Articles

PLANET FORMATION IN STELLAR BINARIES. I. PLANETESIMAL DYNAMICS IN MASSIVE PROTOPLANETARY DISKS

and

Published 2014 December 23 © 2015. The American Astronomical Society. All rights reserved.
, , Citation Roman R. Rafikov and Kedron Silsbee 2015 ApJ 798 69 DOI 10.1088/0004-637X/798/2/69

0004-637X/798/2/69

ABSTRACT

About 20% of exoplanets discovered by radial velocity surveys reside in stellar binaries. To clarify their origin one has to understand the dynamics of planetesimals in protoplanetary disks within binaries. The standard description, accounting for only gas drag and gravity of the companion star, has been challenged recently, as the gravity of the protoplanetary disk was shown to play a crucial role in planetesimal dynamics. An added complication is the tendency of protoplanetary disks in binaries to become eccentric, giving rise to additional excitation of planetesimal eccentricity. Here, for the first time, we analytically explore the secular dynamics of planetesimals in binaries such as α Cen and γ Cep under the combined action of (1) gravity of the eccentric protoplanetary disk, (2) perturbations due to the (coplanar) eccentric companion, and (3) gas drag. We derive explicit solutions for the behavior of planetesimal eccentricity ep in non-precessing disks (and in precessing disks in certain limits). We obtain the analytical form of the distribution of the relative velocities of planetesimals, which is a key input for understanding their collisional evolution. Disk gravity strongly influences relative velocities and tends to push the sizes of planetesimals colliding with comparable objects at the highest speed to small values, ∼1 km. We also find that planetesimals in eccentric protoplanetary disks apsidally aligned with the binary orbit collide at lower relative velocities than in misaligned disks. Our results highlight the decisive role that disk gravity plays in planetesimal dynamics in binaries.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Results of radial velocity surveys demonstrate that ∼20% of exoplanets reside in stellar binaries (Desidera & Barbieri 2007). While most of these binaries have wide separation between stellar components (hundreds of AU), some of them are systems with relatively short binary periods of ∼100 yr. One of the best examples of such a binary is γ Cephei (Hatzes et al. 2003), which consists of two stars with masses of Mp = 1.6 M and Ms = 0.41 M with a semi-major axis of ab = 19 AU (orbital period Pb = 58 yr) and an eccentricity of eb = 0.41. The planet with the projected mass Mplsin i = 1.6 MJ is in orbit around the primary with a semi-major axis of apl ≈ 2 AU. Several more planetary systems within tight (ab ≈ 20 AU) binaries are known at present (Chauvin et al. 2011), including the terrestrial planet around our stellar neighbor α Cen (Dumusque et al. 2012; see Hatzes 2013).

For a long time, theorists struggled to explain the origin of planets in such S-type (according to Dvorak's 1982 classification) systems. The issue lies in the strong dynamical excitation to which any object in a binary is subjected. Gravitational perturbations due to the eccentric companion are expected to adversely affect planet formation already at the stage of planetesimal growth. As first shown by Heppenheimer (1978), companion perturbations drive planetesimal eccentricities to high values, easily approaching 0.1 at 2 AU from the primary. Planetesimals would then be colliding at relative speeds of a couple km s−1, which is much higher than the escape speed from the surface of even a 100 km object (about 100 m s−1). As a result, collisions should lead to planetesimal destruction rather than growth.

A number of possibilities have been explored to at least alleviate this problem. In particular, Marzari & Scholl (2000) studied the dissipative effects of gas drag as the means of damping the relative velocities of planetesimals. These authors have shown that for a circular disk in secular approximation, gas drag induces an alignment of planetesimal orbits such that the periapses of small objects strongly affected by gas drag tend to cluster around 3π/2 with respect to the binary apsidal line. This was originally thought (Marzari & Scholl 2000; Thébault et al. 2004) to assist planetesimal agglomeration since the relative velocities of colliding bodies are reduced by such orbital phasing. However, it was subsequently recognized (Thébault et al. 2008) that the reduction of the relative velocity caused by apsidal alignment is effective only for planetesimals of similar sizes. Objects of different sizes still collide at high speeds, which complicates their growth.

These studies have generally arrived at the same conclusion— the difficulty of planetesimal accretion—despite the different ways in which the gas disk and its interaction with planetesimals was treated. While the early calculations (Thébault et al. 2004, 2006, 2008, 2009; Thébault 2011) typically assumed disk properties to be described by some (semi-)analytic axisymmetric models, several recent studies followed the properties and evolution of gas disks in binaries using direct hydrodynamical simulations (Paardekooper et al. 2008; Kley et al. 2008; Marzari et al. 2009a, 2012; Regály et al. 2011; Müller & Kley 2012; Picogna & Marzari 2013). One of the most important aspects of disk physics that the latter captures is the development of non-axisymmetry in the surface density distribution of the gaseous disk. This non-axisymmetry emerges under the gravitational perturbation of the binary companion, predominantly in the form of non-zero eccentricity of the fluid trajectories (Marzari et al. 2012). Another phenomenon is the disk precession, which sometimes develops in simulations and has a subsequent effect on planetesimal dynamics.

An entirely different way of lowering planetesimal eccentricities in binaries has been pursued by Rafikov (2013, hereafter R13), who demonstrated that planetesimal eccentricities can be considerably lower than previously thought by properly accounting for the gravity of a massive axisymmetric gaseous disk in which planetesimals form. The non-Keplerian potential of the disk drives rapid precession of planetesimal orbits, suppressing the companion's driving of their eccentricity.

Note that massive protoplanetary disks must have been quite natural in γ Cep-like systems since all of the known systems (with the exception of α Cen) harbor Jupiter-like planets with Mplsin i = (1.6–4) MJ (Chauvin et al. 2011). It is natural to expect the parent disk mass to exceed the planet mass by at least a factor of several (this number is very uncertain but is believed to be ∼10 for the Minimum Mass Solar Nebula), which does not make an assumption of a (0.01–0.1) M disk unreasonable despite the fact that sub-millimeter surveys find very low fluxes of thermal dust emission in young binaries with semi-major axes of several tens of AU (Harris et al. 2012).

As discussed above, the assumption of a purely axisymmetric disk may be too simplistic since simulations indicate that protoplanetary disks in binaries often develop significant eccentricities. To that effect, Silsbee & Rafikov (2015, hereafter SR15) presented the first investigation of the secular excitation of planetesimal eccentricities by the simultaneous action of the gravitational perturbations due to both the eccentric gaseous disk and the companion star. They showed that the non-axisymmetric gravitational field of such a disk excites planetesimal eccentricity (in addition to the excitation produced by the companion) and usually does not allow it to drop below the disk eccentricity, which may be rather high as suggested by some simulations (Okazaki et al. 2002; Paardekooper et al. 2008; Kley & Nelson 2008). This would again suppress planetesimal growth. On the other hand, SR15 outlined several ways in which this issue can be alleviated, for example, if the gaseous disk is precessing rapidly or if its own self-gravity is capable of reducing disk eccentricity to low levels. At the same time, SR15 did not include gas drag in their calculations, eliminating the possibility of planetesimal apsidal alignment and their eccentricity damping by drag.

Our current work builds upon the results of previous studies by exploring planetesimal dynamics in disks coplanar with the binary under the combined effects of (1) gravitational perturbations due to the eccentric gaseous disk in which planetesimals are embedded, (2) the gravity of the eccentric companion, and (3) gas drag. While we do not model disks in binaries using hydrodynamical simulations, we still capture their main features, namely their non-axisymmetry and the possibility of precession. Our results are then used in a companion paper (Rafikov & Silsbee 2015, hereafter Paper II) to explore the details of planet formation in binaries.

We thereby extend the existing semi-analytical studies in which only gas drag (and not disk gravity) was accounted for (Thébault et al. 2004, 2006, 2008, 2009; Leiva et al. 2013). We also go beyond the works of R13 and SR15 in which gas drag was neglected and only the gravitational effects of the gaseous disk and binary companion were considered. In addition, we extend the study of Beaugé et al. (2010) devoted to exploring planetesimal dynamics in eccentric, precessing disks by accounting for the gravitational potential of such a disk.

This paper is structured as follows. We discuss our general setup in Section 2 and then derive equations for the evolution of orbital elements of planetesimals in eccentric disks in Section 3. A prescription for the gas-drag-induced eccentricity evolution is described in Section 4. Solutions to the equations of planetesimal dynamics in non-precessing and precessing disks are presented in Sections 5 and 6 (as well as Appendices B and C) correspondingly. The diversity of planetesimal dynamical behavior is discussed in Section 7. We derive the relative velocity distribution of objects of different sizes in Section 8. We provide an extensive discussion of our dynamical results and their applications in Section 10. We compare different approximations for treating planetesimal dynamics in Section 10.1, briefly discuss limitations of this work in Section 10.2, and summarize our main conclusions in Section 11. Finally, some of our analytical derivations use the local approximation for treating elliptical motion, which is reviewed in Appendix A.

