓ(ℓ + 1)(2ℓ + 1), (B6)
′ 1 ′ ′ ′
n(K ) = 4π
δD (m −m) N (a ) f (e ), (A4) along with the bath’s coherence time
recalling that the background populations are assumed to 1 X Z
= Bℓ dK′ n(K′ ) Jℓ2 [K, K′ ], (B7)
be isotropic on average. Tc2 (K)

Assuming that the background populations follow
power law distributions greatly simplifies the computation where the coupling coefficients, Jℓ [K, K′ ], are given in §A
of the coherence time, see §I in Fouvry et al. (2019). In prac- of G20. In practice, these coefficients are pre-computed using
tice, following the notations therein, we pre-computed the the same approach as in Magnan et al. (2022). Equation (B4)
dimensionless function fΓ2 (e) for 0 ≤ γ ≤ 3 and 0 ≤ e ≤ 0.99, also involves the isotropic Elsasser coefficients, EℓLα ℓℓγ , as
on a 300×100 grid, to be interpolated afterwards. This al- spelled out in §B of Fouvry et al. (2019). Finally, we intro-
lows for efficient evaluations of the coherence time of the duced the dimensionless function
Z τ Z τ
background bath, Tc (K), as defined in Eq. (B7). 2 2 √
χ(τ ) = dτ1 dτ2 e−(τ1 −τ2 ) = e−τ −1+ π τ erf(τ ), (B8)
0 0
which captures the transition between the ballistic and dif-
APPENDIX B: VRR NEIGHBOR SEPARATION fusive regimes (Fouvry et al. 2019).
Unfortunately, as argued in §4 of G20, the analytical
We are interested in the stochastic evolution of the angle prediction from Eq. (B2) does not provide a good match to
ϕ(t) between two test stars’ orbits of parameters (K1 , K2 ) the late-time behavior of the dilution of test stars that start
embedded within a background bath characterised by n(K). with very similar initial orientations. To alleviate this diffi-
Figure B1 illustrates typical random walks undergone by culty, we follow G20 and rely on a “piecewise” prediction.
cos ϕ(t) for the simple bath from §C2.

B2 Piecewise prediction
B1 Analytical prediction
Regardless of the initial similarity between the test particles,
Following G20, the dilution of the two test stars can be Eq. (B2) works well on short timescales, i.e. for t ≲ Tc (K).
described via the moments of the correlation function of the As such, we give ourselves the timelapse
stars’ respective orientations, namely

∆t = Min[Tc (K1 ), Tc (K2 )], (B9)
t 7→ Pℓα (cos ϕ(t)) , (B1)
with Pℓα the Legendre polynomials and ⟨ · ⟩ standing for an 1 Equation (21) of G20 is missing a 1/(4π) in its rhs.
ensemble average over realisations of the bath and initial 2 Equation (D10) of G20 is missing a minus sign in its rhs.

MNRAS 000, 000–000 (0000)


6 J.-B. Fouvry, M. J. Bustamante-Rosell, A. Zimmerman
and set out to construct the sequence of angular separations In these expressions, we introduced
X
cos ϕ0 → ... → cos ϕn = cos ϕ(t = n∆t). (B10) Ψ− (K1 , K2 , t) = Bℓ Iℓ− (K1 , K2 , t), (B19)

This prediction is Markovian in the sense that we treat the X
transition cos ϕi → cos ϕi+1 as independent of the past his- Ψ+ (K1 , K2 , t) = Bℓ (Aℓ −2) Iℓ+ (K1 , K2 , t),

tory of the test particles. X
++
Following Eq. (G1) of G20, we can write Ψ (K1 , K2 , t) = Bℓ (Aℓ −2)(Aℓ −6) Iℓ+ (K1 , K2 , t)
Z ℓ



