The following article is Open access

A Comprehensive Perturbative Formalism for Phase Mixing in Perturbed Disks. II. Phase Spirals in an Inhomogeneous Disk Galaxy with a Nonresponsive Dark Matter Halo

, , and

Published 2023 July 18 © 2023. The Author(s). Published by the American Astronomical Society.
, , Citation Uddipan Banik et al 2023 ApJ 952 65 DOI 10.3847/1538-4357/acd641

Download Article PDF
DownloadArticle ePub

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

0004-637X/952/1/65

Abstract

We develop a linear perturbative formalism to compute the response of an inhomogeneous stellar disk embedded in a nonresponsive dark matter (DM) halo to various perturbations like bars, spiral arms, and encounters with satellite galaxies. Without self-gravity to reinforce it, the response of a Fourier mode phase mixes away due to an intrinsic spread in the vertical (Ωz), radial (Ωr), and azimuthal (Ωϕ) frequencies, triggering local phase-space spirals. The detailed galactic potential dictates the shape of phase spirals: phase mixing occurs more slowly and thus phase spirals are more loosely wound in the outer disk and in the presence of an ambient DM halo. Collisional diffusion due to scattering of stars by structures like giant molecular clouds causes superexponential damping of the phase spiral amplitude. The zvz phase spiral is one-armed (two-armed) for vertically antisymmetric (symmetric) bending (breathing) modes. Only transient perturbations with timescales (τP) comparable to the vertical oscillation period (τz ∼ 1/Ωz) can trigger vertical phase spirals. Each (n, l, m) mode of the response to impulsive (τP < τ = 1/(nΩz + lΩr + mΩϕ)) perturbations is power-law (∼τP/τ) suppressed, but that to adiabatic (τP > τ) perturbations is exponentially weak ($\sim \exp \left[-{\left({\tau }_{{\rm{P}}}/\tau \right)}^{\alpha }\right]$) except for resonant (τ ) modes. Slower (τP > τz) perturbations, e.g., distant encounters with satellite galaxies, induce stronger bending modes. Sagittarius (Sgr) dominates the solar neighborhood response of the Milky Way (MW) disk to satellite encounters. Thus, if the Gaia phase spiral was triggered by a MW satellite, Sgr is the leading contender. However, the survival of the phase spiral against collisional damping necessitates an impact ∼0.6–0.7 Gyr ago.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Disk galaxies are characterized by large-scale ordered motion and are therefore highly responsive to perturbations. Following a time-dependent gravitational perturbation, the actions of the disk stars are modified. This in turn causes a perturbation in the distribution function (DF) of the disk known as the response. Over time, the response decays away as the system "relaxes" toward a new quasi-equilibrium via collisionless processes that include kinematic processes like phase mixing (loss of coherence in the response due to different oscillation frequencies of stars) and secular/self-gravitating/collective processes like Landau damping (loss of coherence due to wave–particle interactions; Lynden-Bell 1962). As pointed out by Sridhar (1989) and Maoz (1991), phase mixing is the key ingredient of all collisionless relaxation and re-equilibration.

The timescale of collisionless equilibration is typically longer than the orbital periods of stars. Therefore, disk galaxies usually harbor prolonged features of incomplete equilibration following a perturbation, e.g., bars, spiral arms, warps, and other asymmetries. An intriguing example is the one-armed phase-space spiral, or phase spiral for short, discovered in the Gaia DR2 data (Gaia Collaboration et al. 2018) by Antoja et al. (2018) and discussed in more detail in subsequent studies (e.g., Bland-Hawthorn et al. 2019; Laporte et al. 2019; Li 2021; Li & Widrow 2021; Gandhi et al. 2022). Antoja et al. (2018) plotted the density of stars in the solar neighborhood in the (z, vz ) plane of vertical position, z, and vertical velocity, vz , and noticed a faint spiral pattern that became more pronounced when color coding the (z, vz ) "pixels" by the median radial or azimuthal velocities. The one-armed spiral shows 2–3 complete wraps like a snail shell, and it is interpreted as an indication of vertical phase mixing following a perturbation that is antisymmetric about the midplane (bending mode) and occurred ∼500 Myr ago. More recently, Hunt et al. (2022) used the more extensive Gaia DR3 data to study the distributions of stars in zvz space at different locations in the Milky Way (MW) disk. They found that, unlike the one-armed phase spiral or bending mode at the Solar radius, the inner disk shows a two-armed phase spiral that corresponds to a breathing mode or symmetric perturbation about the midplane. They inferred that, while the one-armed spiral in the Solar neighborhood might have been caused by the impact of a satellite galaxy such as the Sagittarius (Sgr) dwarf, the two-armed spiral in the inner disk could not have been induced by the same, because almost all satellite impacts are far too slow/adiabatic from the perspective of the inner disk. Rather, they suggested that the two-armed phase spiral might haven been triggered by a transient spiral arm or bar.

The phase spiral holds information about the perturbative history and gravitational potential of the disk, and therefore it can serve as an essential tool for galactoseismology (Widrow et al. 2014; Johnston et al. 2017). For a given potential, the winding of the spiral is an indication of the time elapsed since the perturbation, with older spirals revealing more wraps. A one-armed (two-armed) phase spiral corresponds to a bending (breathing) mode. Which mode dominates, in turn, depends on the timescale of the perturbation, with temporally shorter (longer) perturbations (e.g., a fast or slow encounter with a satellite) predominantly triggering breathing (bending) modes (Widrow et al. 2014; Banik et al. 2022).

In addition to depending on the nature of the perturbation, the phase spiral also encodes information about the oscillation frequencies of stars and thus the detailed potential. In particular, the shape of the spiral depends on how the vertical frequencies, Ωz , vary as a function of the vertical action, Iz , which in turn depends on the underlying potential. In Banik et al. (2022, hereafter Paper I), we showed that the amplitude of the phase spiral can damp away due to lateral mixing, with a damping rate that depends on both the spatiotemporal nature of the perturbation and the frequency structure of the galaxy. This damping, though, only affects the response in the coarse-grained sense, i.e., upon marginalization of the response over the lateral degrees of freedom (the action-angle variables). Damping at the fine-grained level requires self-gravity of the response (Landau damping) or collisional diffusion, such as that arising from the gravitational scattering of stars against giant molecular clouds (GMCs), or dark matter (DM) substructure (Tremaine et al. 2023).

Paper I addresses the problem of inferring the nature of the perturbation from the amplitude and structure of the phase spiral using a model of an infinite, isothermal slab for the unperturbed disk. This simple, yet insightful, model provides us with essential physical understanding of the perturbative response of disks without the complexity of modeling a realistic, inhomogeneous disk. However, it suffers from certain glaring caveats: (i) lateral uniformity leading to an incorrect global structure of the response in the lateral direction, (ii) Maxwellian distribution of velocities in the lateral direction that overpredicts lateral mixing and thereby the rate at which the amplitude of the phase spiral damps out, (iii) absence of a DM halo, and (iv) absence of self-gravity of the response. In this paper, we relax the first three assumptions. We consider an inhomogeneous disk characterized by a realistic DF similar to the pseudo-isothermal DF (Binney 2010) that properly captures the orbital dynamics of the disk stars in 3D. In addition, we consider the effect of an underlying DM halo, which for the sake of simplicity we consider to be nonresponsive. This ambient DM halo alters the potential and thus the frequencies of stars, which can in turn affect the shape of the phase spiral and its coarse-grained survival. We also consider the impact of small-scale collisionality on the fine-grained survival of the phase spiral. Because in this paper we are primarily interested in the phase mixing of the disk response that gives rise to phase spirals, we ignore the self-gravity of the response, which to linear order spawns coherent point mode oscillations of the disk as a whole (for treatments of the self-gravitating response of isothermal slabs, see Mathur 1990 and Weinberg 1991) and somewhat enhances the amplitude of phase spirals.

This paper is organized as follows. Section 2 describes the standard linear perturbation theory for galaxies and its application to a realistic disk galaxy embedded in a DM halo that is exposed to a general perturbation. Sections 3 and 4 are concerned with computing the disk response for different perturber models. In Section 3, we compute the disk response and phase spirals induced by bars and spiral arms. We also discuss the impact of collisional diffusion on the fine-grained survivability of the phase spiral. In Section 4, we compute the response to encounters with satellite galaxies. We summarize our findings in Section 5.

2. Linear Perturbation Theory for Galaxies

2.1. Linear Perturbative Formalism

A galaxy, to very good approximation, is devoid of star–star collisions. However, there are other potential sources of collisions, such as scatterings due to gravitational interactions of stars with GMCs or DM substructure. The dynamics of stars in such a system is governed by the Fokker–Planck equation (FPE):

Equation (1)

where f denotes the DF, H denotes the Hamiltonian, square brackets denote the Poisson bracket, and C[f] denotes the collision operator due to small-scale fluctuations, which can be approximated by a Fokker–Planck operator (see Appendix A of Tremaine et al. 2023):

Equation (2)

where ${\boldsymbol{\xi }}=\left({\boldsymbol{q}},{\boldsymbol{p}}\right)$, with q and p denoting the canonically conjugate position and momentum variables, and Dij denotes the diffusion coefficient tensor.

Let the unperturbed steady-state Hamiltonian of the galaxy be H0 and the corresponding DF be given by f0, which satisfies the unperturbed FPE:

Equation (3)

In the presence of a small time-dependent perturbation in the potential, ΦP(t), the perturbed Hamiltonian can be written as

Equation (4)

where Φ1 is the gravitational potential related to the response density, ρ1 = ∫f1 d3 v , via the Poisson equation,

Equation (5)

The perturbed DF can be written as

Equation (6)

where f1 is the linear-order perturbation in the DF. In the weak perturbation limit where linear perturbation theory holds, the time-evolution of f1 is dictated by the following linearized form of the FPE:

Equation (7)

Throughout this paper, we neglect the self-gravity of the disk response, which implies that we set the polarization term, [f0, Φ1] = 0. The implications of including self-gravity are discussed in Paper I.

2.2. Response of a Galactic Disk to a Realistic Perturbation

The dynamics of a realistic disk galaxy like the MW is quasi-periodic, i.e., it can be characterized by oscillations in the azimuthal, radial, and vertical directions. In close proximity to the midplane and under radial epicyclic approximation, the Hamilton–Jacobi equation becomes separable, implying that all stars confined within a few vertical scale heights from the midplane of the disk are on regular, quasi-periodic orbits that are characterized by a radial action IR , an azimuthal action Iϕ , and a vertical action Iz . Hence, the motion of each star is characterized by three frequencies:

Equation (8)

and three angles, ${w}_{i}(t)={w}_{i}(0)+{{\rm{\Omega }}}_{i}t$, with $i=R,\phi $ or $z$. This quasi-periodic nature of the orbits near the midplane is approximately preserved even in the presence of a (non-triaxial) DM halo, because this preserves the axisymmetry of the potential. Typically, as discussed in Section 2.3, the presence of a halo increases the oscillation frequencies of the disk stars.

In terms of these canonical conjugate action-angle variables, using Equation (8), the linearized form of the FPE given in Equation (7) becomes

Equation (9)

Here, we have performed several simplifications of the Fokker–Planck operator (see Appendix A for details). First, the diffusion coefficients are computed using the epicyclic approximation, i.e., small Iz and IR , because only such stars are significantly affected by collisional scattering. In addition, following Tremaine et al. (2023), we consider ${D}_{I}^{(z)}$ and ${D}_{I}^{(R)}$ to be nearly constant, something that is implied by the age–velocity dispersion relation of the MW disk stars. Second, the IR diffusion of the response f1 is negligible because the frequencies do not depend on IR under the radial epicyclic approximation (and only mildly depend on IR without it), and therefore the response does not develop IR gradients. Third, following Binney & Lacey (1988), we have neglected diffusion in Iϕ and wϕ , because the terms involving Dϕ ϕ , Dr ϕ , and Dϕ z are smaller than the Iz and IR diffusion terms by factors of at least σR /vc or σz /vc , which are typically much smaller than unity (σR and σz are radial and vertical velocity dispersions, respectively, and vc is the circular velocity along ϕ). We have retained the wz and wR diffusion terms for the sake of completeness, but as we point out later, the diffusion in angles typically occurs over timescales much longer than that in actions—and hence it is comparatively less important.

