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.

TATOOINE'S FUTURE: THE ECCENTRIC RESPONSE OF KEPLER'S CIRCUMBINARY PLANETS TO COMMON-ENVELOPE EVOLUTION OF THEIR HOST STARS

, , , , and

Published 2016 November 30 © 2016. The American Astronomical Society. All rights reserved.
, , Citation Veselin B. Kostov et al 2016 ApJ 832 183 DOI 10.3847/0004-637X/832/2/183

0004-637X/832/2/183

ABSTRACT

Inspired by the recent Kepler discoveries of circumbinary planets orbiting nine close binary stars, we explore the fate of the former as the latter evolve off the main sequence. We combine binary star evolution models with dynamical simulations to study the orbital evolution of these planets as their hosts undergo common-envelope (CE) stages, losing in the process a tremendous amount of mass on dynamical timescales. Five of the systems experience at least one Roche-lobe overflow and CE stage (Kepler-1647 experiences three), and the binary stars either shrink to very short orbits or coalesce; two systems trigger a double-degenerate supernova explosion. Kepler's circumbinary planets predominantly remain gravitationally bound at the end of the CE phase, migrate to larger orbits, and may gain significant eccentricity; their orbital expansion can be more than an order of magnitude and can occur over the course of a single planetary orbit. The orbits these planets can reach are qualitatively consistent with those of the currently known post-CE, eclipse-time variations circumbinary candidates. Our results also show that circumbinary planets can experience both modes of orbital expansion (adiabatic and nonadiabatic) if their host binaries undergo more than one CE stage; multiplanet circumbinary systems like Kepler-47 can experience both modes during the same CE stage. Additionally, unlike Mercury orbiting the Sun, a circumbinary planet with the same semimajor axis can survive the CE evolution of a close binary star with a total mass of 1 ${M}_{\odot }$.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

"Everything is in motion," Plato quotes Heraclitus (Cratylus, Paragraph 402, section a, line 8), "and nothing remains still." Despite the overwhelming timescales, from a human perspective, the natural life cycle of stars is a prime example of change on the cosmic stage. Following the laws of stellar astrophysics, over millions to billions of years, stars and stellar systems form, evolve, and ultimately die (e.g., Kippenhahn & Wiegert 1990 and references therein). And so do planetary systems, with their fate linked intimately to that of their stellar hosts.

The fate of planets orbiting single stars has been studied extensively, indicating that planets can survive their star's evolution if they avoid engulfment or evaporation during the red giant branch (RGB) and the asymptotic giant branch (AGB) stages (e.g., Livio & Soker 1984; Rasio & Livio 1996; Duncan & Lissauer 1998; Villaver & Livio 2007; also see Veras et al. 2016). There is also accumulating evidence of planetary or asteroidal debris surrounding white dwarfs and polluting their atmospheres (see recent reviews by Farihi 2016 and Veras 2016). A planet's survival depends both on the initial orbital separation and on its mass. For example, even a planet as massive as 15 MJup around a 1 ${M}_{\odot }$ main-sequence star can be destroyed inside the stellar envelope during the AGB phase (Villaver & Livio 2007). An unpleasant prospect for the future of our own planet is that it might not survive the evolving Sun (e.g., Schroder & Connon Smith 2008). Despite the complications, theoretical considerations indicate that planets can indeed survive the evolution of single stars (Villaver & Livio 2007), and there is mounting observational evidence supporting this (e.g., Reffert et al. 2015; Wittenmyer et al. 2016, and references therein7 ).