Pℓ (cos ϕi+1 ) = d(cos ϕi ) ρi (cos ϕi ) Pℓ (cos ϕi+1 ) cos ϕi . In practice, the integrals over (a, e) appearing in Eq. (B19)
(B11) are computed using the standard midpoint rule with
In that expression, ⟨Pℓ (cos ϕi+1 ) | cos ϕi ⟩ follows from Ka = 100 nodes sampled uniformly in logarithm in the
Eq. (B2) evaluated after a time t = ∆t and assuming that domain Max[a1 , a2 ]×10−3 ≤ a ≤ Min[a1 , a2 ]×103 , and using
the test particles are initially separated by the fixed angle ϕi , Ke = 20 nodes uniformly within the domain in e of the con-
i.e. Dℓ = Pℓ (cos ϕi ). Equation (B11) also involves ρi (cos ϕi ), sidered system.
the PDF of cos ϕi , which is unknown. Our goal is therefore
to average over it.
B4 Propagating the moments
We continue the calculation by expanding explicitly the rhs
B3 Perturbative expansion of Eq. (B16) at second order in (1−cos ϕ0 ). We have
We now make progress by performing a perturbative ex-


Pℓα(cos ϕ(t)) cos ϕ0 ≃ aℓα +bℓα P1 (cos ϕ0 )+cℓα P2 (cos ϕ0 ).
pansion of Eq. (B2) for small angles.3 For cos ϕ0 → 1, the (B20)
Legendre polynomials obey the second-order expansion In that expressions, the coefficients (aℓ , bℓ , cℓ ) read
Pℓ (cos ϕ0 ) ≃ 1− 21 Aℓ (1−cos ϕ0 )+ 16
1
Cℓ (1−cos ϕ0 )2 , (B12) 1
Aℓ uℓ vℓ − 6Ψ+ − 8 ,

aℓ = uℓ + 12
bℓ = − 18 Aℓ uℓ vℓ − 4Ψ+ − 6 ,

with Cℓ = Aℓ (Aℓ −2). Let us now expand Eq. (B4) at second
order in 1−cos ϕ0 . We write 1

cℓ = 24 Aℓ uℓ vℓ − 2 , (B21)
Dℓα−Dℓγ 1
 Aℓγ −Aℓα where we introduced
≃ 2
1−cos ϕ0 (B13)
(2ℓα +1)Dℓα 2ℓα +1 1 −
2 Cℓα−Cℓγ +4Aℓα (Aℓγ −Aℓα ) uℓ = e− 2 Aℓ Ψ ; vℓ = Ψ++ + Aℓ (1+Ψ+ )(1+2Ψ+ ), (B22)
1
+ 16 1−cos ϕ0 ,
2ℓα +1 and used the shortened notation Ψ− = Ψ− (K1 , K2 , t), and
where we used Dℓ = Pℓ (cos ϕ0 ). Pursuing the calculation, we similarly for Ψ+ and Ψ++ .
can perform the sum over ℓγ in Eq. (B4) to get We now have everything at our disposal to predict
jointly both moments ⟨P1 (cos ϕi )⟩ and ⟨P2 (cos ϕi )⟩. Let us
X Dℓα−Dℓγ 2 introduce the vector and matrices
EℓLα ℓℓγ ≃ 21 1−cos ϕ0 Aℓα Bℓ (Aℓ −2)

(2ℓα +1)Dℓα      
ℓγ ⟨P1 (cos ϕi )⟩ a1 b1 c 1
Pi = ; A= ; M= , (B23)
1
2 ⟨P2 (cos ϕi )⟩ a2 b2 c 2
+ 16 1−cos ϕ0 Aℓα Bℓ (Aℓ −2)(Aℓα−Aℓ +6). (B14)
so that Eq. (B20), evaluated at time t = ∆t, reads
To get this expression, we used some contraction rules of the
isotropic Elsasser coefficients, namely Eqs. (B3) and (B4) Pi+1 = A+ M Pi . (B24)
in G20, along with −1
The fixed point of this iteration is R = [I−M] A, and can
X 2
1
2ℓα +1
EℓLα ℓℓγ Cℓγ = Aℓα Bℓ be used to construct the generic solution
ℓγ
Pi = R + Mi P0 −R .
 