Because the stars move along quasi-periodic orbits characterized by actions and angles, we can expand the perturbations, ΦP and f1, as discrete Fourier series in the angles as follows:

Equation (10)

where w = (wz , wR , wϕ ) and I = (Iz , IR , Iϕ ). Substituting these Fourier expansions in Equation (9) yields the following differential equation for the evolution of f1nlm :

Equation (11)

This can be solved using the Green's function technique, with the initial condition, f1nlm (ti) = 0, to yield the following closed integral form for f1nlm :

Equation (12)

Here, for ${D}_{I}^{(z)}\ll {\sigma }_{z}^{2}$ (σz is the vertical velocity dispersion), which is typically the case, ${{ \mathcal I }}_{{nlm}}({\boldsymbol{I}},t)$ can be approximately expressed as

Equation (13)

Here, ${{ \mathcal G }}_{{nlm}}(t-\tau )$ is the Green's function (see Appendix A for detailed derivation), given by

Equation (14)

where Ωz1 = ∂Ωz /∂Iz . The sinusoidal factor represents the oscillations of stars at their natural frequencies, which vary with actions, leading to the formation of phase spirals (see Section 3.1.1 for details). The first exponential damping factor indicates the damping of the response due to diffusion in angles, while the second damping factor manifests the damping of the Iz gradients of the response by diffusion in Iz . As discussed in Section 3.1.2, the diffusion in actions is much more efficient than that in angles.

Each (n, l, m) Fourier coefficient of the response acts as a forced damped oscillator with three different natural frequencies, nΩz , lΩR , and mΩϕ , which is being driven by an external time-dependent perturber potential, Φnlm , and damped due to collisional diffusion. A similar expression, albeit without allowing for collisionality, for the DF perturbation has been derived by Carlberg & Sellwood (1985) in the context of spiral arm induced perturbations and radial migrations in the galactic disk, and by other previous studies (e.g., Lynden-Bell & Kalnajs 1972; Tremaine & Weinberg 1984; Carlberg & Sellwood 1985; Weinberg 1989, 1991, 2004; Kaur & Sridhar 2018; Banik & van den Bosch 2021a; Kaur & Stone 2022) in the context of dynamical friction in spherical systems. To obtain the final expression for f1nlm , we need to specify the spatiotemporal behavior of the perturber potential, ΦP, as well as the DF, f0, of the unperturbed galaxy, which is addressed below.

2.3. The Unperturbed Galaxy

Under the radial epicyclic approximation (small IR ), the unperturbed DF, f0, for a rotating MW-like disk galaxy can be well approximated as a pseudo-isothermal DF, i.e., written as a nearly isothermal separable function of the azimuthal, radial and vertical actions. Following Binney (2010), we write

Equation (15)

The vertical structure of this disk is isothermal, while the radial profile is pseudo-isothermal. 3 Here, ${\rm{\Sigma }}={\rm{\Sigma }}(R)\,={\int }_{-\infty }^{\infty }{dz}\,\rho (R,z)$ is the surface density of the disk, Lz is the z-component of the angular momentum, which is equal to Iϕ , Rc = Rc(Lz ) is the guiding radius, Ωϕ is the circular frequency, and $\kappa =\kappa ({R}_{{\rm{c}}})={\mathrm{lim}}_{{I}_{R}\to 0}{{\rm{\Omega }}}_{R}$ is the radial epicyclic frequency (Binney & Tremaine 1987). Θ(x) is the Heaviside step function. Thus, we assume that the entire galaxy is composed of prograde stars with Lz > 0.

The density profile, ρ(R, z), of the disk corresponding to the above DF is the product of a radially exponential profile with scale radius hr and a vertically isothermal (sech2) profile with thickness hz (Equation (B3)). As shown by Smith et al. (2015), this density profile is accurately approximated by a sum of three Miyamoto & Nagai (1975) disks, 4 which has a simple, analytical form for the associated potential. Therefore, we use this 3MN approximation for our disk throughout this work, because this drastically simplifies the computation of orbital frequencies. For the purpose of computing the disk response, we assume typical MW-like parameters for the various quantities, i.e., R = 8 kpc, disk mass Md = 5 × 1010 M, hR = 2.2 kpc, σR (R) = σR,⊙ = 35 km s−1, hz = 0.4 kpc, and ${\sigma }_{z}({R}_{\odot })={\sigma }_{z,\odot }=\sqrt{2\pi {{Gh}}_{z}{\rm{\Sigma }}({R}_{\odot })}=23\,\mathrm{km}\,{{\rm{s}}}^{-1}$ (McMillan 2011; Bovy & Rix 2013). For this set of parameters, the Toomre Q = σR κ/(3.36GΣ) for the stellar disk turns out to be 2.26 in the Solar neighborhood, indicating that the disk is gravitationally stable. The isothermal vertical distribution of disk stars adopted here ignores the potential influence of gas near the midplane, which may increase the shear, $\left|d{{\rm{\Omega }}}_{z}/d\mathrm{ln}{I}_{z}\right|$, for small Iz . This may affect the shape of the phase spiral, making it more tightly wound, but should not substantially impact its amplitude.

The disk is assumed to be embedded in an extended DM halo characterized by a spherical Navarro–Frenk–White (NFW; Navarro et al. 1997) density profile, with virial mass Mvir, concentration c, scale radius rs , and the corresponding potential Φh given by Equation (B6). We adopt Mvir = 9.78 × 1011 M, rs = 16 kpc, and c = 15.3 (Bovy 2015).

The combined potential experienced by the disk stars is simply the sum of disk and halo potentials, i.e.,

Equation (16)

The total energy of a disk star under the radial epicyclic approximation is $E={L}_{z}^{2}/2{R}_{c}^{2}+{{\rm{\Phi }}}_{0}({R}_{{\rm{c}}},0)+\kappa {I}_{R}+{E}_{z}$, where the vertical part of the energy is given by ${E}_{z}={v}_{z}^{2}/2+{{\rm{\Phi }}}_{z}({R}_{{\rm{c}}},z)$, with Rc(Lz ) being the guiding radius given by ${L}_{z}^{2}/{R}_{{\rm{c}}}^{3}=\partial {{\rm{\Phi }}}_{0}/\partial R{| }_{R={R}_{{\rm{c}}}}$. The vertical potential, Φz (Rc, z), is given by

Equation (17)

The vertical action, Iz , can be obtained from Ez as follows:

Equation (18)

where ${{\rm{\Phi }}}_{z}({R}_{{\rm{c}}},{z}_{\max })={E}_{z}$. This implicit equation can be inverted to obtain Ez (Rc, Iz ). The time period of vertical oscillation can then be obtained using

Equation (19)

which yields the vertical frequency, Ωz (Rc, Iz ) = 2π/Tz (Rc, Iz ). The tightness of a vertical phase spiral at a given Iz increases for faster phase mixing, i.e., increasing value of the shear, $\left|d{{\rm{\Omega }}}_{z}/d\mathrm{ln}{I}_{z}\right|$, at that Iz . The phase-mixing timescale, τϕ , is given by the inverse of the shear:

Equation (20)

Figure 1 plots Ωz and τϕ as a function of Iz in the left and right panels, respectively, for four different cases: (i) MW disk plus halo at Rc = 8 kpc (solid red line), (ii) only the MW disk at Rc = 8 kpc (dashed red line), (iii) MW disk plus halo at Rc = 12 kpc (dotted–dashed blue line), and (iv) only the MW disk at Rc = 12 kpc (dotted blue line). The presence of the DM halo deepens and steepens the potential well and thus increases the vertical frequencies. In the presence of the halo and farther out in the disk, the phase-mixing timescale, τϕ , is longer, i.e., phase mixing occurs more slowly, and the phase spiral is more loosely wound. The sensitivity of the tightness of the phase spiral to the detailed MW potential implies that the phase spiral can be used to constrain it.

Figure 1.

Figure 1. Impact of DM halo on vertical phase mixing: the left and right panels, respectively, indicate the vertical frequency, Ωz (in units of σz,⊙/hz ), and the vertical phase-mixing timescale, τϕ (in units of hz /σz,⊙, given by Equation (20)), as a function of the vertical action, Iz (in units of hz σz,⊙). The solid and dashed red lines denote the cases with and without a halo for Rc = R = 8 kpc, while the dotted–dashed and dotted blue lines show the same for Rc = 12 kpc. The vertical dashed line indicates roughly the maximum Iz for which a phase spiral is discernible in the Gaia data. It should be noted that phase mixing occurs the fastest for Iz ∼ 1 and that the inner disk phase mixes faster than the outer disk. We also note that the presence of a DM halo increases Ωz as well as τϕ , leading to slower phase mixing and therefore slower wrapping of the phase spiral. This effect is more pronounced in the outer disk.

Standard image High-resolution image

Substituting the expression of f0 for a MW-like disk given by Equation (15) in Equation (12), we obtain the following integral form for f1nlm :

Equation (21)

As we shall see, the first-order disk response expressed above phase mixes away and gives rise to phase spirals due to oscillations of stars with different frequencies, except when they are resonant with the frequency of the perturber. However, this "direct" response of the disk does not include certain effects. First of all, we ignore the self-gravity of the response. As discussed in Paper I, to linear order, self-gravity gives rise to point mode oscillations of the disk that are decoupled from the phase-mixing component of the response, which is what we are interested in. Second, for the sake of simplicity, we consider the ambient DM halo to be nonresponsive, and therefore we ignore the indirect effect of the halo response on disk oscillations. We leave the inclusion of these two effects in the computation of the disk response for future work.

The spatiotemporal nature of the perturbing potential dictates the disk response. In this paper, we explore two different types of perturbation to which realistic disk galaxies can be exposed and which are thus of general astrophysical interest. The first is an in-plane spiral/bar perturbation with a vertical structure, either formed as a consequence of secular evolution or triggered by an external perturbation. We consider both short-lived (transient) and persistent spirals/bars. The second type of perturbation that we consider is that due to an encounter with a massive object, e.g., a satellite galaxy or DM subhalo.

3. Disk Response to Spiral Arms and Bars

We model the potential of a spiral arm perturbation as one with a vertical profile and a sinusoidal variation along radial and azimuthal directions:

Equation (22)

Here, ΩP is the pattern speed and kR is the horizontal wavenumber of the spiral perturbation. The long wavelength limit, kR → 0, corresponds to a bar. We consider the in-plane part of ΦP to be a combination of an axisymmetric (mϕ = 0) and a two-armed spiral mode (mϕ = 2), and the vertical part to be a combination of antisymmetric/odd and symmetric/even perturbations, respectively, denoted by ${{ \mathcal F }}_{{\rm{o}}}$ and ${{ \mathcal F }}_{{\rm{e}}}$, that are modulated by growth functions, ${{ \mathcal M }}_{{\rm{o}}}(t)$ and ${{ \mathcal M }}_{{\rm{e}}}(t)$, capturing the growth and/or decay of the spiral strength over time. The ratio of the maximum strengths of the antisymmetric and symmetric parts of the perturbation is α. We consider the following two functional forms for ${{ \mathcal M }}_{j}(t)$ (where the subscript j = o or e):

Equation (23)

The first option describes a transient spiral/bar that grows and decays like a Gaussian pulse with a characteristic lifetime τPj ∼ 1/ωj (Paper I). The second form describes a persistent spiral perturbation that grows exponentially on a timescale τGj ∼ 1/γj and then saturates to a constant amplitude. We shall see shortly that these two kinds of spiral perturbations perturb the disk in very different ways.

The vertical part of the perturbation consists of an antisymmetric function, ${{ \mathcal F }}_{{\rm{o}}}(z)$, and a symmetric function, ${{ \mathcal F }}_{{\rm{e}}}(z)$, which, for the sake of simplicity, we take to be the following trigonometric functions:

Equation (24)

Here, ${k}_{z}^{({\rm{o}})}$ and ${k}_{z}^{({\rm{e}})}$ denote the vertical wavenumbers of the antisymmetric and symmetric perturbations, respectively. Because the above functions form a complete Fourier basis in z, any (vertical) perturber profile can be expressed as a linear superposition of ${{ \mathcal F }}_{{\rm{o}}}$ and ${{ \mathcal F }}_{{\rm{e}}}$. The disk response involves the Fourier coefficients of the perturbing potential, Φnlm , which can be obtained by taking the Fourier transform of ΦP given in Equation (22) with respect to the angles, wR , wϕ , and wz , as detailed in Appendix C.

3.1. Computing the Disk Response

The expression for the disk response to bars or spiral arms can be obtained by substituting the Fourier coefficient of the perturber potential given in Equation (C5) in Equation (21) and performing the τ integration with the initial time, ti → − . This yields the modal response, f1nlm (Equation (21)), with ${{ \mathcal I }}_{{nlm}}({\boldsymbol{I}},t)$ given by

Equation (25)

where ${{\rm{\Psi }}}_{{nlm}}^{({\rm{o}})}$ and ${{\rm{\Psi }}}_{{nlm}}^{({\rm{e}})}$, respectively, denote the time-independent parts of the odd and even terms in the expression for Φnlm , and

Equation (26)

which characterizes the temporal evolution of the response. Here, the subscript j = o or e, and the resonance frequency, ${{\rm{\Omega }}}_{\mathrm{res}}$, is given by

Equation (27)

3.1.1. Collisionless Limit

First, we examine the response in the limit of zero diffusion, i.e., ${D}_{I}^{(z)}=0$, where each star acts as a forced oscillator.