2. PROBLEM SETUP

Our general setup is similar to that explored in SR15. We consider an elliptical disk around a primary star in a binary with semi-major axis ab, eccentricity eb, and component masses Mp (primary) and Ms (secondary). We define μ ≡ Ms/(Mp + Ms) and ν ≡ Ms/Mp. Binary, disk, and planetesimal orbits within it are assumed to be coplanar. This distinguishes our work from many other studies focused on the effects of Lidov–Kozai oscillations (Lidov 1962; Kozai 1962) on planetesimal dynamics in systems with inclined companions (Marzari et al. 2009a; Batygin et al. 2011; Zhao et al. 2012).

Non-axisymmetric disk structure is described via the non-zero disk eccentricity, which is a viable approximation given that simulations tend to show the prevalence of the m = 1 azimuthal harmonic of the disk shape distortion (Marzari et al. 2012). Fluid elements in a disk follow elliptical trajectories with an eccentricity eg(ad) that is a function of the semi-major axis ad of a particular ellipse. All of them have the primary star of the binary as a focus. For simplicity, we assume all fluid elliptical trajectories have aligned apsidal lines, uniquely determining disk orientation via a single parameter ϖd—the angle between the disk and binary apsidal lines. The latter is assumed to be fixed in space as the precession of the binary under the gravity of the disk is slower than all other processes. The assumption of apsidal alignment does not affect the qualitative features of planetesimal dynamics and can be easily relaxed using the results of Statler (2001).

Because gas moves on ellipses, its surface density generally varies along the trajectory (Statler 2001; Ogilvie 2001). To obtain the gas surface density Σ(rd, ϕd) at a point in a disk with polar coordinates (rd, ϕd), we specify the gas surface density at periastron for each elliptical trajectory Σp(ad) as a function of the semi-major axis of the corresponding ellipse ad. SR15 show how this and the knowledge of eg(ad) can be used to derive Σ(rd, ϕd) everywhere in the disk. In this work, following SR15, we assume a simple power-law dependence for both eg and Σp:

Equation (1)

where aout is the semi-major axis of the outermost elliptical trajectory of the disk, and Σ0 and e0 are the values of Σp and eg at aout. The gravity of the companion truncates the disk at this outer radius aout, which for eccentric binaries with eb = 0.4 is about (0.2–0.3)ab (Artymowicz & Lubow 1994; Regály et al. 2011). Unless stated otherwise, we will be using aout = 5 AU in this work.

In all calculations of this paper we will be using a disk model with p = 1 and q = −1. Some motivation for singling out these particular values of p and q for circumstellar disks in binaries has been provided in R13 and SR15.

The total disk mass $M_d\approx 2\pi \int ^{a_{\rm out}}_{a_{\rm in}}\Sigma _p(a_d)a_d da_d$ enclosed within aout can be used to express Σp as

Equation (2)

(the numerical estimate is for p = 1), where Md, −2Md/(10−2M), aout, 5aout/(5 AU) and ad, 1ad/AU. In Equation (2) we neglected disk ellipticity and assumed p < 2, so that most of the disk mass is concentrated in its outer part.

3. BASIC EQUATIONS

We are interested in the dynamics of planetesimals orbiting the primary within the disk and coplanar to it. We characterize their orbits with the semi-major axis ap, the eccentricity ep, and the apsidal angle (with regard to the binary apsidal line) ϖp. The orbital evolution of planetesimals is treated in a secular approximation, i.e., neglecting short-term gravitational perturbations (Murray & Dermott 1999). We also assume ep ≪ 1 as well as eg ≪ 1 and introduce, for convenience, the planetesimal eccentricity vector ep = (kp, hp) = ep(cos ϖp, sin ϖp).

In this work, we fully account for gravitational perturbations due to both the binary companion and the eccentric disk using the approach advanced in SR15. For the disk properties described by Equation (1), SR15 calculated an analytic expression for the planetesimal disturbing function accounting for the gravity of both the disk and the secondary. They then derived a set of Lagrange equations (see their Equations (16) and (17)) describing the evolution of ep under the influence of gravitational forces alone.

In addition, we take into account the effects of gas drag on the secular evolution of planetesimal eccentricity. Drag-induced dissipation also results in non-conservation of energy and evolution of ap. However, to zeroth order, we can neglect this since the radial inspiral of planetesimals usually occurs on much longer timescales than their eccentricity evolution (Adachi et al. 1976). As a result, we can concentrate on the behavior of ep at fixed ap and determine the relative velocities of planetesimals and their collisional outcomes.

Gas drag introduces additional terms in the eccentricity evolution equations of SR15, which we re-write in the following form:

Equation (3)

Equation (4)

Here A = Ab + Ad is the planetesimal precession rate. It is contributed both by the gravity of the secondary (Ab) and the disk (Ad), with

Equation (5)

where $n_p=(GM_p/a_p^3)^{1/2}$ is the planetesimal mean rate, Mp, 1Mp/M, ap, 1ap/AU, ab, 20ab/(20 AU), $b_s^{(j)}(\alpha)$ is the standard Laplace coefficient (Murray & Dermott 1999), αbap/ab and the approximation in Equation (5) works for αb ≪ 1. The disk contribution is

Equation (6)

where the numerical estimate is for p = 1 so that ψ1 = −0.5 (SR15). Dimensionless coefficients of order unity ψ1 and ψ2 (see Equation (8)) have been calculated in SR15 and are functions of the disk model and the distance of the planetesimal orbit from the disk edges. One can see that for reasonable assumptions about the disk mass (Md ∼ 10−2M), the planetesimal precession rate at 1 AU is dominated by the disk gravity.

Eccentricity excitation by the binary (Bb) and the disk (Bd) is described by

Equation (7)

Equation (8)

Equation (9)

with the latter explicitly depending on the local value of the disk eccentricity eg(ap).

Note that ϖd in Equations (3) and (4) is not necessarily constant—it can be an explicit function of time, allowing one to treat the case of a precessing disk.

The terms $\dot{h}_p^{\rm drag}$ and $\dot{k}_p^{\rm drag}$, absent in the original version of Equations (3) and (4) in SR15, represent the effect of gas drag on the eccentricity evolution; they are derived in Section 4. The main goal of this work is to see how their introduction affects planetesimal dynamics.

4. DRAG FORCE CALCULATION

Next we derive the expressions for the drag-induced eccentricity evolution terms $\dot{h}_p^{\rm drag}$ and $\dot{k}_p^{\rm drag}$ applicable to the case of an eccentric disk.

Because of our assumption of small eccentricities for both gas and planetesimals, it is reasonable to employ the local (or guiding center) approximation. This approach is often used in studies of planetesimal and galactic dynamics (Binney & Tremaine 2008) and forms a basis of the so-called Hill approximation (Hénon & Petit 1986; Hasegawa & Nakazawa 1990). Local approximation considers planetesimal motion in a local Cartesian xy reference system aligned with the radial and azimuthal directions, respectively. The main features of this approximation are reviewed in Appendix A. In particular, Equations (A8) describe how kp and hp evolve under the effect of external force F.

In our case, F is the drag force arising because of the motion of planetesimals with respect to gas. Adachi et al. (1976) gives the following expression for quadratic drag force appropriate for rapidly moving objects with sizes larger than the mean free path of gas molecules:

Equation (10)

where CD is a constant drag coefficient taken to be 0.5 throughout this paper, dp is the particle size, and ρg is the gas density. The relative particle–gas velocity vr is given by Equations (A5)–(A7) with relative particle–gas eccentricity components

Equation (11)

and eg = (kg, hg) = eg(cos ϖd, sin ϖg) being the local value of the gas eccentricity vector. Using these expressions, we obtain the force components Fx and Fy:

Equation (12)

Equation (13)

where mp is the planetesimal mass, and the relative velocity $v_r^a$ is given by Equation (A7). The prefactor D is given by

Equation (14)

with ρp being the particle bulk density and h = cs/np being the disk scale height (cs = (kTg/μ)1/2).

Now we plug the expressions for Fx, Fy into the first two Equations (A8) and then average them in time t over the planetesimal orbital period (this is the secular, i.e., time-averaged approximation). One can easily see that to get the result to lowest-order in er we do not need to keep the terms O(er, ed, ep) in the expression for ρg. As a result, we find

Equation (15)

Equation (16)

where $E(\sqrt{3}/2)\approx 1.211$ is a complete elliptic integral, and $e_r^2=k_r^2+h_r^2$.

We can rewrite Equations (15) and (16) in the following form:

Equation (17)

where the eccentricity damping time

Equation (18)

Here dp, 1dp/(1 km) and numerical estimate is for p = 1 and ρp = 3 g cm−3; in the case of quadratic drag law (10) τd depends on kp and hp through er; see Equations (11).

5. ECCENTRICITY EVOLUTION

Results of Section 4 allow us to understand the behavior of ep. For simplicity, we start by considering the case of a non-precessing disk, i.e., ϖd =const. Even in this case Equations (3) and (4) with the quadratic drag terms (15) and (16) cannot generally be solved analytically because of the τd dependence on er.