  (B25)
× 12−6Aℓ −6Aℓα+3Aℓ Aℓα +Cℓ +Cℓα . (B15)
To obtain a “continuous” prediction from this discrete solu-
At second-order in (1−cos ϕ0 ), Eq. (B2) becomes tion, we diagonalise M via M = U Diag[λ1 , λ2 ] U−1 , which
gives Mi = U Diag[λi1 , λi2 ] U−1 . To obtain a continuous pre-
Pℓα (cos ϕ(t)) cos ϕ0 ≃ Pℓα (cos ϕ0 ) CℓΩα (t) CℓDα (t), (B16)


diction, we finally make the replacement i → t/∆t, assuming
with that the two eigenvalues (λ1 , λ2 ) are either both positive or
complex conjugate4 .
CℓΩα (t) ≃ exp − 12 Aℓα Ψ− (K1 , K2 , t) ,
 
(B17) In Fig. B2, we compare the initial time-evolution of
and the first two moments, ⟨P1 (cos ϕ(t))⟩ and ⟨P2 (cos ϕ(t))⟩, be-
tween the present kinetic modelling and fiducial numerical
CℓDα (t) ≃ exp − 21 Aℓα (1−cos ϕ0 ) Ψ+(K1 , K2 , t)

(B18) simulations (§C2). While the match is not ideal, we empha-
sise that the prediction still succeeds at predicting the slow
1
A (1−cos ϕ0 )2 + ++ 
− 16 ℓα
AℓαΨ (K1 , K2 , t)−Ψ (K1 , K2 , t) .

4 This is not guaranteed a priori by the present expansion. In


3 Here, we improve upon G20 and perform a second-order expan- practice, for all the background bath and pairs of test particles
sion. Though, the crux of the calculations remains identical. considered, this was, fortunately, always the case.

MNRAS 000, 000–000 (0000)


Constraining IMBHs from SgrA*’s disc 7
〈P1(cos ϕ)〉 μ
1.00 -1 t
10 20 30 40 50
N-body
Prediction
0.95 -2
Fiducial
N-body
Prediction
0.90 t -3
0 10 20 30 40 50 σ
〈P2(cos ϕ)〉 1.0 N-body
1.00 0.8 Prediction

0.6
0.4
0.95
Fiducial 0.2
N-body
Prediction 0.0 t
0.90 t 0 10 20 30 40 50
0 10 20 30 40 50
Figure B4. Time evolution of the log-normal parameters µ (top)
Figure B2. Time evolution of the two first moments ⟨P1 (cos ϕ)⟩ and σ (bottom) for the fiducial simulations from §C2. The solid
(top) and ⟨P2 (cos ϕ)⟩ (bottom) for the fiducial bath from §C2. line is fitted from the numerical simulations, while the dashed line
The solid line is measured in numerical simulations, with the gray is the prediction from §B5.
region highlighting the 16% and 84% levels among the available
realisations. The dashed line is the piecewise prediction from §B4.
B5 Log-normal ansatz
We now have at our disposal a description for the initial
time-evolution of the first two moments ⟨P1 (cos ϕ(t))⟩ and
〈P1(cos ϕ)〉
⟨P2 (cos ϕ(t))⟩. Inspired by §I1 in G20, we now impose the
1.00
ansatz that the random variable ϕ(t) follows a log-normal
PDF at all time. This is a strong assumption. The PDF
0.95 reads
Top-Heavy
1
√ exp − (ln ϕ − µ)2 /(2σ 2 ) , (B26)
 
N-body P (ϕ | µ, σ) =
Prediction ϕ σ 2π
0.90 t [Myr]
0 1 2 3 for some parameters µ, σ.
〈P2(cos ϕ)〉
Unfortunately, our predictions for (⟨P1 ⟩, ⟨P2 ⟩) are not
1.00 guaranteed to be realisable, i.e. there might not ex-
ist a well-defined positive PDF complying with these
0.95 provided moments. Therefore, to estimate (µ, σ) from
Top-Heavy (⟨P1 ⟩, ⟨P2 ⟩), we proceed by minimising the distance
N-body d = (⟨P1 ⟩−P1 [µ, σ])2 +(⟨P2 ⟩−P2 [µ, σ])2 over (µ, σ). In that
Prediction
0.90 t [Myr] expression, the moments Pℓ [µ, σ] are efficiently approxi-
0 1 2 3 mated following eq. (1.3) of Asmussen et al. (2016). More
Figure B3. Same as in Fig. B2, but for the two-population bath precisely, we use
from §C3. Here, ⟨Pℓ (cos ϕ)⟩ is averaged over all the pairs of stars Z +∞ 
in discs composed of 7 stars each. See the text for details. ⟨cos(ℓϕ)⟩ = Re dϕ eiℓϕ P (ϕ | µ, σ)
0
2 √
≃ Re e−w(2+w)/(2σ ) / 1+w ,
 
