This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

Secular Evolution Driven by Massive Eccentric Disks/Rings: An Apsidally Aligned Case

and

Published 2018 August 30 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Irina Davydenkova and Roman R. Rafikov 2018 ApJ 864 74 DOI 10.3847/1538-4357/aad3ba

Download Article PDF
DownloadArticle ePub

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

0004-637X/864/1/74

Abstract

Massive eccentric disks (gaseous or particulate) orbiting a dominant central mass appear in many astrophysical systems, including planetary rings, protoplanetary and accretion disks in binaries, and nuclear stellar disks around supermassive black holes in galactic centers. We present an analytical framework for treating the nearly Keplerian secular dynamics of test particles driven by the gravity of an eccentric, apsidally aligned, zero-thickness disk with arbitrary surface density and eccentricity profiles. We derive a disturbing function describing the secular evolution of coplanar objects, which is explicitly related (via one-dimensional, convergent integrals) to the disk surface density and eccentricity profiles without using any ad hoc softening of the potential. Our analytical framework is verified via direct orbit integrations, which show it to be accurate in the low-eccentricity limit for a variety of disk models (for disk eccentricity ≲0.1–0.2). We find that free precession in the potential of a disk with a smooth surface density distribution can naturally change from prograde to retrograde within the disk. Sharp disk features—edges and gaps—are the locations where this tendency is naturally enhanced, while the precession becomes very fast. Radii where free precession changes sign are the locations where substantial (formally singular) growth of the forced eccentricity of the orbiting objects occurs. Based on our results, we formulate a self-consistent analytical framework for computing an eccentricity profile for an aligned, eccentric disk (with a prescribed surface density profile) capable of precessing as a solid body under its own self-gravity.

Export citation and abstract BibTeX RIS

1. Introduction

Astrophysical disks orbiting in the gravitational potential of a dominant central mass Mc often possess nonaxisymmetric shape. The nonaxisymmetric distortion can be modeled as a manifestation of disk eccentricity. In this picture, to zeroth order, different components of the disk—parcels of gas in fluid (collisional) disks or particles (e.g., stars) in collisionless disks—move on eccentric Keplerian orbits in the field of a central mass. Even if the mass of the disk Md is much smaller than Mc, the self-gravity of the disk can still play a very important role in its dynamics, as well as the orbital evolution of external objects, by driving precession and causing an exchange of angular momentum between different parts of the disk on long (secular) timescales.

Such eccentric disks are encountered in a variety of astrophysical contexts—galactic, stellar, and planetary. One of the closest examples is provided by the eccentric planetary rings, such as the epsilon, α, and β rings4 of Uranus (Elliot & Nicholson 1984), as well as the Titan and Maxwell ringlets of Saturn (Porco 1990). These particulate collisional rings are very narrow, essentially representing limiting cases of eccentric disks with the spread in semimajor axes of their constituent particles Δa ≲ er ar, where ar and er are the mean semimajor axis and eccentricity of the rings. As demonstrated by Goldreich & Tremaine (1979), self-gravity of the rings, coupled with collisional effects (Chiang & Goldreich 2000; Mosqueira & Estrada 2002; Chiang & Culter 2003; Pan & Wu 2016), can counter differential precession driven by the planetary oblateness, allowing rings to precess as a solid body while maintaining a coherent eccentric shape.

A significant number of stellar binaries are known to host exoplanets, orbiting either the whole system (Doyle et al. 2011; Welsh et al. 2014) or one of the binary components (Hatzes et al. 2003; Chauvin et al. 2011). The formation and early dynamics of such planets are significantly complicated by the fact that the nonaxisymmetric binary potential excites the nonzero eccentricity of the protoplanetary disks (Kley et al. 2008; Regály et al. 2011; Miranda et al. 2017), in which the building blocks of these planets—planetesimals—orbit. It has been recently shown (Rafikov 2013a, 2013b; Rafikov & Silsbee 2015a, 2015b; Silsbee & Rafikov 2015a, 2015b) that the gravitational effect of such an eccentric protoplanetary disk plays a key role in planetesimal dynamics for young binaries.

Spectroscopic observations of accretion disks in cataclysmic variables using the technique of Doppler tomography (Marsh & Horne 1988) suggest that a certain type of variability in these systems—the so-called "superhump" (Horne 1984)—is caused by the precession of an eccentric accretion disk around the white dwarf (Lubow 1991). Asymmetric evolving emission lines found in the spectra of compact disks of gaseous debris around some metal-rich white dwarfs (Gänsicke et al. 2006) also provide evidence for their nonzero eccentricity (Dennihy et al. 2018; Miranda & Rafikov 2018).

Finally, optical emission from the central region of the M31 galaxy exhibits a double-nucleus morphology (Bacon et al. 2001; Bender et al. 2005). The best interpretation of the existing photometric and spectroscopic data, first proposed by Tremaine (1995), points at the existence of a highly eccentric (ed ∼ 0.5) stellar disk orbiting the central supermassive black hole. A number of models, both purely kinematic, i.e., not accounting for the disk self-gravity in maintaining its coherence (Peiris & Tremaine 2003; Brown & Magorrian 2013), and fully dynamic (Bacon et al. 2001; Salow & Statler 2001; Sambhus & Sridhar 2002; Salow & Statler 2004), have been put forward to understand this system. Our own Galaxy harbors an eccentric disk of young stars orbiting the central supermassive black hole (Levin & Beloborodov 2003; Bartko et al. 2009), whose gravity may affect its own dynamics (Nayakshin et al. 2006).

In many of these systems, disk gravity plays an important role in disk dynamics, as well as in the orbital evolution of nearby objects (e.g., planetesimals in protoplanetary disks in binaries). This motivated a number of past analytical (Goldreich & Tremaine 1979; Chiang & Goldreich 2000) and numerical (Bacon et al. 2001; Sambhus & Sridhar 2002; Salow & Statler 2004; Nayakshin et al. 2006) studies aimed at clarifying the details of the dynamics driven by the gravity of an eccentric disk. Such calculations inevitably require an efficient method for computing the potential Φd of an eccentric disk at every point. Moreover, since Md ≪ Mc, the disk-driven evolution is typically rather slow, justifying the use of a secular approximation, in which the disk potential is averaged over the orbital motion of an object under consideration. A direct calculation of such an averaged potential, or a secular disturbing function, as it is known in celestial mechanics, in general requires evaluation of three-dimensional integrals (see Equation (21)), which is impractical in many applications.

Silsbee & Rafikov (2015b) presented a calculation of a secular disturbing function for a particular model of a radially extended (i.e., having Δaa), apsidally aligned eccentric disk. They assumed that both the surface density and the eccentricity of the disk vary as power laws of the semimajor axis a of the mass elements comprising the disk. Their resultant disturbing function does not involve multidimensional integration and can be used for efficient analysis of disk-driven orbital dynamics. In particular, it was employed to provide a self-consistent treatment of the secular evolution of planetesimals orbiting in a massive eccentric protoplanetary disk within (or around) a young stellar binary (Rafikov & Silsbee 2015a, 2015b; Silsbee & Rafikov 2015a).

The goal of our present work is to provide a natural but important generalization of the results of Silsbee & Rafikov (2015b). Here we develop a general analytical framework for computing a secular disturbing function for an apsidally aligned,5 eccentric disk with arbitrary radial profiles of the disk surface density Σ and eccentricity ed. We show that this disturbing function can be reduced to a combination of one-dimensional integrals over the radial profiles of Σ and ed, enabling application of our results to a broad range of practical problems (e.g., computation of the structure of a rigidly precessing eccentric disk; see Section 7). We also provide numerical verification of our analytical results using direct orbit integrations.

Our work is organized as follows. We describe our methodology and outline the results of the disturbing function calculation in Section 2; the details of its derivation can be found in Appendix A. We describe the strategy for numerical verification of our analytical calculations in Section 3 and present our findings in Section 4. Having tested our analytical framework, we then describe several of its applications in Section 5, including the derivation of a self-consistent method for calculating the eccentricity distribution of an eccentric, apsidally aligned disk (with a prescribed surface density distribution) that can precess as a solid body while maintaining its overall shape (Section 6). We discuss our results in Section 7 and provide a brief summary in Section 8.

2. Secular Disturbing Function

Our goal is to calculate the secular (i.e., orbit-averaged) gravitational potential felt by a test particle orbiting in the combined gravitational field of central mass Mc and an eccentric disk (fluid or particulate). This particle can be an external object or one of the mass elements comprising the disk (see Section 6). The test particle moves on an eccentric orbit coplanar with the disk, with semimajor axis ap, eccentricity ep, and apsidal angle ϖp.

The disk is purely two-dimensional (i.e., it has zero vertical thickness) and is not warped (i.e., lies in a single plane). It is eccentric and apsidally aligned in the sense that the trajectories of its constituent mass elements (fluid or particle) are confocal Keplerian ellipses, which are apsidally aligned in the direction making an angle ϖd with respect to the reference direction. We define rd to be the distance from the common focus of the eccentric orbits of the constituent particles and φd to be the polar angle with respect to the disk apsidal line; see Figure 1 for illustration.

Figure 1.

Figure 1. Geometry of the problem, showing elliptical trajectories of the test particle (blue) and a mass element of the disk (green). See the text for details.

Standard image High-resolution image

For every trajectory with a semimajor axis a, we can define the disk surface density at the periastron Σd(a) and the eccentricity of the fluid trajectory ed(a), which we will simply call disk eccentricity. The disk mass distribution can also be characterized by the mass per unit semimajor axis μ(a). Using the basic properties of Keplerian dynamics, one can show that Σd and μ are directly related via

Equation (1)

where ζ ≡ d ln ed(a)/d ln  a. In general, Σd(a) (or μ(a)) and ed(a) can be arbitrary functions of the semimajor axis ad, as long as ed(a) varies slowly enough for the particle trajectories to be noncrossing.6 In this work, we choose Σd (rather than μ) to characterize the distribution of mass in the disk.

Note that in the secular approximation relied upon in this study, the energy and semimajor axes of particles (or fluid elements) comprising the disk are the integrals of motion. As a result, the amount of disk mass per unit semimajor axis μ(a) is strictly conserved even if the disk shape changes. Consequently, according to Equation (1), if ed does not change in time, then Σd(a) is also independent of time (this will be important in Section 6).

Ogilvie (2001) and Statler (2001) provided an expression for the two-dimensional surface density Σ(rd, φd) of an eccentric disk in terms of ed(a) and disk mass distribution μ(a). For our present purposes, it is more convenient to write Σ as a function of a and φd, relating it to Σd(a). Using the calculation of Σ(a, φd) in Statler (1999), for an apsidally aligned disk, we can write

Equation (2)

where E(φd) is the eccentric anomaly (E = φd = 0 at pericenter) and Σd and ed are functions of a.

Even though Equation (2) holds for arbitrary ed, in the rest of the paper, we will take the eccentricities of both the disk and test particle to be small, ed(r) ≪ 1 and ep ≪ 1. This is needed for our secular theory (formulated at the lowest order in eccentricity) to provide an accurate description of orbital dynamics. As a consequence of this approximation, Equation (1) also yields Σd(a) ≈ μ(a)/(2π a) to lowest order in ed. Thus, even if ed ≪ 1 varies in time, Σd(a) should still be conserved to O(ed) accuracy in the course of secular evolution.

2.1. Secular (Orbit-averaged) Potential of the Disk

Our calculation of the orbit-averaged disturbing function Rd due to an eccentric disk uses the general mathematical procedure outlined in a seminal paper of Heppenheimer (1980) for the calculation of the disturbing function due to an axisymmetric disk. In this approach, the expansion of the disturbing function in terms of a small parameter—test particle eccentricity—proceeds differently from the classical Laplace–Lagrange theory (Murray & Dermott 1999). The resultant expression for Rd does not contain non-integrable singularities at the particle semimajor axis, resulting in a convergent expression for the disturbing function. In other words, this calculation does not require introduction of an ad hoc softening of the potential. This method was later used by Ward (1981) to study the stability of the early solar system perturbed by an axisymmetric protoplanetary disk.