However, it can be easily shown that solutions of these equations inevitably converge to a steady-state form—the free eccentricity, which depends on initial conditions (R13; SR15), damps out and ep converges to the forced eccentricity vector (Beaugé et al 2010). This is illustrated in Figure 1 where we solve evolution equations numerically. It is clear that starting with arbitrary initial conditions and after initial (sometimes oscillatory) evolution, kp and hp do converge to the same steady state values (depending only on the disk parameters and planetesimal size), which are given by Equations (22) and (23) derived below. This point is additionally illustrated in Figure 2(a) where we plot the trajectory of ep as it evolves in hrkr coordinates. There one can clearly see ep converging to a fixed point solution, in an oscillatory fashion for large planetesimals, and exponentially for small objects, which rapidly couple to the gas disk.

Figure 1.

Figure 1. Time evolution of the components of the relative eccentricity vector er = (kr, hr) = epeg for planetesimals of two different sizes—dp = 0.3 km (red) and 8 km (green) starting with two different sets of initial conditions—er = (0.015, 0.01) (solid) and er = (− 0.03, −0.02) (dashed). Calculations are carried out for the parameters of the γ Cep system at 2 AU in a 0.001 M disk aligned with the binary; disk eccentricity at its outer edge aout = 5 AU is e0 = 0.05, and p = 1, q = −1 (see Equation (1)). After the short initial transient kr and hr converge to the forced values given by Equations (22) and (23).

Standard image High-resolution image
Figure 2.

Figure 2. (a) Planetesimal eccentricity evolution trajectories in krhr space in a non-precessing, aligned (ϖd = 0) disk for the four cases shown in Figure 1. One can see the convergence of trajectories starting at different ep to fixed point solutions (indicated by crosses), which depend on planetesimal radius dp. (b) The same for a disk precessing at a rate of $\dot{\varpi }_d=A$. Evolution trajectories converge to a limit cycle behavior in the precessing disk. See Section 6 for more details. The color and line type scheme is the same for both panels.

Standard image High-resolution image

Damping of the memory of initial conditions can also be demonstrated by solving Equations (3) and (4) analytically in the simplified but qualitatively similar case of a linear drag law, when τd is independent of hp and kp. Such a solution is presented in Appendix B for the general case of a precessing gaseous disk. The non-precessing disk solution is obtained by setting $\dot{\varpi }_d=0$. This clearly demonstrates the convergence of ep to a time-independent, forced value.

SR15 have demonstrated that in the absence gas drag, under the action of only the gravity of the disk and the companion star, the steady state (forced) eccentricity is given by

Equation (19)

Equation (20)

Equation (21)

where eb and ed are forced eccentricity vectors due to the secondary and disk gravity, respectively. Note that the accuracy of the analytical expression (20) for the binary contribution is known to worsen (beyond the ∼10% level) when ap/ab ≳ 0.1 (Thébault et al. 2006; Barnes & Greenberg 2006). More refined calculations of eb are possible (Veras & Armitage 2007; Giuppone et al. 2011) but for the purposes of this work it is sufficient to use Equation (20).

With the gas drag included, the behavior of ep changes. To determine the steady-state values of kp and hp and analyze their properties, we use the prescription (17), set to zero time derivatives in the left-hand sides of Equations (3) and (4) and solve the resulting algebraic system with respect to hp and kp. We find as a result

Equation (22)

Equation (23)

where kb, kd, hb, hd are defined in Equations (20) and (21). These asymptotic results are valid even if τd is a function of er —in that case they simply represent two implicit relations for hp and kp.

Solutions (22) and (23) can be re-written in vectorial form as

Equation (24)

Equation (25)

Equation (26)

where the phase shift ϕ is given by

Equation (27)

In the limit of vanishing drag, Aτd, one finds ϕ → π and the solution (24)–(26) reduces to the non-drag result with no free eccentricity (19)–(21); see SR15.

In the limit of strong drag (Aτd → 0) in a circular disk (i.e., eg = 0) and no disk gravity (i.e., Ad = Bd = 0) one finds hp/kp → −. This means that in this case planetesimal apsidal lines cluster around ϖp = 3π/2, in agreement with Marzari & Scholl (2000). Also, |ep| → Bbτd directly depends on planetesimal size, which implies that in this limit, planetesimals of different sizes collide with non-zero speeds even despite their apsidal alignment (Thébault et al. 2008).

Expressions (24)–(26) clearly show that ep can be split into two distinct components: a contribution ef, b due to the gravity of the binary and a contribution ef, b related to both the gravitational and gas drag effects of the disk. It is also clear that after reaching steady state, planetesimal orbits are generally aligned with neither the disk (ϖp ≠ ϖd) nor the binary (ϖp ≠ 0).

5.1. Relative Particle–Gas Eccentricity

In the case of quadratic drag (10), we can further analyze eccentricity behavior. Using Equations (22) and (23) we express relative particle–gas eccentricity as

Equation (28)

where we introduced a characteristic eccentricity $e_c=|{\bf e}_p^{\rm n/drag} - {\bf e}_g|$ given by

Equation (29)

Plugging this expression for er into Equation (18), one obtains the following bi-quadratic equation for (Aτd):

Equation (30)

where we have introduced a characteristic planetesimal size dc defined as

Equation (31)

All our subsequent results can be formulated completely in terms of ec and dp/dc, underscoring the significance of these variables. A detailed discussion of the characteristic values and general behavior of ec and dc is provided in Sections 7.1 and 7.3.

Solving Equation (30), one finds

Equation (32)

i.e., that |Aτd| is a function of dr/dc only.

Plugging Equation (32) into Equation (28), one also finds the general expression for the relative particle–gas eccentricity

Equation (33)

valid for arbitrary dp/dc.

We illustrate the behaviors of |Aτd| and er given by Equations (32) and (33) in Figure 3. This reveals the meaning of the characteristic size dc: objects with dpdc have |Aτd| ∼ 1, i.e., their stopping time due to gas drag is comparable to their orbital precession period and their relative eccentricity with respect to gas is erec.

Figure 3.

Figure 3. Dependence of er/ec and |Aτd| on planetesimal size dp/dc, given by Equations (28) and (32), respectively. Asymptotic scalings (34) and (38) are also indicated. For dpdc one finds |Aτd| ∼ 1 and erec.

Standard image High-resolution image

It is instructive to further explore general solutions (32), (33) valid for arbitrary dp/dc in the two limits covered next.

5.2. Small Objects, dpdc — Strong Drag (|Aτd|  ≲  1)

In the limit of strong gas drag we expect the damping time τd to be very short and |Aτd| ≪ 1, so that the gas-particle velocity differential is rapidly reduced to zero. According to Equation (32), this regime is valid for small objects with dpdc, when

Equation (34)

From Equation (28), the relative particle–gas eccentricity is

Equation (35)

to leading order in (Aτd).

Equations (22) and (23) become

Equation (36)

Equation (37)

Here brackets encompass the leading order subdominant terms, compared to the zeroth order terms outside brackets.

It is clear from these asymptotic expressions that in the case of strong drag, the eccentricity vector of planetesimals tends toward the eccentricity vector of the gas, epeg. It is only weakly sensitive to gravitational perturbations due to either the companion or the disk. Thus, to leading order, the value of the eccentricity vector is independent of particle size (which enters only through τd).

5.3. Big Objects, dpdc — Weak Drag (|Aτd|  ≳  1)

In the opposite limit of weak drag or long damping time |Aτd| ≫ 1 valid for large objects with dpdc, Equation (32) yields

Equation (38)

while the relative particle–gas eccentricity is

Equation (39)

see Equation (28). Thus, in the weak drag regime, er saturates at the value independent of the size of the object.

Equations (22) and (23) reduce in this limit to

Equation (40)

Equation (41)

Again, terms in brackets are subdominant compared to the leading terms (outside brackets).

This solution shows that in the limit of weak drag ${\bf e}_p\rightarrow {\bf e}_p^{\rm n/drag}$, i.e., the behavior of the particle eccentricity vector is determined predominantly by the gravitational effects of the secondary and the disk. Thus, ep is again almost independent of the particle size.

6. PRECESSING DISKS

So far we have assumed the orientation of the disk to be fixed in the binary frame. However, some simulations find disks in binaries to precess (e.g., Marzari et al 2009b; Müller & Kley 2012). We now study how planetesimal dynamics change in the case of a disk uniformly precessing at a constant rate of $\dot{\varpi }_d$. Figure 2(b) displays the evolution of ep for the same parameters as in panel (a) of that figure, but in a disk precessing at the rate $\dot{\varpi }_d=A$. One can see that the main difference compared to the non-precessing case is that in the long run ep converged to limit cycle behavior (Beaugé et al. 2010) rather than to a fixed point, as in panel (a). The sizes and shapes of the asymptotic limit cycles depend on both the planetesimal size dp and the disk precession rate $\dot{\varpi }_d$, as discussed in detail in Appendix C and shown in Figure 9. This certainly complicates planetesimal dynamics.

To gain additional insights, in Appendix B, we derive a full time-dependent solution for ep in a precessing disk for the case of linear gas drag, when τd is independent of the relative particle–gas eccentricity er. This solution fully accounts for the gravitational and gas drag effects of the precessing disk as well as for the gravity of the binary companion.

