You are on page 1of 14

Astronomy & Astrophysics manuscript no.

f1timedep ©ESO 2020


December 15, 2020

Perturbed distribution functions with accurate action


estimates for the Galactic disc
H. Al Kazwini1 , Q. Agobert1 , A. Siebert1 , B. Famaey1 , G. Monari1 , S. Rozier1 , P. Ramos1 , R. Ibata1 ,
S. Gausland2 , C. Rivière2 , and D. Spolyar3

1
Université de Strasbourg, CNRS UMR 7550, Observatoire astronomique de Strasbourg, 11 rue de l’Université, 67000
Strasbourg, France
2
Université de Strasbourg, UFR de Mathématique et d’Informatique, 7 rue René Descartes, 67084 Strasbourg, France
3
The Oskar Klein Centre for Cosmoparticle Physics, Department of Physics, Stockholm University, AlbaNova, 10691
arXiv:2012.06597v1 [astro-ph.GA] 11 Dec 2020

Stockholm, Sweden

Received; accepted

ABSTRACT

In the Gaia era, understanding the effects of perturbations of the Galactic disc is of major importance in the context
of dynamical modelling. In this theoretical paper, we extend previous work in which, making use of the epicyclic
approximation, the linearized Boltzmann equation had been used to explicitly compute, away from resonances, the
perturbed distribution function of a Galactic thin disc population in the presence of a non-axisymmetric perturbation
of constant amplitude. Here we improve this theoretical framework in two distinct ways in the new code that we present.
First, we use better estimates for the action-angle variables away from quasi-circular orbits, computed from the AGAMA
software, and we present an efficient routine to numerically re-express any perturbing potential in these coordinates
with an accuracy well below the percent level. The use of more accurate action estimates allows us to identify resonances
such as the outer 1:1 bar resonance at larger azimuthal velocities than the outer Lindblad resonance, and to extend our
previous theoretical results well above the Galactic plane, where we explicitly show how they differ from the epicyclic
approximation. In particular, the displacement of resonances in velocity space as a function of height can in principle
constrain the 3D structure of the Galactic potential. Second, we allow the perturbation to be time-dependent, thereby
allowing us to model the effect of transient spiral arms or of a growing bar. The theoretical framework and tools
presented here will be useful for a thorough analytical dynamical modelling of the complex velocity distribution of disc
stars as measured by past and upcoming Gaia data releases.
Key words. Galaxy: kinematics and dynamics – Galaxy: disc – Galaxy: solar neighborhood – Galaxy: structure –
Galaxy: evolution – galaxies: spiral

1. Introduction details the fine structure of stellar action space (e.g., Trick
et al. 2019; Monari et al. 2019b,a). While the existence of
The natural canonical coordinate system of phase-space for moving groups of dynamical origin had been known for
Galactic dynamics and perturbation theory is the system of a long time in local velocity space around the Sun (e.g.,
action-angle variables (Binney & Tremaine 2008). In an ax- Dehnen 1998; Famaey et al. 2005), Gaia revealed their
isymmetric system in equilibrium, Jeans’ theorem implies structure in exquisite details (Ramos et al. 2018), as well as
that the phase-space distribution function (DF) of any stel- an estimate of their age distribution (Laporte et al. 2020),
lar (or dark matter) component can be expressed solely as a together with the shape of the global velocity field away
function of the actions, that are labelling the actual orbits from the Sun within the Galactic disc (Gaia Collaboration
(e.g., Binney & Piffl 2015; Cole & Binney 2017). However, et al. 2018b). One additional major finding of Gaia has been
the effect of various perturbers of the potential (e.g., the the existence of a local phase-spiral in vertical height vs.
Galactic bar and spiral arms) must be included in this pro- vertical velocity in the Solar neighbourhood (Antoja et al.
cess, together with the response of the distribution function. 2018) which might be related to a vertical perturbation of
Within the resonant regions, to fully capture the behaviour the disc by, e.g., the Sagittarius dwarf galaxy (e.g., Laporte
of the DF, one needs to construct for each perturber new et al. 2019; Binney & Schönrich 2018; Bland-Hawthorn &
orbital tori, complete with a new system of action-angle Tepper-Garcia 2020).
variables (e.g., Monari et al. 2017a; Binney 2020a,b). Away
In previous theoretical work, Monari et al. (2016) –
from resonances, however, one can simply use the linearized
hereafter M16 – explicitly computed the response of an ax-
Boltzmann equation. This also allows us to accurately iden-
isymmetric DF in action space, representing a Milky Way
tify the location of resonances.
thin disc stellar population, to a quasi-stationary three-
This is particularly important in the context of the inter- dimensional spiral potential, using the epicyclic approxi-
pretation of recent data from the Gaia mission (Gaia Col- mation to model the actions, which is only a valid approxi-
laboration et al. 2018a, 2020), which revealed in exquisite mation for quasi-circular orbits in the thin disc. It was no-
Article number, page 1 of 14
A&A proofs: manuscript no. f1timedep