Silsbee & Rafikov (2015b) extended the method of Heppenheimer (1980) to the case of nonaxisymmetric, apsidally aligned, eccentric disks with Σ given by Equation (2). However, their work was restricted to disks with power-law profiles of Σd(a) and ed(a). Here we generalize this calculation even further to cover the disks with arbitrary behaviors of Σd(a) and ed(a).

As a result of a rather lengthy derivation, the mathematical details of which are presented in Appendix A, we arrive at the following expression for the disturbing function due to an apsidally aligned disk:

Equation (3)

where np = (GMc/ap3)1/2 is the particle mean motion, with the coefficients Ad and Bd (having dimensions of [s−1] and discussed in more detail in Section 2.2) given by Equations (45)–(49). The key underlying simplifications making this calculation possible are that ed ≪ 1, as well as d ed/d ln a ≪ 1; see Equation (22).

Introducing a two-component eccentricity vector ep =(kp, hp) = ep(cosϖp, sin ϖp) for a test particle, as well as the auxiliary vector

Equation (4)

Equation (3) can be rewritten as

Equation (5)

Note that Bd is different from the disk eccentricity vector ed = ed(cosϖd, sin ϖd); in fact, Bd involves a convolution of ed(a) with a complicated kernel (see Equation (48)) over the radial extent of the disk.

The mathematical structure of Rd in Equation (3) is identical to that of the conventional Laplace–Lagrange secular disturbing function (Murray & Dermott 1999). The difference between them lies in the explicit dependence of the coefficients Ad and Bd on ap. The behavior of Ad(ap) and Bd(ap) is determined, eventually, by the radial profiles of the disk surface density Σd and eccentricity ed. The exploration of this behavior is the major focus of our study.

2.2. Mathematical Properties of Ad and Bd

Equation (45) for Ad consists of two parts. One of them, ${A}_{d}^{\mathrm{bulk}}$, involves integral convolution over the disk semimajor axis a of a (prescribed) disk surface density ${{\rm{\Sigma }}}_{d}(a)$, as well as its radial derivatives up to second order, all of which enter linearly; see Equation (46). The convolution kernel involves a Laplace coefficient ${b}_{1/2}^{(0)}(\alpha )$, which features a weak (logarithmic) singularity arising as α = a/ap → 1 (i.e., for disk annuli close to the particle orbit). This singularity is, however, fully integrable and vanishes upon integration over the radial extent of the disk (see Appendix A). As always, the Laplace coefficients are defined as

Equation (6)

In addition, whenever the disk has sharp edges (or discontinuous transitions in surface density), Equation (45) for Ad features boundary terms ${A}_{d}^{\mathrm{edge}}$, given by Equation (47). These terms involve ${b}_{1/2}^{(0)}(\alpha )$ and its derivative ${b}_{1/2}^{(0)^{\prime} }(\alpha )$, both evaluated at α = ae/ap, where ae is the semimajor axis of the disk edge. As the semimajor axis of a test particle ap approaches the disk edge, both ${b}_{1/2}^{(0)}(\alpha )$ and ${b}_{1/2}^{(0)^{\prime} }(\alpha )$ diverge: ${b}_{1/2}^{(0)}(\alpha )\sim \mathrm{ln}| 1-\alpha | $, while ${b}_{1/2}^{(0)^{\prime} }(\alpha )\sim | 1-\alpha {| }^{-1}$, where in this case $\alpha ={a}_{e}/{a}_{p}\to 1$. As a result, boundary terms, as well as Ad, diverge at the sharp disk edge. This singularity of the secular disturbing function is further explored in Section 4.

The divergence of Ad at the disk edge does not arise if ${{\rm{\Sigma }}}_{d}(a)$ goes to zero at the boundary in a sufficiently smooth fashion. Indeed, ${b}_{1/2}^{(0)^{\prime} }(\alpha )$ in the boundary terms in Equation (47) is multiplied by Σd at the edge; ${b}_{1/2}^{(0)}(\alpha )$ is multiplied by both Σd and its radial derivative ${{\rm{\Sigma }}}_{d}^{{\prime} }$. As a result, boundary terms vanish whenever ${{\rm{\Sigma }}}_{d}\propto | a-{a}_{e}{| }^{\kappa }$ and $\kappa \gt 1$ near the edge at ae. Thus, in disks with Σd smoothly (faster than linearly in $| a-{a}_{e}| $) turning to zero at a finite semimajor axis ae, the coefficient Ad has no boundary contributions and, hence, does not diverge at the boundary. The same is also true for disks without edges, which have their surface density smoothly declining to zero as a → 0 and $a\to \infty $. Both possibilities will be explored further in Section 4.

Similarly, Equation (45) for Bd involves radial integrals over the product of Σd(a) and the disk eccentricity ed(a), as well as their derivatives up to second order (contribution ${B}_{d}^{\mathrm{bulk}}$ given by Equation (48)), in addition to the boundary terms ${B}_{d}^{\mathrm{edge}}$ represented by Equation (48). Once again, ${B}_{d}^{\mathrm{edge}}$ vanishes whenever Σd goes to zero at the disk edges sufficiently rapidly, e.g., as ${{\rm{\Sigma }}}_{d}\sim | a-{a}_{e}{| }^{\kappa }$ with κ > 1.

Other features of Ad(ap) and Bd(ap) behavior and their effect on secular dynamics will be discussed in Section 47.

3. Numerical Verification

Having derived the analytical framework embodied in Equations (3) and (45)–(49), we also provide its numerical verification. We do this by comparing the eccentricity evolution of test particles computed based on our analytical results with the results of direct orbit integration in the potential of an eccentric disk (computed numerically). Details of both approaches, as well as the disk models used in this comparison, are outlined below.

3.1. (Semi-)analytical Orbital Evolution

One approach to calculating orbital evolution in the potential of an eccentric disk uses Lagrange equations for the evolution of orbital elements, ${\dot{k}}_{p}=-{({a}_{p}^{2}{n}_{p})}^{-1}\partial {R}_{d}/\partial {h}_{p}$, ${\dot{h}}_{p}={({a}_{p}^{2}{n}_{p})}^{-1}\partial {R}_{d}/\partial {k}_{p}$, where we use the secular expression of Equation (3) for the disturbing function Rd. As a result, one finds

Equation (7)

with the general solution given by the superposition of the free and forced eccentricity vectors (Murray & Dermott 1999):

Equation (8)

Equation (9)

Equation (10)

Here the forced eccentricity is

Equation (11)

and constants efree > 0 and ϖ0 are such that at t = 0, Equation (8) satisfies the initial condition ${{\boldsymbol{e}}}_{p}(0)=(k(0),h(0))={e}_{p}(0)(\cos {\varpi }_{p}(0),\sin {\varpi }_{p}(0))$, where ${e}_{p}(0)$ and ${\varpi }_{p}(0)$ are the initial eccentricity and periastron angle of an object.

In particular, an object starting on a circular orbit (ep(0) = 0) has efree = eforced, and its motion is described by

Equation (12)

where ϖp stays in the range (0, 2π). We will use this solution for numerical verification of our semi-analytical results, although any alternative solution corresponding to different initial conditions would work just as well.

Full determination of ep requires calculating Ad and Bd for the given profiles of the disk surface density Σd(a) and eccentricity ϖd(a). We do this with the help of Equations (45)–(49) by numerically evaluating the corresponding integrals. Despite the weak (logarithmic) singularity of the integrands, the integrals themselves are convergent and calculated directly without resorting to any type of softening of the integrand near its singular point (which is often done in studies of disk dynamics; see, e.g., Tremaine 2001; Hahn 2003; Touma et al. 2009). The lack of an ad hoc parameter (softening length) in our calculation is its important distinctive feature.

Once Ad and Bd are calculated, Equations (12) and (13) (or, more generally, Equations (8)–(11)) provide a full semi-analytical description of the particle motion in the disk potential.

3.2. Direct Orbit Integration

An alternative method to compute particle motion uses the direct orbit integrator MERCURY (Chambers 1999), which employs the Bulirsch–Stoer algorithm (Press et al. 1992). This integration uses as its input the gravitational acceleration g due to the disk potential, which is calculated on a grid of (245, 100) points7 in (r, φ). Values of g on the grid are then interpolated to provide accurate accelerations everywhere in the disk.

At each grid point, g is computed by direct summation of ∇ Φd (see Equation (21)) over the disk surface, with the surface density given by Equation (2) in its exact form, i.e., not performing small-ed expansion. This two-dimensional integral is convergent in the Cauchy principal value sense, even though the integrand diverges at the point where acceleration is calculated. To avoid this mathematical singularity, our numerical evaluation is performed at a small height (10−3 au or smaller) above the disk; we make sure that the result is convergent with respect to the value of this height. Note that this procedure is not equivalent to the introduction of softening in our theoretical calculation.

We integrate particle orbits starting with ep = 0. This allows us to directly compare the results with Equations (12) and (13) following from our semi-analytical calculation, thus directly verifying the accuracy of our framework for the secular evolution in the potential of an eccentric disk.

3.3. Summary of the Disk Models Used

In our comparison effort, we use the following model distributions of the disk eccentricity ed(a):

Equation (13)

Equation (14)

and surface density at periastron Σd(a),

Equation (15)

Equation (16)

Equation (17)

with a1 = 0.1 au, a2 = 5 au, and ac = 1.5 au and e0 and Σ0 being the normalization factors that we vary in our models. In our calculations, we always assume the mass of the central star to be 1 M.

Figure 2 illustrates the behavior of these profiles of Σd and ed. In the majority of our calculations, we use the linear eccentricity profile given by Equation (13), corresponding to a disk that is everywhere eccentric. Disks with the eccentricity profile given by Equation (14) studied in Section 4.3.2 have circular inner and outer edges but are eccentric in between.

Figure 2.

Figure 2. Profiles of the disk surface density Σd(a) and eccentricity ed(a) considered in this work. Panel (a) shows the following Σd(a) profiles: (i) Equation (15) with Σ0 ≈ 1100 g cm−2, (ii) Equation (16) with Σ0 =100 g cm−2, and (iii) Equation (17) with Σ0 = 20 g cm−2. Panel (b) shows the ed(a) profiles as follows: (i) Equation (14) with e0 = 2 and (ii) Equation (13) with e0 = 0.1.

Standard image High-resolution image

The surface density profile given by Equation (15) corresponds to a disk in which Σd smoothly goes to zero at the edges at a1 and a2. Based on the discussion in Section 2.2, we expect Ad and Bd not to diverge at the edges of such a disk, which is verified in Section 4.1.1. Equation (16) describes a disk without edges (Section 4.1.2), which has its surface density exponentially decreasing for both a ≪ ac and a ≫ ac. This disk also does not feature boundary terms (as it has no boundaries). Finally, the Σd profile given by Equation (17) describes a disk with discontinuous drops of the surface density at the edges. In this disk, we expect Ad and Bd to diverge at the boundaries, which is verified in Section 4.1.3.

Figure 3 illustrates the 2D distributions of the surface density obtained via Equation (2) with some of the Σd and ed profiles listed above. One can see that Σ(r, φ) can have a rather complicated structure, depending on the particular disk model used. The contours of constant Σ(r, φ) (left) are very different from the elliptical trajectories (right) of the mass elements giving rise to the surface density distribution in the disk.

Figure 3.

Figure 3. Maps of the disk surface density Σ(r, φ) (color indicates the amplitude of Σ) with overlaid contours of Σ (thin curves; left panels) and eccentric trajectories of the mass elements comprising the disk (gray; right panels). The disk is always oriented with the apsidal lines pointing to the right. Panels (a) and (b) are drawn for Σd(a), given by Equation (15) with Σ0 ≈ 1100 g cm−2, and ed(a), given by Equation (13) with e0 = 0.2 (this high e0 was chosen to better illustrate the Σ distribution). Panels (c) and (d) are drawn for the same Σd(a) but a different disk eccentricity profile given by Equation (14) with e0 = 0.5. Note the substantial difference in Σ(r, φ) caused by just the difference in the ed(a) profiles.