Transient spirals and bars—First, we consider the case of transient spiral arm or bar perturbations that grow and decay in strength over time, i.e., the temporal modulation ${{ \mathcal M }}_{j}(t)$ is given by the first of Equations (23). In this case,

Equation (28)

The error function describes the growth and transient oscillations of the response amplitude; over time, the transients die away, and in the limit t, the response saturates to a constant amplitude (in the absence of collisional diffusion).

The left-hand panel of Figure 2 plots the amplitude of the steady-state disk response to transient spiral/bar perturbations, relative to the unperturbed DF, as a function of the modulation/pulse frequency, ωj (j = o and e for bending and breathing modes, respectively), for different modes indicated in different colors. Solid and dashed lines correspond to the n = 1 bending modes and the n = 2 breathing modes, respectively. We adopt ΣP = 5.5 M pc−2, ΩP = 12 km s−1 kpc−1, ${k}_{z}^{(o)}={k}_{z}^{(e)}=1\,\mathrm{kpc}$, kR = 10 kpc, and Iz = Iz,⊙hz σz,⊙ = 9.2 km s−1, and marginalize the response over IR . We set α = 1, implying equal maximum strengths for the bending and breathing modes. As evident from this figure, and also from Equation (28), the long-term strength of the disk response (after the initial transients have died out like ${e}^{-{\omega }_{j}^{2}{t}^{2}}$) scales as ∼1/ωj in the impulsive (large ωj ) limit, but is superexponentially suppressed ($\sim \exp \left[-{{\rm{\Omega }}}_{\mathrm{res}}^{2}/4{\omega }_{j}^{2}\right]$) in the adiabatic (small ωj ) limit away from resonances, i.e., for ${{\rm{\Omega }}}_{\mathrm{res}}\ne 0$. The adiabatic suppression scales differently with ωj for other functional forms of ${{ \mathcal M }}_{j}(t)$; e.g., for ${{ \mathcal M }}_{j}(t)=1/\sqrt{1+{\omega }_{j}^{2}{t}^{2}}$, the response strength is exponentially suppressed $(\sim \exp [-{{\rm{\Omega }}}_{\mathrm{res}}/{\omega }_{j}])$. However, the response of resonant modes (${{\rm{\Omega }}}_{\mathrm{res}}=0$) does not undergo adiabatic suppression, and it scales as ∼1/ωj throughout, becoming nonlinear in the adiabatic regime. Because there are many resonance modes, the cumulative response in the adiabatic limit of all modes combined is only suppressed as a power law, rather than an exponential, in ωj (Weinberg 1994a, 1994b).

Figure 2.

Figure 2. MW disk response to transient bars/two-armed spirals with Gaussian temporal modulation in absence of collisional diffusion: Left panel shows the steady-state (t ) amplitude of the disk response, f1,nlm /f0, in the solar neighborhood, computed using Equations (25) and (28) in presence of an ambient DM halo, as a function of the pulse frequency, ωj , where the subscripts j = o and e for vertically antisymmetric (odd n) and symmetric (even n) perturbations. Solid (dashed) lines indicate the n = 1 bending (n = 2 breathing) modes and different colors denote (l, m) = (0, − 2), (0, 0) and (0, 2), respectively. We consider Iz = Iz,⊙hz σz,⊙ and marginalize the response over IR . It should be noted that the response peaks at intermediate values of ωj , which are different for different modes, and it is suppressed like a power law in the impulsive (large ωj ) limit and superexponentially in the adiabatic (small ωj ) limit. Right panel shows the breathing-to-bending ratio, f1,200/f1,100, as a function of ωe and ωo, the pulse frequencies of the bending and breathing mode perturbations, respectively. The dashed, solid, dotted–dashed, and dotted contours correspond to breathing-to-bending ratios of 0.1, 1, 5, and 10, respectively. The breathing-to-bending ratio rises and falls with increasing ωe at fixed ωo, while the reverse occurs with increasing ωo at fixed ωe, leading to a saddle point at (ωe, ωo) ≈ (9, 7).

Standard image High-resolution image

The sinusoidal factor, $\exp \left[-i(n{{\rm{\Omega }}}_{z}+l\kappa +m{{\rm{\Omega }}}_{\phi })t\right]$, in ${{ \mathcal P }}_{{nlm}}^{(j)}$ describes the oscillations of stars with three different frequencies, Ωz , κ, and Ωϕ , along the vertical, radial, and azimuthal directions, respectively. Due to the dependence of these frequencies on the actions, that of Ωz on Iz and of κ and Ωϕ on Iϕ = Lz , the response integrated over actions eventually phase mixes away. This manifests as phase spirals in the ${I}_{z}\cos {w}_{z}-{I}_{z}\sin {w}_{z}$ and ${I}_{\phi }\cos \phi -{I}_{\phi }\sin \phi $ phase spaces, which are proxies for the zvz and $\phi \mbox{--}\dot{\phi }$ phase spaces, respectively. As is evident from Equation (26), ${{ \mathcal P }}_{{nlm}}^{(j)}\sim \exp \left[-{im}{{\rm{\Omega }}}_{{\rm{P}}}t\right]$ in the adiabatic limit (ωj → 0); hence, in this limit, the sinusoidal factor $\exp \left[-i(n{{\rm{\Omega }}}_{z}+l\kappa +m{{\rm{\Omega }}}_{\phi })t\right]$ is absent from the response, which implies that phase spirals only occur for sufficiently impulsive perturbations. As shown in Paper I, n = 1 bending modes involve a dipolar perturbation in the vertical phase space (${I}_{z}\cos {w}_{z}-{I}_{z}\sin {w}_{z}$) distribution immediately after the perturbing pulse reaches its maximum strength. This dipolar distortion is subsequently wound up into a one-armed phase spiral because Ωz is a function of Iz . Breathing modes, on the other hand, involve an initial quadrupolar perturbation in the phase-space distribution, which is subsequently wrapped up into a two-armed phase spiral. Because Ωz , Ωϕ , and ΩR all depend on Lz , the amplitude of the ${I}_{z}\cos {w}_{z}-{I}_{z}\sin {w}_{z}$ phase spiral damps out over time due to mixing between stars with different Lz . The modal response, f1nlm , when marginalized in a narrow bin of size ΔLz around Lz , damps out as follows:

Equation (29)

Because the frequencies vary with Lz , marginalizing over Lz mixes phase spirals that differ slightly in phases, giving way to a ∼1/t damping accompanied by a beat-like modulation with a characteristic lateral mixing timescale:

Equation (30)

This explains why the density contrast of the Gaia phase spiral is enhanced upon color coding by vϕ , or equivalently, Lz (Antoja et al. 2018; Bland-Hawthorn et al. 2019). Radial phase mixing is also present, but is typically much weaker because none of the frequencies depend on IR under the radial epicyclic approximation and only mildly depend on IR without it. Hence, due to ordered motion, the phase spiral amplitude in a realistic disk galaxy damps out at a much slower rate, as ∼1/t (in absence of collisional diffusion), than the lateral mixing damping in the isothermal slab case considered in Paper I, which arises from the unconstrained lateral velocities of the stars and exhibits a Gaussian temporal behavior.

It is worth emphasizing that not all frequencies undergo phase mixing. In fact, the resonant frequencies, for which

Equation (31)

do not phase mix away. Hence, parts of the phase space closer to a resonance take longer to phase mix away. Moreover, as is manifest from the adiabatic suppression factor, $\exp [-{{\rm{\Omega }}}_{\mathrm{res}}^{2}/4{\omega }_{j}^{2}]$, the near-resonant modes with ${{\rm{\Omega }}}_{\mathrm{res}}\ll 2{\omega }_{j}$ have much larger amplitude than those with ${{\rm{\Omega }}}_{\mathrm{res}}\gg 2{\omega }_{j}$ that are far from resonance. Therefore, the long-term disk response consists of stars in (near) resonance with the perturbing bar or spiral arm. Most of the strong resonances are confined to the disk plane, including the corotation resonance (n, l, m) = (0, 0, m), the Lindblad resonances (0, ±1, ±2), the ultraharmonic resonances (0, ±1, ±4), and so on. For thin disks with hz hR , the vertical degrees of freedom are generally not in resonance with the radial or azimuthal ones, because Ωz is much larger than Ωϕ or κ. Hence, the vertical oscillation modes (n ≠ 0) such as the n = 1 bending or n = 2 breathing modes undergo phase mixing and give rise to phase spirals. However, if the disk has significant thickness, then the vertical degrees of freedom can be in resonance with the horizontal ones, e.g., banana orbits (Ωz = 2Ωr ) in barred disks.

The excitability of the bending and breathing modes is dictated by the perturbation timescale, or more precisely by the ratio of the pulse frequency, ωj , and the resonant frequency, ${{\rm{\Omega }}}_{\mathrm{res}}$. The right panel of Figure 2 shows the breathing-to-bending ratio, f1,200/f1,100, as a function of ωe and ωo, with blue (yellow) shades indicating low (high) values. In general, the breathing-to-bending ratio rises steeply and falls gradually with ωe at fixed ωo, while the trend is reversed as a function of ωo at fixed ωe, resulting in a saddle point at (ωe, ωo) ≈ (9, 7). This owes to the superexponential suppression in the adiabatic (${\omega }_{j}\ll {{\rm{\Omega }}}_{\mathrm{res}}$) limit and the power-law suppression in the impulsive (${\omega }_{j}\gg {{\rm{\Omega }}}_{\mathrm{res}}$) limit. Along the ωo = ωe line, the bending (breathing) modes dominate in the adiabatic (impulsive) limit, which is also evident from the left panel of Figure 2. All this suggests that bending modes dominate over breathing modes when (i) the antisymmetric perturbation is more impulsive, i.e., evolves faster than the symmetric one, or (ii) both symmetric and antisymmetric perturbations occur over comparable timescales but more slowly than the stellar vertical oscillation period.

Persistent spirals and bars—Next, we consider perturbations caused by a persistent spiral arm or bar that grows exponentially until it saturates at a constant strength. The corresponding temporal modulation ${{ \mathcal M }}_{j}(t)$ is given by the second form of Equation (23). In this case, as shown by Equation (19) of Banik & van den Bosch (2021a):

Equation (32)

Up to t = 0, when the perturber amplitude stops growing, the response from all modes oscillates with the pattern speed ΩP and grows hand-in-hand with the perturber. Subsequently, as the perturbation attains a steady strength, the disk response undergoes temporary phase mixing due to the oscillations of stars at different frequencies, giving rise to phase spirals. These transients, however, are quickly taken over by long-term oscillations driven at the forcing frequency ΩP.

For a slowly growing spiral/bar, i.e., in the "adiabatic growth" limit (γ → 0), the entire disk oscillates at the driving frequency, ΩP, i.e.,

Equation (33)

This has two major implications. First of all, because all stars, both resonant and nonresonant, are driven at the pattern speed of the perturbing spiral/bar, transient phase mixing does not occur and thus no phase spiral arises. Second, the response is dominated by the resonances, ${{\rm{\Omega }}}_{\mathrm{res}}=0$. In fact, the resonant response diverges, reflecting the failure of (standard) linear perturbation theory near resonances. The adiabatic invariance of actions is partially broken near these resonances, causing the stars to get trapped in librating near-resonant orbits. A proper treatment of the near-resonant response can be performed by working with "slow" and "fast" action-angle variables (Tremaine & Weinberg 1984; Lichtenberg & Lieberman 1992; Banik & van den Bosch 2022; Chiba & Schönrich 2022; Hamilton et al. 2022), which are uniquely defined for each resonance as linear combinations of the original action-angle variables. The fast actions remain nearly invariant while the fast angles oscillate with periods comparable to the unperturbed orbital periods of stars. The slow action-angle variables, on the other hand, undergo large-amplitude oscillations about their resonance values over a libration timescale that is typically much longer than the orbital periods. For example, at corotation resonance (n = l = 0), angular momentum behaves as the slow action while the radial and vertical actions behave as the fast ones.