(B27)
with w = W [−iℓσ 2 eµ ], W [z] the Lambert function, and the
initial onset of separation, i.e. the plateau that the naive Legendre moments naturally follow. This same expression
application of the analytical formula from Eq. (B2) cannot can also be used to compute the gradients ∂Pℓ [µ, σ]/∂(µ, σ),
recover. which are used by the optimiser. In practice, the opti-
In Fig. B3, we illustrate the same measurement for the misation is performed using the LBFGS algorithm from
more realistic Top-Heavy model (§C3). In that case, every Optim.jl (Mogensen & Riseth 2018).
disc is composed of 7 stars and sustains therefore 21 indepen- In Fig. B4, we compare the time-evolution of the two
dent pairs of stars. In Fig. (B3), ⟨Pℓ (cos ϕ)⟩ stands therefore lognormal parameters, (µ, σ), as obtained from the kinetic
for the average over all these pairs of stars. To compute the prediction and directly fitted from the numerical simula-
kinetic prediction, following Eq. (C2), we drew 100 inde- tions. The match for µ is quite good, while the VRR for-
pendent initial conditions for 100 discs, following Eq. (C2). malism somewhat fails at predicting σ.
These are then used as initial conditions to compute the Finally, in Fig. B5, we illustrate the PDF of the an-
prediction from §B4. Finally, we average the Legendre mo- gles ϕ at various times, and compare it with the predicted
ments over this large sample of pairs. Reassuringly, Figs. B2 log-normal ansatz. In that figure, we note that using the
and B3 exhibit similar trends. predicted µ and fitting σ from the PDF offers a very sat-

MNRAS 000, 000–000 (0000)


8 J.-B. Fouvry, M. J. Bustamante-Rosell, A. Zimmerman
P(ϕ)

× Iℓ+ (K1 , K2 , t) , (B29)
t = 10 N-body
12 (μpred , σpred )
(μpred , σfit ) to stitch the piecewise and analytical predictions.
8 Following all these manipulations, for a given back-
ground bath, n(K), a given pair of test particles, (K1 , K2 ),
4 and a given initial angular separation, ϕ0 , we have at our
disposal a prediction for the time evolution of the PDF of
0 ϕ
0.0 0.1 0.2 0.3 the angular separation, t 7→ P (ϕ | t). It is this PDF that is
P(ϕ) used in the likelihood estimations.
8 t = 20 N-body
(μpred , σpred )
6 (μpred , σfit )
APPENDIX C: NUMERICAL SIMULATIONS
4
In order to assess the validity of our approach, it is important
2 to test it on numerical simulations of background baths of
increasing complexity.
0 ϕ
0.0 0.1 0.2 0.3 0.4 0.5