Single stars, however, do not have a monopoly over planetary systems. Nearly half of solar-type stars are members of binary and higher order stellar systems (Raghavan et al. 2010), and an increasing number of planets have been discovered in such systems. While the presence of a distant stellar companion will have little effect on a planet around one (evolving) member of a wide binary stellar system (≳100 au), a planet orbiting around both members of a close binary system (separation ∼10 au or less) will experience qualitatively and quantitatively different stellar evolution. Namely, where a single star loses mass and expands on timescales of millions to billions of years (Veras 2016) while on the main sequence (MS), the RGB, and the AGB—and its planets react accordingly—close binary stars can experience events of dramatic mass loss, orbital shrinkage on timescales of years, and common-envelope (CE; see Appendix for abbreviations and parameters) stages where the two stars share (and quickly expel) a common atmosphere (Paczynski 1976; Hilditch 2001). During the in-spiraling CE phase, the two stars can strongly interact with each other by transferring mass, smoothly coalesce or violently collide, or even explode as a supernova (SN). Both the amount of energy released during this stage and the timescale of the release are staggering: a close binary star can lose an entire solar mass over the course of just a few months (e.g., Livio & Soker 1984; Rasio & Livio 1996; Ricker & Taam 2008, 2012; Passy et al. 2012; Ivanova et al. 2013; Nandez et al. 2014, but also see Sandquist et al. 1998 and De Marco et al. 2009 for longer timescales); in the extreme case of an SN, the binary will be completely disrupted. Overall, the evolution of close binary stars is much richer than single stars, as the separation, eccentricity, and metallicity of the binary add additional complications to an already complex astrophysical process. Interestingly, to date there are nine close MS binary systems harboring confirmed circumbinary planets (CBPs; Doyle et al. 2011; Orosz et al. 2012a, 2012b; Welsh et al. 2012, 2015; Kostov et al. 2013, 2014; Schwamb et al. 2013). It is reasonable to assume that the reaction of these planets to the violent CE phase of their host binary stars will be no less dramatic.

Previous studies of the dynamical response of planets in evolving multiple stellar systems have focused on post-CE (PCE) CBP candidates (e.g., Portegies Zwart 2013; Mustill et al. 2013; Volschow et al. 2014), on circumprimary planets in evolving wide binary systems (Kratter & Perets 2012), and on the survival prospects of CB planets (Veras et al. 2011; Veras & Tout 2012). In particular, Portegies Zwart (2013) and Mustill et al. (2013) start with the final outcome of close binary evolution—PCE systems with candidate CBP companions (HU Aqu and NN Ser, respectively)—and reconstruct the initial configuration of their progenitors.

Inspired by recent discoveries of planets in multiple stellar systems, here we tackle the reverse problem: we start with the nine confirmed CBP Kepler systems with known initial binary configurations (using data from NASA's Kepler  mission), and we study their dynamical response as their stellar hosts evolve. We combine established, publicly available binary star evolution code (BSE; Hurley et al. 2002) with direct N-body integrations (using REBOUND, Rein & Liu 2012), allowing for stellar mass loss and orbital shrinkage on dynamical timescales. For simplicity, here we focus explicitly on dynamical interactions only (e.g., no stellar tides) and assume that the CBPs do not interact with the ejected stellar material. We expand on the latter limitation in Section 5.1.

We note that Mustill et al. (2013) assume an adiabatic mass-loss regime (Hadjidemetriou 1963) during the CE stage for the evolution of NN Ser. In this case, the orbital period of the CBP is much shorter than the CE mass-loss timescale (hereafter TCE), and the planet's orbit expands at constant eccentricity.8 Volschow et al. (2014) study the two analytic extremes of mass-loss events for the case of NN Ser: the adiabatic regime and the instantaneous regime where ${T}_{\mathrm{CE}}$  is much shorter than the orbital period of the planets.9 The Kepler  CBP systems, however, occupy a unique parameter space where the planets' periods are comparable to the ${T}_{\mathrm{CE}}$  used by Portegies Zwart (2013) and suggested by Ricker & Taam (2008, 2012) for binary systems similar to Kepler  CBP hosts. Thus the adiabatic approximation may not be an adequate assumption when treating the CE phases of these systems. Specifically, the CE phase can occur on timescales as short as one year (Paczynski 1976; Passy et al. 2012), and during this year a binary can expel ∼2 ${M}_{\odot }$ worth of mass (Ricker & Taam 2008, 2012; Iaconi et al. 2016). Interestingly, such timescales are well within a single orbit of CBPs like Kepler-1647 with PCBP = 3 years (Kostov et al. 2016). Such a dramatic restructuring will have a profound impact on the dynamics of the system—for example, a CBP can achieve very high eccentricity—and may even result in the ejection of the planet. Similar to Portegies Zwart (2013), here we tackle such complications by directly integrating the equations of motion during the CE phases for each Kepler  CBP system.

The transition from adiabatic to nonadiabatic regime can be characterized through a mass-loss index ψ. For a planet orbiting a single star, Veras & Tout (2012) define

Equation (1)

 where ${\alpha }_{M}$  is the mass-loss rate, ap,0 is the initial semimajor axis of the planet, and μ = Mstar + Mp is the total mass of the system.10 If Ψ ≪ 1, then the evolution of the planet's orbit is in the adiabatic regime and its semimajor axis grows at constant eccentricity.