Based on the above discussion, we infer that phase spirals can only be excited in the galactic disk by transient spiral/bar perturbations whose amplitude changes over a timescale comparable to the vertical oscillation periods of stars. Persistent spirals or bars rotating with a fixed pattern speed cannot give rise to phase spirals. Rather, they trigger stellar oscillations at the pattern speed itself, which manifests in phase space as a steadily rotating dipole or quadrupole, depending on whether the n = 1 or 2 mode dominates the response. Thus, a phase spiral is necessarily always triggered by a transient perturbation.

3.1.2. Impact of Collisions on the Disk Response

In the above section, we discussed the characteristics of the disk response in the absence of collisions. However, in a real galaxy like the MW disk, small-scale collisionality can potentially damp away any coherent response to a perturbation. Collisional diffusion arises not from star–star collisions, which are typically negligible, but from gravitational scattering with other objects, such as GMCs, DM substructure, etc. As discussed in Section 2.2, the impact of collisional diffusion is mainly captured by the diffusion coefficients ${D}_{I}^{(z)}$ and ${D}_{I}^{(R)}$. Following Tremaine et al. (2023), we assume that the disk stars have gained their mean vertical and radial actions over the age of the disk, Tdisk = 10 Gyr, due to collisional heating, which implies that ${D}_{a}=\left\langle {I}_{a}\right\rangle /{T}_{\mathrm{disk}}$, where a is either z or R and $\left\langle {I}_{a}\right\rangle =\int {{dI}}_{a}\,{I}_{a}\,{f}_{0}\,/\int {{dI}}_{a}\,{f}_{0}$.

For a transient bar/spiral with pulse frequency ωj , ${{ \mathcal P }}_{{nlm}}^{(j)}$ is given by Equation (26). In the impulsive limit (ωj ), we have that ${{ \mathcal M }}_{j}(t-\tau )\to {\omega }_{j}\delta (t-\tau )$. Upon absorbing ωj in the prefactor, the expression for ${{ \mathcal P }}_{{nlm}}^{(j)}$ then simplifies to

Equation (34)

This demonstrates that, in the impulsive limit, the disk response instantaneously grows and spawns phase spirals whose amplitude decays due to collisional diffusion, manifest in the exponential damping terms. The first and second exponential factors, respectively, characterize the diffusion in vertical angle and action, which occur over the following timescales:

Equation (35)

Of these, the timescale for the diffusion in angles, ${\tau }_{{\rm{D}}}^{({\rm{w}})}$, typically exceeds that for the diffusion in actions, ${\tau }_{{\rm{D}}}^{({\rm{I}})}$, by at least an order of magnitude, implying that angle diffusion is negligible. Hence, collisional diffusion mainly causes the abatement of action gradients in the phase-space structure of the response (arising from the action dependence of the frequencies, i.e., Ωz1 ≠ 0). The left (right) panel of Figure 3 plots the diffusion timescale, ${\tau }_{{\rm{D}}}^{({\rm{I}})}$, as a function of Iz (Rc) for three different values of Rc (Iz ) as indicated. We note that ${\tau }_{{\rm{D}}}^{({\rm{I}})}$ diverges in the small Iz limit, attains a minimum around Iz ∼ 0.2 − 0.5 hz σz,⊙, and increases as ${I}_{z}^{\beta }$ with β < 1 at large Iz . As a function of Rc, ${\tau }_{{\rm{D}}}^{({\rm{I}})}$ shows an approximately exponential rise. This is owed to the fact that $\left\langle {D}_{I}^{(z)}\right\rangle \sim {h}_{z}\,{\sigma }_{z}({R}_{{\rm{c}}})/{T}_{\mathrm{disk}}\,\sim \exp \left[-{R}_{{\rm{c}}}/2{h}_{R}\right]/{T}_{\mathrm{disk}}$ for the 3MN profile adopted for the MW disk. At Rc = R = 8 kpc, ${\tau }_{{\rm{D}}}^{(I)}\sim 0.6\mbox{--}0.7\,\mathrm{Gyr}$, in agreement with Tremaine et al. (2023). Hence, we see that collisional diffusion in action space is fairly efficient, and thus that phase spirals are short-lived features.

Figure 3.

Figure 3. Timescale at which the disk response damps away due to collisional diffusion, i.e., small-scale scatterings of stars with structures like GMCs, is plotted as a function of Iz (Rc ) for three different values of Rc (Iz ) as indicated, in the left (right) panel. Typically, collisional diffusion occurs faster for smaller Iz and smaller Rc .

Standard image High-resolution image

Figure 4 plots the amplitude of the disk response (for IR = 0) to a transient spiral of pulse frequency, ωo = ωe = 0.5σz,⊙/hz , computed using Equations (21), (25), and (26), as a function of time. Dashed and solid lines show the results with and without collisional diffusion, respectively. The rows correspond to different values of Rc, while the columns denote different values of ${I}_{z}/({h}_{z}{\sigma }_{z}({R}_{c}))$ as indicated. The blue and red lines denote the response for the (n, l, m) = (1, 0, 0) and (2, 0, 0) modes, respectively, and the dotted gray line represents the Gaussian pulse strength. The response for both bending and breathing modes initially grows hand in hand with the perturbing pulse. Following the point of maximum pulse strength, the response follows the decaying pulse strength before saturating to the steady-state amplitude given in Equation (28) in the collisionless limit. In the presence of collisional diffusion, however, the response continues to damp out as $\sim \exp [-{(t/{\tau }_{{\rm{D}}}^{({\rm{I}})})}^{3}]$ after temporarily saturating at the collisionless steady state. We note that the collisional damping is faster for smaller Rc and smaller Iz . In addition, n = 2 breathing modes damp out faster than the n = 1 bending modes, due to the n−2/3 dependence of ${\tau }_{{\rm{D}}}^{({\rm{I}})}$.

Figure 4.

Figure 4. MW disk response to transient bars/two-armed spirals with Gaussian temporal modulation of pulse frequency, ωo = ωe = 0.5 σz,⊙/hz : the amplitude of the disk response, f1n00/f0, is plotted as a function of time. The rows and columns, respectively, denote different values of Rc and Iz as indicated. Blue and red lines indicate the n = 1 and 2 modes, while the solid and dashed lines, respectively, denote the cases with and without collisional diffusion (due to interactions of stars with structures like GMCs). The disk response initially rises and falls hand in hand with the perturbing pulse (indicated by the gray dotted line), before saturating to a steady state in the collisionless case and undergoing superexponential damping in the collisional case. Collisional damping is faster for smaller Iz , smaller Rc , and larger n modes.

Standard image High-resolution image

To summarize, we have shown that phase spirals can be triggered by impulsive perturbations resulting from transient spiral arms or bars, but are subject to superexponential damping due to collisional diffusion that is likely to be dominated by scattering against GMCs. This collisional damping is more efficient in the inner disk, for stars with smaller Iz , and for modes of larger n.

4. Disk Response to Satellite Encounter

In addition to the spiral arm/bar perturbations considered above, we also consider disk perturbations triggered by encounters with a satellite galaxy. For the sake of brevity, we only compute the disk response in the collisionless limit. In the case of impulsive encounters, the impact of collisional diffusion is simply expressed by multiplying the collisionless response by the collisional damping factor $\exp [-{(t/{\tau }_{{\rm{D}}}^{({\rm{I}})})}^{3}]$, with ${\tau }_{{\rm{D}}}^{({\rm{I}})}$ given by Equation (35).

For simplicity, we assume that the satellite is moving with uniform velocity vP along a straight line, impacting the disk at a galactocentric distance Rd with an arbitrary orientation, specified by the angles θP and ϕP, which are, respectively, defined as the angle between v P and the z-axis and that between the projection of v P on the midplane and the x-axis (see Figure 5). Thus, the position vector of the satellite with respect to the galactic center can be written as

Equation (36)

while that of a star is given by

Equation (37)

We consider the satellite to be a Plummer sphere of mass MP and size ε, such that its gravitational potential at location r is given by

Equation (38)

Here, the first term is the "direct" term and the second is the "indirect" term that accounts for the reflex motion of the disk and the fact that the disk center is accelerated by the satellite and is thus noninertial. Typically, the first one dominates over the second.

Figure 5.

Figure 5. Illustration of the geometry of a satellite galaxy with mass MP and scale radius ε impacting a disk galaxy with uniform velocity vP along a straight line. The impact occurs at a galactocentric distance Rd. The orientation of v P is specified by θP, the angle between v P and the z-axis, and ϕP, the angle between the projection of v P on the midplane and the x-axis.

Standard image High-resolution image

In order to compute the disk response to this external perturbation, we need to compute its Fourier coefficients, which is challenging. Rather, we first evaluate the τ-integral in Equation (21), setting ti → − , and then compute the Fourier transform of the result, as worked out in Appendix D.1. For IR ≈ 0 (this is justified because we adopt the radial epicyclic approximation in this paper), this yields a modal response, f1nlm (Equation (21)), with ${{ \mathcal I }}_{{nlm}}({\boldsymbol{I}},t)$ given by

Equation (39)

where Ω is given by

Equation (40)

Here, ${{ \mathcal R }}_{{\rm{c}}}={ \mathcal R }({R}_{{\rm{c}}})$ and ${{ \mathcal S }}_{{\rm{c}}}={ \mathcal S }({R}_{{\rm{c}}})$, with ${ \mathcal R }$ and ${ \mathcal S }$ given by Equation (D5). K0i is given by Equation (D3), which asymptotes to the modified Bessel function of the second kind, ${K}_{0}\left(\left|{\rm{\Omega }}\right|\sqrt{{{ \mathcal R }}_{{\rm{c}}}^{2}+{\varepsilon }^{2}}/{v}_{{\rm{P}}}\right)$, in the large time limit. A more precise expression for ${{ \mathcal I }}_{{nlm}}$ that is valid for higher values of IR is given by Equation (D7) of Appendix D.1.

The expression for ${{ \mathcal I }}_{{nlm}}$ given in Equation (39) exhibits several key features of the disk response to satellite encounters. The $\exp \left[-i{\rm{\Omega }}t\right]$ factor encodes the phase mixing of the response due to oscillations at different frequencies, giving rise to phase spirals. The $\exp \left[i\left({\rm{\Omega }}\cos {\theta }_{{\rm{P}}}/{v}_{{\rm{P}}}\right)z\right]$ and $\exp \left[i\left({\rm{\Omega }}\sin {\theta }_{{\rm{P}}}\cos (\phi -{\phi }_{{\rm{P}}})/{v}_{{\rm{P}}}\right){R}_{{\rm{c}}}\right]$ factors, respectively, indicate that the satellite induces wave-like perturbations in the disk with two characteristic wavenumbers: the vertical wavenumber, ${k}_{z}\approx {\rm{\Omega }}\cos {\theta }_{{\rm{P}}}/{v}_{{\rm{P}}}$ and the horizontal wavenumber, ${k}_{R}\approx {\rm{\Omega }}\sin {\theta }_{{\rm{P}}}/{v}_{{\rm{P}}}$. Therefore, the disk response will be vertically (horizontally) stratified in case of a perpendicular (planar) impact of the satellite. As shown in Appendix D.2, the response in the simple case of a satellite having a face-on, perpendicular, impulsive encounter through the center of the disk can be obtained from an asymptotic analysis of the general response to satellite encounters, given by Equations (21) and (39).

4.1. Asymptotic Behavior of the Response

It is instructive to study the two extreme cases of encounter speed, the impulsive limit (large vP) and the adiabatic limit (small vP). Using the asymptotic form of the K0 Bessel function that appears in Equation (39), we obtain the following approximate asymptotic behavior of f1nlm at large time:

Equation (41)

where b is the impact parameter of the encounter, defined as the perpendicular distance of the nearest star on the midplane from the satellite's (straight) orbit, and expressed as

Equation (42)

It is clear from these limits that the disk response is most pronounced for intermediate velocities, vP ∼ Ωb. For impulsive encounters, the response is suppressed as a power law in vP, whereas in the adiabatic limit, the response is exponentially suppressed, except at resonances, Ω = nΩz + l κ + mΩϕ = 0. In this limit, far from the resonances, the perturbation timescale, b/vP, is much larger than Ω−1, and the net response is washed away due to many oscillations during the perturbation (i.e., the actions are adiabatically invariant), a phenomenon known as adiabatic shielding (Weinberg 1994a, 1994b; Gnedin & Ostriker 1999).

4.2. Response of the Milky Way Disk to Satellites

The MW halo harbors several fairly massive satellite galaxies that repeatedly perturb the MW disk. Here, we use existing data on the phase-space coordinates of those MW satellites to compute the disk response to satellite encounters that occurred in the past few hundred million years, which we may expect to have triggered phase spirals that have survived to the present day.