We use this solution as a basis for understanding planetesimal dynamics in a precessing disk in the more complicated but realistic case of quadratic gas drag. This regime, which does not admit a general analytical solution even for the long-term behavior, is explored in Appendix C. There we show that planetesimal dynamics with drag law (10) depend on the relative role played by the binary companion, as described in the next section.

6.1. Strong Binary Perturbation Case

The results in Appendices B and C show that whenever binary gravity dominates ep excitation and the condition

Equation (42)

is fulfilled, planetesimal dynamics proceed as if the disk were not precessing: neither the gas eccentricity eg nor the eccentricity driven by disk gravity ed, Equation (21), is significant compared to the forced eccentricity due to the binary eb = Bb/A (note that both the binary and disk gravity contribute to A).

In this case, ep is close to the relative planetesimal-gas eccentricity er and is approximately constant. As a result, the planetesimal orbit maintains a roughly fixed orientation with respect to the binary orbit and

Equation (43)

with kb defined by Equation (20). Planetesimal orbits are aligned with the binary (ϖp → 0) for |Aτd| → (weak drag), but in the case of strong drag, |Aτd| → 0, the planetesimal apsidal line points at ϖd = 270°, which agrees with Marzari & Scholl (2000) despite the disk precession.

Interestingly, even though gas eccentricity eg does not appear in these expressions (and neither does the precession rate $\dot{\varpi }_d$, at the lowest order), the effect of the gas drag is explicitly present via the non-trivial τd dependence. Thus, our precessing disk results obtained in the limit (42) apply equally well to planetesimal dynamics in a purely axisymmetric (eg = 0) gaseous disk, extending the results of R13 to the case of non-zero gas drag—note that A in Equations (43) and in the definition of kb is the full precession rate due to both the binary and the disk.

The value of er in the regime (42) is given by Equations (28) and (33) with dc and |Aτd| computed using eceb = |Bb/A| (i.e., Equation (29) in the limit Bd → 0, eg → 0); see Equations (31) and (32).

6.2. Weak Binary Perturbation Case

In the opposite case of weak driving of ep by the binary companion, we combine solutions (C5) and find the relative particle–gas eccentricity to be

Equation (44)

replacing Equation (28) in the case of the precessing disk. Here we defined the characteristic eccentricity as

Equation (45)

which, according to SR15, is the relative particle–gas forced eccentricity in the no drag (τd) and no binary (Bb → 0) case. As $\dot{\varpi }_d\rightarrow 0$, one finds $|e_c^{\rm pr}|\rightarrow e_c$ given by Equation (29) with kb = hb = 0; also, Equation (44) reduces to the non-precessing disk result (28).

Plugging this expression for er into Equation (18) one finds

Equation (46)

with a new characteristic planetesimal size

Equation (47)

These expressions are different from Equations (31) and (32) in that they use $A-\dot{\varpi }_d$ instead of A and $|e_c^{\rm pr}|$ instead of ec. It is then clear that whenever a precessing disk dominates the planetesimal dynamics Equation (33) also holds provided that we replace $d_c\rightarrow d_c^{\rm pr}$ and $e_c\rightarrow |e_c^{\rm pr}|$. The same is true for our asymptotic results on ep behavior presented in Sections 5.2 and 5.3 if we also take kb → 0.

In the limit $\dot{\varpi }_d\rightarrow 0$, the value of ef reduces to ef, d given by Equation (26), but when $|\dot{\varpi }_d|\gg |A|$, rapid disk precession suppresses excitation of planetesimal eccentricity by the disk gravity, i.e., the first term in Equation (45).

It is worth noting that the results of Appendix B for the case of linear drag suggest that neglecting binary gravity in the case of a precessing disk might require a condition different from the direct opposite to the constraint (42). Indeed, the asymptotic solution (B5) for the relative eccentricity of planetesimals in the case of weak drag ($\tau _{d,1},\tau _{d,2}\gg |A-\dot{\varpi }_d|^{-1}$) shows that the term proportional to kb can be neglected only when

Equation (48)

which is a more stringent criterion whenever $|\dot{\varpi }_d|\gg |A|$. The same constraint may be needed in the case of quadratic drag. However, in practice, one often finds $|\dot{\varpi }_d|\lesssim |A|$; see Paper II in which case Equation (48) is just the opposite of the condition (42).

7. DIVERSITY OF PLANETESIMAL DYNAMICS

Results of Section 5 demonstrate that the steady state value of the eccentricity vector ep is fully determined by just two key parameters—the characteristic eccentricity ec and the critical planetesimal size dc; see Equation (33). The eccentricity ec sets the overall scale of ep, while dc is the planetesimal size at which planetesimal coupling to gas changes from weak to strong. We now explore the behavior of these variables as a function of system parameters to elucidate some important features of planetesimal dynamics.

7.1. Behavior of ec

In Figures 4(a) and (b) we show ec computed for the γ Cep system at 2 AU—the semi-major axis of its planet—as a function of disk mass Md and eccentricity e0, for two disk orientations—aligned (ϖd = 0) and anti-aligned (ϖd = π) with the apsidal line of the binary.

Figure 4.

Figure 4. Map of the characteristic eccentricity ec as a function of e0 and Md (upper panels) for two different disk orientations—ϖd = 0 (a) and (b)—and as a function of ap and Md for two values of disk eccentricity e0 at aout (lower panels). The calculation is done for γ Cep system at ap = 2 AU (the observed semi-major axis of the planet). The dashed red line corresponds to Mpsin i for the observed planet in the γ Cephei system. The purple line is where |Ad| = |Ab| and the blue line is where |Bd| = |Bb|. See the text for details.

Standard image High-resolution image

One can immediately see a feature common to both panels—a narrow valley of high ec (white because of saturation at high ec) at almost constant Md. It appears because at this value of disk mass, Ad = −Ab and A = 0, giving rise to a secular resonance. According to Equations (19)–(21) and (29), ec gets driven to high values as A → 0. This resonance has been previously discussed in R13 and SR15.

Equations (5) and (6) predict that at a given distance from the primary ap, this resonance occurs for the disk mass

Equation (49)

where aout, 5aout/(5 AU), ap, 2ap/(2 AU), and ab, 20ab/(20 AU). This estimate agrees with Figures 4(a) and (b) for the γ Cep parameters and a disk with p = 1 and ψ1(p = 1) = −0.5 (SR15).

The existence of this resonance is independent of the relative disk–binary orientation because planetesimal precession rates Ab and Ad are determined by the axisymmetric components of the binary and disk gravitational potentials. For this reason, Md, A = 0 is the same for all disk orientations. To the right of the secular resonance, disk gravity dominates the planetesimal precession rate and suppresses ec if the disk eccentricity is small (R13).

At high disk eccentricity, typically e0 ≳ 0.05, this suppression vanishes because for large Md ≳ 10−3M, disk gravity starts to dominate ep excitation. This statement is true above the blue line |Bb| = |Bd| in Figures 4(a) and (b) (the origin of the low-ec band at small Md and high e0 in Figure 4(a) is discussed in Section 7.2). Further increase of the disk mass in this region does not affect ec because the planetesimal dynamics switch to the so-called DD regime (SR15) in which ep(ap) ≈ |ψ21|eg(ap), independent of Md. As a result, high eg leads to high ep.

In Figures 4(c) and (d) we explore the dependence of ec on the distance from the binary ap and Md for two different values of the disk eccentricity e0 = 0.1 and 0.01. Here we look only at an aligned disk case. Again, an obvious feature of these maps is the secular resonance around the blue dashed curve for |Ad| = |Ab|, where ec is very large and collisional growth is impossible. In Figure 4(d) there is also a "valley" of low ec to the right of the blue line |Bb| = |Bd|, whose origin is discussed in Section 7.2.

These maps make it clear that ec becomes independent of Md (at a given separation ap) when the disk mass becomes large enough. This is a direct consequence of the planetesimal dynamics switching into the DD regime (SR15) when both the eccentricity excitation and apsidal precession of planetesimals are dominated by the disk gravity with a negligible contribution from the binary companion. In the high-Md regime, ec decreases as ap decreases. This is a consequence of our adopted disk model, in which edap and the fact that eced in the DD regime.

7.2. Valley of Stability in Aligned Disks

Figures 4(a) and (b) show that irrespective of the disk orientation, ec is low for high Md ≳ 10−2M and small disk eccentricity, e0 ≲ 10−2. Outside this corner of phase space, ec is much higher, which makes planetesimal growth problematic there. At the same time, in the case of an aligned disk (ϖd = 0), low values of ec are also possible in a narrow "valley" stretching toward high e0 and low Md. Since this feature may have interesting implications for planet formation in binaries (see Paper II for details), we discuss its origin in more detail.