tably shown that the first order moments of the perturbed or with a large vertical amplitude. More precise ways of
DF then give rise to non-zero radial and vertical bulk flows determining the action and angle coordinates have been
(breathing modes). However, to treat perturbations away devised. They typically differ depending on whether one
from the disc, which is particularly important in the Gaia wishes to (i) transform angles and actions to positions and
context, one cannot make use of the epicyclic approxima- velocities, or to (ii) make the reverse transformation from
tion to compute action-angle variables. Moreover, it is well- positions and velocities to actions and angles. In the former
known that spiral modes in simulations can be transient, case, a very efficient method is the Torus Mapping method
remaining stationary for only a few rotations (Sellwood & first introduced by McGill & Binney (1990) – see also Bin-
Carlberg 2014), and the response of the disc should be dif- ney & McMillan (2016) for a recent overview –, while in
ferent during the period of rising or declining amplitude. the latter case, a “Stäckel fudge” is generally used (Binney
The same can be true for the bar, whose amplitude and 2012; Sanders & Binney 2016).
pattern speed can also grow or vary with time (e.g., Chiba The general idea of the Torus Mapping is to first express
et al. 2020; Hilmi et al. 2020). the Hamiltonian in the action-angle coordinates (J T , θ T ) of
Here, we improve on this previous modelling of M16 a toy potential, for which the transformation to positions
in two ways. First, we will use a better estimate than the and velocities is fully known analytically. The algorithm
epicyclic approximation for the action-angle variables, re- then searches for a type-2 generating function G(θ T , J ) ex-
lying on the “Torus Mapping" method of McGill & Binney pressed as a Fourier series expansion on the toy angles θ T ,
(1990) to convert from actions and angles to positions and for which the Fourier coefficients are such that the Hamilto-
velocities, as well as on the “Stäckel fudge" (Binney 2012; nian remains constant at constant J . This generating func-
Sanders & Binney 2016) for the reverse transformation. tion fully defines the canonical transformation from actions
This will allow us to extend previous results to eccentric and angles to positions and velocities. For the inverse trans-
orbits and orbits wandering well above the Galactic plane. formation, an estimate based on separable potentials can be
The routines developed and presented in this paper will also used. Such potentials are known as Stäckel potentials (e.g.,
be of fundamental importance to study the vertical pertur- Famaey & Dejonghe 2003), for which each generalized mo-
bations of the Galactic disc in further works. Second, we mentum depends on its conjugated position through three
will let the perturbation evolve with time, thereby allowing isolating integrals of the motion. These potentials are ex-
us to model the effect of a growing bar. pressed in spheroidal coordinates defined by a focal dis-
The paper is organized as follows. In Sect. 2, after a tance. For a Stäckel potential, this focal distance of the
short reminder of the theoretical framework of perturbed coordinate system is related to the first and second deriva-
DF – already exposed in detail in M16 –, we present the tives of the potential. One can thus use the true potential at
method used to expand in Fourier series the perturbing any configuration space point to compute the equivalent fo-
potential within the new action-angle coordinate system. cal distance as if the potential was a Stäckel one, and from
Then, a comparison with the results in the epicyclic case is there compute the corresponding isolating integrals of the
made in Sect. 3. In Sect. 4, we explore the temporal treat- motion, and the actions. In the following, we will make use
ment of the DF for an analytic growth of the amplitude of of both types of transformations, namely the Torus Map-
the perturber. Finally, we conclude and discuss the possible ping to express the potential in action-angle coordinates,
future applications of these theoretical tools in Sect. 5. An and the Stäckel fudge to represent distribution functions in
appendix is dedicated to the presentation of the code. velocity space at a given configuration space point.
Now, let H0 (J ) be the Hamiltonian of the axisymmetric
and time-independent zeroth order gravitational potential
2. The perturbing potential and perturbed DF Φ0 of the Galaxy. The equations of motion connecting ac-
tions J and the canonically conjugate angles θ are then
2.1. Action-angle variables
simply
The canonical phase-space action-angle variables (J ,θ) of
∂H0 ∂H0
an integrable system are obtained from a canonical trans- θ̇ = = ω(J ), J̇ = − = 0, (1)
formation implicitly using Hamilton’s characteristic func- ∂J ∂θ
tion as a type-2 generating function. The actions J are with ω being the fundamental orbital frequencies. Thus, for
defined as new generalized momenta corresponding to a a given orbit, the actions J are constant in time, defining
closed path integral of the velocities along their corre- an orbital torus on which the angles θ evolve linearly with
spondingH canonically conjugate position variable, namely time, according to θ(t) = θ 0 + ωt. Jeans’ theorem then
Ji = vi dxi /(2π). Since this does not depend on time, tells us that the phase-space distribution function (DF) of
these actions are integrals of motion, and the Hamiltonian an axisymmetric potential f = f0 (J ) is in equilibrium. In
can be expressed purely as a function of these. other words, f0 is a solution of the collisionless Boltzmann
It turns out that Galactic potentials are close enough to equation:
integrable systems that actions can be estimated for them.
For quasi-circular orbits close to the Galactic plane, with df
= 0. (2)
separable motion in the vertical and horizontal directions, dt
one can locally approximate the radial and vertical mo-
tions of an orbit with harmonic motions, which is known as 2.2. Perturbed distribution functions
the epicyclic approximation. The radial and vertical actions
then simply correspond to the radial and vertical energies Let now Φ1 be the potential of a small perturbation to
divided by their respective (radial and vertical) epicyclic the axisymmetric background potential Φ0 of the Galaxy,
frequency. However, the epicyclic approximation is not valid with an amplitude relative to this axisymmetric background
anymore when considering orbits with higher eccentricity,   1. Then the total potential is Φ = Φ0 + Φ1 and the
Article number, page 2 of 14
Al Kazwini et al.: Perturbed distribution functions with accurate action estimates

JR = 50 kpc kms 1 JR = 50 kpc kms 1 JR = 50 kpc kms 1

60
40 1
80
2

120 (km 2 s 2)
60
020 (km 2 s 2)

120 (km 2 s 2)
100

120 3
80
140
4
160 100
5
180
120
1400 1500 1600 1700 1800 1900 2000 1400 1500 1600 1700 1800 1900 2000 1400 1500 1600 1700 1800 1900 2000
J (kpc kms 1) J (kpc kms 1) J (kpc kms 1)
J = 1760 kpc kms 1 J = 1760 kpc kms 1 J = 1760 kpc kms 1
82.5
0 0.0
85.0
10
87.5 0.5

120 (km 2 s 2)
20
020 (km 2 s 2)

120 (km 2 s 2)

90.0
1.0
92.5 30
95.0
40 1.5
97.5
50
100.0 2.0
102.5 60
2.5
0 20 40 60 80 100 0 20 40 60 80 100 0 20 40 60 80 100
JR (kpc kms 1) JR (kpc kms 1) JR (kpc kms 1)

Fig. 1. Variations of a few Fourier coefficients φjml (J ) of the bar potential of Sect. 2.4, as JR or Jϕ increase separately at Jz = 0.
The actions on the abscissa axis are in kpc km s−1 . The curves are very smooth, which justifies our use of cubic splines method to
interpolate.

DF becomes, to first order in , f = f0 + f1 , which is still a the perturbing potential. In the case of an m-fold in-plane
solution of the collisionless Boltzmann equation. Inserting perturbation, it is sufficient to take k = m. The main goal
f = f0 + f1 in Eq. (2), and keeping only the terms of order of this section will be to express typical non-axisymmetric
, leads to the linearized collisionless Boltzmann equation perturbing potentials originally expressed in configuration
space as such a Fourier series in action-angle space. The al-
df1 gorithm that we present in Sect. 2.3 can however be applied
+ [f0 , Φ1 ] = 0. (3)
dt to any perturbing potential, including non-plane symmetric
vertical perturbations.
where [,] is the Poisson bracket. Integrating Eq. (3) over
Once the potential is expressed as a function of angles
time, from −∞ to the time t, then leads to
and actions as in Eq. (5), then Eq. (4) becomes
Z t
∂f0 ∂Φ1 0 0 0 X Z t
dt0 0 (J 0 ) · (4)
 
f1 (J , θ, t) = (J , θ , t ), ∂f0 0 0

−∞ ∂J ∂θ 0 f1 (J , θ, t) = Re i (J ) · n dt0 φn (J 0 , t0 ) ein·θ (t ) .
∂J n −∞
where J 0 and θ 0 correspond to the action and angles in the (6)
unperturbed case as a function of time (i.e. constant actions
J 0 and angles evolving linearly). In M16, assuming φn (J 0 , t0 ) = g(t0 ) h(t0 ) φn (J ), with
0
Since any function of the angles is 2π-periodic in the h(t0 ) = e−imΩp t , and the amplitude of the perturbing po-
angles, the perturbing potential Φ1 can be expanded in a tential constant in time at present time (g(t) = 1), and zero
Fourier series as and constant in time at −∞, led to the following explicit
X  solution for f1 (J , θ, t),
Φ1 (J , θ, t) = Re φn (J , t) ein·θ . (5)
eiθs,n
 