Figure B5. Illustration of the PDF of the angle ϕ(t) in the fidu- C1 Numerical simulations
cial simulations from §C2 for t = 10 (top) and t = 20 (bottom).
The full lines correspond to the prediction using both (µ, σ) pre-
The numerical simulations were performed using a multi-
dicted in §B5, while the dashed lines only use the predicted µ and pole algorithm similar to the one presented in Fouvry et al.
subsequently fit σ from the empirical histograms. (2022), here adapted to the simpler case of VRR. Here, we
only outline the main elements of the algorithm.
All the radial orbit averages appearing in the coupling
isfactory match. This motivates the marginalisation over σ coefficients Jℓ are replaced by discrete sampling with K
when performing likelihood estimations (§D1). nodes in eccentric anomaly. The interaction Hamiltonian is
truncated to the maximum harmonic ℓmax . Inspired by Fou-
vry et al. (2022), the rates of change, dL/dt,
b are com-
puted in O(N Kℓ2max ) operations owing to prefix sums. We
B6 Stitching rewrite the equations of motion as precession equations of
Unfortunately, because it relies on a perturbative expan- the form dL/dt
b = Ω× L.
b The dynamics is integrated using
sion for small angles (§B3), the piecewise prediction from the structure-preserving MK2 scheme (see Fouvry et al. 2022)
Eq. (B25) cannot be used for arbitrarily large angular sep- with a constant timestep h. This ensures the exact conser-
aration, see, e.g., fig. 6 in G20. To circumvent this is- vation of |L|
b = 1 throughout the evolution.
sue, we use Eq. (B25) only if the following conditions are
all met: (i) ⟨cos ϕ⟩ ≥ ⟨cos ϕcut ⟩ = 0.8, i.e. the two test par-
C2 One-population bath
ticles are close neighbors; (ii) both ∂⟨P1 (cos ϕ(t))⟩/∂t and
∂⟨P2 (cos ϕ(t))⟩/∂t are negative, i.e. the correlations decay. Similarly to §F of G20, we consider a single-population back-
In any other situations, we revert back to the straightfor- ground bath. Fixing the units to G = M• = 1, we assume that
ward application of the analytical result from Eq. (B2). We it is composed of N = 103 stars of individual mass m = 1.
now detail how this “stitching” between the piecewise and Their distribution in a is proportional to a2−γ with γ = 1.5,
analytical predictions can be made in practice. and limited to 1 ≤ a ≤ 100. Finally, their distribution in ec-
Equation (B3) is straightforward to evaluate for centricities is proportional to e and limited to 0 ≤ e ≤ 0.3.
ℓα = 1, 2. To evaluate Eq. (B4), we first estimate the val- For the test stars, we consider pairs of independent test
ues of Dℓ = ⟨Pℓ (cos ϕ)⟩ using Eq. (B27). We then use the particles with (a1 , e1 ) = (9.5, 0.15) and (a2 , e2 ) = (10.5, 0.05).
exclusion rules of the isotropic Elsasser coefficients, EℓLα ℓℓγ , For each realisation, we inject 500 pairs of test particles with
see Eq. (B2) in G20, to perform the sum over ℓγ in Eq. (B4). an initial angular separation set by ϕ0 = 3.0◦ . We performed
More precisely, for the particular cases ℓα = 1 (resp. ℓα = 2), a total of 500 realisations. Following §C1, we used K = 20,
the range of this sum reduces to ℓγ = ℓ (resp. ℓγ = ℓ−1, ℓ+1). ℓmax = 10, h = 10−3 , and integrated up to tmax = 100. Each re-
We also rely on the simple relations alisation required ∼ 0.5 h of computation on two cores, with
L 2 final relative errors in the total energy (resp. total angular
E1ℓℓ = 6Bℓ , momentum) of order 10−10 (resp. 10−10 ).
L 2 45
E2,ℓ,ℓ−1 = 15Bℓ − 8π Aℓ ,
L 2 45
E2,ℓ,ℓ+1 = 15Bℓ + 8π Aℓ . (B28) C3 Two-population bath

We ultimately obtain the explicit expressions We consider a more involved bath better mimicking SgrA*.
 X  This is the Top-Heavy model from Generozov & Madigan
2(D1 −Dℓ ) + (2020), also reproduced in eq. (14) of Tep et al. (2021). We
C1D (t) = exp − Bℓ Iℓ (K1 , K2 , t) ,

D1 consider the same BH mass as in §A. The bath is composed
 X
6Bℓ of two populations, namely stars and IMBHs, with
C2D (t) = exp − ℓ−1 ℓ+2

D2 −Dℓ−1 2ℓ+1 −Dℓ+1 2ℓ+1
D2 γ⋆ , m⋆ , M⋆ (< a0 ) = 1.5, 1 M⊙ , 7.9×103 M⊙ ,
   

MNRAS 000, 000–000 (0000)


Constraining IMBHs from SgrA*’s disc 9
γ• , m• , M• (< a0 ) = 1.8, 50 M⊙ , 38×103 M⊙ ,
   