Standard image High-resolution image

4. Results

We start our comparison by providing an illustration of the orbital evolution caused by the gravity of an underlying eccentric disk. Figure 4 displays the variation of the test particle eccentricity ep (left) and apsidal angle ϖp (right) in time, computed at several values of the semimajor axis ap for a particular disk model—Σd(a) given by Equation (15) with Σ0 ≈ 1100 g cm−2 and ed(a) given by Equation (13) with e0 = 10−2. One can see that the agreement between the direct integration (green) and our secular prediction (red) is very good at all radii. Both ep(t) and ϖp(t) follow the predictions of Equations (12) and (13) very closely, agreeing both in the amplitude of the eccentricity oscillations and in their phase (or period).

Figure 4.

Figure 4. Verification of our analytical calculation of the disk disturbing function Rd using direct numerical integration of particle orbits in the (numerically computed) disk potential with MERCURY. The disk model has ${{\rm{\Sigma }}}_{d}(a)$ given by Equation (15) with ${{\rm{\Sigma }}}_{0}\approx 1100$ g cm−2 and ed(a) given by Equation (13) with e0 = 0.01. The time evolution of the particle eccentricity ep (left) and apsidal angle ϖp (right) is shown for different values of the particle semimajor axis ap, as labeled on the panels. In all cases, test particles start with zero eccentricity, which results in pronounced secular oscillations of ep, well described by the solution of Equation (12). The free eccentricity vector precesses at a steady rate, resulting in ϖd evolution characterized by the solution of Equation (13).

Standard image High-resolution image

From the behavior of ϖp(t) at different semimajor axes, one can immediately see that free precession of the particle orbit can be prograde in some parts of the disk and retrograde in others. The change of sign of the precession occurs for orbits fully enclosed within the disk. It is also clear that not only the sign but also the period of the precession is a function of the location in the disk.

Evolutionary time series of ep(t), such as the one presented in Figure 4, allow us to easily measure the maximum amplitude of the eccentricity oscillations, emp, which should be compared to the theoretical values of $2{B}_{d}/{A}_{d}$; see Equation (12). Similarly, the period Psec of each ep oscillation yields the corresponding free precession rate as ${\dot{\varpi }}_{\sec }=2\pi /{P}_{\sec }$, which in this figure agrees very well with the theoretical Ad. The value of ${\dot{\varpi }}_{\sec }$ can also be independently inferred from the slope of the numerical $\dot{\varpi }(t)$ curves, which should be equal to Ad/2; see Equation (13).

Close inspection of Figure 4 reveals additional features in the ep(t), ϖp(t) evolution curves beyond the large-scale secular oscillations, which are well described by Equations (12) and (13). These features manifest themselves as small-amplitude, short-period oscillations around the purely secular solution, most pronounced near the outer edge of the disk. The period of these oscillations is equal to the local orbital period 2π/np. Their amplitude scales linearly with the disk mass, depends on the semimajor axis of test particle ap, and is independent of the disk eccentricity (we observe oscillations with the same amplitude even when the disk eccentricity is zero). We interpret these oscillations as resulting from the nonelliptical shape of the particle orbits in the combined potential of the star and disk. As the numerical integration outputs osculating orbital elements (effectively fitting a pure ellipse at every point of a truly nonelliptical trajectory), oscillations of ep (as well as ϖp and ap) on a local dynamical timescale naturally arise. We did verify that the angular momentum of a test particle is strictly conserved through these oscillation cycles when ed = 0. A similar effect was discussed by Georgakarakos (2002) in application to hierarchical triples.

In our subsequent presentation, we will focus on the behavior of ${e}_{p}^{{\rm{m}}}$ and ${\dot{\varpi }}_{\sec }$, derived from data similar to those shown in Figure 4, in different disk models.

4.1. Radial Variation for Different Σd Models

We now test the accuracy of our secular theory in disks with different Σd(a) profiles. This allows us to examine the different possible behaviors of the secular coefficients Ad and Bd, as well as to see how sensitive the agreement with the numerical results is to the different features of the Σd(a) distributions.

4.1.1. Σd Smoothly Vanishing at the Edges

We start by looking at the secular effect of a disk, in which Σd smoothly goes to zero at the boundaries, i.e., the one with Σd given by Equation (15). In Figure 5, we display the theoretical radial profiles of Ad, Bd, and 2Bd/Ad for such a disk. The curves of Ad(ap) are compared with the numerically determined ${\dot{\varpi }}_{\sec }$ (blue dots), while the theoretical $2{B}_{d}({a}_{p})/{A}_{d}({a}_{p})$ is compared against the numerical ${e}_{p}^{{\rm{m}}}$. This particular calculation uses the linear eccentricity profile ed(a) given by Equation (13) with relatively low e0 = 0.01, so that the maximum value of the disk eccentricity (reached at its inner edge) is 0.02. Thus, the requirement ed ≪ 1 necessary for the secular results of Equations (3) and (45)–(49) to apply is fulfilled.

Figure 5.

Figure 5. Characterization of secular oscillations of test particles (with orbits fully enclosed within the disk) driven by the disk gravity. Shown as a function of the semimajor axis of test particle ap are (a) the amplitude ${e}_{p}^{{\rm{m}}}$ of the eccentricity oscillations (blue dots) compared with theoretical $2{e}_{\mathrm{forced}}=2{B}_{d}/{A}_{d}$ (red curve), (b) the frequency of secular oscillations ${\dot{\varpi }}_{\sec }$ (blue dots) compared against Ad (red curve), and (c) the coefficient Bd of the theoretical secular disturbing function. This calculation uses the disk model with Σd distribution (Equation (15)) and Σ0 ≈1100 g cm−2, so that Σd goes to zero at finite semimajor axes (0.1 and 5 au), but smoothly, so that the boundary terms in the expressions for Ad and Bd do not arise; see Section 2.2. The radial profile of the disk eccentricity ed(a) is given by Equation (13) with e0 = 0.01. Circled dots correspond to the four values of ap used in Figure 4. Vertical dotted lines mark the locations where Ad = 0. Gray regions near the disk edges correspond to the edge noncrossing constraints ${a}_{p}(1+{e}_{p}^{{\rm{m}}})\lt {a}_{2}$ at apastron and ${a}_{p}(1-{e}_{p}^{{\rm{m}}})\gt {a}_{1}$ at periastron. One can see excellent agreement between the secular theory and the direct numerical integration.

Standard image High-resolution image

One can see an almost perfect match between the numerical results and theory in terms of both the amplitude of the disk-induced particle eccentricity and the period of the associated orbital precession. Theoretical calculation easily reproduces even the very subtle features of the ${e}_{p}^{{\rm{m}}}(a)$ behavior, including the variations happening on very short radial scales manifesting themselves as sharp features in Figure 5. There are several features to note in this figure.

First, Ad has a different sign in different parts of the disk: precession of the free eccentricity vector is prograde near the boundaries of the disk, while it is retrograde away from the edges. This change of sign of ${\dot{\varpi }}_{\sec }$ was clear already in Figure 4 (which is drawn for the same disk model as Figure 5, at the locations highlighted with circles in the latter), but Figure 5 provides a much more detailed representation of the different characteristics of the secular effect of the disk.

Second, Bd also changes sign as ap varies. As a result of sign variations of both Ad and Bd, the forced eccentricity vector ${{\boldsymbol{e}}}_{\mathrm{forced}}$ can be aligned with the apsidal line of the disk in some intervals of ap and anti-aligned with it in others.

Third, both numerical and analytical ${e}_{p}^{{\rm{m}}}$ exhibit formal singularity at two distinct locations in this disk, a ≈ 1.54 and ≈4.39 au. The origin of these singularities can be traced to Equation (11) for the forced eccentricity (recall that our theory predicts ${e}_{p}^{{\rm{m}}}=2{e}_{\mathrm{forced}}$), which has Ad in the denominator, and the fact that Ad crosses zero (i.e., the direction of efree precession changes) at these locations; see Figure 5(b). We will discuss these singularities in more detail in Section 4.2.

Fourth, at certain values of the semimajor axis (at ≈0.92 and ≈4.39 au), the disk-induced forced eccentricity vanishes. This happens because Bd changes sign at these locations so that Bd → 0; see Figure 5(c) and Equation (11). Interestingly, one of the semimajor axes where ${e}_{p}^{{\rm{m}}}\to 0$ (${a}_{p}\approx 4.39$ au) is located in the immediate proximity of the singularity of ${e}_{p}^{{\rm{m}}}$. This occurs because in this part of the disk, Ad and Bd go through zero at almost the same (but still slightly different) values of ap.

Fifth, for the surface density profile given by Equation (15), we find that Ad and Bd remain finite everywhere in the disk, including the boundaries. This is in line with the expectations outlined in Section 2.2 for the disks with Σd smoothly going to zero at the edges, which do not result in divergent boundary terms in Ad and Bd.

4.1.2. Disk without Boundaries

Next we consider secular dynamics in the potential of a Gaussian ring with Σd given by Equation (16), which does not feature well-defined edges. We plot the behavior of the corresponding ${e}_{p}^{{\rm{m}}}$, Ad, and Bd in Figure 6.

Figure 6.

Figure 6. Same as Figure 5 but for a disk (Gaussian ring) model with different Σd(a) distribution (Equation (16)), with Σ0 = 100 g cm−2; this model does not have boundaries at finite radii. The disk eccentricity profile is still given by Equation (13) with e0 = 0.01. The inset in panel (b) shows the behavior of Ad far from the main body of the ring.

Standard image High-resolution image

One can see that many features of secular dynamics present in the case studied in Section 4.1.1 are present here as well: both Ad and Bd are finite, vary with ap, and change sign, while ${e}_{p}^{{\rm{m}}}$ exhibits both singularities and nulls. This is likely related to the fact that these two Σd profiles are morphologically similar: they exhibit no discontinuities, are single-peaked, and smoothly decay to zero away from the peak. The only obvious difference is that in the Gaussian case in Figure 6, the nulls of ${e}_{p}^{{\rm{m}}}$ are located very close to both singularities of ${e}_{p}^{{\rm{m}}}$. However, the significance of this difference is unclear.

4.1.3. Σd Sharply Truncated at the Edges

We also examined secular behavior for the disk model in which surface density displays a discontinuous drop to zero at the inner and outer edges, namely, the one represented by Equation (17). Sharply truncated Σ distributions are rather typical in planetary rings and other astrophysical disks, and it is important to study the effect of their gravity on the dynamics of embedded objects.

Figure 7 illustrates secular dynamics in the potential of such a disk. It again shows the variation of sign of Ad, which results in the emergence of two singularities of ${e}_{p}^{{\rm{m}}}$—one very close to the inner edge of the disk at ap = 0.11 au and another at ${a}_{p}\approx 1.28\,\mathrm{au}$. However, for this disk model, Bd does not change sign—it always stays positive. As a result, there are no nulls of ${{\boldsymbol{e}}}_{p}^{{\rm{m}}}$ for orbits in the potential of such a disk.

Figure 7.

Figure 7. Same as Figure 5 but for the disk model with Σd(a) distribution (Equation (17)) and Σ0 = 20 g cm−2, which has sharp edges at finite semimajor axes (0.1 and 5 au). The radial profile of the disk eccentricity is still given by Equation (13), but now with the overall normalization e0 = 0.005. Note the sharp increase of the amplitude of Ad near the outer edge of the disk, accurately matched by our theory.

Standard image High-resolution image

An even more dramatic difference with the cases explored in Sections 4.1.14.1.2 is the behavior of Ad and Bd near the disk edges. Unlike the two previous cases, in which both coefficients remained finite everywhere at all radii, in a sharply truncated disk, Ad and Bd exhibit singularity as the disk edge is approached. This is very clearly seen at the outer8 boundary of the disk in Figure 7, where both theoretical curves and the results of orbit integrations exhibit divergent behavior. Unfortunately, we cannot probe this divergence in great detail numerically, as the orbits of test particles start crossing the edge of the disk.