∂f0 X
n f1 (J , θ, t) = Re (J ) · nφn (J ) , (7)
∂J n
ωs,n
Hereafter, we will consider in-plane perturbations such as
spiral arms, meaning that we can write the time-varying where we defined
Fourier coefficients in a non-rotating frame as φn (J , t) =
g(t) h(t) φn (J ), where g(t) controls the amplitude of the θs,n = n · θ − mΩp t, (8)
perturbation and h(t) controls its pattern speed, with
h(t) = e−imΩp t , where Ωp is the pattern speed of the per-
turbation and m its azimuthal wavenumber, i.e. its mul- ωs,n = n · ω − mΩp . (9)
tiplicity. We will mostly consider hereafter m = 2 pertur-
bations. The vector index n is a triplet of scalar integer The subscript "s" stands for "slow", because in proximity
indices (j, k, l) running in principle from −∞ to ∞, but in of a resonance of the type ωs,n = 0, the angle θs,n evolves
practice limited to a given range sufficient to approximate very slowly. One can also immediately see that the above
Article number, page 3 of 14
A&A proofs: manuscript no. f1timedep

Fig. 2. Top panel: estimate of the central bar potential of Fig. 3. Top panel: estimate of the spiral arms potential of
Sect. 2.4 at the Sun’s position in the plane with 41 complex Sect. 2.5 at the Sun’s position in the plane with 41 complex
Fourier coefficients in Eq. 10 and the cubic splines reconstruc- Fourier coefficients in Eq. 10 and the cubic splines reconstruc-
tion. The vertical line denotes the true value. Bottom panel: rel- tion. The vertical line denotes the true value. Bottom panel:
ative accuracy. The potential is estimated at the same configura- relative accuracy. The accuracy is again well below the percent
tion space location (the Sun’s position) for different velocities, level, although with a slight bias towards higher amplitudes in
with a global accuracy well below the percent level, although the reconstruction w.r.t. to the input spiral potential.
with a slight bias towards higher amplitudes in the case of the
spiral potential.
2.3. The perturbing potential in actions and angles
linearized solution is valid only away from such resonances, M16 worked in the epicyclic approximation as it provides
since it diverges for these orbits. Orbits near these reso- an analytical expression for evaluating actions and angles
nances are actually trapped, and for them the determina- from cylindrical coordinates. The Fourier coefficients of the
tion of the linearly perturbed DF becomes inappropriate. Fourier series expansion of the perturbing potentials were
A specific treatment for these resonant regions is required, then also determined analytically within this approxima-
which was addressed in, e.g., Monari et al. (2017a) within tion. Approximating the vertical component of the perturb-
the epicyclic approximation, and by Binney (2020a) in a ing potential by an harmonic oscillator, the 9 Fourier coef-
more general context. ficients φjml were then limited to the range j = {−1, 0, 1},
Using the epicyclic approximation, the Fourier coeffi- corresponding to the θR oscillations of the potential, and
cients of a spiral potential have been computed analytically l = {−2, 0, 2}, corresponding to the θz oscillations of the
in M16 with indices n = (j, k, l) running over the values potential close to the Galactic plane.
j = {−1, 0, 1}, k = m = 2, and l = {−2, 0, 2}, and the However, the epicyclic approximation is not valid any-
perturbed distribution function away from resonances was more when considering eccentric orbits. Thus, apart from
then computed. nearly circular orbits, this approximation is not really ef-
In the following, we are going to extend the results of ficient to expand the potential in Fourier series. Hereafter,
M16 to a more general estimate of the action-angle vari- the transformations from angles and actions to positions
ables, through the Torus Mapping method. The resulting and velocities (and reciprocally) are all evaluated numer-
DF will be plotted in velocity space by making use of the ically with AGAMA (Vasiliev 2018). The code makes use of
Stäckel fudge. For both transformations, we will use the the Torus Mapping to go from actions-angles to positions-
AGAMA code (Action-based Galaxy Modelling Architecture, velocities, and uses the Stäckel fudge for the inverse trans-
Vasiliev 2019, 2018). formation. Our goal now is to obtain the Fourier coefficients
Article number, page 4 of 14
Al Kazwini et al.: Perturbed distribution functions with accurate action estimates

Fig. 4. Local stellar velocity distribution at axisymmetric equilibrium from Eq. 16 in the (u, v) plane at (R, z, ϕ) = (R0 , 0, 0).
Left panel: epicyclic approximation. Right panel: Stäckel fudge with AGAMA.

of a known perturbing potential using these numerically 2.4. Bar potential


computed actions (instead of epicyclic).
The potential we choose for the bar is a simple quadrupole
We proceed in the following way for evaluating Fourier potential (Weinberg 1994; Dehnen 2000) with
coefficients of the perturbing potential in Eq. (5). The first
step is to choose a set of actions within a range represent-
 
Φ1,b (R, z, ϕ, t) = Re Φa,b (R, z)eim(ϕ−ϕb −Ωb t) , (11)
ing all the orbits of interest in the axisymmetric background
configuration, each triplet of actions representing one of the
orbits. For each orbit, we then define an array of angles where m = 2, Ωb is the pattern speed of the bar (expressed
(θR , θϕ , θz ). These actions and angles can then all be con- hereafter in multiples of the angular frequency at the Sun
verted to positions thanks to the Torus machinery in AGAMA. Ω0 = v0 /R0 , where v0 is the local circular velocity at the
For each triplet of actions, a range of positions (R, ϕ, z) is Galactocentric radius of the Sun R0 ), and the azimuth is
covered by the angles, and we look for the best-fitting co- defined with respect to a line which would correspond to
efficients φjml (JR , Jz , Jϕ ), satisfying the following equation the Galactic Center-Sun direction in the Milky Way, ϕb
(setting t = 0 for the time being): thus being the angle between the Sun and the long axis of
the bar. We also choose
  −3
r
3  2  R ≥ Rb ,
v02
 