(C1) with 1 the indicator function inspired by the range in σ effec-
tively observed in Fig. B4. Injecting Eq. (D3) into Eq. (D1),
fixing the scale radius to a0 = 100 mpc. The distribution of
we can perform the integral over dµdσ to get
eccentricities for both populations is proportional to e and
limited to the range 0 ≤ e ≤ 0.99. P (ϕij | θ, Ki , Kj ) = A I(ϕij | µ
e[θ, Ki , Kj ]), (D5)
In practice, we can only simulate a finite number of
where we introduced
background particles, hence a finite range in semi-major
Γ[0, 21 (ln ϕ−µ)2 ]
Z
axes. The minimum (resp. maximum) semi-major axis of the I(ϕ | µ) = dσ P (ϕ | µ, σ) P (σ | µ) = √ , (D6)
disc stars is a⋆min ≃ 45 mpc (resp. a⋆max ≃ 109 mpc) as given 2ϕ 2π
by S67 (resp. S87). We fix the range of semi-major axes for with Γ[0, x] the incomplete gamma function. If one has a
the bath populations to amin ≤ a ≤ amax with amin = a⋆min /η, sample {ϕa }1≤a≤n of distinct pairs of stars at one’s disposal,
amax = a⋆max
R η, and η = 5. The number of background stars is with the corresponding (Ki , Kj ) implicit, Eq. (D5) becomes
a
then N⋆ = a maxda N (a), with N (a) following Eq. (A1). In n
min
practice, we find N⋆ = 83 136 and N• = 5 466.
Y
P ({ϕa } | θ) = A I(ϕa | µ
e[θ]), (D7)
In each realisation, we inject a total of 100 independent a=1
discs composed of 7 test stars with orbital parameters fol-
i.e. pairs are treated as independent from one another, and
lowing §A. The disc stars’ orientations are drawn from a Von
we marginalise over σ for each. The larger P ({ϕa } | θ), the
Mises-Fischer PDF (see §C2 in G20)
more likely the data for the given parameters.
b ∝ eκL0 ·L
P (L)
b b
(C2) We apply this method to the fiducial simulations
from §C2, considered at time T⋆ = 20. To obtain the like-
of concentration κ = 2/(1−cos ϕ0 ) with ϕ0 = 3◦ , and L
b 0 the
lihoods presented in Fig. 2, we fix the bath’s parameters
disc’s mean orientation drawn uniformly on the unit sphere. to their fiducial values (§C2) except for (m, γ) or (ϕ0 , T⋆ ),
We performed a total of 500 realisations using K = 20, which we respectively try to infer in the two panels of Fig. 2.
ℓmax = 10, h = 0.4 kyr, and integrated up to tmax = 10 Myr. To mimick the observed data, for some given parameters,
Each realisation required ∼ 36 h of computation on two we pick, at random, a total of 21 independent pairs of sim-
cores, with final relative errors in the total energy (resp. ulated test stars whose angular separations, {ϕa }, are used
total angular momentum) of order 10−7 (resp. 10−8 ). in Eq. (D7). For every set of parameters, we repeat this pro-
cedure 100 times and subsequently average the likelihood
over this sample. This has the effect of taking the expecta-
APPENDIX D: LIKELIHOOD ESTIMATION tion value of the likelihood over an ensemble of realisations
of the observations, and reduces the variation due to the
Ultimately, our goal is to constrain some parameters, θ, for
stochastic evolution of the pairs. In that figure, the data
the background bath. Let us first assume that we have at
was computed on a 500×500 grid of parameters, and sub-
our disposal one observed angular separation, ϕij , between
sequently smoothed with a Gaussian filter of standard de-
two test stars. We can write the likelihood of this angular
viation 6. Finally, a contour labeled x% corresponds to the
separation given the bath parameters θ and the conserved
level line P (θ)
R = p where p follows from G(p)/G(p = 0) = x%
orbital parameters (Ki , Kj ) as
Z with G(p) = P (θ)≥p dθP (θ).
P (ϕij | θ, Ki , Kj ) = A dµdσ P (ϕij | µ, σ) P (µ, σ | θ, Ki , Kj ),
(D1) D2 Dilution of a disc
with A a normalisation constant and P (ϕ | µ, σ) the log-
In practice, we are interested in the dilution of a disc com-
normal PDF from Eq. (B26). In addition, we also introduced
posed of 7 stars of individual orientations {L b i }1≤i≤7 . From
P (µ, σ | θ) = δD [µ−e
µ(θ, Ki , Kj )] δD [σ−e
σ (θ, Ki , Kj )], (D2) these, we can construct n = 21 different pairs of distinct stars
of respective angular separations
with δD the Dirac delta, and (e µ(θ, Ki , Kj ), σ
e(θ, Ki , Kj ))
ϕa 1≤a≤n = cos−1 L
  