Alternatively, if Ψ ≫ 1, then the planet sees an "instantaneous" mass loss, and its orbital evolution is in a "runaway" regime. In this scenario, the fate of the planet depends on the ratio between the final and initial mass of the system, that is, β = μfinal/μinit, and on the orbital phase of the planet at the onset of the mass-loss event. For example, if a highly eccentric planet is at pericenter at the beginning of the CE stage, it becomes unbound. Overall, ejection occurs when the system loses sufficient mass by the end of the mass-loss event. To first order, the critical mass ratio for ejection βeject is given by Equation (39) of Veras et al. (2011), which we reproduce here for completeness:

Equation (2)

where ep,0 is the eccentricity of the planet at the beginning of the mass-loss event. A planet on a circular orbit or at pericenter is ejected from the system if β < βeject.11 A planet at apocenter remains bound even in the runaway regime, and its orbit expands to arunaway,apocenter and circularizes or reaches an eccentricity of e(post-circ). If a planet remains bound in the runaway regime, its semimajor axis is given by Equations (43) and (44) of Veras et al. (2011):

Equation (3)

where the signs represent initial pericenter and apocenter, respectively.

According to the above prescription, the dynamical evolution of a CBP's orbit for Ψ ∼ 0.1–1 lies between these two regimes. For both Ψ ∼ 1 and Ψ ≫ 1, the evolution is not adiabatic, and eCBP can vary (increase or decrease) as aCBP  varies. Thus the planet may become unbound only if Ψ ≳ 1, with the caveat that the transition is not clear-cut and the Ψ ∼ 0.1–1 regime is more complex (Veras et al. 2011; Veras & Tout 2012; Portegies Zwart 2013; Veras 2016). As we show below, the fact that the central object in a CBP system is itself a binary instead of a single star adds yet another layer of complication to these theoretical considerations.

Based on Equation (1), and assuming a CE timescale of ${T}_{\mathrm{CE}}$  ∼ 103 years, we would a priori expect that all Kepler CBPs remain bound to their hosts during their primary stars' CE stage (e.g., see Figure 3, Veras & Tout 2012). However, for the rapid CE timescale of ${T}_{\mathrm{CE}}$  ∼ 1 year suggested by recent results (e.g., Ricker & Taam 2008, 2012), some of the Kepler  CBP systems are within a factor of 5–10 of the transition to the nonadiabatic regime during the primary or secondary CE (i.e., Ψ ∼ 0.1–0.2), and within a factor of 2 of the transition during their respective secondary CE stage (i.e., Ψ ∼ 0.5). We might therefore expect significant variations in the respective CBPs' orbital eccentricities, as well as deviations from the final semimajor axes expected from adiabatic mass loss.

Because the transition between adiabatic and nonadiabatic regimes is difficult to study analytically, here we examine the dynamical evolution of all Kepler CBP systems numerically, using the adaptive-time-step, high-order integrator IAS15 (Rein & Spiegel 2015), which is available as part of the modular and open-source REBOUND package (Rein & Liu 2012). REBOUND is written in C99 and comes with an extensive Python interface.

This paper is organized as follows. In Section 2 we describe the algorithm we use for the evolution of the Kepler  CBP hosts. Section 3 details our implementation of an N-body code in exploring the dynamical reaction of said CBPs to the respective CE stages. We present the results for each system in Section 4, and we discuss our results and draw conclusions in Section 5.

2. BINARY STAR EVOLUTION

The evolution of a gravitationally bound system of two stars depends primarily on its initial configuration. If the binary is wide enough, the stars will evolve in isolation, according to the prescription of single-star evolution theory. Alternatively, if the initial separation is sufficiently small, then the stars will influence each other's evolution. The evolution of such systems, described as close (or interacting) binary stars (Hilditch 2001), is more complex and includes a number of additional processes, such as tidal interactions and mass transfer. The nine Kepler CBP-harboring binary systems fall in the latter category, and we use the established, open-source binary star evolution code (Hurley et al. 2002) to study their evolution. Briefly, the code works as follows.

While the binary is in a detached state, the code evolves each star individually using a single-star evolutionary code (Hurley et al. 2000), which includes tidal and braking mechanisms and wind accretion. When the stars begin to interact, BSE uses a suite of binary-specific features, such as mass transfer and accretion, CE evolution, collisions and mergers, and angular momentum loss mechanisms. The evolution algorithm allows the specific processes to be turned on and off, or even modified based on custom requirements. By default, all stars are assumed to be initially on the zero-age main sequence (ZAMS), but any possible evolutionary state, such as corotation with the orbit, can be the starting point of the simulation. BSE uses an evolution time step small enough to prevent the stellar mass and radius from changing too much (not more than 1% in mass and 10% in radius), which allows identification of the time when, and if, a star first fills its Roche lobe.