R0 R
X 
Rb

i(jθR +mθϕ +lθz ) Φa,b (R, z) = −αb
Φ1 (R, ϕ, z) = Re φjml (JR , Jz , Jϕ )e .  3
3 Rb r  r
j,l  2−
 R < Rb ,
(10) Rb
(12)

where r2 = R2 + z 2 is the spherical radius, Rb is the length


This is performed with the method of linear least squares by of the bar and αb represents the maximum ratio between
use of singular value decomposition as proposed in chapter the bar and axisymmetric background radial forces at the
15.4 of Press et al. (1992). We then interpolate the value Sun’s Galactocentric radius R = R0 . We use hereafter,
of the coefficient φjml with cubic splines, also as proposed as a representative example, Rb = 0.625 R0 , ϕb = 25o
in Press et al. (1992), chapter 3.3. The number of Fourier and αb = 0.01. We will also consider two typical pattern
coefficients is chosen to be high enough to ensure that all speeds: Ωb = 1.89 Ω0 and Ωb = 1.16 Ω0 .
orbits passing through a given configuration space point
yield the same value of the potential at this point within a The bar potential is quite easy to reproduce using
relative accuracy of less than 1%. Fourier coefficients since it varies smoothly on orbits. Thus,
for a study in the Galactic plane, only 41 complex Fourier
Concretely, we apply this hereafter to the potential of a coefficients for each triplet of actions are sufficient to ap-
central bar and of a 2-armed spiral pattern. The background proximate the value of the potential with a (much) better
axisymmetric potential is chosen to be Model I from Binney than 1% accuracy. The Fourier coefficients themselves vary
& Tremaine (2008). This potential has a bulge described smoothly, as illustrated on Fig. 1, which shows the vari-
by a truncated oblate spheroidal power-law, a gaseous disc ations of a few Fourier coefficients as JR and Jϕ increase
with a hole at the center, a stellar thin disc and a stellar separately, justifying the use of cubic-splines interpolation
thick disc, both with a scale-length of 2 kpc, and a dark halo to get the value of the potential at a specific position. Fig. 2
with an oblate two-power-law profile. The Galactocentric demonstrates the accuracy of our method in reproducing
radius of the Sun is set at R0 = 8 kpc, and the local circular the bar potential in the Solar neighborhood for different
velocity is v0 = 220 km s−1 . values of the local velocities. The potential is estimated at
Article number, page 5 of 14
A&A proofs: manuscript no. f1timedep

the same configuration space location (the Sun’s position) where Rg , Ω, κ and ν are all functions of Jϕ , and
for the whole range of relevant velocities, with a global accu-  
racy well below the percent level. This tool is of course not σ̃R (Rg ) = σ̃R (R0 ) exp −
Rg − R0
,
limited to any specific form of the potential, the only ad- hσR
justable parameter being the number of Fourier coefficients 
Rg − R0

necessary to recover a given perturbing potential with a σ̃z (Rg ) = σ̃z (R0 ) exp − (18)
percent-level accuracy. hσz
For the thin disc DF fthin , we choose hR = 2 kpc, z0 =
2.5. Spiral potential 0.3 kpc, hσR = hσz = 10 kpc, σ̃R (R0 ) = 35 km s−1 and
σ̃z (R0 ) = 15 km s−1 . For the thick disc DF fthick , we choose
The potential we use for the spiral arms is the following hR = 2 kpc, z0 = 1 kpc, hσR = 10 kpc, hσz = 5 kpc,
(Cox & Gómez 2002; Monari et al. 2016) σ̃R (R0 ) = 50 km s−1 and σ̃z (R0 ) = 50 km s−1 .
n o The background axisymmetric potential is chosen to be
Φ1,sp (R, z, ϕ, t) = Re Φa,sp (R, z) eim(ϕ−ϕsp −Ωsp t) , (13) Model I from Binney & Tremaine (2008), in which the above
equilbrium DF f0 is a good representation of the thin and
where m = 2, Ωsp is the pattern speed of the spiral arms, thick disc components. In this model, one has R0 = 8 kpc
and and v0 = 220 km s−1 .
  β Fig. 4 displays the (u, v)-plane in the Solar neighbour-
A ln(R/Rsp ) Kz hood within the z = 0 plane for this f0 axisymmetric
Φa,sp (R, z) = − eim tan(p) sech , (14)
Rsp KD β background, where u = −vR and v = vϕ − v0 , obtained
by converting velocity-space into action-space through the
where epicyclic approximation and the Stäckel fudge from AGAMA.
The velocity distributions are globally similar.
2
K(R) = , β(R) = K(R)hsp [1 + 0.4K(R)hsp ],
R sin(p) 3.2. Resonant zones
1 + K(R)hsp + 0.3[K(R)hsp ]2 In the case of a perturbation with quasi-static amplitude
D(R) = . (15)
1 + 0.3K(R)hsp that has reached its plateau, once the Fourier coefficients
representing the perturbing potential have been computed
Here, Rsp = 1 kpc is the length parameter of the logarith-
– from the epicyclic approximation or from Eq. 10 –, the
mic spiral potential, hsp = 0.1 kpc the height parameter,
expression for the perturbed DF can be simply expressed
p = −9.9o the pitch angle, ϕsp = −26o the phase and
away from resonances with Eq. 7 as
A = 683.4 km2 s−2 the amplitude. Hereafter, we will adopt
a pattern speed Ωsp = 0.84 Ω0 , placing the main resonances n
 X 
away from (or at high azimuthal velocities in) the solar f1 (J , θ, t) = Re φjml Fjml e i[jθR +m(θϕ −Ωp t)+lθz ]
,
neighbourhood. j,l=−n
We note that the spiral potential tends to vary more
(19)
than the bar potential along a specific orbit, especially when
the orbit is not restricted to the plane, which requires us to with n the order of the Fourier series (in this paper, m = 2
increase the number of Fourier coefficients to approximate it in both the bar and spiral cases), and
correctly in 3D. Within the Galactic plane, Fig. 3 shows our
reconstruction of the spiral potential at the Sun’s position, ∂f0
j ∂J ∂f0
+ m ∂J ∂f0
+ l ∂J
again with 41 complex Fourier coefficients. The accuracy is Fjml = R ϕ z
, (20)
again below the percent level as in the bar case, although jωR + m(ωϕ − Ωp ) + lωz
in the spiral case there is a (small) bias towards higher where ωR , ωϕ and ωz can be approximated as epicyclic
amplitudes w.r.t. the input spiral potential. This bias is frequencies in the epicyclic case, or can be determined with
however very small and will not affect our results. AGAMA. The denominator of Fjml may lead to a divergence
in the DF when it approaches zero. Following our notation
3. Results and comparison with the epicyclic in Eq. 9, let us name it
approximation ωs,jml (JR , Jϕ , Jz ) = jωR + m(ωϕ − Ωp ) + lωz . (21)
3.1. Background equilibrium
The amount of such resonances is limited in the epicyclic
From here on, we work with a background axisymmetric case because, by construction, indices run only over the val-
DF f0 as a sum of two quasi-isothermal DFs (Binney & ues j = {−1, 0, 1} and l = {−2, 0, 2}, but they can be much
McMillan 2011) for the thin and thick disc: more numerous in the more accurate AGAMA case. For the
bar potential of Eq. 11 and Eq. 12, and choosing a pattern
f0 (JR , Jz , Jϕ ) = fthin + 0.075fthick . (16) speed Ωp = 1.89Ω0 like for our fiducial bar model, we ex-
plore in Fig. 5 and Fig. 6, the values of ωs,jml (JR , Jϕ , Jz ) in
The form of each DF is: action space, when varying the pair of integer indices (j, l).
Ω exp(−Rg /hR )