Equation (29) implies that in an aligned disk hg = hd = 0 so that

Equation (50)

For massive disks, to the right of the vertical |Ab| = |Ad| line in Figure 4(a), one can set AAd and relate it to Bd via Equations (6) and (8). As a result, Equation (50) becomes

Equation (51)

For the disk model considered here (p = 1, q = −1), one has ψ1 = −0.5, ψ2 = 1.5 and $1+2\psi _1\psi _2^{-1}=1/3$ so that ec ≈ |Ad|−1|Bb + Bd/3|. Also Bd > 0 while Bb is always negative; see Equations (7) and (8). Given that Bde0Md it is then obvious that one can make ec ≈ 0 by choosing e0Md such that |Bb| ≈ |Bd|/3. Thus, in the case of an aligned disk, a "valley" of low ec is described by the relation $e_0\propto M_d^{-1}$ as long as |Ad| ≳ |Ab| (i.e., for massive disks).

From this discussion we see that ec ≈ 0 for values of e0 and Md that are close to the curve

Equation (52)

on which |Bb| = |Bd|; see Equations (7) and (8) in which we took p = 1, q = −1. This relation is shown by the blue line in Figure 4 and is quite close to the valley of low ec.

Note that according to Equation (51), the value of ec can be lowered globally in a massive disk if its structure is such that $1+2\psi _1\psi _2^{-1}=0$. However, this is not the case for the disk model used in this work.

The situation is different for the low mass, aligned disks to the left of the |Ad| = |Ab| (blue dashed) line in Figure 4(a). Here AAb and Bb dominates over Bd for low enough Md at a fixed e0, which, in the terminology of SR15, corresponds to Case BB of planetesimal excitation. In this regime, Equation (50) shows that

Equation (53)

Our adopted radial scaling of eg in the form Equation (1) with q = −1 results in a particular value of

Equation (54)

for which ec → 0. This critical value of e0 is independent of Md explaining why the valley of low ec starts going almost horizontally for MdMd|A = 0 in Figure 4(a).

Moreover, $e_0|_{e_c\rightarrow 0}$ is also independent of ap, which means that ec → 0 globally when $e_0\rightarrow e_0|_{e_c\rightarrow 0}$ in parts of the disk where |Ab| ≳ |Ad| and |Bb| ≳ |Bd|. This is the reason why in the upper left corner of Figure 4(c) ec is considerably lower than in the same region of Figure 4(d), despite e0 being an order of magnitude higher in the former case. Indeed, according to Equation (54), e0 = 0.1 used in Figure 4(c) is very close to $e_0|_{e_c\rightarrow 0}$ for the adopted system parameters. As a result of this coincidence, ec is strongly suppressed in the BB regime in a rather eccentric (e0 = 0.1) disk.

A narrow region of low ec stretching along the blue curve |Bb| = |Bd| in Figures 4(c) and (d) is the same valley of stability, but now revealing itself in Mdap coordinates.1 It may lie inside (for low e0) as well as outside (for high e0) of the secular resonance. Note that in Figure 4(c) the |Ad| = |Ab| and |Bd| = |Bb| curves fall almost on top of each other, which is a coincidence caused by our choice of e0 = 0.1 in this case. Because of that, the valley of stability appears as a very narrow band of low ec just to the left of the |Bd| = |Bb| curve in this panel.

If the disk is not aligned with the binary orbit and ϖd is not small, then both hd and hg are nonzero and contribute to ec; see Equation (29). Moreover, for disks that are close to being anti-aligned with the binary, kb and kd have the same sign, eliminating the possibility of their mutual cancellation. As a result, the low-ec valley at high e0 and low Md disappears as long as |ϖd − ϖb| ≳ 10°.

To summarize, the valley of stability creates favorable conditions for locally lowering planetesimal velocity in aligned disks around some particular locations even in low-mass disks with Md ≲ 10−2M.

7.3. Behavior of dc

Next we discuss the behavior of the characteristic size dc at which planetesimals of similar (but not equal) mass collide at the highest relative velocity ∼ecvK. Equation (31) shows that for a given value of ec, critical size dc scales inversely with the planetesimal precession rate|A|. If planetesimal precession is dominated by the potential of the secondary, then A = Ab, and one finds

Equation (55)

where the numerical estimate is for a p = 1 disk and ρp = 3 g cm−3.

In the opposite case, when precession is dominated by the disk gravity and A = Ad, one obtains

Equation (56)

independent of the disk mass. It is obvious that in the disk-dominated case, dc is much smaller than in the binary-dominated case for ap ≲ 1AU, a fact predicted in R13.

This difference can be easily seen in Figure 8, where the situation depicted in panel (a) corresponds to the DD regime where Equation (56) applies. As a result, the planetesimal size for which the low-er "waist" in this figure is narrowest is around 1 km. In constrast, Figure 8(b) shows a situation in which the disk gravity has been turned off, so the dynamics are in the BB regime and Equation (55) applies. Not surprisingly, this pushes the characteristic dp at the narrowest point of the waist to be about 30 km.

Using this reasoning, one might expect the critical "dangerous" size dp at which erec for objects of comparable size to be smaller for more massive disks in which |Ad| ≫ |Ab|. However, this logic directly applies only if ec were kept the same. In reality, changing A also directly affects the value of ec; see Equation (29). Figure 5 shows that in practice, the behavior of dc largely reflects that of ec, with all the features of ec maps (e.g., valleys of low dc) present in dc maps as well. In particular, the valley of stability shows up prominently in Figures 5(a) and (d).

Figure 5.

Figure 5. Same as Figure 4 but for the behavior of the characteristic size dc given by Equation (31).

Standard image High-resolution image

The only noticeable difference with Figure 4 is the increase of dc with decreasing ap in the high-Md (DD) regime; see Figures 5(c) and (d), a behavior that is predicted by Equation (56). Also, in agreement with Equation (55), dc decreases with increasing ap in the outer disk for small Md (upper left in Figures 5(c) and (d)) even though ec varies weakly there. In this region, planetesimal dynamics are determined predominantly by the binary companion (BB regime of SR15) and Equation (55) applies.

8. DISTRIBUTION OF RELATIVE PLANETESIMAL VELOCITIES

Our next step is to study the behavior of the relative approach velocity v12 between planetesimals with sizes d1 and d2. It is this velocity that determines the outcome of their collision.

We now provide a calculation of the distribution df12/dv12 of v12 between the two planetesimal populations, one with the eccentricity vector ep(d1) and another with ep(d2). In previous sections, we have shown that after the initial transient period, when the free eccentricity damps out, the value of ep becomes time-independent and is uniquely determined by the planetesimal size. Then the only additional orbital parameter that can give rise to the variation of the relative velocity v12 is the difference in semi-major axes b12 between approaching particles; see Equation (A7) of Appendix A. Using Equations (A4) and (A7), this relation can be written as

Equation (57)

where a is the mean semi-major axis of both planetesimals, and the condition of close approach x12 = 0 is used. Note that in this expression we ignored the contribution of particle inclination to the velocity. This is a reasonable assumption since we expect eccentricity excitation in the binary plane to dominate over the out-of-plane excitation.

Ida et al. (1993) consider encounters between the two populations of objects with fixed eccentricity vectors e1 = (k1, h1) and e2 = (k2, h2). They derive the following expression for the flux of objects with an eccentricity e2 approaching a given object with an eccentricity e1 with random orbital phases, having a separation of their semi-major axes b12 in the range (b12, b12 + db12):

Equation (58)

Here e12 = [(h1h2)2 + (k1k2)2]1/2 is the relative eccentricity between the two particle populations, i12 is their relative inclination, and Σ2 is the surface density of objects with eccentricity e2.

Using Equation (57), we can express db in Equation (58) via dv12, resulting in a differential particle flux per unit v12

Equation (59)

We now express b12 via v12 using Equation (57) and introduce

Equation (60)

Then it is clear that vmin < v12 < vmax and we can re-write(59) as

Equation (61)

From this we find that the distribution of relative velocities df12/dv12 of different planetesimals normalized to unity is given by the following expression:

Equation (62)

Particle sizes enter into this expression only through e12 via Equations (60).

This distribution of relative velocities is shown in Figure 6. It diverges at both v = vmin and v = vmax, but the total particle flux is finite and given by

Equation (63)

With the distribution function (61), one finds that the mean relative velocity 〈v12〉 ≈ 0.81vmax = 0.81e12npap, while the rms velocity is given by $v_{\rm rms}=\langle v_{12}^2\rangle ^{1/2}=0.828 e_{12} n_p a_p$.

Figure 6.

Figure 6. Distribution of the relative approach velocity v12 of colliding planetesimals given by Equation (62). The relative velocity is normalized by its maximum value vmax = e12npr, where e12 is the relative eccentricity of the two planetesimals, which is a function of their sizes, see Sections 9.19.3. The minimum approach velocity is vmax/2.

Standard image High-resolution image

9. RELATIVE VELOCITY BETWEEN PLANETESIMALS

The results of the previous section clearly demonstrate that the relative velocity with which two planetesimals with sizes d1 and d2 approach each other prior to collision is determined by their relative eccentricity e12 = |ep(d1) − ep(d2)|. Using solutions (22) and (23), it is trivial to show that

Equation (64)

where τd, i ≡ τd(di), i = 1, 2. According to the results of Section 5, Aτd and, subsequently, e12, are functions of (1) the sizes of the colliding planetesimals d1, 2 and (2) the binary parameters and local disk properties, which set the values of both ec and dc; see Equations (29) and (31). We already explored the latter in Section 7 and now we turn our attention to understanding e1, 2(d1, d2).