To compute the disk response to the MW satellites, we proceed as follows. As in Paper I, we adopt the galactocentric coordinates and velocities computed and documented by Riley et al. (2019) (Table A2; see also Li et al. 2020) and Vasiliev & Belokurov (2020) as initial conditions for the MW satellites. We then simulate their orbits in the combined gravitational potential of the MW halo, disk plus bulge, 5 using a second-order leapfrog integrator. For each individual orbit, we record the times, tcross, and the galactocentric radii, Rd, corresponding to disk crossings. We also register the corresponding impact velocities, ${v}_{{\rm{P}}}=\sqrt{{v}_{z}^{2}+{v}_{R}^{2}+{v}_{\phi }^{2}}$, and the angles of impact, ${\theta }_{{\rm{P}}}={\cos }^{-1}({v}_{z}/{v}_{{\rm{P}}})$ and ${\phi }_{{\rm{P}}}={\tan }^{-1}({v}_{\phi }/{v}_{R})$. We substitute these quantities in Equation (D7) and compute the disk response (integrated over IR ) following the satellite encounter, using Equations (21) and (D7). Results are summarized in Table 1 of Appendix D.1. Figure 6 plots the amplitude of the solar neighborhood (for which Rc(Lz ) = R = 8 kpc) bending mode response, f1,n=1/f0 (top panel), and breathing-to-bending ratio, f1,n=2/f1,n=1 (bottom panel), as a function of tcross. Here, we only show the responses for (l, m) = (0, 0) modes and consider stars with Iz = Iz,⊙ = 9.2 kpc km s−1.

Figure 6.

Figure 6. Steady-state MW disk response to satellite encounter in the collisionless limit: bending mode strength, f1,n=1/f0 (upper panel), and the corresponding breathing vs. bending ratio, f1,n=2/f1,n=1 (lower panel) for the (l, m) = (0, 0) modes, in the solar neighborhood for the MW satellites, as a function of the disk-crossing time, tcross, in Gyr, where tcross = 0 marks today. The previous two and the next impacts are shown. Here, we consider Iz = hz σz,⊙, with fiducial MW parameters, and marginalize over IR . The effect of the (nonresponsive) ambient DM halo on the stellar frequencies is taken into account. The estimates of tcross are very sensitive to the detailed potential of the MW system, while the response estimates are fairly robust (see text for details). In the upper panel, the region with bending mode response, f1,n=1/f0 < 10−4, has been grayscaled, indicating that the response from the satellites in this region is far too weak and adiabatic to be detected by Gaia. We note that the response is dominated by that due to Sgr, followed by Hercules, Leo II, Segue 2, and the Large Magellanic Cloud (LMC). We also note that the previous two and next impacts of all the satellites excite bending modes in the solar neighborhood.

Standard image High-resolution image

Table 1. Steady-state Response of the MW Disk to Encounters with Satellites in the Collisionless Limit, for the (n, l, m) = (1, 0, 0) Mode and for Stars with Iz = Iz,⊙ = hz σz,⊙ in the Solar Neighborhood

MW SatelliteMass f1,n=1/f0 tcross f1,n=1/f0 tcross f1,n=1/f0 tcross
Name(M) (Gyr) (Gyr) (Gyr)
  PenultimatePenultimateLastLastNextNext
(1)(2)(3)(4)(5)(6)(7)(8)
Sagittarius109 2.7 × 10−1 −1.014.9 × 10−8 −0.351.3 × 10−1 0.03
Hercules7.1 × 106 8.4 × 10−8 −3.782.4 × 10−3 −0.52.4 × 10−3 3.18
Leo II8.2 × 106 −3.861.6 × 10−3 −1.783.2 × 10−3 2.31
Segue 25.5 × 105 6.2 × 10−4 −0.848.5 × 10−4 −0.256.3 × 10−5 0.28
LMC1.4 × 1011 5.1 × 10−2 −7.63−2.672.3 × 10−2 0.11
SMC6.5 × 109 2.8 × 10−5 −3.32−1.448.7 × 10−6 0.22
Draco I2.2 × 107 −2.469.9 × 10−5 −1.247.1 × 10−6 0.24
Bootes I107 1.8 × 10−7 −1.673.7 × 10−5 −0.350.88
Willman I4 × 105 1.6 × 10−8 −0.661.2 × 10−6 −0.219.3 × 10−6 0.41
Ursa Minor2 × 107 −2.281.7 × 10−5 −1.172.6 × 10−6 0.29
Ursa Major II4.9 × 106 5.8 × 10−6 −2.122.5 × 10−6 −0.090.97
Coma Berenices I1.2 × 106 9.2 × 10−7 −2.583.7 × 10−8 −0.250.71
Sculptor3.1 × 107 −2.743.4 × 10−8 −0.461.48

Notes. We have marginalized the response over IR . Columns (1) and (2) list the name and dynamical mass of each satellite. The latter is taken from the literature (Simon & Geha 2007; Bekki & Stanimirović, 2009; Łokas 2009; Erkal et al. 2019; Vasiliev & Belokurov 2020), except for Sagittarius, for which we adopt a mass of 109 M. There is a discrepancy between its estimated mass of ∼4 × 108 M (Vasiliev & Belokurov 2020) and the mass required (109–1010 M) to produce detectable phase spiral signatures in N-body simulations (see, for example, Bennett et al. 2022). Columns (3) and (4), respectively, denote the penultimate bending mode response assuming our fiducial MW parameters and the penultimate disk-crossing time. Columns (5) and (6) indicate the same for the last disk crossing, while columns (7) and (8) show it for the next one. Only satellites that induce a bending mode response, f1,n=1/f0 ≥ 10−8, in at least one of the three cases are shown. Any response weaker than 10−8 is considered negligible and is indicated with a horizontal dash.

Download table as:  ASCIITypeset image

It is noteworthy that the responses in the realistic MW disk computed here are ∼1–2 orders of magnitude larger than those evaluated for the isothermal slab model shown in Figure 7 of Paper I. This is owed to the reduced damping of the phase spiral amplitude due to lateral mixing, which is more pronounced in the isothermal slab with unconstrained lateral velocities than in the realistic disk with constrained, ordered motion. From the lower panel of Figure 6, it is evident that, as in the isothermal slab case, almost all satellites trigger a bending mode response in the solar neighborhood, resulting in a one-armed phase spiral in qualitative agreement with the Gaia snail. However, as is evident from the upper panel, only five of the satellites trigger a detectable response in the disk, with ${f}_{1,n=1}/{f}_{0}\gt {\delta }_{\min }\equiv {10}^{-4}$ (see Appendix C of Paper I for a derivation of this approximate detectability criterion for Gaia). The response to encounters with the other satellites is weak, either because they have insufficient mass or because the encounter with respect to the Sun is too slow and adiabatically suppressed. Sgr excites the strongest response by far; its bending mode response, f1,n=1/f0, is at least 1–2 orders of magnitude above that for any other satellite. Its penultimate disk crossing, about the same time as its last pericentric passage ∼1 Gyr ago, triggered a strong response of f1,n=1/f0 ∼ 0.3 in the solar neighborhood. For comparison, the response from its last disk crossing, which nearly coincides with its last apocentric passage about 350 Myr ago, triggered a very weak, adiabatically suppressed response (∼5 × 10−8) that falls below the lower limit of Figure 6. Its next disk crossing in about 30 Myr is estimated to trigger a strong response with f1,n=1/f0 ∼ 0.1. Besides Sgr, the satellites that excite a detectable response, ${f}_{1,n=1}/{f}_{0}\gt {\delta }_{\min }$ are Hercules, Segue 2, Leo II, and the LMC. The imminent crossing of LMC is estimated to trigger f1,n=1/f0 ∼ 2 × 10−2, which is an order of magnitude below Sgr. Only for Iz /Iz,⊙ ≳ 4.5 (${z}_{\max }\gtrsim 3.4{h}_{z}$) does the LMC response dominate over Sgr. This exercise therefore suggests that Sgr is the leading contender, among the MW satellites considered here, for triggering the Gaia snail in the solar neighborhood, in agreement with several previous studies (Antoja et al. 2018; Binney & Schönrich 2018; Laporte et al. 2018, 2019; Bland-Hawthorn et al. 2019; Darling & Widrow 2019a; Bland-Hawthorn & Tepper-García 2021; Hunt et al. 2021).

We caution that the above estimates of the disk response are obtained under the assumption of nearly straight-line orbits of the satellites. This assumption is well-justified as long as the satellite orbit is sufficiently eccentric. However, realistic orbits with low eccentricities can trigger a substantially different disk response. Also, there is a large uncertainty in Sgr's orbit (as well as that of the other satellites), due to the uncertainty in its observed phase-space coordinates and the MW parameters. These can explain the apparent discrepancy between our results and those of Bennett et al. (2022): they find in their N-body simulations that Sgr triggers a weaker response in the Solar neighborhood than what we compute in this paper. This is probably attributable to the fact that Sgr's orbit induces an adiabatic perturbation in the solar neighborhood in their simulations. To integrate the satellite orbits for our analysis, we pick the median values of the observed phase-space coordinates quoted by Riley et al. (2019) and Vasiliev & Belokurov (2020) as the initial conditions. Other values within the margin of error may lead to substantially different orbits, implying different Rd, vP, θP and ϕP, and therefore different responses. Moreover, some of these orbits can be mildly eccentric, for which the straight-line orbit approximation adopted in this paper is not well-justified. A perturbative analysis of the disk response to satellites along general orbits is beyond the scope of this paper and deserves future investigation. We also ignore the effect of dynamical friction on the satellite orbits and thus on the resulting disk response. Moreover, the disk-crossing times are sensitive to the satellite orbits and therefore to the detailed MW potential and the current phase-space coordinates of the satellites. For example, a heavier MW model with a total mass of 1.5 × 1012 M leaves the relative amplitudes of the satellite responses (in the collisionless limit) nearly unchanged, but it makes the satellites more bound, bringing most of the disk-crossing times closer to the present day. In particular, the last disk passage of Sgr that triggers a significant response now occurs ∼600 Myr ago (as opposed to 1 Gyr ago in the fiducial case), which is closer to the winding time of ∼500 Myr inferred from the phase spiral observed in the solar neighborhood (Bland-Hawthorn et al. 2019).

In this section, we have computed the responses in the collisionless limit. In reality, collisional diffusion due to interactions of stars with GMCs, etc. would damp away the response superexponentially over a timescale that is ∼0.6–0.7 Gyr in the solar neighborhood (see Section 3.1.2). This would almost completely wash away the response to any satellite encounter that occurred ≳1 Gyr ago. For example, the present-day response to the last pericentric passage of Leo II that occurred ∼1.8 Gyr ago would be completely erased. If the last disk crossing of Sgr that induced a strong response occurred ∼1 Gyr ago, as in the fiducial MW model, the response would have been damped out by ∼2 orders of magnitude by today, meaning Sgr is unlikely to be the agent behind the Gaia snail. However, as discussed above, the disk-crossing times are sensitive to the satellite orbits. The heavier MW model with a total mass of 1.5 × 1012 M implies a Sgr crossing time of ∼0.6 Gyr instead of 1 Gyr. In this case, the response would only have been damped by a factor of ∼0.4. Therefore, the collisionality argument suggests that, if the Gaia snail was indeed triggered by Sgr, the impact causing it must have happened within ∼0.6–0.7 Gyr from the present day.

4.3. Exploring Parameter Space

Having computed the MW disk response to its satellites, we now investigate the sensitivity of the response to the various encounter parameters. In Figure 7, we plot the amplitude of the solar neighborhood response, f1,nlm /f0 (marginalized over IR ), as a function of the impact velocity, vP (in units of the circular velocity at Rc = R), for the (n, l, m) = (1, 0, 0) bending and (n, l, m) = (2, 0, 0) breathing modes, shown in the left and right columns, respectively. The top, middle, and bottom rows show the results for varying Iz , θP, and ϕP, respectively, assuming the fiducial parameters to be those for Sgr (mass MP = 109 M, scale radius ε = 1.6 kpc) during its penultimate disk crossing (most relevant for the Gaia snail), i.e., impact radius Rd = 17 kpc, impact velocity vP = 340 km s−1, and angles of impact, θP = 21° and ϕP = 150°. In Figure 8, we plot the bending and breathing mode response amplitudes (in the solar neighborhood) as a function of vP for different (l, m) modes, with the fiducial parameters again corresponding to Sgr. The left and right columns, respectively, indicate the n = 1 bending and n = 2 breathing modes, while the top and bottom rows correspond to l = 1 and l = 2, respectively. The different lines in each panel denote the responses for m = −2, −1, 0, 1, and 2. Figure 9 shows the ratio of the bending and breathing response amplitudes as a function of vP for the dominant mode (l, m) = (0, − 2). Different lines indicate breathing-to-bending ratios for different values of θP, while the left and right columns, respectively, indicate the ratios observed at Rc = 8 and 12 kpc.