JR κ Jz ν
 The actions are renormalized by the radial velocity disper-
f (JR , Jz , Jϕ ) = 2 σ̃ z
exp − 2 − 2 sion of the thin disc, circular velocity, and vertical velocity
2 (2π)3/2 κ σ̃R z 0 σ̃R σ̃z dispersion of the thin disc at the Sun, respectively, to only
(17) display a relevant range of actions. Exploring indices in the
Article number, page 6 of 14
Al Kazwini et al.: Perturbed distribution functions with accurate action estimates

Fig. 5. Values of log(ωs,jml ) in the (JR , Jϕ ) plane with fixed Jz = 10 kpc km s−1 , for a few examples of combinations of (j, l)
indices giving rise to resonant zones in action space (remember that m = 2). The pattern speed Ωp is here 1.89 Ω0 . The two actions
are renormalized by the radial velocity dispersion of the thin disc and circular velocity at the Sun, respectively. The deep blue
lines correspond to resonant zones. For instance, the (1, 0) case corrsponds to the traditional OLR (for a non-zero Jz ). Most other
low-order combinations of indices did not give rise to any relevant resonant zone in the region of interest.

Fig. 6. Values of log(ωs,jml ) in the (JR ,Jz ) plane with fixed Jϕ = 1759 kpc km s−1 . The pattern speed Ωp is the one of our fiducial
central bar fixed at 1.89 Ω0 . The two actions are renormalized by the radial velocity dispersion and vertical velocity dispersion of
the thin disc at the Sun, respectively. The deep blue lines correspond to resonance zones. Most combinations of indices explored
did not give rise to any relevant resonant zone in the region of interest.

range [−4, +4], it appears clearly that most combinations resonances (and thus their location in velocity space) can
do not induce a resonance which is relevant to the dynam- clearly be identified with this linear perturbation method.
ics of the Solar neighbourhood. We only display on Fig. 5
and Fig. 6 the combination of indices for which a resonant
zone appears in the plotted region of action space. It ap- 3.3. Comparing the perturbed DF for different action
pears clearly that very few low order resonances are indeed estimates
present in the range of actions that are really relevant for We are now in a position to compare the linear deforma-
the solar neighbourhood. As displayed on Fig. 7 and Fig. 8, tion of local velocity space for different action estimates,
the same is true for a lower pattern speed Ωp = 0.84 Ω0 , namely the epicyclic case used in previous works and the
corresponding to the pattern speed of our fiducial spiral more accurate AGAMA action estimates.
potential. Fig. 9 displays the f0 + f1 linearly perturbed distribu-
tion function at the position of the Sun in the Galactic plane
While a specific treatment is needed in these resonant for the bar potential of Sect. 2.4 and two different pattern
zones (e.g., Monari et al. 2017a), their sparse presence in the speeds, as well as for the spiral potential of Sect. 2.5. As
relevant zones of action space guarantees us to get meaning- in Monari et al. (2017b), whenever f1 > f0 , we cap f1 at
ful results. Note also that, while the form of the DF is not the value of f0 to roughly represent the resonant zone. The
well estimated in the resonant zones, the signature of the more rigorous approach, which we leave to further works in
Article number, page 7 of 14
A&A proofs: manuscript no. f1timedep

Fig. 7. Same as Fig. 5, with the combinations of indices giving rise to resonant zones for Ωp = 0.84 Ω0 .

Fig. 8. Same as Fig. 6, with some combinations of indices giving rise to resonant zones for Ωp = 0.84 Ω0 .

Fig. 9. The DF of Fig. 4 in velocity space at the Solar position within the Galactic plane, now perturbed to linear order by a
bar (perturbing potential of Sect. 2.4) with pattern speeds Ωb = 1.16 Ω0 (left) and Ωb = 1.89Ω0 (middle), or by a spiral pattern
(perturbing potential of Sect. 2.5) with pattern speed Ωsp = 0.84 Ω0 (right). Top row: Epicyclic approximation. Bottom row:
Stäckel fudge.

the context of AGAMA actions (Al Kazwini et al., in prep.), linear deformation due to the bar is generally stronger in
is to treat the DF with the method of Monari et al. (2017a) the AGAMA case, and that due to the spiral is weaker in the
in these regions. However, while the DF within the reso- AGAMA case. This means that reproducing the effect of spi-
nant zone is not well modelled by the present method, the ral arms on the local velocity distribution might require a
location of resonances is accurately reproduced. The linear higher amplitude when considering an accurate estimate of
deformation outside of the resonant zones is very accurately the action-angle variables rather than the epicyclic approx-
described by our method as well. Interestingly enough, the imation.

Article number, page 8 of 14


Al Kazwini et al.: Perturbed distribution functions with accurate action estimates

Fig. 10. Local stellar velocity distribution perturbed to linear order at the Solar Galactocentric radius and azimuth at three
different heights (left: z = 0 kpc, middle: z = 0.3 kpc, right: z = 1 kpc), when perturbed by a bar (perturbing potential of
Sect. 2.4) with pattern speed Ωb = 1.89Ω0 . Top row: Epicyclic approximation. Bottom row: Stäckel fudge.