At the same time, the test particle eccentricity ${e}_{p}^{{\rm{m}}}$ remains finite at the disk edges, even though both Ad and Bd are singular there. This is because both coefficients of the disturbing function given by Equation (3) diverge in similar fashion at the edge, so that ${e}_{\mathrm{forced}}$ remains finite there.

4.2. Singularities of ${e}_{p}^{{\rm{m}}}$

A common feature for all ${{\rm{\Sigma }}}_{d}$ distributions examined in Section 4.1 is the emergence of multiple singularities of ${e}_{p}^{{\rm{m}}}$. At these locations, ${e}_{p}^{{\rm{m}}}$ formally diverges, and the assumption ${e}_{p}\,\ll 1$ used in deriving our secular disturbing function (Equations (3)–(5)) breaks down. The way in which this happens is illustrated in Figure 8, where we compare the numerical and analytical results in the vicinity of one of the singularities (near ap = 1.54 au) in a disk with ${{\rm{\Sigma }}}_{d}(a)$ given by Equation (15); see Figure 5.

Figure 8.

Figure 8. Zoom-in on a part of Figure 5 in the vicinity of a secular singularity at $a\approx 1.5\,\mathrm{au}$. One can see how particle motion starts to deviate from our lowest-order secular theory as ep grows to values of order unity.

Standard image High-resolution image

One can see that as the theoretical singularity is approached, the behavior of ${\dot{\varpi }}_{\sec }$ starts to deviate from the prediction of Equations (45)–(47). As a result, ${\dot{\varpi }}_{\sec }$ goes through zero at a location slightly different from the one where Ad = 0. Note that even at the point where ${\dot{\varpi }}_{\sec }=0$, particle eccentricity remains finite (even though it reaches values close to 1). This means that Equation (11) is no longer valid when ${e}_{p}\sim 1$ and that additional, higher-order terms become important in addition to the lowest-order secular potential contribution (Equation (3)). This discrepancy could be at least partly ameliorated by including higher-order (in ep) terms in the calculation of Rd, as was done recently in Sefilian & Touma (2018) for the case of power-law disks.

As evidenced by Figures 5 and 6, ${e}_{p}^{{\rm{m}}}$ singularities often occur in the immediate vicinity of nulls of Bd. This results in a characteristic shape of these singularities, with ${e}_{p}^{{\rm{m}}}$ sharply dropping to zero in close proximity to the singularity. This leads to a dramatic difference in the eccentricities of particles with almost identical semimajor axes, resulting in their orbits crossing. Such locations thus provide a natural environment for particle collisions.

It is not clear why in some cases conditions Ad = 0 and Bd = 0 get realized at almost the same value of a. Inspection of the integrands in Equations (46) and (48) does not reveal an obvious reason for that to be the case. Interestingly, we find such "null-singularity" pairs only in the two disks without sharp edges (Section 4.1.14.1.2), for which the boundary terms ${A}_{d}^{\mathrm{edge}}$ and ${B}_{d}^{\mathrm{edge}}$ in Equation (45) vanish. The disk with sharply truncated ${{\rm{\Sigma }}}_{d}$ (see Section 4.1.3 and Figure 7) has singularities without neighboring nulls of Bd (in fact, Bd does not change sign in this disk). Whether this outcome is due to the nontrivial boundary terms in this disk model is not clear. These curious properties of ${e}_{p}^{{\rm{m}}}$ singularities deserve further investigation.

4.3. Sensitivity to the Disk Eccentricity ed

Next we examine how our secular theory fares against changes of the disk eccentricity. First, we explore the effect of uniformly varying just the amplitude of eccentricity (Section 4.3.1), keeping the radial profile of ed(a) the same. We then look at the effect of a different radial profile of ed(a) on the agreement between the theory and numerical calculations (Section 4.3.2).

4.3.1. Variation of the Disk Eccentricity Amplitude

In Figure 9, we plot ${e}_{p}^{{\rm{m}}}(a)$ for the disk model with ${{\rm{\Sigma }}}_{d}$ and ed given by Equations (15) and (13), where we set ${{\rm{\Sigma }}}_{0}\,\approx 1100$ g cm−2 but vary the eccentricity normalization e0 as indicated in the figure (which is very similar to Figure 5(a)).

Figure 9.

Figure 9. Agreement between our secular theory and direct orbit integrations as a function of the disk eccentricity amplitude. Shown is the amplitude of secular oscillations found in direct orbit integrations, as well as theoretical values of $2{e}_{p}^{\mathrm{forced}}$ shown as a function of both the distance from the star (horizontal axis) and the overall normalization e0 of the disk eccentricity ed (different curves). This calculation assumes surface density profile in the form of Equation (15) with ${{\rm{\Sigma }}}_{0}\approx 1100$ g cm−2 and the eccentricity profile given by Equation (13). Continuous curves of different colors (corresponding to different values of eccentricity normalization e0 in Equation (13), as shown in the panel) represent analytical results based on our secular theory. Dots of different colors display the corresponding numerical results. Other notation is the same as in Figure 5.

Standard image High-resolution image

One can see that our theory works surprisingly well in predicting ${e}_{p}^{{\rm{m}}}(a)$, even when particle eccentricity reaches values in excess of ep = 0.5, when one would naively expect the description based on the lowest-order secular disturbing function given by Equation (3) to break down. For the ed profile given by Equation (13), the maximum value of ed, reached at the inner boundary of the disk, is 2e0. Thus, even for disks with inner eccentricities reaching ${e}_{d}({a}_{1})=0.4$, our theory performs quite well.

The amplitude of eccentricity oscillations ${e}_{p}^{{\rm{m}}}(a)$ is just one metric by which the performance of our theory can be judged. Another obvious one is the free precession rate ${\dot{\varpi }}_{\sec }$. In Figure 10, we plot the ratio of ${\dot{\varpi }}_{\sec }$ to its analytical counterpart Ad as a function of the disk eccentricity normalization e0. In the framework of our secular calculation, the precession rate should not depend on e0 and must be equal to Ad.

Figure 10.

Figure 10. Relative deviation between the numerical (${\dot{\varpi }}_{\sec }$) and theoretical (Ad) precession rates ${\dot{\varpi }}_{\sec }/{A}_{d}$, plotted as a function of the eccentricity amplitude e0 of the linear disk eccentricity profile given by Equation (13). Different panels correspond to different semimajor axes of the test particle: (a) ap = 0.5, (b) 1.25, and (c) 2.5 au. The calculation assumes the ${{\rm{\Sigma }}}_{d}$ model given by Equation (15) with ${{\rm{\Sigma }}}_{0}\approx 1100$ g cm−2.

Standard image High-resolution image

Figure 10 shows that this is not really the case and ${\dot{\varpi }}_{\sec }$ does deviate from Ad when e0 is nonnegligible. The agreement between these two frequencies is somewhat less impressive than for ${e}_{p}^{{\rm{m}}}(a)$, with ${\dot{\varpi }}_{\sec }$ deviating from Ad by tens of percent already for e0 = 0.1. Note that the discrepancy between the numerical and analytical secular frequencies is a strong function of the semimajor axis ap, with the largest deviation occurring in the vicinity of the ${\dot{\varpi }}_{\sec }$ singularity (just interior of it) at 1.3 au. Thus, one should be careful when applying our lowest-order secular calculation to characterize the orbital evolution of test particles at certain locations in the disk with ${e}_{d}\gtrsim 0.1$.

4.3.2. Variation of the Disk Eccentricity Profile

Next we study the effect of changing the radial profile of the disk eccentricity ed(a). Figure 11 is the analog of Figure 9 but made for a disk with an eccentricity profile given by Equation (14).

Figure 11.

Figure 11. Same as Figure 9 but for a different eccentricity profile given by Equation (14), with different e0 corresponding to this profile. Comparing the results with Figure 9, we see that variation of the ed(a) profile changes the behavior of the maximum particle eccentricity, although the general topology of the curves remains roughly the same. The agreement between theory and direct orbit integrations is somewhat worse in this case compared to Figure 9.

Standard image High-resolution image

The first thing to note is that the radial profile of the theoretical ${e}_{p}^{{\rm{m}}}(a)$ still shows two clear singularities at the same radial locations as in Figure 9. This is easy to understand, since ${e}_{p}^{{\rm{m}}}$ diverges at radii where ${A}_{d}\to 0$, and Ad is independent of the disk eccentricity profile (it depends only on the ${{\rm{\Sigma }}}_{d}(a)$ profile). As a result, singularities of ${e}_{p}^{{\rm{m}}}$ stay at fixed locations even though ed(a) varies.

Second, the detailed shape of ${e}_{p}^{{\rm{m}}}(a)$ is notably different from that shown in Figures 5 and 9, despite the fact that the ${{\rm{\Sigma }}}_{d}(a)$ profile is the same in both cases. While previously, ${e}_{p}^{{\rm{m}}}$ was dropping to zero right next to the outer singularity at ${a}_{p}\approx 4.39\,\mathrm{au}$, in Figure 11, this happens near the inner singularity at $\approx 1.5\,\mathrm{au}$. Another null of ${e}_{p}^{{\rm{m}}}$, more distant from the singularity, has also swapped its location and now lies in the outer part of the disk in Figure 11.

Third, the agreement between the theory and direct orbit integrations is somewhat worse for the disk with an ed(a) profile (Equation (14)). This is especially noticeable to the right of the inner singularity. For example, the ${e}_{p}^{{\rm{m}}}$ differs by $\approx 50 \% $ from the theoretical expectation at ${a}_{p}=2\,\mathrm{au}$ for e0 = 1.75 (black curve, which corresponds to the maximum eccentricity in the disk of $\approx 0.1$). Away from the region $1.5\mbox{--}2\,\mathrm{au}$, the agreement is generally better, even though the particle eccentricity excited by the disk potential can be quite high, ${e}_{p}\gtrsim 0.1\mbox{--}0.2$.

We hypothesize that the reduced accuracy with which our secular theory predicts secular dynamics in the case of a disk with an ed(a) profile given by Equation (14) is caused by the fact that this disk features a rather nonaxisymmetric surface density distribution, ${\rm{\Sigma }}(r,\varphi )$. Indeed, Figures 3(c) and (d) show that the disk with ed(a) given by Equation (14) features a well-defined concentration of mass in the outer region near the apastra of the constituent trajectories. As a result, for certain values of a, surface density ${\rm{\Sigma }}({r}_{d},{\varphi }_{d})$ exhibits substantial variations along the eccentric trajectories of the mass elements comprising the disk; see Figure 3(d). According to Equation (2), this is only possible if $\zeta {e}_{d}$ is not small for these values of a, meaning that the assumption $| \zeta {e}_{d}| \ll 1$ underlying our expansion given by Equation (22) is not fulfilled, which explains the deviations seen for this particular disk model.

By contrast, the disk with the eccentricity profile given by Equation (13) shows a more axisymmetric distribution of ${\rm{\Sigma }}(r,\varphi )$ (see Figures 3(a) and (b)), meaning smaller $| \zeta {e}_{d}| $ and higher accuracy of the expansion given by Equation (22). Thus, we expect that our secular theory should perform better for disks without highly nonuniform azimuthal features in ${\rm{\Sigma }}({r}_{d},{\varphi }_{d})$.

4.4. Motion Outside the Disk

Our derivation of the secular disturbing function in Appendix A explicitly assumes that the particle orbits fully within the disk. In other words, in the case of a disk with ${{\rm{\Sigma }}}_{d}$ dropping to zero at some finite a1 and a2, the particle semimajor axis ap satisfies ${a}_{1}\lt {a}_{p}\lt {a}_{2}$, and its eccentric orbit does not cross the boundaries of the disk. Obviously, if the disk has no edges, as in, e.g., the case of a Gaussian ring given by Equation (16), then the particle orbit is always fully enclosed within the disk.