At the onset of Roche-lobe overflow (RLOF), the two stars either come into contact and coalesce or initiate a CE stage. CE evolution is a complex mechanism that typically occurs when an evolved, giant star transfers mass to an MS star on a dynamical timescale.12 When the giant star overfills the Roche lobe of both stars, its core and the MS star share a single envelope. This envelope rotates slower than the orbiting "cores" within due to expansion, and the resulting friction causes in-spiraling and energy transfer to the envelope, described by an efficiency parameter αCE. The outcome of this phase is either ejection of the envelope (assumed to be isotropic) if neither core fills its Roche lobe, leaving a close white dwarf–MS pair (assumed to be in corotation with the orbit), or coalescence of the cores. A challenging process to study numerically, BSE recognizes the CE phase—beginning with RLOF—and forces the system through it on an instantaneous timescale, allowing the evolution to continue according to the appropriate PCE parameters. The code resolves the outcome of the CE phase based on the initial binding energy of the envelope and on the initial orbital energy of the two cores.13

On input, BSE requires a number of parameters. Namely, the CE efficiency parameter αCE (in the range of 0.5–10), the masses and stellar types of both the primary and secondary stars (in the range of 0.1–100 ${M}_{\odot }$), the binary orbital period (in days), the eccentricity (0.0–1.0), the maximum time for the evolution (here we choose 15 Gyr, roughly the age of the universe, unless a particular Kepler CBP system stops evolving—or the stars coalesce—much sooner, at which point we stop the simulations), and the metallicity (0.0001–0.03) of the system. The code also checks whether tidal evolution is included, noted as "on" (hereafter TCP for tidal circularization path) or "off" (NTCP for no tidal circularization path). Given the still-uncertain mechanism of the CE stage, the parameters describing the efficiency of envelope ejection and the treatment of tidal decay—and their associated uncertainties—are likely the dominant source of error in our BSE results.

Depending on αCE, a binary system may exit a CE stage as a merged star, as a very tight PCE binary star, or may even trigger a supernova explosion. For example, a higher αCE  accounts for energy sources other than orbital energy, and the system then requires less energy to dissipate the envelope. The BSE code also allows an alternate CE model to be used, the de Kool CE evolution model (de Kool 1990), which first introduced the CE evolution binding energy factor λ.

Likewise, tides are integral to the evolution of close binary stars. The strength of tidal dissipation will affect the orbital separation and eccentricity when mass transfer begins and hence the future evolution under RLOF or a CE stage. BSE achieves orbital circularization mainly through tides, but also by accounting for more mass accretion at pericenter than apocenter of an eccentric orbit. Another effect is the exchange of orbital angular momentum with the component spins due to tidal synchronization, thus causing the evolution to proceed in a similar way to closer binary systems if tides were ignored. In addition, primary spin–orbit corotation can be forced within the code to avoid unstable RLOF, which gives two possible outcomes for each system. To account for the effect of tides on binary evolution, we use both the tides "on" (TCP) or "off" (NTCP) options in BSE. As a result, the binary starts the first RLOF at either ebin,RLOF = 0 and abin,RLOF < abin,0 (for TCP), or at ebin,RLOF = ebin,0 and abin,RLOF = abin,0 (for NTCP).

Additionally, while the binary periods of the Kepler  CBP systems are known to very high precision, their stellar masses and metallicities have nonnegligible uncertainties.14 For example, the primary masses for Kepler-38 and Kepler-64 have 1σ errors of 0.05 MSun and 0.1 MSun, respectively, and the 1σ errors on the metallicities of half of the systems are as large as, or even larger than, the measured values. To better characterize the impact of these uncertainties on the evolution of the Kepler  CBP systems, we obtain BSE results for both the best-fit mass and metallicity values and also for their corresponding 3σ range.15

The entire parameter space we explored, shown in Table 1, includes a range of CE parameters as well and consists of 22,680 BSE simulations. Our default configurations are denoted in boldface in the table; for the rest of the input parameters, we use the default values from model A of Hurley et al. (2002), described as the most favorable and effective. Model A uses all default wind-related parameters and a Reimers mass-loss coefficient of η = 0.5.