Fig. 11. Same as Fig. 10, in the Stäckel fudge case, but now for joint perturbation by a bar (perturbing potential of Sect. 2.4)
with pattern speed Ωb = 1.89Ω0 and a spiral pattern (perturbing potential of Sect. 2.5) with pattern speed Ωsp = 0.84 Ω0 .

The case of the pattern speed of the bar being 1.89 Ω0 distribution is affected by a larger asymmetric drift at large
would correspond to a configuration where the ‘Hercules’ heights, and the location of the outer Lindblad resonance
stream at negative u and negative v corresponds to the of the bar in the uv-plane is also displaced to lower az-
linear deformation below the 2 : 1 outer Lindblad reso- imuthal v at larger heights. This is because at fixed Jϕ , the
nance of the bar (e.g., Dehnen 2000; Minchev et al. 2007; azimuthal and radial frequencies computed with AGAMA are
Monari et al. 2017b; Fragkoudi et al. 2019). Interestingly, lower at higher z, meaning that one needs to reach lower
this feature is less squashed in the more realistic AGAMA Jϕ (corresponding to orbits whose guiding radii are in the
case. Moreover, a resonance unnoticed within the epicyclic inner Galaxy) to reach the resonance. This trend is already
approximation appears at high azimuthal velocities: we can visible at z = 0.3 kpc where the ‘Hercules’ feature moves
identify this resonance as the outer 1 : 1 resonance of the to lower azimuthal v in the AGAMA case, and becomes dra-
bar (Dehnen 2000). In the spiral case, the resonant ridge at matic at z = 1 kpc, where the epicyclic approximation is
large azimuthal velocities can be identified as the corotation not representing accurately the velocity distribution at all.
of the spiral pattern. Interestingly, comparing the displacement with height of
Fig. 10 displays the linear deformation due to the bar, the OLR in the case of a bar with pattern speed 1.89 Ω0
in the case of the pattern speed being 1.89 Ω0 , at different with that of the corotation in the case of a 1.16 Ω0 pat-
heights above the Galactic plane, both in the epicyclic and tern speed, we noted that the corotation location in the
AGAMA cases. As can be seen on this figure, the epicyclic uv-plane is more displaced than the OLR. This is because
approximation quickly becomes very imprecise at large the corotation only depends on the azimuthal frequency,
heights, because it implies a drastic falloff of the density while the OLR depends on a combination of the azimuthal
with height while not changing the azimuthal velocity dis- and radial frequencies. However, this different behaviour
tribution due to the hypothesis of complete decoupling of clearly depends on the location in the velocity plane: at
vertical motions. In the AGAMA case, the azimuthal velocity high-azimuthal velocities, the corotation of the spiral arm

Article number, page 9 of 14


A&A proofs: manuscript no. f1timedep

is less displaced with z than the outer 1 : 1 resonance of To calculate the second part of the integral, since
the bar, causing them to almost merge at z = 0.3 kpc dg(t)/dt = π/(2tf ) sin(πt/tf ), we write,
when linearly adding the effect of the bar and spiral on
Fig. 11. The two resonances are however separated again t t
πt0
Z Z  
at z = 1 kpc. At this height, other resonances involving dg 0 π 1 0
(t )η(t0 )dt0 = sin eiθs,n (t ) dt0 .
the radial and vertical frequencies also appear in velocity 0 dt0 2tf iωs,n 0 tf
space. All these subtle variations however depend strongly (27)
on the background Galactic potential. This means that,
once identifying the resonances potentially responsible for
We look for a primitive G of sin(πt/tf )eiθs,n (t) of the form
moving groups in the Solar neighbourhood, studying their
position in the uv-plane as a function of z can in princi-     
ple be a powerful way to constrain the 3D structure of the πt πt
G(t) = A cos + B sin eiθs,n (t) . (28)
Galactic potential. This cannot be done within the epicyclic tf tf
approximation.
Deriving G(t) with respect to t, and equating it to the in-
4. Adding the temporal evolution tegrand in Eq. 27 we get

Up to now, we always considered a constant amplitude for π π


B + A iωs,n = 0 and B iωs,n − A = 1, (29)
the perturbing potential in order to determine an analyt- tf tf
ical expression for the perturbed DF. In this section, we
investigate the time dependence of the DF by choosing a which leads to
time-varying amplitude for the perturbing potential.
π/tf −iωs,n
A= 2
and B= . (30)
4.1. Time-varying amplitude function ωs,n − (π/tf )2 2
ωs,n − (π/tf )2
The expression we will use for the time dependence function
g controlling the amplitude of the perturbation during its Substituting the Eqs. (26) and (28) into Eq. (25) results
growth is in the following expression for the perturbed DF
1 − cos (πt/tf )
(22)
(
g(t) = , ∂f0
2
X
f1 (J , θ, t) = Re (J ) · n φn (J ) ×
where tf is the time at which the perturbation is completely ∂J n
formed, expressed in Gyr. We will consider tf = 0.5 Gyr. "    iθs,n
1 πt e π 1 1
The motivation for this choice of growth function is its 1 − cos − ×
2 tf ωs,n 2 − (π/t )2
2tf ωs,n ωs,n
analytic simplicity, having a function starting from exactly f
zero at the origin, and smooth over the whole considered     
range. The first derivative, [π/(2tf )] sin(πt/tf ), assures the π πt πt π
cos − iωs,n sin eiθs,n − ei(θs,n −ωs,n t
continuity at 0 and tf with both stages, fixed at 0 for t ≤ 0 tf tf tf tf
and at 1 for t ≥ tf (the first derivative is thus equal to 0 at (31)
0 and tf ).
Note that we do not exactly recover the static case at
4.2. Time-dependent perturbed distribution function t = tf because all derivatives of g(t) are not strictly zero at
We now take the integral of Eq. (6), restricted to [0, t] (be- the initial and final time, as assumed in M16. If reaching
cause the g function is equal to 0 on ]−∞, 0]) and inte- a true plateau after tf in an analytic fashion, one would
grate by parts. We take φn (J 0 , t0 ) = g(t0 ) h(t0 ) φn (J ), with nevertheless converge towards the static case very quickly.
0 Now we can study analytically how the linear response
h(t0 ) = e−imΩp t , and we define
to a fiducial bar with Ωb = 1.89Ω0 evolves with time. As
eiθs,n (t) before, the method is not strictly valid at resonances, where
η(t) ≡ → dη = eiθs,n (t) dt, (23) the treatment of Monari et al. (2017a) must be applied. It
iωs,n
is nevertheless interesting to see on Fig. 12 how the linear
allowing us to rewrite Eq. (6) as deformation of the velocity plane evolves with time near
 Z t  resonances (in a patch co-moving with the bar, hence at
∂f0 dη
g(t0 ) 0 (t0 ) dt0 . constant azimuthal angle to the bar) while the amplitude
X
f1 (J , θ, t) = Re i (J ) · nφn (J )
∂J n 0 dt of the perturbation grows. The effect of the OLR appears
(24) as soon as the perturbation starts to grow. As it progres-
sively grows, the two linear modes in the DF separate, and
We can now integrate by parts lead to a velocity plane already very much resembling the
Z t