In Figure 7 we map out e12(d1, d2) (as well as the relative velocity v12 = e1, 2vK) at the location of the planet ap = 2 AU in the γ Cephei system for different characteristics of the disk, for which a model (1) with p = 1, q = −1 is adopted. We vary disk mass Md, eccentricity at its outer edge e0, and its orientation with respect to the binary orbit ϖd, one at a time, keeping other disk parameters fixed. All panels clearly show several key invariant features.

Figure 7.

Figure 7. Relative approach velocity (right color bar) and relative eccentricity (left color bar) of planetesimals with sizes d1 and d2 experiencing close approach. The calculation is done for the γ Cephei system at 2 AU assuming an eccentric disk with p = 1, q = −1 and other disk parameters— Md, e0, ϖd—varying as indicated on the panels. Eccentricity and planetesimal size scales ec and dc in different panels can be inferred from Figures 4(a) and (b) and 5(a) and (b).

Standard image High-resolution image

First, there is a critical size of order dc, around d1 = d2 ∼ (0.1–1) km, at which maps exhibit a "waist," in which e12 is small for collisions of equal size bodies. Second, e12 becomes small for encounters between both the small bodies, with d1, d2dc, and for large objects with d1, d2dc. Third, e12 saturates at a value roughly independent of d1 or d2 for collisions of particles with very different sizes, i.e., when d1dcd2, and vice versa.

These gross features, as well as the variations of the overall velocity scale seen in these maps, are addressed below using the results of Section 5. Given that particles can be in different drag regimes—strong or weak—we will consider several possibilities separately.

9.1. Strong–Strong Encounters

When both planetesimals are in the strong drag regime, d1, d2dc, both |Aτd, 1| ≪ 1 and |Aτd, 2| ≪ 1. Then Equation (64) predicts that

Equation (65)

Equation (66)

where we used Equation (34) to express |Aτd| in terms of planetesimal sizes. Since d1, 2dc in the strong drag limit, one finds that $e_r^{\rm ss}\lesssim e_c$, which explains the low values of er in the lower left corner in the maps in Figure 7.

Physically, in this regime, the relative velocity of two planetesimals is considerably lower than their individual velocities because of the apsidal alignment of their orbits by gas drag, see Marzari & Scholl 2000) and similar magnitudes of ep.

9.2. Weak–Weak Encounters

When both planetesimals are in the weak drag regime |Aτd, 1| ≫ 1 and |Aτd, 2| ≫ 1, one finds using Equation (64) that

Equation (67)

Equation (68)

where Equation (38) has been used. Since d1, 2dc in the weak drag limit, one again finds that $e_r^{\rm ww}\lesssim e_c$, explaining the low relative eccentricity in the upper right corner in the maps in Figure 7.

In this case, apsidal alignment is again at work, lowering er compared to ep(d1), ep(d2). However, now it is caused by the disk+binary gravity, which affects planetesimals in the same way when they are weakly coupled to gas. This is because the gas damps the free eccentricity, but is not strong enough to significantly change the forced eccentricity.

9.3. Weak–Strong Encounters

When one of the planetesimals (e.g., of size d1) is in the strong drag regime, |Aτd, 1| ≪ 1, while the other is in the weak drag regime, |Aτd, 2| ≫ 1, Equation (64) shows that their relative eccentricity e12 is just

Equation (69)

One can see that e12 is roughly independent of the sizes of particles participating in an encounter.

9.4. Overall e12 Scale as a Function of Disk Parameters

The overall scale of e12 in each of the maps shown in Figure 7 is characterized by e12 in one of the high-velocity corners. According to Section 9.3, this scale is just ec, which allows us to use the results of Section 7 to understand how the typical e12 varies as we change the disk parameters.

A comparison of panels (a) and (b) of Figure 7 shows that disk mass Md plays an important role in setting e1, 2: planetesimals in low-mass disks (Md = 4 × 10−4M) collide with much higher speeds than in higher mass (Md = 2 × 10−2M) disk. This is because for the chosen value of e0 = 0.05 the low-mass disk is in the BB regime and the value of ec ≈ 0.05 is high; see Figure 4(a). Increasing Md as in panel (a) brings the disk into the DD regime and also close to the valley of stability. For that reason, in a higher mass disk with Md = 0.02 M, one gets a much lower ec ≈ 0.008.

Lowering e0 for a high-mass disk as in panel (c) reduces the relative velocity scale even more simply because for e0 = 0.007 the system gets even deeper into the valley of stability, where the corresponding ec ≈ 1.5 × 10−3; see Figure 4(a).

A comparison of panels (a) and (d) shows that changing disk orientation also strongly affects er: there is no valley of stability in the misaligned disk and the characteristic eccentricity scale becomes ec ≈ 0.014. As a result, particles in a misaligned disk collide at higher speeds than in the aligned disk.

10. DISCUSSION

Our work extends and complements existing results on planetesimal dynamics in binaries in several important ways.

First, for the first time, our solutions for ep in Section 5 simultaneously account for a number of key physical ingredients needed for a complete description of the secular dynamics of planetesimals in binaries: the gravity of both the eccentric disk and the eccentric companion as well as the gas drag, which causes orbital phasing of planetesimals and reduces their relative eccentricity in certain regimes.

Second, we provide a rigorous derivation of the equations of eccentricity evolution due to gas drag (15)–(18) in an eccentric disk. Previously, Adachi et al. (1976) derived analogous equations for the case of a circular disk, while Beaugé et al. (2010) proposed a set of empirical equations similar to Equations (15) and (16) but without proper calculation of the constant pre-factors.

Third, we derive an analytic expression (62) for the relative velocity distribution function df12/dv12 for locally homogeneous populations of objects with fixed eccentricity vectors, which is appropriate in the limit |ep| ≪ 1 in the presence of gas drag. We also provide an in-depth analysis of e12 behavior for objects of different sizes in systems with different parameters (Section 9). Previously the distribution of planetesimal encounter velocities has been explored only numerically by following a large number of trace particles in simulations of different kinds (Thébault et al. 2006, 2008, 2009; Paardekooper et al. 2008; Fragner et al. 2011). Thus, our derivation of df12/dvr represents an important analytical step in understanding planetesimal dynamics.

We now provide a more detailed comparison of our results with previous studies and discuss the limitations of this work.

10.1. Comparison of Different Dynamical Approximations

The main novelty of our study is the extension of the line of analytical investigation of disk gravity effects, which was started in R13 and SR15 for axisymmetric and non-axisymmetric disks, respectively, by including gas drag. Previous (semi-)analytical studies of planetesimal dynamics in binaries neglected the gravitational effect of the disk.

Our calculations account for both the precession of planetesimal orbits due to the axisymmetric part of the disk potential and the eccentricity excitation due to its non-axisymmetric component. Disk non-axisymmetry is modeled via its nonzero eccentricity, i.e., m = 1 distortion, which can be a function of radius. We expect this approximation to capture the key effect of the disk asymmetry, as higher-m distortions of the disk shape are relatively small (Marzari et al. 2012).

In Table 1 we summarize some existing (semi-)analytical treatments of planetesimal dynamics including this work (note that this list is not exhaustive), classified according to the physical ingredients that are taken into account. We primarily focus on studies of secular effects to put our work in proper context. Our current results cover all dynamical regimes listed in this table in appropriate limits. The majority of previous studies considered planetesimal dynamics in the presence of gas drag with only the direct binary gravitational perturbations taken into account (Marzari & Scholl 2000; Thébault et al. 2004, 2006, 2008, 2009; Paardekooper et al. 2008). As shown in SR15, this approximation is unwarranted as long as the disk mass Md ≳ 10−2M since then the disk potential dominates gravitational perturbation.

Table 1. Different Approximations for Planetesimal Dynamics in Binaries

Gravitational Effects W/O Gas Drag With Gas Drag
Included
Binary companion only 2, 3 4, 5, 6, 8, 9, 10
Axisymmetric disk 7    1
and binary companion    
Non-axisymmetric disk 8    1
and binary companion    

Notes. (1) This work; (2) Giuppone et al. 2011; (3) Heppenheimer 1978; (4) Marzari & Scholl 2000; (5) Paardekooper et al. 2008; (6) Beaugé et al. 2010; (7) Rafikov 2013; (8) Silsbee & Rafikov 2015; (9) Thébault et al. 2006; (10) Xie & Zhou 2008.

Download table as:  ASCIITypeset image

We also provide full analytical solutions for test particle dynamics in a general precessing or non-precessing disk without companion perturbation; see Equations (44)–(47). Previously, Beaugé et al. (2010) studied this regime for a precessing disk but did not account for the gravitational effect of the disk (i.e., only gas drag was taken into account). In Figure 8 we illustrate the differences in various descriptions of planetesimal dynamics. It shows the relative eccentricity as a function of planetesimal sizes d1 and d2 at 1 AU in an aligned disk of Md = 10−2Mp and e0 = 0.1 around a primary of γ Cephei in four different limits. Panel (a) presents a full calculation with all physical ingredients (gas drag, gravity of both the eccentric disk and the binary companion) accounted for using the solutions obtained in Section 5.