However, close examination of our derivation of Rd in Appendix A demonstrates that it should also apply equally well outside the disk with sharp edges, as long as the particle orbit does not cross the disk boundaries. For example, in the case of a particle orbiting outside the outer edge of the disk (${a}_{p}\gt {a}_{2}$), first, one drops the contribution of the outer disk (i.e., $a\gt {a}_{p}$, as there is no disk material there) when computing Rd, and, second, the integration in the inner disk runs not up to $\alpha =1$ but only up to $\alpha ={a}_{2}/{a}_{p}\lt 1$. As a result, the resultant Equations (45)–(49) for the coefficients Ad and Bd apply without modification, even when ${a}_{p}\lt {a}_{1}\lt {a}_{2}$ or ${a}_{1}\lt {a}_{2}\lt {a}_{p}$.

To verify this claim, in Figures 12 and 13, we show radial profiles of ${e}_{p}^{{\rm{m}}}$ and ${\dot{\varpi }}_{\sec }$, as well as those of $2{e}_{\mathrm{forced}}$, Ad, and Bd for particles orbiting inside and outside (correspondingly) the radial extent of the disk. Disk parameters are the same as in Figure 7; in particular, ${{\rm{\Sigma }}}_{d}$ is given by Equation (17) with sharp edges at ${a}_{1}=0.1\,\mathrm{au}$ and ${a}_{2}=5\,\mathrm{au}$.

Figure 12.

Figure 12. Characterization of the secular behavior in the potential of a disk with the same parameters as in Figure 7 but computed for a test particle orbiting inside the inner hole of the disk at 0.1 au (i.e., outside the main body of the disk with sharp edges). One can see that our theory still works very well even beyond the radial extent of the disk.

Standard image High-resolution image
Figure 13.

Figure 13. Same as Figure 12 but for a test particle orbiting outside the outer edge of the disk at 5 au.

Standard image High-resolution image

One can see that, as in the previous cases illustrated in Figures 57, there is excellent agreement between our theory and direct orbit integrations, as long as the disk and particle eccentricities are low. Both the theoretical and the numerical values of the particle eccentricity extrapolated to the disk edges match the corresponding values extrapolated from inside the disk; see Figure 7. At the same time, both Ad and Bd diverge as ${a}_{p}\to {a}_{1}-0$ and ${a}_{p}\to {a}_{2}+0$, mirroring the singularity of these coefficients identified previously (Section 2.2).

These results extend the applicability of our calculation of Rd to (coplanar) particles having arbitrary semimajor axes relative to the disk, as long as their orbits do not cross the edge of the disk where ${{\rm{\Sigma }}}_{d}$ discontinuously drops to zero. However, it can be shown that even this latter constraint can be removed, extending the applicability of our results even further. We do not dwell on this point here,9 deferring it to a future study.

5. Applications

The results presented in Section 4 demonstrate the validity and accuracy of the secular theory developed in this work in the low-e limit. This motivates us to use this theory to further explore several aspects of secular motion in the potential of an eccentric disk.

5.1. Edge Effects

Some astrophysical disks are known to have very sharp edges. For example, using Voyager 2 occultation data, Graps et al. (1995) demonstrated that the epsilon ring of Uranus has an ${{\rm{\Sigma }}}_{d}$ steeply dropping to zero at the ring boundaries. The edges of the Saturn rings are also known to be very sharp (Tiscareno 2013). Outside the solar system, eclipse data reveal the sharpness of the edge of the circumbinary ring around the young star KH 15D (Winn et al. 2006).

Our calculations predict that sharp edges result in the divergence of the secular potential of a razor-thin disk, leading to a divergence in the precession rate near these locations (Section 4.1.3). This outcome, resulting from the nonvanishing boundary terms, was previously pointed out in Silsbee & Rafikov (2015a) for truncated power-law disks, and now we generalize it for other models of eccentric disks with edges. This prediction is nicely confirmed by the direct integration of particle orbits in a particular disk model with sharp edges; see Figure 7 and Section 4.1.3.

The divergence of ${\dot{\varpi }}_{\sec }$ (or Ad) near the sharp edge of a zero-thickness disk can be traced to the fact that the (in-plane) gravitational acceleration in this region behaves as ${g}_{d}\propto \mathrm{ln}| {\rm{\Delta }}r| $, where ${\rm{\Delta }}r$ is the separation from the edge. Specializing to the case of an axisymmetric disk, one then finds the free precession rate (Fontana & Marzari 2016)

Equation (18)

(it should be ${g}_{{\rm{K}}}={{GM}}_{\star }/{r}^{2}$) near the edge at the leading order. The divergent behavior ${\dot{\varpi }}_{\sec }\propto {({\rm{\Delta }}r)}^{-1}$ coincides with the scaling of the boundary terms in the expression (Equation (47)) for ${A}_{d}^{\mathrm{edge}}$; see Section 2.2. Analogous singularities should arise at any radius in the disk where ${{\rm{\Sigma }}}_{d}(a)$ exhibits a discontinuity.

A disk with ${{\rm{\Sigma }}}_{d}$ dropping to zero smoothly over a narrow but finite range of a would not have Ad diverging there, as the boundary terms (${A}_{d}^{\mathrm{edge}}$ and ${B}_{d}^{\mathrm{edge}}$) vanish for smooth ${{\rm{\Sigma }}}_{d}$ profiles. Nevertheless, Ad still exhibits a nontrivial behavior in this region. This is illustrated in Figure 14, where we plot ${\dot{\varpi }}_{\sec }={A}_{d}(a)$ (top panel) for rings with several ${{\rm{\Sigma }}}_{d}(a)$ profiles10 of different degree of steepness near the boundaries (bottom panel). One can see that near the edge, ${\dot{\varpi }}_{\sec }$ exhibits very rapid variation, changing from large negative values just inside the edge to large positive values just outside the edge.

Figure 14.

Figure 14. Illustration of the divergence of the free precession rate Ad (panel (a)) near the sharp edges of the disk. A massive circular ring with three different profiles of ${{\rm{\Sigma }}}_{d}$ given by Equation (50) and shown in panel (b), varying in sharpness of ${{\rm{\Sigma }}}_{d}$ drop near the edges (regulated by the parameter ν, as shown in panel (a)), is considered. One can see that the more abrupt the variation of ${{\rm{\Sigma }}}_{d}$ at the edges, the higher the amplitude of Ad that is reached in these regions. In the limit of a discontinuity in ${{\rm{\Sigma }}}_{d}$, one would find ${A}_{d}\to \infty $, in agreement with our theory.

Standard image High-resolution image

This behavior can be understood by noticing that Equation (46) for ${A}_{d}^{\mathrm{bulk}}={A}_{d}$ contains a second derivative of the surface density ${{\rm{\Sigma }}}_{d}$ inside the integral. Very close to the boundary, where ${{\rm{\Sigma }}}_{d}$ is still high, ${{\rm{\Sigma }}}_{d}^{{\prime\prime} }(a)$ is large and negative. Due to the logarithmic singularity of ${b}_{1/2}^{(0)}(\alpha )$ as $\alpha \to 1$, the radial convolution in Equation (46) enhances the contribution of this region (with large ${{\rm{\Sigma }}}_{d}^{{\prime\prime} }\lt 0$) to Ad, resulting in rapid retrograde precession in this part of the disk. On the other hand, just slightly away from the boundary, where ${{\rm{\Sigma }}}_{d}$ is low, ${{\rm{\Sigma }}}_{d}^{{\prime\prime} }(a)$ is large and positive, driving fast prograde precession there.

As the sharpness of the ${{\rm{\Sigma }}}_{d}$ profile near the boundaries increases, so does the magnitude of ${{\rm{\Sigma }}}_{d}^{{\prime\prime} }(a)$. As a result, the amplitude of ${\dot{\varpi }}_{\sec }={A}_{d}$ on both sides of the edge grows. In the limit of the infinitely sharp ${{\rm{\Sigma }}}_{d}$ transition at the edge, the behavior of Ad becomes singular, with the change of sign at the edge, in agreement with the expectation of Equation (18). In our formalism, this is accounted for mathematically by the appearance of the boundary terms of Equation (47), whereas ${{\rm{\Sigma }}}_{d}^{{\prime\prime} }$ in the integral of Equation (46) remains finite.

In a disk with small but finite vertical thickness $h\ll a$, the behavior of the coefficients of Rd would be slightly different. In such a disk, the rise of Ad and Bd would saturate at a finite value ∝h−1 as the edge is approached. This transition is easy to understand by setting ${\rm{\Delta }}r\sim h$ in Equation (18).

The divergence of Ad near the sharp edges can have important implications for, e.g., the dynamics of planetary rings. Even though ${{\rm{\Sigma }}}_{d}$ remains finite at the edge, our results demonstrate that particle orbits should precess very rapidly and at a rapidly changing rate as their semimajor axes get closer to the edge. This naturally leads to particle orbit crossing resulting in their collisions, helping redistribute angular momentum near the disk edge (Chiang & Goldreich 2000; Chiang & Culter 2003).

5.2. Free Precession in Disks with Gaps

Disk gaps present another example of sharp surface density gradients. Gaps could form naturally as a result of gas clearing by the gravitational torques due to massive planets orbiting within the disk.

Ward (1981) looked at the effect of gaps on the free precession rate of test particles in axisymmetric disks with a power-law profile of ${{\rm{\Sigma }}}_{d}$, finding ${\dot{\varpi }}_{\sec }$ to be negative within the disk but positive within the gap. Ward (1981) modeled the gap by setting ${{\rm{\Sigma }}}_{d}$ to zero within a range of semimajor axes, which makes his result quite natural. Indeed, a disk with such a gap can be viewed as a combination of two disjoint disks with sharp edges. An object orbiting within a gap is thus moving exterior to an inner truncated disk and interior to an outer truncated disk. Our results in Sections 4.1.3, 4.4, and 5.1 demonstrate that interaction with each of the disks drives prograde free precession of such an object, with the combined effect being simply a linear superposition (also prograde) of the two contributions.

Using our theoretical formalism, we can explore how the results of Ward (1981) change for more realistic (i.e., less sharp) gap profiles. We consider a disk, which is a combination of a wide ring with flat ${{\rm{\Sigma }}}_{d}$ given by Equation (50), and a gap of width w and relative depth d (changing from d = 0 for no gap to d = 1 for ${{\rm{\Sigma }}}_{d}=0$ in the gap center), with the profile given by Equation (51). In Figure 15, we show how ${A}_{d}={\dot{\varpi }}_{\sec }$ varies as we change the gap depth. One can see that Ad inside the gap region, which is negative in the absence of a ${{\rm{\Sigma }}}_{d}$ depression, gradually decreases in magnitude, crosses zero, and becomes large and positive as the gap depth is increased. At the edges of the gap, Ad exhibits a nontrivial structure reminiscent of that seen in Figure 14. If our gap had a more abrupt drop of ${{\rm{\Sigma }}}_{d}$ at its edges, we would have converged to the case explored by Ward (1981).

Figure 15.

Figure 15. Variation of the free eccentricity precession rate Ad (panel (a)) in a disk with a gap. An underlying profile (Equation (50)) of ${{\rm{\Sigma }}}_{d}$ is modified by imposing a gap of different relative depth d (values are shown in panel (a)) according to the prescription of Equation (51), as illustrated in panel (b). The width of the gap ${\rm{w}}=1.5\,\mathrm{au}$ is fixed. Note the evolution of the precession rate in the gap from negative to positive as the gap depth is increased.

Standard image High-resolution image

Figure 16 looks at the effect of variation of the gap width, keeping its depth constant. One can see that narrower gaps have a more pronounced effect on free precession in the center of the gap. This is because the edges of such gaps are sharper and also closer to the orbit of a precessing object. In this case, the intuition developed in Figure 14 again suggests that Ad should be large and positive, as is observed in Figure 16. To summarize, the effect of a gap on free precession seems to be determined more by the sharpness of the ${{\rm{\Sigma }}}_{d}$ gradient than by the gap width or depth separately.