Figure 7.

Figure 7. Steady-state MW disk response to a satellite encounter in the collisionless limit: each panel shows the behavior of the disk response amplitude, f1,n00/f0 (evaluated using Equations (21) and (D7)) and marginalized over IR ), as a function of the impact velocity, vP, in the solar neighborhood, i.e., Rc = R = 8 kpc, in presence of an ambient DM halo. The left and right columns, respectively, indicate the response for the n = 1 bending and n = 2 breathing modes. The top, middle, and bottom rows, respectively, show the same for different values of Iz (in units of Iz,⊙), θP, and ϕP as indicated, with the fiducial parameters corresponding to Iz,⊙ and the parameters for Sgr impact, the response amplitude for which is indicated by the red circle. We note that the response is suppressed as ${v}_{{\rm{P}}}^{-1}$ in the impulsive (large vP) limit but exponentially suppressed in the adiabatic (small vP) regime, and it peaks at an intermediate velocity, vP ∼ 2−3 vcirc(R) (which is very similar to the encounter speed of Sgr). The peak of the response shifts to smaller vP for larger Iz , because Ωz decreases with Iz . The response depends only very weakly on ϕP but is quite sensitive to θP; an increase in planar encounters, i.e., an increase in θP, triggers stronger responses.

Standard image High-resolution image
Figure 8.

Figure 8. Steady-state MW disk response to satellite encounter in the collisionless limit: each panel shows the behavior of the disk response amplitude, f1,nlm /f0 (marginalized over IR ), as a function of the impact velocity, vP, in the solar neighborhood, in presence of an ambient DM halo. Different lines correspond to different m modes as indicated. The top and bottom rows show the response for l = 0 and 1 while the left and right columns indicate it for the n = 1 bending and n = 2 breathing modes. The fiducial parameters correspond to Iz = Iz,⊙ and the parameters for Sgr impact, the response amplitudes for which are indicated by the red circles in each panel. The response is dominated by the (n, l, m) = (1, 0, −2) mode or the two-armed warp at small vP and the (2, 0, −2) mode or the two-armed spiral at large vP. Typically, the m = −2 and −1 responses dominate over m = 0, 1, and 2, while the l = 0 response is more pronounced than l = 1.

Standard image High-resolution image
Figure 9.

Figure 9. MW disk response to satellite encounter: breathing-to-bending ratio or the relative strength of the n = 2 and n = 1 modes of disk response to a Sgr-like impact is plotted as a function of the impact velocity, vP, at Rc = R = 8 kpc and Rc = 1.5R = 12 kpc, shown in the left and right columns, respectively, for the (l, m) = (0, − 2) mode, which typically dominates the response. Different lines correspond to different values of θP as indicated. We consider Iz = Iz,⊙ and the fiducial parameters to correspond to those for Sgr encounter, for which the breathing-to-bending ratio is denoted by the red circle. Bending modes dominate over breathing modes at small vP and vice versa at large vP. Breathing modes are relatively more pronounced than bending modes in the outer disk, closer to the Sgr impact radius, Rd = 17 kpc. More planar (perpendicular) encounters trigger larger breathing-to-bending ratios farther away from (closer to) the impact radius.

Standard image High-resolution image

From Figures 7 and 8, it is evident that, as shown in Equation (41), the disk response is suppressed like a power law ($\sim {v}_{{\rm{P}}}^{-1}$) in the high-velocity/impulsive limit and exponentially ($\sim \exp \left[-{\rm{\Omega }}b/{v}_{{\rm{P}}}\right]$) suppressed in the low-velocity/adiabatic limit. The response is the strongest for intermediate velocities, vP ∼ 2−3 vcirc(R), where the frequencies of the vertical, radial, and azimuthal oscillations of the stars are nearly commensurate with the encounter frequency, ${v}_{{\rm{P}}}/\sqrt{{b}^{2}+{\varepsilon }^{2}}$. The ${v}_{{\rm{P}}}^{-1}$ and K0i factors in Equation (39) conspire to provide the near-resonance condition for maximum response,

Equation (43)

where b is the impact parameter of the encounter, given by Equation (42). From the top panels of Figure 7, it is clear that the peak response shifts to smaller vP with increasing Iz . This is easy to understand from the fact that the corresponding vertical frequency, Ωz , decreases with increasing Iz , making the encounter more impulsive for larger actions. The middle and bottom panels show that the response depends strongly on the polar angle of the encounter, θP, but very mildly on the azimuthal angle, ϕP. Moreover, the middle panels indicate that more planar encounters (larger θP) induce stronger responses. This is a consequence of the fact that satellites spend more time near the disk during more planar encounters.

The in-plane structure of the disk response depends on the relative contribution of the different (l, m) modes. From Figure 8, it is evident that a typical Sgr-like encounter predominantly excites (l, m) = (0, −1) and (l, m) = (0, −2) in the solar neighborhood. The dominant mode for slower encounters is (n, l, m) = (1, 0, −2), while that for faster ones is (n, l, m) = (2, 0, − 2). Because, for a Sgr-like encounter, f1,nlm /f0 ≳ 1 for these dominant modes, the response to the impact by Sgr is in fact nonlinear in the solar neighborhood. Either way, a satellite encounter is typically found to excite strong m = −2 modes, i.e., two-armed warps (n = 1) and spirals (n = 2). This is due to a quadrupolar tidal distortion of the disk by the satellite, which manifests as a stretching of the disk in the direction of the impact and a compression perpendicular to it.

Figure 9 elucidates that the bending mode response dominates for slower encounters, i.e., smaller vP, and at guiding radii far from the impact radius, Rd. More planar impacts trigger larger breathing-to-bending ratios farther away from the impact radius, while this trend reverses closer to it. This is because more planar encounters cause more vertically symmetric perturbations farther away from the impact radius. The predominance of bending modes for low-vP encounters versus that of breathing modes for high-vP ones has been observed by Widrow et al. (2014) and Hunt et al. (2021) in their N-body simulations of satellite–disk encounters. As demonstrated by Widrow et al. (2014), slower encounters provide energy to the stars near one of the vertical turning points while draining energy from those near the other turning point, thereby driving bending wave perturbations that are asymmetric about the midplane. On the other hand, fast satellite passages are impulsive and impart energy to the stars near both the turning points, thus triggering symmetric breathing waves.

The predominance of breathing (bending) modes closer to (farther away from) the impact radius is qualitatively similar to the observation by Hunt et al. (2021) in their simulations of MW–Sgr encounter that the outer part of the MW disk that is closer to the impact radius shows a preponderance of two-armed phase spirals or breathing modes. This can be understood within the framework of our formalism by noting that the impact parameter, b, and therefore the encounter timescale, $\sim \sqrt{{b}^{2}+{\varepsilon }^{2}}/{v}_{{\rm{P}}}$, decreases with increasing proximity to the point of impact; hence, the impact is faster than the vertical oscillations of stars near the point of impact, driving stronger breathing mode perturbations. However, contrary to these predictions for the MW–Sgr encounter, Hunt et al. (2022), using Gaia DR3 data, revealed two-armed phase spirals, and therefore breathing modes, in the inner disk (Rc ∼ 6–7 kpc). Our analysis suggests that none of the MW satellites could have caused this. Using N-body simulations of an isolated MW system, Hunt et al. (2022) suggested that a transient spiral arm or bar could be a potential trigger for breathing modes in the inner disk. However, such a transient perturbation would have to be sufficiently impulsive, i.e., occur over a timescale that is comparable to or smaller than the vertical oscillation timescale in the inner disk (see Section 3.1.1), in order to produce two-armed phase spirals with density contrast as strong as in the data. Such short timescales are unlikely to arise from the secular evolution of the disk alone, and they may instead require forcing of the inner disk by perturbations in the MW halo. Another possible trigger of this feature is the recent passage of one or more dark satellites through the inner disk. The true origin of this feature is, however, unclear. Hence, we conclude that the presence of two-armed phase spirals in the inner disk is rather unexpected, and that its origin poses an intriguing conundrum.

5. Conclusion

In this paper, we have developed a linear perturbative formalism to analyze the response of a realistic disk galaxy (characterized by a pseudo-isothermal DF) embedded in an ambient spherical DM halo (modeled by an NFW profile) to perturbations of diverse spatiotemporal nature: bars, spiral arms, and encounters with satellite galaxies. Adopting the radial epicyclic approximation, we perturb the FPE up to linear order (in action-angle space) in the presence of a perturbing potential, ΦP, to compute the post-perturbation linear response in the DF, f1. Without self-gravity to reinforce the response, the oscillations in the response phase mix away due to an intrinsic spread in the frequencies of stars, giving rise to spiral features in the phase-space distribution known as phase spirals. Depending on the timescale of ΦP, different modes of disk oscillation, corresponding to different phase spiral structures, are excited. We summarize our conclusions as follows:

  • 1.  
    Following an impulsive perturbation, the (n, l, m) mode of the disk response consists of stars oscillating with frequencies, nΩz , lΩR l κ, and mΩϕ , along vertical, radial, and azimuthal directions, respectively. Because the frequencies depend on the actions, primarily on the vertical action Iz and the angular momentum Lz , the response phase mixes away, spawning phase spirals. The dominant modes of vertical oscillation are the antisymmetric bending (n = 1) and symmetric breathing (n = 2) modes, which induce initial dipolar and quadrupolar perturbations in the zvz or ${I}_{z}\cos {w}_{z}-{I}_{z}\sin {w}_{z}$ phase space. Over time, these features are phase wrapped into one- and two-armed phase spirals, respectively, due to the variation of Ωz with Iz .
  • 2.  
    Because Ωz and Ωϕ both depend on Lz , the amplitude of the ${I}_{z}\cos {w}_{z}-{I}_{z}\sin {w}_{z}$ phase spiral damps away over time, typically as ∼1/t (Equation (29)), on a coarse-grained level, i.e., upon marginalization over Lz . Therefore, in a realistic disk with ordered motion, lateral mixing causes phase spirals to damp out much more slowly than in the isothermal slab with unconstrained lateral velocities discussed in Paper I, where it occurs like a Gaussian in time.
  • 3.  
    Collisional diffusion due to scatterings of stars by GMCs, DM substructure, etc., damps away the disk response to a perturbation, and therefore the phase spiral amplitude, on a fine-grained level. Typically, the diffusion in actions is much more efficient than that in angles. The action gradients of the response, which predominantly arise from the action dependence of the oscillation frequencies, are erased by collisional diffusion, causing a superexponential damping of the response over a timescale, ${\tau }_{{\rm{D}}}^{({\rm{I}})}$, which is ∼0.6–0.7 Gyr in the solar neighborhood. The diffusion timescale is shorter in the inner disk, for stars with smaller Iz , and for higher-n modes.
  • 4.  
    The response to a bar or spiral arm with a fixed pattern speed, ΩP, is dominated by the near-resonant stars (${{\rm{\Omega }}}_{\mathrm{res}}=n{{\rm{\Omega }}}_{z}+l\kappa +m({{\rm{\Omega }}}_{\phi }-{{\rm{\Omega }}}_{{\rm{P}}})\approx 0$), especially in the adiabatic regime (slowly evolving perturber amplitude). Moreover, phase mixing occurs gradually in the near-resonant parts of phase space. Most of the strong resonances are confined to the disk plane, such as the corotation (n = l = 0) and Lindblad (n = 0, l = ± 1, m = ± 2) resonances. For a transient bar or spiral arm whose amplitude varies over time as $\sim \exp \left[-{\omega }_{0}^{2}{t}^{2}\right]$, the response is maximal when ${\omega }_{0}\sim {{\rm{\Omega }}}_{\mathrm{res}}$. In the impulsive limit (${\omega }_{0}\gg {{\rm{\Omega }}}_{\mathrm{res}}$), the response is power-law suppressed, while in the adiabatic limit (${\omega }_{0}\ll {{\rm{\Omega }}}_{\mathrm{res}}$), it is suppressed (super)exponentially.
  • 5.  
    For a thin disk, because Ωz is very different from Ωϕ and κ, the vertical modes (n ≠ 0) are generally not resonant with the radial and azimuthal ones and thus undergo phase mixing. The strength of a vertical mode primarily depends on the nature of the perturbing potential, most importantly its timescale. Slower pulses trigger mainly bending (n = 1) modes, while faster pulses excite more pronounced breathing (n = 2) modes. Therefore, a transient bar or spiral arm with amplitude $\sim \exp \left[-{\omega }_{0}^{2}{t}^{2}\right]$ triggers a bending (breathing) mode when the pulse frequency, ω0, is smaller (larger) than Ωz . The response to very slow perturbations (ω0 ≪ Ωz ) is, however, heavily suppressed (adiabatic shielding).
  • 6.  
    For a persistent bar or spiral arm with a fixed pattern speed, ΩP, that grows and saturates over time, the response initially develops a phase spiral. However, this transient response is quickly taken over by coherent oscillations at the driving frequency, ΩP, which manifest in phase space as a steadily rotating dipole (quadrupole) for the bending (breathing) mode. Therefore, a transient (pulse-like) perturbation, such as a bar or spiral arm whose amplitude varies over a timescale comparable to the vertical oscillation period, Tz hz /σz , is essential for the formation of a phase spiral in zvz space.
  • 7.  
    The above analysis suggests that, if the recently discovered two-armed Gaia phase spiral (breathing mode) in the inner disk of the MW was indeed induced by a spiral arm/bar as suggested by Hunt et al. (2022) using N-body simulations, the spiral arm/bar was probably a transient one with a predominantly symmetric vertical profile whose amplitude varied over a timescale comparable to the vertical oscillation period. However, it remains to be seen whether such a rapid excitation and decay of a spiral arm/bar perturbation is realistic.
  • 8.  
    We have computed the response of the MW disk, embedded in an extended DM halo, to disk-crossing perturbations by several of its satellite galaxies. We find that the response in the solar neighborhood is dominated by the perturbations due to Sgr, followed by those due to the LMC, Hercules, and Leo II. This implies that, if the Gaia snail near the solar radius was indeed triggered by a MW satellite (which is still subject to debate), Sgr is the leading contender (see also Paper I). However, if that is the case, then the impact (disk crossing) must have happened within the last ∼0.6–0.7 Gyr in order for the response to have survived damping due to collisional diffusion.
  • 9.  
    The amplitude of the response (at a fixed guiding radius Rc) to satellite encounters scales as ${v}_{{\rm{P}}}^{-1}$ in the impulsive (large vP) limit, but is exponentially suppressed in the adiabatic (small vP) limit, a phenomenon known as adiabatic shielding. The resonant modes with nΩz + l κ + mΩϕ = 0 are not suppressed but rather become nonlinear in the adiabatic regime. The peak response of a mode (with nΩz + l κ + mΩϕ ≠ 0) is achieved at intermediate velocities for which the encounter frequency is commensurate with the oscillation frequencies of the stars, i.e., the near-resonance condition given by Equation (43) is satisfied.
  • 10.  
    The response of a disk to an encounter with a satellite galaxy depends primarily on three parameters: (i) impact velocity vP, (ii) polar angle of impact θP, and (iii) position on the disk relative to the point of impact where the satellite crossed the disk. Slower (faster) encounters excite predominantly n = 1 bending (n = 2 breathing) modes. More planar encounters (those with larger θP) typically result in larger breathing-to-bending ratios farther away from the impact radius, while this trend gets reversed closer to it. In general, breathing modes dominate over bending modes closer to the point of impact, in agreement with N-body simulations of the MW–Sgr encounter (Hunt et al. 2021). Because the impact velocities of the MW satellites are all fairly similar to the local circular velocity, the decisive factor for breathing versus bending modes is not so much the impact velocity but rather the distance from the point of impact.
  • 11.  
    The (n, m) = (1, −2) and (−1, 2) modes generally dominate the response for slower satellite encounters, e.g., that of Sgr with respect to the solar neighborhood, due to the tidal distortion of the disk by the satellite. The in-plane spatial structure of the disk response therefore generally resembles a two-armed warp (n = 1) or spiral (n = 2).
  • 12.  
    The presence of an extended DM halo causes phase mixing to occur more slowly and thus makes the phase spiral less wound. Hence, provided that the phase spiral was triggered by a single, impulsive perturbation, the detailed shape of the phase spiral can in principle be used to constrain not only the time elapsed since the perturbation (Darragh-Ford et al. 2023) but also the total (disk+halo) potential. If the phase spiral has been triggered and/or impacted by multiple, overlapping perturbations, the situation is less clear. In future work, we intend to investigate the constraining power of phase spirals for different kinds of perturbation.

This paper focused on the analysis of the phase-mixing component (phase spirals) of the "direct" disk response to various perturbations such as bars, spiral arms, and satellite galaxies. However, this leaves out some other potentially important features of the disk response. First of all, we considered the ambient DM halo to be nonresponsive. In reality, the DM halo will also be perturbed, for example by an impacting satellite, and this halo response, which can be enhanced by self-gravity, can indirectly perturb the disk. A preliminary, perturbative analysis based on the N-body simulation of the MW–Sgr encounter by Hunt et al. (2021) suggests that the indirect disk response to halo perturbations (triggered by Sgr) is comparable, but subdominant, to the direct response to Sgr. However, a more detailed analysis is warranted, which we leave for future work. Second, we have neglected the self-gravity of the disk response. As discussed in Paper I, the dominant effect of self-gravity is to cause coherent point mode oscillations (Mathur 1990; Weinberg 1991) of the disk, which in linear theory are decoupled from the phase-mixing component of the response. The point modes manifest as coherently rotating features in the phase space, which can make the phase spiral look less wound than in the non-self-gravitating case (Darling & Widrow 2019b; Widrow 2023), before the point modes Landau damp away. Usually, there is one dominant point mode. In linear theory, one can always transform to the momentarily comoving reference frame of the dominant point mode, or in other words subtract away the point mode response, to obtain the pure phase-mixing contribution to the phase spiral. Besides triggering coherent point mode oscillations that make the phase spiral less wound, self-gravity can also enhance the amplitude of the phase-mixing component of the response, thus increasing the density contrast of the phase spiral. Recent work by Dootson & Magorrian (2022) has shed some light on the self-gravitating response of razor-thin disks to bar perturbations, while that by Widrow (2023) has investigated the impact of self-gravity on vertical phase spirals in a shearing box for impulsive perturbations. However, a more generic theoretical description of the self-gravitating response of inhomogeneous, thick disks to general perturbations (bars, spiral arms, satellite galaxies, etc.) is still lacking. We hope to include the effects of self-gravity on disk perturbations in future work.

Acknowledgments

The authors are grateful to Kathryn Johnston, Jason Hunt, Adrian Price-Whelan, Scott Tremaine, Chris Hamilton, Elise Darragh-Ford, James Binney, John Magorrian, and Elena D'Onghia for insightful discussions and valuable suggestions. F.C.v.d.B. is supported by the National Aeronautics and Space Administration through grant No. 19-ATP19-0059 issued as part of the Astrophysics Theory Program. M.D.W. is supported by the National Science Foundation through grant No. AST-1812689.

Appendix A: Perturbative Solution of the Fokker–Planck Equation

To solve the FPE in the action-angle space, the collision operator, C[f], given in Equation (2) first needs to be expressed in terms of the action-angle variables. For this purpose, we follow Tremaine et al. (2023) and invoke the epicyclic approximation in both vertical and radial directions. This is justified by the fact that stars with smaller Iz and IR are more strongly scattered by structures like giant molecular clouds, because these are concentrated toward the midplane of the MW disk and are mostly corotating with the disk stars. This implies that $z\approx \sqrt{2{I}_{z}/\nu }\sin {w}_{z}$, ${p}_{z}\approx \sqrt{2\nu {I}_{z}}\cos {w}_{z}$, $R\approx {R}_{c}({L}_{z})+\sqrt{2{I}_{R}/\kappa }\sin {w}_{z}$, and ${p}_{R}\approx \sqrt{2\kappa {I}_{R}}\cos {w}_{R}$, where ν and κ are, respectively, the vertical and radial epicyclic frequencies. This also implies that ${D}_{{zz}}={D}_{I}^{(z)}/\nu $, ${D}_{{p}_{z}{p}_{z}}=\nu {D}_{I}^{(z)}$, ${D}_{{RR}}={D}_{I}^{(R)}/\kappa $, and ${D}_{{p}_{R}{p}_{R}}=\kappa {D}_{I}^{(R)}$, with ${D}_{I}^{(z)}$ and ${D}_{I}^{(R)}$ being the first-order diffusion coefficients in Iz and IR , respectively. Substituting these in Equation (2), the expression for

Equation (A1)

simplifies to

Equation (A2)

We have made several approximations here. First, we have adopted the epicyclic approximation. Second, we have neglected the zpz and RpR cross terms, for simplicity. Third, following Binney & Lacey (1988), we have neglected diffusion in ϕ and pϕ , because the terms involving Dϕ ϕ , Dr ϕ , and Dϕ z are smaller than the Iz and IR diffusion terms by factors of at least σR /vc or σz /vc , which are typically much smaller than unity (σR and σz are radial and vertical velocity dispersions, respectively, and vc is the circular velocity along ϕ). These approximations together imply the form for the collision operator given in Equation (A2). A close inspection of this tells us that the diffusion coefficients in the action-angle variables satisfy the relations: ${D}_{{II}}^{(z)}=2{D}_{I}^{(z)}{I}_{z}$, ${D}_{{ww}}^{(z)}={D}_{I}^{(z)}/(2{I}_{z})$, ${D}_{{II}}^{(R)}=2{D}_{I}^{(R)}{I}_{R}$, and ${D}_{{ww}}^{(R)}={D}_{I}^{(R)}/(2{I}_{R})$. These diffusion coefficients approximately preserve the form of the unperturbed DF (Equation (15)) of the MW disk (Binney & Lacey 1988).

From the age–velocity dispersion relation of the MW disk stars, we deem ${D}_{I}^{(z)}$ and ${D}_{I}^{(R)}$ as constants and approximate them as ${D}_{I}^{(z)}=\left\langle {I}_{z}\right\rangle /{T}_{\mathrm{disk}}$ and ${D}_{I}^{(R)}=\left\langle {I}_{R}\right\rangle /{T}_{\mathrm{disk}}$, with Tdisk being the age of the disk (Tremaine et al. 2023). Here, $\left\langle {I}_{a}\right\rangle =\int {{dI}}_{a}\,{I}_{a}\,{f}_{0}\,/\int {{dI}}_{a}\,{f}_{0}$, where a is either z or R. This yields

Equation (A3)

Substituting this form of the collision operator in Equation (7) and neglecting the IR diffusion term—because the response does not develop strong IR gradients, due to the mild dependence of the stellar frequencies on IR —we obtain the evolution equation for the response, f1, given in Equation (9). Using the Fourier series expansions of f1 and ΦP given in Equations (10) in Equation (9), the evolution equation for the Fourier mode of f1 (Equation (11)) can be obtained and rearranged to yield the following inhomogeneous differential equation for f1nlm :

Equation (A4)

If ${D}_{I}^{(z)}$ is much smaller than ${\sigma }_{z}^{2}$, which is typically the case, then one can assume the Iz diffusion term to be small, i.e., the effect of diffusion to be localized in Iz . More specifically, one can look for a solution in the neighborhood of Iz = Iz0, i.e., for Iz = Iz0 + ΔIz , where ΔIz Iz0 (Tremaine et al. 2023). Thus, Equation (A4) can be rewritten as

Equation (A5)

The above inhomogeneous differential equation can be solved using the Green's function technique. With the initial condition, f1nlm (t = ti ) = 0, and in the limit of ΔIz → 0, one can obtain the following solution for f1nlm :

Equation (A6)

where ${{ \mathcal G }}_{{nlm}}({\boldsymbol{I}},t-\tau )$ is the Green's function, i.e., the solution of the homogeneous equation,

Equation (A7)

The Green's function can be computed as a function of ΔIz , and later evaluated in the ΔIz → 0. The computation proceeds as follows. First, we expand ${{ \mathcal G }}_{{nlm}}$ as a continuous Fourier series:

Equation (A8)

This reduces Equation (A7) to the following differential equation for gnlm :

Equation (A9)

where Ωz1 = ∂Ωz /∂Iz evaluated at Iz = Iz0. The above equation can be written more concisely as

Equation (A10)

where

Equation (A11)

The differential Equation (A10) is known as the Airy equation, which has the general solution

Equation (A12)

with c1 and c2 being constants, and Ai and Bi being the Airy functions. The long time behavior of ${{ \mathcal G }}_{{nlm}}$ is captured by the Ai part of gnlm . Ai (z) is given by the following integral form:

Equation (A13)

Upon substituting x and k from Equations (A11) in ${g}_{{nlm}}\approx {A}_{i}\left({(-k)}^{1/3}x\right)$, and performing the inverse Fourier transform of gnlm , we obtain

Equation (A14)

where

Equation (A15)

Integrating over s, we obtain the following form for ${{ \mathcal G }}_{{nlm}}$:

Equation (A16)

which, in the limit ΔIz → 0, reduces to

Equation (A17)

Appendix B: The Unperturbed Galaxy

Under the radial epicyclic approximation (small IR ), the unperturbed DF, f0, for a rotating MW-like disk galaxy can be well approximated as a pseudo-isothermal DF, i.e., written as a nearly isothermal separable function of the azimuthal, radial and vertical actions. Following Binney (2010), we write

Equation (B1)

The vertical structure of this disk is isothermal, while the radial profile is pseudo-isothermal. Here, Σ = Σ(R) is the surface density of the disk, Lz is the z-component of the angular momentum, which is equal to Iϕ , Rc = Rc(Lz ) is the guiding radius, Ωϕ is the circular frequency, and $\kappa =\kappa ({R}_{{\rm{c}}})={\mathrm{lim}}_{{I}_{R}\to 0}{{\rm{\Omega }}}_{R}$ is the radial epicyclic frequency (Binney & Tremaine 1987). If L0 is sufficiently small, then we can further approximate the above form for f0 as

Equation (B2)

where Θ(x) is the Heaviside step function. Thus, we assume that the entire galaxy is composed of prograde stars with Lz > 0.