Z t
dg 0 stationary form of the perturbed DF after 0.25 Gyr, that
t
g(t0 ) 0 (t0 ) dt0 = [g(t0 )η(t0 )]0 − 0
(t ) η(t 0
) dt 0
. (25) is when g(t) = 0.5 and the perturbation is half-formed. In
0 dt 0 dt the absence of a pattern speed variation, it is therefore not
Since g(0) = 0, necessarily obvious to disentangle the effect of a bar whose
amplitude is growing from that of a fully formed bar with
t
[g(t0 )η(t0 )]0 = g(t)η(t). (26) larger and constant amplitude.
Article number, page 10 of 14
Al Kazwini et al.: Perturbed distribution functions with accurate action estimates

Fig. 12. Local stellar velocity distribution perturbed to linear order in the Galactic plane by the bar of Sect. 2.4 with Ωb = 1.89Ω0
with the Stäckel fudge, and an amplitude of the bar growing as described in Sect. 4.1. The first eleven panels display the temporal
evolution of the perturbation. The last panel displays the stationary case. The amplitude of the bar goes from 0 at t = 0 to its
plateau (g(t) = 1) at t = 0.5 Gyr.

5. Conclusion vertical motions. In the AGAMA case instead, the locations


of resonances are displaced to lower azimuthal v at larger
Starting from the formalism exposed in M16, we proposed a heights. This is already visible at z = 0.3 kpc but becomes
more accurate way to determine the DF of the Galactic disc a dramatic effect at z = 1 kpc. Thus, the position of moving
perturbed to linear order by a non-axisymmetric perturba- groups in the uv-plane as a function of z can be a powerful
tion, using a more accurate action-angle coordinate system. way to constrain the 3D structure of the Galactic potential.
First, we used the Torus Mapping from AGAMA to numeri- The key to exploring this will be the DR3 of Gaia (Brown
cally compute the perturbing potential in action-angle co- 2019) with its ∼ 35 millions radial velocities allowing to
ordinates, as a Fourier series expansion over the angles. We better probe the z-axis above and below the Milky Way
showed that we could estimate typical non-axisymmetric plane.
perturbing potentials with an accuracy well below the per- Finally, the temporal treatment is also an improvement
cent level. The algorithm can be applied to any perturbing over M16. We applied it to the case of a bar of growing
potential, including non-plane symmetric vertical perturba- amplitude, with an analytic evolution of the amplitude. As
tions, which will be particularly important when studying the bar progressively grows, the two linear modes in the DF
the vertical perturbations of the disc with similar methods separate, and lead to a velocity plane already very much
(Rozier et al., in prep.). resembling the stationary form of the perturbed DF once
We then compute the DF perturbed to linear order by the perturbation is half-formed. In the absence of a pattern
a typical bar or spiral potential (or a linear combination speed variation, it is therefore not necessarily obvious to
of both), and compute the local stellar velocity distribu- disentangle the effect of a bar whose amplitude is growing
tion by converting velocities to actions and angles through from that of a fully formed bar with larger and constant
the Stäckel fudge implemented in AGAMA. The results are amplitude. Note that we explored here a peculiar form of
compared to those obtained by using the epicyclic approx- the growth function motivated by its analytic simplicity. If
imation. The linear deformation due to the bar is gener- the perturbation grows by a linear instability, an exponen-
ally stronger in the AGAMA case, and that due to the spi- tial growth would be more realistic. Numerical experiments
ral is weaker in the AGAMA case.This means that reproduc- are usually well fitted by a logistic function (exponential
ing the effect of spiral arms on the local velocity distri- growth at the beginning, and saturation to the plateau).
bution might require a higher amplitude when considering One problem for our treatment is that the logistic func-
an accurate estimate of the action-angle variables rather tion is never strictly equal to 0. Besides, there is hope that
than the epicyclic approximation. Most importantly, the similar analytical simplifications as those for the amplitude
epicyclic approximation is inadequate at large heights above growth studied here can also be made with this function,
the plane, because it implies a drastic falloff of the density which we will investigate in the future.
with height while not changing the azimuthal velocity dis- While the form of the DF is not well estimated in the
tribution due to the hypothesis of complete decoupling of resonant zones with the linear perturbations presented in
Article number, page 11 of 14
A&A proofs: manuscript no. f1timedep

this paper, the signature of the resonances (and thus their Monari, G., Famaey, B., Fouvry, J.-B., & Binney, J. 2017a, MNRAS,
location in velocity space) can clearly be identified with this 471, 4314
linear perturbation method. The more rigorous approach, Monari, G., Famaey, B., & Siebert, A. 2016, MNRAS, 457, 2569
Monari, G., Famaey, B., Siebert, A., et al. 2019a, A&A, 632, A107
which we leave to further works in the context of AGAMA ac- Monari, G., Famaey, B., Siebert, A., et al. 2017b, MNRAS, 465, 1443
tions (Al Kazwini et al., in prep.), is to treat the DF with Monari, G., Famaey, B., Siebert, A., Wegg, C., & Gerhard, O. 2019b,
the method of Monari et al. (2017a) in these regions, patch- A&A, 626, A41
ing these results over the linear deformation computed here. Press, W. H., Teukolsky, S. A., Vetterling, W. T., & Flannery, B. P.
1992, Numerical Recipes in C: Second Edition (Cambridge Univer-
Another caveat is that the Torus Mapping was used for ex- sity Press)
pressing the perturbing potential in actions and angles, but Ramos, P., Antoja, T., & Figueras, F. 2018, A&A, 619, A72
for the estimate of the local stellar velocity field, we made Sanders, J. L. & Binney, J. 2016, MNRAS, 457, 2107
use of the less precise Stäckel fudge method. Therefore, an- Sellwood, J. A. & Carlberg, R. G. 2014, ApJ, 785, 137
other promising way for improvement would be to use the Trick, W. H., Coronado, J., & Rix, H.-W. 2019, MNRAS, 484, 3291
Vasiliev, E. 2018, AGAMA: Action-based galaxy modeling framework,
new ACTIONFINDER deep-learning algorithm (Ibata et al. Astrophysics Source Code Library
2020) to make the reverse transformation. Vasiliev, E. 2019, MNRAS, 482, 1525
The tools presented in this paper will be useful for a Weinberg, M. D. 1994, ApJ, 420, 597
thorough analytical dynamical modelling of the complex
velocity distribution of Milky Way disc stars as measured by
past and upcoming Gaia data releases. These tools will also
be useful for fully self-consistent treatments of the response
of the disc to external perturbations. The ultimate goal
will be to adjust models to the exquisite data from Gaia,
which cannot be done properly with N-body simulations
due to the vast parameter space to explore. The theoretical
tools and the new code presented in this paper therefore
represent a useful step in this direction.
Acknowledgements. AS, BF, GM, SR, PR and RI acknowledge fund-
ing from the Agence Nationale de la Recherche (ANR project ANR-18-
CE31-0006 and ANR-19-CE31-0017) and from the European Research
Council (ERC) under the European Union’s Horizon 2020 research
and innovation programme (grant agreement No. 834148).