the kinetic predictions of the log-normal parameters for bi ·L
bj . (D8)
1≤i<j≤7
some given bath, as detailed in §B5. The likelihood
We use this sample of separations in Eq. (D7).
P (ϕij | θ, Ki , Kj ) is directly proportional to the posterior
Let us now apply our likelihood estimation to the sim-
probability P (θ | ϕij , Ki , Kj ) derived from the measured pa-
ulations from §C3, which we consider at time T⋆ = 1.5 Myr.
rameters (ϕij , Ki , Kj ), provided we assume flat priors. In
For the likelihoods shown in Fig. 3, we fix the bath’s pa-
this study, we use the likelihood as a proxy for the posterior
rameters to their fiducial values (§C3) except for (m• , γ• ) or
measurements of the bath’s parameters.
(ϕ0 , T⋆ ) (with κ = 2/[1−cos ϕ0 ]), which we respectively try
to infer. For a given set of parameters, we pick, at random,
D1 Marginalising over σ one simulated disc. The initial conditions used in the pre-
diction from §B are obtained by drawing 7 stars according
In Fig. B4, we noted that our prediction for σ
e(θ, Ki , Kj ) is to Eq. (C2) and following Eq. (D8) to get the pairs’ initial
poor. As a consequence, we marginalise over σ. We write separation. For every parameters, we repeat the procedure
P (µ, σ | θ, Ki , Kj ) = P (σ | µ) P (µ | θ, Ki , Kj ), (D3) 100 times and average P ({ϕa } | θ) for every pair over this
sample, again giving us the expected likelihood function for
with P (µ | θ, Ki , Kj ) = δD [µ− µ
e(θ, Ki , Kj )] and assuming a
this synthetic observation. The associated inferences are il-
flat prior for σ, namely
lustrated in Fig. 3, using the same grid size and smoothing
P (σ | µ) = 10≤σ≤1 , (D4) approach as in §D1.

MNRAS 000, 000–000 (0000)