The corresponding density profile can be written as a product of an exponential radial profile and an isothermal (sech2) vertical profile, i.e.,

Equation (B3)

where hR is the radial scale height and and hz is the vertical thickness. Throughout, we adopt the thin disk limit, i.e., hz hR . The surface density profile is given by

Equation (B4)

where Σc = ρc hz is the central surface density of the disk. We assume a radially varying vertical velocity dispersion, σz , satisfying ${\sigma }_{z}^{2}(R)=2\pi {{Gh}}_{z}{\rm{\Sigma }}(R)$ (Binney & Tremaine 2008). We assume a similar profile for ${\sigma }_{R}^{2}$ such that the ratio, σR /σz , is constant throughout the disk (Binney 2010) and equal to the value at the solar vicinity.

Throughout this paper, for the ease of computation of the frequencies (because of a simple analytic form of the potential), we approximate the above density profile by a combination of three Miyamoto & Nagai (1975) disk profiles (Smith et al. 2015), i.e., the 3MN profile as implemented in the Gala Python package (Price-Whelan 2017; Price-Whelan et al. 2020). The corresponding disk potential is given by

Equation (B5)

where Mi , a, and bi , with i = 1, 2, and 3, are the mass, scale radius, and scale height corresponding to each of the MN profiles.

The MW disk is believed to be embedded in a much more extended DM halo, which we model using a spherical NFW (Navarro et al. 1997) profile with potential

Equation (B6)

Here, Mvir is the virial mass of the halo, rs is the scale radius, c = Rvir/rs is the concentration (Rvir is the virial radius), and $f(c)=\mathrm{ln}\left(1+c\right)-c/(1+c)$. The combined potential experienced by the disk stars is thus given by

Equation (B7)

Appendix C: Fourier Coefficients of Spiral Arm or Bar Perturbing Potential

An essential ingredient of the disk response to spiral arm or bar perturbations is the Fourier component of the perturber potential, Φnlm . This can be computed as follows:

Equation (C1)

To evaluate this, first we need to calculate r = (z, R, ϕ) as a function of $\left({\boldsymbol{w}},{\boldsymbol{I}}\right)=({w}_{z},{w}_{\phi },{w}_{R},{I}_{z},{I}_{\phi },{I}_{R})$ where Iϕ = Lz , the angular momentum. Under the epicyclic approximation, R can be expressed as a sum of the guiding radius and an oscillating epicyclic term, i.e.,

Equation (C2)

and the azimuthal angle, wϕ , is given by

Equation (C3)

The vertical distance z from the midplane is related to Rc(Lz ) and $\left({w}_{z},{I}_{z}\right)$, according to

Equation (C4)

where Ωz (Rc, Iz ) = 2π/Tz (Rc, Iz ), with Tz (Rc, Iz ) given by Equation (19). The above equation can be numerically inverted to obtain z(Rc, wz , Iz ).

Upon substituting the above expressions for R, ϕ, and z in terms of ( w , I ) in the expression for ΦP given in Equation (22), we obtain

Equation (C5)

where Jl is the lth-order Bessel function of the first kind,

Equation (C6)

and ${{\rm{\Phi }}}_{n}^{({\rm{o}})}({I}_{z})$ and ${{\rm{\Phi }}}_{n}^{({\rm{e}})}({I}_{z})$ are given by

Equation (C7)

In deriving Equation (C5), we have used the Hansen–Bessel formula, which provides the following integral representation for Bessel functions of the first kind,

Equation (C8)

and the identity for expansion in products of Bessel functions given in equation (8.530.2) of Gradshteyn & Ryzhik (1965; see also Section 6.1 of Binney & Lacey 1988). We have also used the identity

Equation (C9)

Appendix D: Perturbation by Encounter with Satellite Galaxy

D.1. Computation of the Disk Response

To evaluate the disk response to satellite encounters using Equation (21), we first evaluate the τ integral (with ti → − ) of the satellite potential given in Equation (38) and then compute the Fourier transform of the result. This yields the expression for the response in Equation (21) with

Equation (D1)

where

Equation (D2)

We perform the inner τ integral of ΦP to obtain

Equation (D3)

Here, K0i is defined as

Equation (D4)

which asymptotes to the zeroth order modified Bessel function of the second kind, ${K}_{0}\left(\left|\alpha \right|\right)$, in the limit β . ${ \mathcal R }$ and ${ \mathcal S }$ are, respectively, the projections perpendicular and parallel to the direction of v P of the vector connecting the point of observation, (z, R, ϕ), with the point of impact, and are given by

Equation (D5)

In deriving Equation (D3), we have only considered the direct term in the expression for ΦP given in Equation (38); the indirect term turns out to be subdominant.

In the large time limit, i.e., $t\gg { \mathcal S }/{v}_{{\rm{P}}}$, K0i asymptotes to ${K}_{0}\left(\left|{\rm{\Omega }}\right|\sqrt{{{ \mathcal R }}^{2}+{\varepsilon }^{2}}/{v}_{{\rm{P}}}\right)$. We substitute the expressions for R and z in terms of ( w , I ) given in Equations (C2) and (C4) in the above expressions for ${ \mathcal R }$ and ${ \mathcal S }$. Further substituting the resultant τ integral from Equation (D3) in Equation (D1), substituting wϕ in terms of ϕ using Equation (C3), adopting the small IR limit, and performing the wR integral, we obtain

Equation (D6)

where ${{ \mathcal R }}_{{\rm{c}}}={ \mathcal R }(R={R}_{{\rm{c}}})$ and ${{ \mathcal S }}_{{\rm{c}}}={ \mathcal S }(R={R}_{{\rm{c}}})$. Here, we have used the integral representation of Bessel functions of the first kind given in Equation (C8) and the identity given in Equation (8.530.2) of Gradshteyn & Ryzhik (1965).

The expression for ${{ \mathcal I }}_{{nlm}}$ given in Equation (D6) consists of the leading-order expansion in $\sqrt{2{I}_{R}/\kappa }$. A more precise expression that is accurate up to second order in $\sqrt{2{I}_{R}/\kappa }$ is given, in the large time limit, as

Equation (D7)

where

Equation (D8)

and

Equation (D9)

with

Equation (D10)

Here, each prime denotes a single derivative of the function with respect to its argument.

Substituting the expression for ${{ \mathcal I }}_{{nlm}}$ given in Equation (D7) in the expression for the disk response given in Equation (21), and adopting the fiducial parameters for the MW galaxy and those corresponding to satellite encounters (as detailed in Section 4.2), we compute the response of the MW disk to past and future encounters with its satellite galaxies. Results for the steady-state disk response (in the collisionless limit) of the (n, l, m) = (1, 0, 0) mode, corresponding to Iz = Iz,⊙ = hz σz,⊙ and Rc (Lz ) = 8 kpc and marginalized over IR , are summarized in Table 1.

D.2. Special Case: Disk Response for Face-on Impulsive Encounters

The disk response in the general case, expressed by Equation (39), depends on several encounter parameters (Rd, θP, and ϕP), and it is complicated to evaluate. Therefore, as a sanity check, here we compute the response as well as corresponding energy change for the special case of a satellite undergoing an impulsive, perpendicular passage through the center of the disk.

As shown in van den Bosch et al. (2018) (see also Banik & van den Bosch 2021b), the total energy change due to a head-on encounter of velocity vP with a Plummer sphere of mass MP and size ε is given by

Equation (D11)

where

Equation (D12)

Knowing that the enclosed mass profile of a Plummer sphere is given by ${M}_{{\rm{P}}}(R)={M}_{{\rm{P}}}{R}^{3}{\left({R}^{2}+{\varepsilon }^{2}\right)}^{-3/2}$, we find that I0(R) = R2/(R2 + ε2), which yields

Equation (D13)

Now we compute the disk response to the face-on satellite encounter using Equations (21) and (D7)–(D10). For a perpendicular face-on impact through the center of the disk, we have Rd = 0 and θP = 0, implying that ${{ \mathcal R }}_{{\rm{c}}}$ becomes Rc. The corresponding response is greatly simplified. In the large time and small IR limit, it is given by Equation (21) with

Equation (D14)

where the ϕ integral only leaves contribution from the axisymmetric m = 0 mode. The wR integrand can be expanded as a Taylor series, and the wR integral can be performed to yield the following leading-order expression for ${{ \mathcal I }}_{{nlm}}$:

Equation (D15)

In the impulsive limit, vP , this becomes

Equation (D16)

which can be substituted in Equation (21) to yield

Equation (D17)

with f0 given by Equation (15). Hence, the response is given by

Equation (D18)

which shows that the satellite passage introduces a relative overdensity, ${f}_{1}\left({\boldsymbol{w}},{\boldsymbol{I}},t\right)/{f}_{0}({\boldsymbol{I}})$, that scales as $\sim {R}_{{\rm{c}}}/\left({\varepsilon }^{2}+{R}_{c}^{2}\right)$, which increases from zero at the center, peaks at Rc = ε, and asymptotes to zero again at large Rc. The $\cos ({w}_{R}-\kappa t)$ term describes the radial epicyclic oscillations in the response.

To compute the energy change due to the impact, we note that dE/dt = ∂E/∂ I · d I /dt, where ∂E/∂ I = Ω = (Ωz , ΩR , Ωϕ ) and d I /dt = − ∂ΦP/∂ w from Hamilton's equations of motion. Thus, the total phase-averaged energy injected per unit of phase space can be obtained as follows:

Equation (D19)

We can substitute the Fourier series expansions of ΦP and f1 given in Equations (10) in the above expression and integrate over w to obtain (Weinberg 1994a, 1994b)

Equation (D20)

We can now substitute the form of ΦP for a Plummer perturber given in Equation (38), with r P and r given by Equations (36) and (37). The time integral can thus be written as

Equation (D21)

Using Equations (C2) and (C4) to express R and z in terms of ( w , I ), and substituting the form for f1nlm ( I , t) from Equation (D17), we can perform the above integrals over w and t. Substituting the result in Equation (D20), we obtain

Equation (D22)

The total energy, ΔEtot, imparted into the disk by the impulsive satellite passage can be computed by integrating the above expression over I and w (which simply introduces a factor of ${\left(2\pi \right)}^{3}$, because $\left\langle {\rm{\Delta }}E\left({\boldsymbol{I}}\right)\right\rangle $ is already phase-averaged), using Equation (15) and transforming from Lz to Rc using the Jacobian dLz /dRc = Rc κ2/2Ωϕ . This yields

Equation (D23)

This is indeed the expression for ΔEtot derived under the impulse approximation given by Equation (D13).

Footnotes

  • 3  

    In the limit of small IR , the radial energy ER can be approximated as κ IR . In this case, the isothermal form of the unperturbed DF, $\exp \left[-{E}_{R}/{\sigma }_{R}^{2}\right]$, reduces to $\exp \left[-\kappa {I}_{R}/{\sigma }_{R}^{2}\right]$, which is known as a pseudo-isothermal distribution.

  • 4  

    The 3MN profile as implemented in the Gala Python package (Price-Whelan 2017; Price-Whelan et al. 2020).

  • 5  

    The bulge is modeled as a spherical Hernquist (1990) profile with mass Mb = 6.5 × 109 M and scale radius rb = 0.6 kpc.

Please wait… references are loading.
10.3847/1538-4357/acd641