References
Antoja, T., Helmi, A., Romero-Gómez, M., et al. 2018, Nature, 561,
360
Binney, J. 2012, MNRAS, 426, 1324
Binney, J. 2020a, MNRAS, 495, 886
Binney, J. 2020b, MNRAS, 495, 895
Binney, J. & McMillan, P. 2011, MNRAS, 413, 1889
Binney, J. & McMillan, P. J. 2016, MNRAS, 456, 1982
Binney, J. & Piffl, T. 2015, MNRAS, 454, 3653
Binney, J. & Schönrich, R. 2018, MNRAS, 481, 1501
Binney, J. & Tremaine, S. 2008, Galactic Dynamics: Second Edition
(Princeton University Press)
Bland-Hawthorn, J. & Tepper-Garcia, T. 2020, arXiv e-prints,
arXiv:2009.02434
Brown, A. G. A. 2019, in The Gaia Universe, 18
Chiba, R., Friske, J. K. S., & Schönrich, R. 2020, MN-
RAS[arXiv:1912.04304]
Cole, D. R. & Binney, J. 2017, MNRAS, 465, 798
Cox, D. P. & Gómez, G. C. 2002, ApJS, 142, 261
Dehnen, W. 1998, AJ, 115, 2384
Dehnen, W. 2000, AJ, 119, 800
Famaey, B. & Dejonghe, H. 2003, MNRAS, 340, 752
Famaey, B., Jorissen, A., Luri, X., et al. 2005, A&A, 430, 165
Fragkoudi, F., Katz, D., Trick, W., et al. 2019, MNRAS, 488, 3324
Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2018a, A&A,
616, A1
Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2020, arXiv
e-prints, arXiv:2012.01533
Gaia Collaboration, Katz, D., Antoja, T., et al. 2018b, A&A, 616,
A11
Hilmi, T., Minchev, I., Buck, T., et al. 2020, MNRAS, 497, 933
Ibata, R., Diakogiannis, F., Famaey, B., & Monari, G. 2020, arXiv
e-prints, arXiv:2012.05250
Laporte, C. F. P., Famaey, B., Monari, G., et al. 2020, A&A, 643, L3
Laporte, C. F. P., Minchev, I., Johnston, K. V., & Gómez, F. A. 2019,
MNRAS, 485, 3134
McGill, C. & Binney, J. 1990, MNRAS, 244, 634
Minchev, I., Nordhaus, J., & Quillen, A. C. 2007, ApJ, 664, L31

Article number, page 12 of 14


Al Kazwini et al.: Perturbed distribution functions with accurate action estimates

Appendix A: PERDIGAL code presentation


All the results shown in this article were obtained with a
single code, written in C++ and Python, that calculates
both the perturbing potential in action-angle coordinates
and the perturbed DF. The code is named PERDIGAL (PER-
turbed DIstribution functions for the GALactic disc), and
will be made available on request, although it may even-
tually be embedded in a larger Galactic dynamics toolkit.
Launching the code without argument gives the explana-
tions shown in Fig. A.1.

The calculation of the perturbed DF consists in five steps:


the creation of a file storing the positions of the DF (DF-
pos), the creation (via AGAMA) of several files containing
positions from a lot of orbits (FCorb) required for the fol-
lowing step, the determination of Fourier coefficients for
the perturbing potential to create the grid (FCgrid), the
determination of the Fourier coefficients for each position
which the DF will be calculated at (FCforDF) and finally
the calculation of the perturbed DF (PertDF). The mode
ALL is processing the three above steps in one time, with-
out saving Fourier coefficients from FCgrid and FCforDF in
files. Then, the mode ALL regroups in one the three modes:
FCgrid, FCforDF and PertDF. However, this mode should
not be used without certainty about the decomposition of
the potential. Indeed, this process depends a lot on the ini-
tial conditions imposed to the code, and a check after each
part of the calculation is recommended. Finally, PDFdisp
displays the perturbed DF using Matplotlib.

All useful parameters are stored in two particular files


and can be modified without compiling the code. They con-
tain parameters for the calculation of Fourier coefficients,
parameters for the perturbing potentials and others for
whether the DF is determined at first or second order, or
with a time-dependent amplitude for the perturbing poten-
tial. Figure A.2 is a logigram reminding the procedure of
the calculation of the perturbed DF.

Article number, page 13 of 14


A&A proofs: manuscript no. f1timedep

Fig. A.1. Guide for the use of the code. The first three modes deal with the determination of the Fourier coefficients when
expanding in series the perturbing potential (FCgrid and FCforDF), and the calculation of the perturbed distribution function
(PertDF). PDFdisp allows this function to be displayed as a distribution of velocities. ALL regroups the three above modes in one,
but is not recommended because of the need to check the Fourier decomposition. EPI calculates directly the perturbed distribution
function under the epicyclic approximation, as explained in M16. The FUNCTIONs correspond to different operations required
before using the MODEs. DFpos creates the file storing the positions at which the distribution function is going to be determined.
FCorb creates several files containing positions from a lot of orbits, needed to calculate the Fourier coefficients in the FCgrid step.

Fig. A.2. Logigram reminding the procedure of the calculation of the perturbed DF. The different steps correspond respectively
to the mode DFpos, FCorb, FCgrid, FCforDF and PertDF. There are two cases where we need to check the result of the code,
after the third step and after the fourth step. In the first case, we have to control if the decomposition is correct and in the second
case if the perturbing potential is well reproduced.

Article number, page 14 of 14

You might also like