Figure 16.

Figure 16. Same as Figure 15 but showing the effect of the gap width w (indicated in panel (a)) on the behavior of Ad. The gap depth in Equation (51) is kept fixed at d = 0.6.

Standard image High-resolution image

5.3. Free Precession in Smooth Disks

The nontrivial behavior of ${\dot{\varpi }}_{\sec }$ discussed in Sections 5.15.2 is caused by the localized sharp features in ${{\rm{\Sigma }}}_{d}(a)$. However, our results in Sections 4.1.1 and 4.1.3, confirmed by direct orbit integration, clearly demonstrate that Ad can also easily change sign within a disk with a smooth distribution of ${{\rm{\Sigma }}}_{d}$. Since, according to Equation (9), ${A}_{d}={\dot{\varpi }}_{\sec }$ is the rate of precession of the free eccentricity vector, this implies that free precession of a test particle can be both prograde and retrograde, depending on its location within the disk. The possibility of a change of sign of free precession may appear rather surprising, given that the simple truncated power-law models usually employed to study the gravitational effect of thin disks tend to predict ${\dot{\varpi }}_{\sec }\lt 0$, i.e., retrograde free precession (Heppenheimer 1980; Ward 1981; Batygin et al. 2011). Prograde precession is (locally) possible in disks featuring gaps, i.e., sharp drops of ${{\rm{\Sigma }}}_{d}$ (Ward 1981; see also Section 5.2), but the disks explored in Sections 4.1.1 and 4.1.2 have rather smooth profiles of ${{\rm{\Sigma }}}_{d}(a)$.

On the other hand, in Silsbee & Rafikov (2015a), it was shown that even the truncated power-law disks can exhibit prograde free precession far from the disk edges for certain values of the density slope $p=-d\mathrm{ln}{{\rm{\Sigma }}}_{d}/d\mathrm{ln}a$, e.g., when $p\lt 0$, so that ${{\rm{\Sigma }}}_{d}$ grows with a. Given that near the edges (but still within the disk), one expects ${A}_{d}\lt 0$ (see Section 5.1), the direction of free precession must change sign at some location inside such a disk. In light of this observation, our current results simply show that the change of sign of free precession is a rather common phenomenon for arbitrary profiles of ${{\rm{\Sigma }}}_{d}$.

Unfortunately, predicting the direction of the free precession (i.e., the sign of Ad) at a specific semimajor axis a in a disk with a given profile ${{\rm{\Sigma }}}_{d}(a)$ is not an easy task. Even in a disk without sharp boundaries, when the boundary term ${A}_{d}^{\mathrm{edge}}$ vanishes and ${A}_{d}={A}_{d}^{\mathrm{bulk}}$ is represented by Equation (46), it is generally not straightforward to predict a priori the sign of this integral term. Indeed, a smooth, continuous ${{\rm{\Sigma }}}_{d}(a)$ gradually decaying to zero at finite (e.g., given by Equation (15)) or infinite (e.g., Equation (16)) boundaries would necessarily have ${{\rm{\Sigma }}}_{d}^{{\prime\prime} }(a)$ changing sign within the disk. Integration over a provides a nontrivial, nonlocal mapping between the global behavior of ${{\rm{\Sigma }}}_{d}^{{\prime\prime} }(a)$ and the value (and sign) of Ad.

As discussed in Sections 4.1 and 4.2, a change of sign of Ad inside the disk is also important because it gives rise to very high (formally divergent) values of the test particle eccentricity at radii where Ad = 0, as long as the disk eccentricity ed is nonzero. Previously, a similar effect—a localized singularity of ep—was identified in studies of planet formation within stellar binaries, both analytically (Rafikov 2013a, 2013b; Rafikov & Silsbee 2015a, 2015b; Silsbee & Rafikov 2015b, 2015a) and numerically (Meschiari 2014). Its origin could be traced to a secular resonance caused by the cancellation of prograde precession due to the binary companion and retrograde precession driven by the disk gravity in the presence of the nonzero binary torque (and disk torque, if the disk is eccentric). The importance of these singularities for planet formation in binaries lies in the fact that high values of ep lead to very energetic collisions between planetesimals, resulting in their destruction and hampering planetary accretion. In this work, we show that the same mechanism works even without a binary companion—disk torque is always present when ${e}_{d}\ne 0$ and results in divergent ep whenever ${A}_{d}\to 0$, which naturally happens in our disks.

As shown in Rafikov & Silsbee (2015b) and Silsbee & Rafikov (2015b), an opposite effect is also possible in disks in binaries—at certain locations, the eccentricities of test particles affected by the combined potential of a companion and an eccentric disk become very small, as a result of the cancellation of the corresponding torques.11 Again, in our case, this happens even without a binary companion—at the locations where Bd = 0, one naturally finds ${e}_{p}\to 0$, as a result of the cancellation of torques arising from different parts of the same disk. This can be seen, e.g., at $\approx 0.9\,\mathrm{au}$ in Figures 5 and 9 and at $\approx 4\,\mathrm{au}$ in Figure 11. At these locations, the relative velocities of colliding objects naturally become very small, promoting their agglomeration (rather than fragmentation) and growth.

6. Self-consistent Models of Self-gravitating, Rigidly Precessing Disks

We now use our results to assess the possibility of constructing self-consistent models of long-lived, self-gravitating eccentric disks orbiting massive central objects. Such models could describe, for example, the eccentric nuclear stellar disks around supermassive black holes observed in the centers of some galaxies.

We will assume that the surface density distribution in the disk is given by Equation (2), which essentially implies that for each semimajor axis, there is a single, unique value of the disk eccentricity ed(a), and that eccentric orbits of particles at all semimajor axes have the same orientation ${\varpi }_{d}$. This orientation cannot be fixed in time, as the disk's own non-Newtonian potential causes the orbits of individual particles to precess. In other words, ${\varpi }_{d}={\varpi }_{d}(t)$. Given the distribution of the surface density ${{\rm{\Sigma }}}_{d}(a)$ at the pericenter, the question we ask is whether one can determine the profile of ed(a) that the disk must have to precess coherently as a solid body at a constant rate ${\dot{\varpi }}_{d}$ (i.e., ${\varpi }_{d}(t)={\dot{\varpi }}_{d}t$). This arrangement, obviously, requires ${\dot{\varpi }}_{d}$ to be independent of a, since otherwise, differential precession would lead to disk twisting (apsidal misalignment of different parts of the disk), destroying its coherence.

Introducing for convenience the complex eccentricity ${E}_{p}={k}_{p}+{{ih}}_{p}={e}_{p}{e}^{i{\varpi }_{p}}$, we can combine Equation (7) into a single evolution equation for Ep:

Equation (19)

This equation is valid for any object, including the particles or fluid elements comprising the disk and contributing to its potential. Provided that solid-body precession is the only secular effect of the disk self-gravity, i.e., that the disk remains stationary in the frame precessing at the rate ${\dot{\varpi }}_{d}$, we look for solutions with ${\dot{e}}_{p}=0$ (i.e., ${e}_{p}(a,t)={e}_{p}(a)$) and ${\varpi }_{p}={\varpi }_{d}={\dot{\varpi }}_{d}t$. Also, by our assumption, at each point the disk eccentricity ed is the same as the eccentricity of its constituent particles passing through this point, meaning that we need to identify ep = ed. Plugging the ansatz ${E}_{p}={e}_{d}(a){e}^{i{\dot{\varpi }}_{d}t}$ into Equation (19), one arrives at the following master equation:

Equation (20)

This equation represents a self-consistent mathematical framework for determining the radial profile of ed that an eccentric disk needs to have to be able to precess as a solid body (without changing its shape) under the action of its own self-gravity. The precession rate ${\dot{\varpi }}_{d}$ plays the role of an eigenvalue of the problem. Equations (45)–(49) provide explicit dependencies of Ad and Bd on ${{\rm{\Sigma }}}_{d}$, ed, and a. The dependence is such that Equation (20) is an integral equation for ed. It is linear in ed and is essentially a Fredholm equation of the second type. Solving this integral equation, we obtain a set of eigenvalues (precession rates ${\dot{\varpi }}_{d}$), as well as the corresponding eigenfunctions (radial profiles of ed; the normalization of ed remains unconstrained because of the linear nature of Equation (20)).

This calculation uses the radial distribution of ${{\rm{\Sigma }}}_{d}$ as an input. As mentioned in Section 2, when ed(a) does not change in the course of evolution (or whenever ${e}_{d}\ll 1$), the radial profile of the surface density at periastron ${{\rm{\Sigma }}}_{d}(a)$ remains fixed in the course of secular evolution.

We defer the detailed exploration of Equation (20) for disks with different ${{\rm{\Sigma }}}_{d}$ profiles to future work. We will simply note here that some of our findings—the divergent behavior of Ad and Bd near the sharp disk edges, the changes of signs of Ad and Bd inside the disk, etc.—make finding the solutions of this equation rather nontrivial.

7. Discussion

Our results allow one to efficiently compute the effect of the gravity of an eccentric disk on the secular evolution of astrophysical objects coplanar with the disk. This work provides a natural generalization of the earlier calculation of Silsbee & Rafikov (2015a), in which the secular potential was computed for eccentric disks with ${{\rm{\Sigma }}}_{d}(a)$ and ed(a) given by power laws of a only. Even prior to that, Heppenheimer (1980) and Ward (1981) derived the disturbing function for a particular case of axisymmetric power-law disks. Our present results extend these calculations for arbitrary behaviors of ${{\rm{\Sigma }}}_{d}(a)$ and ed(a), allowing a much broader range of applications of our framework.

The accuracy with which our lowest-order theory works depends on both the behavior of the disk eccentricity and the particle eccentricity. The results of Section 4.3 demonstrate that our secular theory becomes inaccurate when ${e}_{d}\gtrsim 0.2$ or so; the actual value of ed at which this happens depends on both the location in the disk (see Figure 10) and the radial profile of ed (see Figure 11). Moreover, our results clearly show that even for nearly circular disks, the behavior of ep can be singular at certain locations (at least for ${{\rm{\Sigma }}}_{d}$ profiles considered in this work), resulting in very high ${e}_{p}\sim 1$, even at semimajor axes where ${e}_{d}(a)\ll 1$. As evidenced by Figure 8, this naturally leads to deviations from our theory at these locations, even for low-ed disks.

Real astrophysical disks have rather different values of ed. For example, the stellar disk in M31 has a substantial eccentricity, ${e}_{d}\sim 0.5$ (Tremaine 1995; Peiris & Tremaine 2003; Brown & Magorrian 2013). Our secular theory is unlikely to provide a good description of the secular dynamics in this system, as even its topology of the phase space should look very different from that corresponding to the disturbing function given by Equation (3). A higher-order extension of our approach, such as that presented recently in Sefilian & Touma (2018) to generalize the Silsbee & Rafikov (2015a) calculations to fourth order in ep, may provide a better tool for studying disks with nonnegligible ed.

On the other hand, disk eccentricity can be low enough in gaseous protoplanetary disks in binary stellar systems. Simulations demonstrate that under certain circumstances (moderately high binary eccentricity), a disk orbiting one of the binary components can have a rather low eccentricity, at the level of several percent (Marzari et al. 2009; Regály et al. 2011; Müller & Kley 2012). The same is true for circumbinary protoplanetary disks on au scales (Meschiari 2014). In such systems, our theory should be well suited for describing both the effect of the disk gravity on planetesimal motion and planet formation in binary systems (Rafikov & Silsbee 2015a, 2015b), as well as for studying the self-consistent dynamics of the gaseous component of the disk (Ogilvie 2001) driven by its self-gravity, pressure forces, etc.

Eccentric planetary rings typically have ${e}_{d}\sim {10}^{-3}\mbox{--}{10}^{-2}$ (Elliot & Nicholson 1984), which is also low enough for our theory to work well in describing the effect of the ring gravity on the secular motion of the adjacent objects (including the ring particles themselves; see Section 6).