10 J.-B. Fouvry, M. J. Bustamante-Rosell, A. Zimmerman
D3 Incorporating measurement errors Levin Y., Beloborodov A. M., 2003, ApJ, 590, L33
Lu J. R., Ghez A. M., Hornstein S. D., Morris M. R., Becklin
When applying the present method to the observed S-stars, E. E., Matthews K., 2009, ApJ, 690, 1463
we must also account for the measurement errors in the Magnan N., Fouvry J.-B., Pichon C., Chavanis P.-H., 2022, MN-
stars’ orbital parameters. Fortunately, table 3 of Gillessen RAS, 514, 3452
et al. (2017) provides us with measurement errors for the Máthé G., Szölgyén Á., Kocsis B., 2023, MNRAS, 520, 2204
stars’ semi-major axes, eccentricities, inclinations and longi- Merritt D., 2013, Dynamics and Evolution of Galactic Nuclei.
tudes of the ascending nodes. For simplicity, we assume that Princeton University Press
(i) these errors are independent from one another; (ii) to ac- Mogensen P. K., Riseth A. N., 2018, J. Open Source Softw., 3,
615
commodate the physical ranges of each variables, we assume
Neumayer N., Seth A., Böker T., 2020, A&ARv, 28, 4
that the semi-major axes follow Gamma distributions, the
Paumard T., et al., 2006, ApJ, 643, 1011
eccentricities Beta distributions, and the inclinations and Portegies Zwart S. F., McMillan S. L. W., 2002, ApJ, 576, 899
longitudes Gaussian distributions, with the reported means Pott J. U., et al., 2018, in Ground-based and Airborne Instru-
and variances. For a given realisation, the “observed” pa- mentation for Astronomy VII. p. 1070290
rameters {Ki , L b i }1≤i≤7 are drawn from these distributions, Rauch K. P., Tremaine S., 1996, New Astron., 1, 149
and used in the likelihood estimation. For each set of param- Roupas Z., Kocsis B., Tremaine S., 2017, ApJ, 842, 90
eters, we repeat the procedure 100 times, and average the Szölgyén Á., Kocsis B., 2018, Phys. Rev. Lett., 121, 101101
likelihood over this sample. This has the effect of marginalis- Szölgyén Á., Máthé G., Kocsis B., 2021, ApJ, 919, 140
ing the likelihood over the measurements uncertainties, un- Takács Á., Kocsis B., 2018, ApJ, 856, 113
Tep K., Fouvry J.-B., Pichon C., Heißel G., Paumard T., Perrin
like the previous averaging procedures which produce the
G., Vincent F., 2021, MNRAS, 506, 4289
expectation value of the likelihood for the simulated baths.
Touma J., Tremaine S., Kazandjian M., 2019, Phys. Rev. Lett.,
For SgrA* we have only a single realisation of the dynamics 123, 021103
of the S-stars, and this is a major source of the uncertainty Yelda S., Ghez A. M., Lu J. R., Do T., Meyer L., Morris M. R.,
in our inferences: this can only be reduced by the inclusion Matthews K., 2014, ApJ, 783, 131
of additional observed stars. von Fellenberg S. D., et al., 2022, ApJ, 932, L6
For the inferences presented in §4, we systematically
fix (γ⋆ , γ• ) = (1.5, 1.8) (see §C3), as well as the age of the
disc, T⋆ = 7.1 Myr, and the initial angular size of the disc set
by ϕ0 = 3◦ . In Fig. 4, we also fix m⋆ = 1 M⊙ and the sum
M⋆ (< a0 )+M• (< a0 ) = 45.9×103 M⊙ for a0 = 100 mpc, while
varying m• and the mass fraction of IMBHs. For Fig. 5,
we fix [M⋆ (< a0 ), M• (< a0 )] = [7.9×103 M⊙ , 38×103 M⊙ ] and
vary m⋆ and m• /m⋆ . Finally, we use the same grid size and
smoothing procedure as in §D1.

REFERENCES
Alexander T., 2017, ARA&A, 55, 17
Asmussen S., Jensen J. L., Rojas-Nandayapa L., 2016, Methodol.
Comput. Appl. Probab., 18, 441
Bartko H., et al., 2009, ApJ, 697, 1741
Davies R., et al., 2018, in Ground-based and Airborne Instrumen-
tation for Astronomy VII. p. 107021S
Do T., et al., 2019, BAAS, 51, 530
Fouvry J.-B., Bar-Or B., Chavanis P.-H., 2019, ApJ, 883, 161
Fouvry J.-B., Dehnen W., Tremaine S., Bar-Or B., 2022, ApJ,
931, 8
Fragione G., Loeb A., 2022, ApJ, 932, L17
GRAVITY Collaboration et al., 2019, A&A, 625, L10
GRAVITY Collaboration et al., 2020, A&A, 636, L5
Generozov A., Madigan A.-M., 2020, ApJ, 896, 137
Ghez A. M., et al., 2008, ApJ, 689, 1044
Gillessen S., et al., 2017, ApJ, 837, 30
Ginat Y. B., Panamarev T., Kocsis B., Perets H. B., 2022, arXiv,
2211.14784
Giral Martı́nez J., Fouvry J.-B., Pichon C., 2020, MNRAS, 499,
2714
Gruzinov A., Levin Y., Zhu J., 2020, ApJ, 905, 11
Habibi M., et al., 2017, ApJ, 847, 120
Hamers A. S., Bar-Or B., Petrovich C., Antonini F., 2018, ApJ,
865, 2
Kocsis B., Tremaine S., 2011, MNRAS, 412, 187
Kocsis B., Tremaine S., 2015, MNRAS, 448, 3265
Levin Y., 2022, arXiv, 2211.12754

MNRAS 000, 000–000 (0000)

You might also like