Figure 8.

Figure 8. Comparison of different approximations for describing planetesimal dynamics (indicated on panels), as reflected in the map of the relative eccentricity of planetesimals er of different sizes; see the text for details. Maps are drawn for an aligned disk in γ Cep at 1 AU (note the different semi-major axis compared to other figures).

Standard image High-resolution image

In panel (b) we show how things change if disk gravity is completely switched off by setting Ad = Bd = 0—an approximation common to a number of previous studies (Marzari & Scholl 2000; Thébault et al. 2004, 2006, 2008, 2009; Paardekooper et al. 2008; Beaugé et al. 2010; Leiva et al. 2013). One can see that without disk gravity, relative planetesimal velocities go up by a factor of several. Moreover, the "waist" between the two high-e12 regions in panel (b) is narrowest at d1d2 ∼ 102 km, which is considerably larger than in panel (a) where this happens for d ∼ 0.3 km objects. This difference is in complete agreement with Equations (55) and (56).

In panel (c) we account for the gravitational effect of a non-axisymmetric disk but neglect gas drag (τd), i.e., use Equations (19)–(21), as was done in SR15. In the absence of gas drag there is no apsidal alignment of planetesimal orbits and they approach each other at random phases. Also, er is independent of d1 and d2 (the size-dependent drag is absent) explaining uniform color in Figure 8(c). The absence of gas drag results in rather high relative velocities of planetesimals, making their survival in collisions problematic. Thus, apsidal alignment of planetesimal orbits and eccentricity suppression due to gas drag are very important for the proper description of their dynamics.

Finally, in panel (d) we retain only the axisymmetric component of the disk potential, neglecting the eccentricity excitation by the disk, i.e., Bd = 0 but Ad ≠ 0. In this limit, also neglecting gas drag (accounted for here), R13 predicted a dramatic lowering of ep. A comparison with panel (a) clearly shows this is not the case when gas drag is included, which can be understood by noting that ec in Equation (29) can significantly deviate from $|{\bf e}_p^{\rm n/drag}|$ because of the eg contribution. This is why lowering $|{\bf e}_p^{\rm n/drag}|$ by setting Bd = 0 and increasing |A| does not necessarily result in smaller ec, as expected in R13.

To summarize, simultaneously accounting for all the physical processes affecting planetesimals—gas drag, disk, and secondary gravity—is very important for understanding planetesimal growth. Omission of even a single physical ingredient can significantly affect the conclusions drawn from the dynamical calculations.

Previously, Kley & Nelson (2007) and Fragner et al. (2011) numerically explored planetesimal dynamics in gaseous disks that were evolved using direct hydrodynamical simulations. They accounted for the effect of disk gravity on planetesimal motion and at least some of their calculations assumed coplanarity of the disk and the binary. However, even though the setup of these studies is very similar to that of our present work, some subtle differences prevent direct comparison of their results. In particular, when estimating the relative velocities of planetesimals based on their orbit crossing, Kley & Nelson (2007) do not take into account the apsidal phasing of their orbits (Marzari & Scholl 2000), clearly obvious in their Figure 10. As a result they find very high relative speeds even between equal-size planetesimals, which we believe is an artifact of their neglect of apsidal phasing. Fragner et al. (2011) study the case of a circular binary in which apsidal phasing is naturally absent, resulting in high relative speeds of planetesimals. As a result, the applicability of calculations using circular binaries to understanding planetesimal dynamics in eccentric systems like γ Cep is not obvious.

10.2. Limitations of this Work

Finally, we discuss the limitations of our study. Some of them have to do with the adoption of secular, i.e., orbit-averaged, approximation. While averaging over the planetesimal orbit is justified because $n_p^{-1}$ is always much shorter than other periodicities (e.g., of planetesimal apsidal precession), when averaging over the longer binary period, one may overlook important dynamical features of the systems possessing very massive disks. Indeed, Equation (6) suggests that for Md ∼ 0.1 M, the planetesimal precession rate |Ad| becomes comparable to the binary angular frequency—np ≈ 0.1 yr−1 for γ Cephei. In these conditions, averaging over the latter is not justified and new effects, such as the possibility of evection resonance (Touma & Wisdom 1998) inside the disk, may also affect planetesimal dynamics.

Other effects omitted in our study, such as the density waves or higher-m contributions to the azimuthal mass distribution in the disk and short-term fluctuations of the disk potential, may also affect planetesimal dynamics. They may account for some of the difference between the results of this work, which uses a secular, time-averaged description of the disk and binary potential, and direct numerical studies of Kley & Nelson (2007) and Fragner et al. (2011). Planetesimal eccentricity can be additionally excited by the stochastic gravitational perturbations due to the turbulence in the disk. This issue has been previously investigated for disks around single stars (Ida et al. 2008; Yang et al. 2009, 2012) and for circumbinary disks (Meschiari 2012).

Coplanarity of the disk and the binary orbit is another restriction that can be easily eliminated in future studies. We believe that small but non-zero inclination (Xie & Zhou 2009) would not affect our solutions for the behavior of planetesimal eccentricity. However, as shown in Xie et al. (2010), such non-zero inclination has a strong effect on planetesimal collision rates.

There is also room for improvement within the framework of our model. Some approximations that we adopt, such as the power-law behavior of Σ(a) and ed(a) and constant2 ϖd(a), are dictated by our desire to obtain analytical solutions using the results of SR15 whenever possible. Also, we did not investigate the conditions under which our model (1) represents a steady-state solution for a fluid disk perturbed by a companion (Statler 2001). More refined semi-analytical or numerical calculations using improved disk models are certainly desirable but are unlikely to seriously affect our results and conclusions.

11. SUMMARY

We studied secular dynamics of planetesimals and explored prospects for planet formation around one of the components of an eccentric binary. We believe that our study includes most, if not all, of the important physical ingredients relevant for this problem—perturbations due to the binary, gas drag, and gravitational effects of an eccentric disk. This is the first time planetesimal dynamics in binaries have been studied analytically in such generality. The analytical nature of our solutions for planetesimal dynamical variables allowed us to explore their dependence on system parameters in great detail.

Our main results can be summarized as follows.

  • 1.  
    We find that under the action of gas drag as well as the gravitational effects of the binary companion and the eccentric disk, the planetesimal eccentricity vector ep converges to a constant value depending on the planetesimal size and the disk and binary properties. We obtained complete analytical solutions for ep in the case of a non-precessing disk and analyzed them in detail, extending the results of previous studies.
  • 2.  
    We showed that relative particle–gas (Equation (33)) and particle–particle velocities can be expressed as simple functions of only two key parameters—the characteristic eccentricity ec and planetesimal size d/dc in units of characteristic size dc, given by Equations (29) and (31). The behavior of these variables has been explored in detail in Section 7.
  • 3.  
    We show that in massive disks containing enough gas to form giant planets (Md ≳ 10−2M), planetesimal dynamics are always in the regime where the apsidal precession of planetesimal orbits is dominated by disk gravity, i.e., in the DB or DD regimes in the classification of SR15. Significantly eccentric (e0 ≳ 10−3) disks also dominate the eccentricity excitation of planetesimals by their gravity (DD regime). This emphasizes the key role of the disk gravity in relation to planet formation in binaries.
  • 4.  
    We derive the explicit form of the relative velocity distribution between the populations of planetesimals with different sizes and show that it depends only on the relative eccentricity e12 of the approaching objects.
  • 5.  
    In disks aligned with the binary, planetesimals collide with lower velocities than in misaligned disks. Thus, planetesimal growth favors disk–binary apsidal alignment.
  • 6.  
    We also present analytical results for the dynamics of planetesimals in precessing disks in certain limits.

Our results will be used in Paper II to understand planet formation in small separation binaries, such as γ Cep and α Cen. They can also be used to understand the circumbinary planet formation.

We are grateful to Jihad Touma for useful discussions.

APPENDIX A: LOCAL APPROXIMATION

Here we review local (or guiding center) approximation, which is often used in studies of planetesimal and galactic dynamics (Binney & Tremaine 2008) and forms the basis of the so-called Hill approximation (Hénon & Petit 1986; Hasegawa & Nakazawa 1990). In this approach, eccentric motion of a planetesimal is considered in a locally Cartesian frame (xp, yp), with ex, ey pointing in the radial and horizontal directions, correspondingly. The origin of this frame is in circular Keplerian motion at some characteristic semi-major axis a0, which is close to the planetesimal semi-major axis ap, so that bp ≡ |apa0| ≪ ap. Equations of motion for a particle of mass mp subject to external force F = (Fx, Fy) can be reduced to

Equation (A1)

Provided that ep ≪ 1, one can represent planetesimal motion unperturbed by external forces as

Equation (A2)

where ψp is a constant and ep = (kp, hp). This is an exact solution of Equations (A1) with F = 0 and is a superposition of linear shear and epicyclic motion.