However, when applying our framework to real planetary rings, a word of caution is in order. One of the underlying assumptions used in deriving the linearized Equation (22) is that $\zeta {e}_{d}={{de}}_{d}/d\mathrm{ln}a\ll 1$. At the same time, the epsilon ring of Uranus exhibits a change of eccentricity ${\rm{\Delta }}{e}_{d}\approx 7.11\times {10}^{-4}$ over a width ${\rm{\Delta }}a\approx 58.1$ km of the ring with a mean semimajor axis a = 51,149 km (French et al. 1991). We can evaluate $\zeta {e}_{d}\,\approx a{\rm{\Delta }}{e}_{d}/{\rm{\Delta }}a\approx 0.6$, which is certainly not small compared to unity. This means that Σ exhibits strong variation along the eccentric orbits of ring particles. As a result, both the expansion (Equation (22)) and our resultant framework (Equations (3) and (45)–(49)) may become inaccurate when applied to rings with $\zeta {e}_{d}\sim 1$. A similar issue was previously discussed in Section 4.3.2.

The cost involved in calculating secular evolution according to our approach is relatively low; it requires computation of only one-dimensional integrals involving ${{\rm{\Sigma }}}_{d}(a)$ and ed(a) (to obtain coefficients Ad and Bd). This is to be contrasted with the direct approach to computing secular potential, embodied by Equation (21), in which one first needs to carry out the two-dimensional integration over the full disk to obtain the potential at every point and then an additional integration to average it over the particle trajectory. Our procedure is clearly less numerically intensive and reproduces direct calculations very well in the low-e limit. This allowed us to use it for exploring different characteristics of secular motion in the disk potential, which we did in Sections 5.16.

The secular dynamics of self-gravitating disks is often explored by modeling them as collections of narrow adjacent rings coupled via the softened secular gravitational potential (Tremaine 2001; Hahn 2003; Touma et al. 2009). In this approach, the secular potential is approximated by the modified version of the classical Laplace–Lagrange disturbing function (Murray & Dermott 1999), which is regularized via softening to avoid the singularity that arises when the semimajor axes of the rings overlap. This procedure inevitably introduces an ad hoc softening parameter into the calculation, which inevitably leads to ambiguity of the results, since a physical justification for a particular choice of softening is not obvious.

Our approach, ascending to the framework developed in Heppenheimer (1980), Ward (1981), and Silsbee & Rafikov (2015a), does not suffer from this ambiguity. Even though the integrand of Equations (46) and (48) for Ad and Bd contains Laplace coefficients ${b}_{1/2}^{(0)}(\alpha )$ and ${b}_{1/2}^{(1)}(\alpha )$, which diverge as $\alpha \to 1$ (i.e., for particles with semimajor axes inside the radial extent of the disk), their singularity is weak (logarithmic in α). As a result, the integrals in Equations (46) and (48) are fully convergent without introducing any ad hoc softening. This makes our calculation of Rd more robust and self-consistent compared to some other treatments. The only possible divergences that remain in our case may arise at the sharp edges of the disk and are truly physical in their nature (see Section 5.1).

Our calculation of the disturbing function explicitly assumes the disk to have zero thickness and be coplanar with the particle orbit. It is easy to show that our results would not change qualitatively if the disk were to have a small vertical thickness h, as long as h is small compared to the semimajor axis a in the disk. Quantitative corrections to our results due to the nonzero h should be of order $O(h/a)$ at most. The only substantial difference will arise close to the edges of the disks with sharply truncated ${{\rm{\Sigma }}}_{d}$; see Section 5.1.

We also explicitly assumed the surface density of the disk Σ to have a specific form given by Equation (2), valid when the mass elements comprising the disk move around the central mass on eccentric orbits, which are apsidally aligned and have a unique value of eccentricity for a given semimajor axis. This assumption is natural for fluid disks, in which orbit crossing is impossible in the absence of shocks, as well as for highly collisional planetary rings. However, particulate disks of low optical depth (e.g., nuclear stellar disks) may have a more complicated form of Σ because orbit crossings are possible in such systems. We can naturally extend our approach to treat such cases by using the additive nature of gravity: if a more general disk structure can be represented as a superposition of multiple (sub)disks, each with Σ in the form of Equation (2) but different ${{\rm{\Sigma }}}_{d}(a)$, ed(a), and apsidal orientation, then the resultant disturbing function Rd will be a sum of individual contributions in the form of Equation (3) produced by each of the subdisks.

8. Summary

We explored the secular effect of a massive, razor-thin, eccentric, apsidally aligned disk on the motion of coplanar objects in the combined potential of such a disk (considered as a perturbation) and a central point mass. This problem is of great importance for many astrophysical disks (both gaseous and particulate), ranging from planetary rings to nuclear stellar disks in centers of galaxies harboring supermassive black holes. Our main findings are briefly summarized below.

  • 1.  
    We developed a general analytical framework for computing the secular disturbing function due to the gravity of an aligned eccentric disk. This disturbing function has the conventional Laplace–Lagrange form with coefficients that contain one-dimensional integrals over the semimajor axis only and do not involve softening of any form. It is valid for arbitrary radial profiles of the disk surface density ${{\rm{\Sigma }}}_{d}$ and eccentricity ed and works for coplanar objects orbiting both inside and outside the disk (for disks with edges).
  • 2.  
    We verified the accuracy of our analytical calculation using direct orbit integrations, finding excellent agreement in the limit of low eccentricity (both disk and particle) for all radial profiles of ${{\rm{\Sigma }}}_{d}$ and ed that we considered. Our framework is accurate at the $\lesssim 10 \% $ level for disk eccentricities ${e}_{d}\lesssim 0.1\mbox{--}0.2$. However, these figures strongly depend on both the location in the disk and the disk model, e.g., eccentricity profile ed(a).
  • 3.  
    Our calculations demonstrate that free precession of particle orbits in the potential of a smooth eccentric disk can naturally change from prograde to retrograde, and vice versa. Thus, the retrograde free precession previously found for certain types of disks (with a power-law dependence of density on radius) does not hold in general. At the locations where the free precession rate changes sign, forced particle eccentricities reach very high values (formally diverge).
  • 4.  
    Sharp features in the disk surface density distribution, such as disk edges or gaps, inevitably change the sense of free precession (e.g., prograde in the gaps, retrograde in the disk). In disks with sharp edges (where ${{\rm{\Sigma }}}_{d}$ discontinuously drops to zero), the free precession rate formally diverges at the edge.
  • 5.  
    Using our results, we formulated a general framework for computing the eccentricity profile of a disk (with a prescribed surface density profile ${{\rm{\Sigma }}}_{d}(a)$), which is required for it to be precessing rigidly due to its self-gravity alone, while maintaining stationary structure in the frame of precession.

In the future, we plan to extend our approach for understanding the secular dynamics of twisted (i.e., apsidally misaligned) eccentric disks (I. Davydenkova & R. R. Rafikov 2018, in preparation), as well as for exploring the inclination dynamics in the case of a non-coplanar (warped) disk. Exploration of the modal solutions (i.e., long-lived, rigidly precessing disk models discussed in Section 6) based on our secular framework is another logical application of our results.

We are grateful to Ryan Miranda, Antranik Sefilian, Kedron Silsbee, and Scott Tremaine for useful discussions. Financial support for this study has been provided by the NSF via grant AST-1409524 and NASA via grant 15-XRP15-2-0139.

Appendix A: Calculation of the Secular Potential of the Disk

In this section, we present a calculation of the disturbing function Rd due to an eccentric disk with a surface density in the form of Equation (2) in the low-eccentricity limit. We do not assume any specific forms for the functions ${{\rm{\Sigma }}}_{d}(a)$ and disk eccentricity ed(a), apart from requiring them to be twice differentiable. Following the recipe presented in Silsbee & Rafikov (2015b), we first write the disturbing function Rd due to the disk ${\mathbb{D}}$ as

Equation (21)

where brackets $\langle ...\rangle $ represent time averaging over planetesimal orbital motion, ${{\boldsymbol{r}}}_{p}$ is the instantaneous planetesimal radius vector (${r}_{p}=| {{\boldsymbol{r}}}_{p}| $, making an angle ${\varphi }_{p}$ with the planetesimal apsidal line), ${{\boldsymbol{r}}}_{d}$ is the radius vector of a surface element of the disk (${r}_{d}=| {{\boldsymbol{r}}}_{d}| $, making an angle ${\varphi }_{d}$ with the disk apsidal line), and $\theta ={\varphi }_{d}+{\varpi }_{d}-{\varphi }_{p}-{\varpi }_{p}$ is the angle between the two; see Figure 1.

Note that the indirect potential vanishes identically for a disk composed of mass elements moving on purely Keplerian orbits around the central mass. This is easy to understand based on Kepler's second law, which in our case means that the mass elements on opposite sides of a given elliptical trajectory (with respect to its focus) exert equal and opposite forces on the central mass. As a result, mass elements with orbits within a given semimajor axis interval da exert no net force on the central mass and do not cause its reflex motion, meaning that the indirect potential is zero.

We need to evaluate the integral of Equation (21) to second order in eccentricity, i.e., keeping the terms $O({e}_{p}^{2})$ and $O({e}_{p}{e}_{d})$. At the same time, we do not need to keep the terms not involving ep (e.g., $O({e}_{d})$ or $O({e}_{d}^{2})$), as those will vanish upon substitution into the Lagrange equations (Murray & Dermott 1999). Nor do we need to keep terms of higher order in ed than those listed above. Given that some of the variables entering Equation (2) are functions of ap and not rp, we need to express them through rp via ${a}_{p}\approx {r}_{p}(1+{e}_{d}({r}_{p})\cos {\varphi }_{d})+O({e}_{d}^{2})$. With this in mind, we can expand Equation (2) to lowest order in ${e}_{d}\ll 1$ and ${{de}}_{d}/d\mathrm{ln}{r}_{d}\ll 1$ as (Silsbee & Rafikov 2015b)

Equation (22)

As will become obvious later, the integral of Equation (21) over the last term in Equation (22) does not contribute to Rd at the required level of accuracy, so we drop it from consideration from now on.

Following Silsbee & Rafikov (2015b), we split Equation (21) into three integrals over the regions ${{\mathbb{D}}}_{c}$, ${{\mathbb{D}}}_{i}$, and ${{\mathbb{D}}}_{o}$, as shown in Figure 17. We define ${{\mathbb{D}}}_{c}$ to be a circular region with an inner radius equal to the periastron distance of the inner edge of the disk, ${r}_{d,\mathrm{in}}={a}_{1}(1-{e}_{d}({a}_{1}))$, and the outer radius equal to the periastron distance of the outer edge of the disk, ${r}_{d,\mathrm{out}}={a}_{2}(1-{e}_{d}({a}_{2}))$. Region ${{\mathbb{D}}}_{i}$ is the crescent region confined between the inner circle of ${{\mathbb{D}}}_{c}$ and the inner ellipse of the integration region ${\mathbb{D}};$ ${{\mathbb{D}}}_{o}$ is the crescent region confined between the outer circle of ${{\mathbb{D}}}_{c}$ and the outer ellipse of the integration region ${\mathbb{D}}$. In other words,

Equation (23)

Figure 17.

Figure 17. Decomposition of the disk into three distinct regions (${{\mathbb{D}}}_{c}$, ${{\mathbb{D}}}_{i}$, ${{\mathbb{D}}}_{o}$) over which potential is integrated. The gray area corresponds to the integration area, that is, the physical disk ${\mathbb{D}}={{\mathbb{D}}}_{o}+{{\mathbb{D}}}_{c}-{{\mathbb{D}}}_{i}$. The hatched region ${{\mathbb{D}}}_{c}$ is circular. See text for details.

Standard image High-resolution image

We now consider the three contributions to $R({\mathbb{D}})$ separately.

A.1. Calculation of the Part of the Disturbing Function Due to the Circular Region ${{\mathbb{D}}}_{c}$

We start by computing the integral over the circular annulus ${{\mathbb{D}}}_{c}$. When using Equation (22), it is natural to divide $R({{\mathbb{D}}}_{c})$ into two parts: one containing $\cos {\varphi }_{d}$ (we call it I1) and one independent of ${\varphi }_{d}$ (I2). We evaluate the ${\varphi }_{d}$-independent contribution first:

Equation (24)

This integral has a singularity at rp = rd, so we break it into two parts, ${r}_{p}\lt {r}_{d}$ and ${r}_{p}\gt {r}_{d}$, while also rewriting the inner integral as a Laplace coefficient ${b}_{1/2}^{(0)}$; see Equation (6):

Equation (25)

Now let us make a change of variables in both integrals so as to get rid of the time dependence in the Laplace coefficients, entering via rp. In the first integral, we set $\alpha ={r}_{d}/{r}_{p}\lt 1$, while in the second, we use $\alpha ={r}_{p}/{r}_{d}\lt 1$. We will also approximate ${r}_{d,\mathrm{in}}\approx {a}_{1}$ and ${r}_{d,\mathrm{out}}\approx {a}_{2}$, which, as we will see, introduces error of negligible order. Now we have

Equation (26)

Our goal is to get rid of triangular brackets, that is, to time average the integrals. The time dependence is hidden in rp and has the form ${r}_{p}(t)={a}_{p}(1-{e}_{p}\cos E(t))$, where ${e}_{p}\ll 1$ and E is the eccentric anomaly of the particle orbit. Since we are only interested in terms that are no more than quadratic in eccentricity (of both particle and disk), we next expand the integrals in a Taylor series over ${r}_{p}-{a}_{p}=-{a}_{p}{e}_{p}\cos E(t)$ up to second order and average over the planetesimal orbit using the relations $\langle \cos E\rangle =-{e}_{p}/2$, $\langle {\cos }^{2}E\rangle =1/2$. Note that the dependence on rp is also present in the limits of integration, which, upon this Taylor expansion, give rise to the nonintegral boundary terms. We omit the tedious process of writing out the Taylor expansion and subsequent averaging and present just the result,

Equation (27)

where we dropped the terms independent of ep. Such terms vanish when substituted into the Lagrange evolution equations. For ${I}_{1}^{(1)}$, we similarly have

Equation (28)

Now we turn to calculation of the integral contribution I2 containing ${\varphi }_{d}=\theta +v$, where we defined $v={\varphi }_{p}+{\varpi }_{p}-{\varpi }_{d}$. Let us introduce a new auxiliary function, $Q({r}_{d})=d[e({r}_{d}){{\rm{\Sigma }}}_{d}({r}_{d})]/{{dr}}_{d}.$ Then,

Equation (29)

where we expanded $\cos (\theta +v)$ and took into account that the term proportional to $\sin \theta $ evaluates to zero.

Next, we divide the integral over drd into two parts, one for ${r}_{d}\lt {r}_{p}$ and another for ${r}_{d}\gt {r}_{p}$, just like it was done for I1. We again approximate ${r}_{d,\mathrm{in}}\approx {a}_{1}$ and ${r}_{d,\mathrm{out}}\approx {a}_{2}$, as the linear corrections to these expressions would lead to a contribution, which is third order in eccentricity. We then again make a change of variables in both resulting integrals, setting $\alpha ={r}_{d}/{r}_{p}\lt 1$ in the first one and $\alpha ={r}_{p}/{r}_{d}\lt 1$ in the second. As a result, using the definition of the Laplace coefficient ${b}_{1/2}^{(1)}$, we obtain the following expression:

Equation (30)

There are two main differences in our subsequent expansion over rp − ap compared to the previous case. First, $Q({r}_{d})\sim O({e}_{d})$; thus, in the expansion, we only need to retain terms up to first order in ep. Second, now we have the time dependence of the integrand, also through $\cos v(t)=\cos ({\rm{\Delta }}\varpi +{\varphi }_{p}(t))$ $=\,\cos ({\rm{\Delta }}\varpi +E(t)+{e}_{p}\sin E(t))$, where we defined ${\rm{\Delta }}\varpi ={\varpi }_{p}-{\varpi }_{d}$ and used the relation ${\varphi }_{p}=E+{e}_{p}\sin E$. The easiest way to deal with the calculation is to consider rp as a function of ep, ${r}_{p}={a}_{p}(1-{e}_{p}\cos E),$ and to expand the integrals directly over ep. We start with ${I}_{2}^{(0)}$:

Equation (31)

We need the zeroth and first-order (in ${e}_{p}\ll 1$) terms of the Taylor expansion of this expression at ep = 0. We will omit the technicalities of writing out the expansion and present the result:

Equation (32)

Using the averages of the trigonometric functions of E,

Equation (33)

based on $\langle \sin E\rangle =0$, $\langle \sin E\cos E\rangle =0$, and $\langle {\sin }^{2}E\rangle =1/2$, the final result for ${I}_{2}^{(0)}$ becomes

Equation (34)

Following a similar procedure, for ${I}_{2}^{(1)}$, we have

Equation (35)

Equations (27), (28), (34), and (35) provide the desired contribution to the disturbing function from the circular annulus ${{\mathbb{D}}}_{c}$, since $R({{\mathbb{D}}}_{c})={I}_{1}^{(0)}+{I}_{1}^{(1)}+{I}_{2}^{(0)}+{I}_{2}^{(1)}$.

A.2. Calculation of the Part of the Disturbing Function Due to ${{\mathbb{D}}}_{i}$

Now let us turn to the inner crescent ${{\mathbb{D}}}_{i}$. As the eccentricity of our disk is small, the width of this crescent is already $O({e}_{d})$; thus, we only need to keep the terms up to first order in ep in our expansion. Following the recipe in Silsbee & Rafikov (2015b), we write

Equation (36)

where $\xi ({r}_{d})=\arccos (({a}_{1}-{r}_{d})/{e}_{1}{a}_{1})$ determines the azimuthal extent of the crescent for a given rd (Silsbee & Rafikov 2015b), and $\alpha ={r}_{d}/{r}_{p}$.

For the function in the inner integral, we can use the Fourier expansion,

Equation (37)

letting us calculate the inner integral as

Equation (38)

The next step is to remember that we only need to keep terms up to $O({e}_{d})$ in Equation (36) and to notice that the interval over which radial integration is performed is already $O({e}_{d})$. As a result, in Equation (36), we can set ${r}_{d}\approx {a}_{1}$ and ${{\rm{\Sigma }}}_{d}({r}_{d})\approx {{\rm{\Sigma }}}_{d}({a}_{1})$ and take them out of the integral. We can also set $\alpha \approx {a}_{1}/{r}_{p}$ in Equation (38) for the same reason. As a result, we find

Equation (39)

where we have used the fact that

Equation (40)

where ${\delta }_{{ij}}$ is the Kronecker delta.

Now we just need to expand Equation (39) up to first order in ep and then orbit average the resultant expression, which is straightforward using Equation (33)

Equation (41)

As a result, by dropping the first term not containing ep, one finally finds

Equation (42)

A.3. Calculation of the Part of the Disturbing Function due to ${{\mathbb{D}}}_{o}$

This case has only minor differences from the previous one, so we will omit the details:

Equation (43)

Thus, again dropping the first, ep-independent term, we eventually arrive at

Equation (44)

A.4. The Full Result

The full expression resulting from evaluating the integral in Equation (21) is given by ${R}_{d}={I}_{1}^{(0)}+{I}_{1}^{(1)}+{I}_{2}^{(0)}+{I}_{2}^{(1)}+R({{\mathbb{D}}}_{i})+R({{\mathbb{D}}}_{o})$ and can be written out explicitly using Equations (27), (28), (34), (35), (42), and (44). When doing this, we introduce two important modifications.

First, it is possible to combine the different terms corresponding to the outer and inner disks into a single expression using certain properties of the Laplace coefficients. For example, we can combine ${I}_{1}^{(0)}$ and ${I}_{1}^{(1)}$ by changing the variable $\alpha \to {\alpha }^{-1}$ in Equation (28) for ${I}_{1}^{(1)}$ and then manipulate the result using the fact that ${b}_{s}^{(j)}({\alpha }^{-1})={\alpha }^{2s}{b}_{s}^{(j)}(\alpha )$. This procedure turns ${I}_{1}^{(1)}$ in Equation (28) into the expression analogous to Equation (27) but with the limits of integration running from 1 to ${a}_{2}/{a}_{p}$. This allows us to combine it with ${I}_{1}^{(0)}$ to get an expression analogous to Equation (27) with integration running from ${a}_{1}/{a}_{p}\lt 1$ to ${a}_{2}/{a}_{p}\gt 1$ (for orbits within the disk, ${a}_{1}\lt {a}_{p}\lt {a}_{2}$). The integrand of this final expression has a weak (logarithmic) singularity at $\alpha =1$, but the integral itself is fully convergent. This procedure is then repeated to combine ${I}_{2}^{(0)}$ and ${I}_{2}^{(1)}$. In addition, differentiating the identity ${b}_{s}^{(j)}({\alpha }^{-1})={\alpha }^{2s}{b}_{s}^{(j)}(\alpha )$, we obtain the relation for ${b}_{s}^{(j)^{\prime} }({\alpha }^{-1})$, which allows us to also merge $R({{\mathbb{D}}}_{i})$ and $R({{\mathbb{D}}}_{o})$ into a single expression.

Second, in all these expressions, we switch from d/drp to derivatives with respect to the argument (denoted by prime), e.g., ${dQ}(\alpha {r}_{p})/{{dr}}_{p}=\alpha {Q}^{{\prime} }(a){| }_{a=\alpha {r}_{p}}$.

As a result of performing these steps, one finds that Rd can be written in the form of Equation (3) with the coefficients Ad and Bd given by the following expressions:

Equation (45)

Equation (46)

Equation (47)

Equation (48)

and

Equation (49)

where $F(a){| }_{a={a}_{1}}^{a={a}_{2}}\equiv F(a={a}_{2})-F(a={a}_{1})$. If we consider a disk model with power-law profiles of ${{\rm{\Sigma }}}_{d}(a)\propto {a}^{-p}$ and ${e}_{d}(a)\propto {a}^{-q}$, Equations (45)–(49) reduce to the corresponding formulae in Silsbee & Rafikov (2015b).

Appendix B: Gap

The ring profile used in Section 5.1 to illustrate the effects of sharp disk edges on the free precession rate is given by ${{\rm{\Sigma }}}_{d}(a)={{\rm{\Sigma }}}_{0}(a,\nu )$ with

Equation (50)

where $x=a/\mathrm{au}$ and ν is a parameter controlling the sharpness of the disk edges; see Figure 14.

The presence of a gap of radial width w and relative depth $d\lt 1$ is modeled in Section 5.2 using the surface density profile ${{\rm{\Sigma }}}_{d}(a)={{\rm{\Sigma }}}_{0}(a,2)\times \mathrm{Gap}(a,d,{\rm{w}})$, where

Equation (51)

Footnotes

  • The latter two are also significantly inclined with respect to the equatorial plane of the planet.

  • The assumption of apsidal alignment is retained here for simplicity. It is relaxed in a subsequent work of I. Davydenkova & R. R. Rafikov (2018, in preparation).

  • This requires ded/d ln a < (1−ed) (Ogilvie 2001; Statler 2001).

  • We also tried denser grids but have not found any difference in the outcomes.

  • Numerical issues prevent us from demonstrating the divergent behavior of Ad and Bd at the inner edge of the disk.

  • Verification of this statement by direct orbit integrations can be tricky because of the formal logarithmic divergence of the acceleration at the edge of the disk; see Section 5.1.

  • 10 

    The explicit expression for ${{\rm{\Sigma }}}_{d}(a)$ is given by Equation (50).

  • 11 

    This often requires a particular relative orientation of the disk and binary apsidal lines (Rafikov & Silsbee 2015b).

Please wait… references are loading.
10.3847/1538-4357/aad3ba