Assuming that fluid in a gaseous disk also moves on eccentric Keplerian orbits, motion of the gas can be represented by analogous equations

Equation (A3)

Relative motion of a particular fluid element and a particle is described using relative coordinates xr = xpxg, yr = ypyg. According to Equations (A2)

Equation (A4)

where krkpkg, hrhphg are the components of the relative eccentricity vector, brbpbg is the semi-major axis separation between the particle and fluid element, and ψr ≡ ψp − ψg. We have also used the fact that aga0ap and switched from a0 to ap.

Velocity of Keplerian motion in the local approximation is obtained by differentiating Equations (A4) with respect to time. In particular, relative particle–gas velocity is given by

Equation (A5)

Analogous formulae apply to the relative motion of two planetesimals with sizes d1 and d2, with the replacement ere12, brb12, (xr, yr) → (x12, y12), and so on. In particular, Equations (A4) shows that two objects with |b12| < ape12 can experience close approaches. When this happens, x12 = y12 = 0 and b12 can be eliminated from Equations (A5) giving

Equation (A6)

(here e12 = (k12, h12)) so that the relative approach velocity (i.e., the velocity unaffected by the mutual gravitational attraction of particles) is

Equation (A7)

Whenever a particle is affected by forces other than the stellar gravity, i.e., F ≠ 0, solutions (A2) are no longer strictly valid. However, one can still represent particle motion via these solutions, assuming that orbital elements osculate, i.e., evolve, in time. Hasegawa & Nakazawa (1990) derived equations for the orbital element evolution, in particular

Equation (A8)

For a given force expression F, these equations, after averaging over the orbital period, represent the extra terms entering the Equations (3) and (4).

APPENDIX B: PLANETESIMAL ECCENTRICITY IN A PRECESSING DISK IN THE CASE OF LINEAR DRAG

Here we derive the full time-dependent solution for planetesimal eccentricity starting with arbitrary initial conditions and assuming that the gas drag is linear, i.e., τd in Equation (17) is a constant independent of ep. We also include a possibility of uniform disk precession so that $\varpi _d(t)=\dot{\varpi }_d t+\varpi _{d0}$. Then Equations (3) and (4) represent a linear system of equations which can be trivially solved to give

Equation (B1)

where the first term represents the free eccentricity, with efree and ϖ0 being constant, while the second is the forced eccentricity ef = (kf, hf) = ef, b + ef, d, where ef, b is given by Equation (25) and

Equation (B2)

In the limit of slow precession $|\dot{\varpi }_d|\ll |A|$ one finds that ef is given by expressions (24)–(27). Generally, the relative planetesimal–gas eccentricity er = efeg is

Equation (B3)

The first forced term ef, b results from excitation by the binary companion. It is constant in time and is independent of $\dot{\varpi }_d$. The second term is induced by the disk via both its gravitational potential and gas drag. This contribution to ef circulates at the disk precession frequency $\dot{\varpi }_d$ and its amplitude is sensitive to $\dot{\varpi }_d$.

Independent of the initial conditions (i.e., the values of efree and ϖ0), the free eccentricity contribution damps out on a characteristic timescale τd. As a result, in the long run, ep inevitably converges to ef.

In the limit of strong gas drag, τd → 0, one finds that efeg as expected, since drag is strong enough to align planetesimal orbits with fluid trajectories. In this limit, the relative eccentricity between planetesimals of different sizes having different damping times τd, 1 and τd, 2 is

Equation (B4)

where $e_c^{\rm pr}$ is given by Equation (45).

In the opposite extreme, τd (weak drag), one finds ϕ → π and ef reduces to the forced eccentricity value (with disk precession) obtained in SR15. The relative velocity becomes

Equation (B5)

Note that in this expression, kb is multiplied by a factor different from that in Equation (B4). However, it is clear that in both limiting cases, e12e1, e2, i.e., the relative planetesimal eccentricity is much less than the individual eccentricities e1 and e2, a result that remains valid in a precessing disk.

APPENDIX C: PLANETESIMAL ECCENTRICITY IN A PRECESSING DISK IN THE CASE OF QUADRATIC DRAG

In the case of quadratic drag (10), Figure 2(b) clearly shows the phenomenon of ep convergence to a quasi-stationary limit cycle behavior, similar to the results of Appendix B. This behavior is further illustrated in Figure 9, where we show the dependence of the limit cycles on planetesimal size dp and disk precession rate $\dot{\varpi }_d$. Because of the nonlinear drag law, the shapes of the limit cycles in general deviate from ellipses.

Figure 9.

Figure 9. Limit cycles to which the relative gas–planetesimal eccentricity vector er converges in precessing disks. Panel (a) shows evolution of the limit cycles as a function of planetesimal size dp, while in panel (b) we vary the disk precession rate $\dot{\varpi }_d$. Calculations have been performed at ap = 2.5 AU in a standard aligned disk with Md = 10−3Mp, e0 = 0.04, aout = 5 AU in γ Cep system. These parameters place planetesimal dynamics in the strong binary perturbation regime; see Section 6.1. Crosses mark the centers of the limit cycles computed according to Equation (43). Note the evolution of the positions and shapes of the limit cycles as dp and $\dot{\varpi }_d$ are varied.

Standard image High-resolution image

Nevertheless, their gross features still can be understood in our linear solution (B3). In particular, limit cycles are not centered on (kr, hr) = 0 because of the binary companion perturbations, i.e., non-zero ef, b varying as dp (and τd) change. The amplitude of the limit cycles goes down for smaller dp because τd is also smaller, which, according to Equation (B3), reduces the oscillating contribution to er. As we vary $\dot{\varpi }_d$ in Figure 9(b), the binary contribution stays unchanged and all limit cycles stay centered on the same point in hr-kr space.

Their sizes vary with $\dot{\varpi }_d$ as predicted by Equation (B3). They shrink at high $|\dot{\varpi }_d|\sim |A|$, in agreement with Equation (B3). For slow precession, $|\dot{\varpi }_d|\ll |A|$ limit cycles converge to the trajectory for the non-precessing disk solution (24) in which ϖd is set to vary as $\varpi _d(t)=\dot{\varpi }_d t+\varpi _{d0}$. Note that such a convergence to solution (24) is obvious only in the case of $|\dot{\varpi }_d \tau _d|\ll 1$, i.e., when gas drag allows ep to quickly re-adjust to a new "quasi-static" solution as ϖd changes. This is the case shown in Figure 9(b). In the opposite case of $|\dot{\varpi }_d \tau _d|\gg 1$ (and $|\dot{\varpi }_d|\ll |A|$), this convergence is not obvious as the disk precession constantly drives free eccentricity, while the gas drag is not strong enough to quickly damp it. We leave detailed exploration of such topics to a future study.

Now, let us rewrite Equations (3) and (4) in terms of the relative particle–gas eccentricity components kr and hr:

Equation (C1)

Equation (C2)

with τd given by Equation (18) and dependent upon er.

The explicit time dependence of the last terms in these nonlinear equations precludes us from finding their general analytical solutions even in the case of the limit-cycle behavior. However, we can still obtain analytical results for planetesimal eccentricity in the two limiting cases, reviewed next.

First, one can assume that the binary companion dominates eccentricity forcing, which implies that the condition (42) is fulfilled. Then one can drop the last ϖd-dependent terms in Equations (C1) and (C2), removing the explicit time dependence from them. This is essentially equivalent to neglecting both the gravitational effect of the disk, i.e., |ed| → 0, and the gas eccentricity, eg, compared to |eb|. As a result, we find a steady-state solution (43) for krkp and hrhp, which is essentially the Equations (22) and (23) with |ed|, |eg| → 0. Then planetesimal dynamics are described by the analytical results of Section 5 with ec ≈ |Bb/A|.

In the opposite extreme of weak eccentricity excitation by the binary companion we introduce new coordinates Hkghrhgkr, Khghr + kgkr (see Beaugé et al. 2010 for a similar treatment). Then the evolution of H and K is given by

Equation (C3)

Equation (C4)

When the eccentricity excitation by the companion is small, we can drop the Bb terms in these equations, removing the explicit time-dependence, which appears because of circulating kg and hg. As a result, we find the steady state solutions for H and K in the implicit form

Equation (C5)

where τd is a function of the relative particle–gas eccentricity $e_r=e_g^{-1}(K^2+H^2)^{1/2}$ given by Equation (44). This solution corresponds to the eccentricity vector ep fixed in a disk frame, which uniformly precesses at the rate $\dot{\varpi }_d$.

Using these solutions, it is trivial to show that epef, d given by Equation (B2) with Bb set to zero. That in the weak binary perturbation regime we find the same expression for ep as in the case of linear drag is not surprising: with Bb = 0 one finds that |er| is constant in time, so that τd is also constant. Then Equations (C3) and (C4) are the same as in the linear drag case and have the same steady state solutions (C5).

Footnotes

  • Curves of |Ad| = |Ab| and |Bd| = |Bb| run parallel to each other in Figures 4(c) and (d) because edap in our disk model; see Equations (5) and (8).

  • Variable ϖd(a) can be used to describe disks with density waves.

Please wait… references are loading.
10.1088/0004-637X/798/2/69