Table 1.  BSE Parameter Space

Parameter Explored Values
M1 M1, M1+ 3σ, M1 − 3σ
M2 M2, M2+ 3σ, M2 − 3σ
Metallicity Z Z, Z + 1σ, Z + 2σ, Z + 3σ, Z − 1σ, Z − 2σ, Z − 3σ
αCE 0.5, 1.0, 3.0, 5.0, 10.0
Tides OFF, ON
de Kool CE Model OFF, ON
Force Corotation OFF, ON

Note. Values denoted in boldface represent the default parameters used in our simulations.

Download table as:  ASCIITypeset image

As BSE evolves the stars, their types change, and each type is checked throughout the code to ensure the correct evolutionary tracks are used for both the individual stars and the system as a whole. These numerical values, as well as what they represent, are listed in Table 2. At each significant event, such as at the beginning and end of the RLOF and of the CE phase, when the stars coalesce or collide, the code also produces a "type" label. These labels, along with their meaning, are shown in Table 3.

Table 2.  BSE Stellar Types

Stellar Type Description
0 Deeply or fully convective low-mass MS star
1 Main-sequence star
2 Hertzsprung gap (HG)
3 First giant branch (GB)
4 Core helium burning (CHeB)
5 First/early asymptotic giant branch (EAGB)
6 Second/thermally pulsing asymptotic giant branch (TPAGB)
7 Main-sequence naked helium star (HeMS)
8 Hertzsprung gap naked helium star (HeHG)
9 Giant branch naked helium star (HeGB)
10 Helium white dwarf (HeWD)
11 Carbon/oxygen white dwarf (COWD)
12 Oxygen/neon white dwarf (ONeWD)
13 Neutron star (NS)
14 Black hole (BH)
15 Massless supernova/massless remnant

Download table as:  ASCIITypeset image

Table 3.  BSE Output Evolution Stages

Evolutionary Stage Label Description
INITIAL Initial configuration
KW CHNGE Stellar type change
BEG RCHE Begin Roche lobe overflow
END RCHE End Roche lobe overflow
CONTACT Contact system
COELESCE Coalescence of stars
COMENV Common-envelope system
GNTAGE New giant star from CE; appropriate age and initial mass set to match core mass and stellar mass
NO REMNT No remnant
MAX TIME Max evolutionary time reached; end of program
DISRUPT System is disrupted
BEG SYMB Begin symbiotic system
END SYMB End symbiotic system
BEG BSS Begin blue stragglers

Download table as:  ASCIITypeset image

An example of the BSE output for Kepler-38, αCE  = 0.5, NTCP, and TCP is shown in Table 4. The BSE results for all Kepler  systems can be found in an online supplement to Table 4.

Table 4.  BSE Evolution of Kepler-38 for αCE  = 0.5

Timea M1 M2 Stell. Type Stell. Type abin  ebin  R1/RRoche R1/RRoche Evol. Stage
(Gyr) (${M}_{\odot }$) (${M}_{\odot }$)     (R)        
Kepler-38b                
0.0 0.95 0.25 1 0 31.6 0.10 0.05 0.03 INITIAL
12.36 0.95 0.25 2 0 31.6 0.10 0.10 0.03 KW CHNGE
13.01 0.95 0.25 3 0 31.61 0.10 0.14 0.03 KW CHNGE
13.72 0.94 0.25 3 0 31.86 0.10 1.00 0.03 BEG RCHE
13.72 0.75 0.25 3 15 0.51 0.00 1.00 0.03 COMENV
13.76 0.47 0.00 10 15 0.00 −1.00 0.00 −1.00 KW CHNGE
15.00 0.47 0.00 10 15 0.00 −1.00 0.00 −1.00 MAX TIME
Kepler-38c                
0.0 0.95 0.25 1 0 31.59 0.10 0.05 0.03 INITIAL
12.36 0.95 0.25 2 0 31.60 0.10 0.10 0.03 KW CHNGE
13.01 0.95 0.25 3 0 31.45 0.09 0.14 0.03 KW CHNGE
13.70 0.94 0.25 3 0 24.58 0.00 1.00 0.04 BEG RCHE
13.70 0.91 0.25 3 15 0.37 0.00 1.00 0.04 COMENV
13.75 0.66 0.00 4 15 0.00 −1.00 0.00 −1.00 KW CHNGE
13.88 0.63 0.00 5 15 0.00 −1.00 0.00 −1.00 KW CHNGE
13.89 0.55 0.00 6 15 0.00 −1.00 0.00 −1.00 KW CHNGE
13.89 0.52 0.00 11 15 0.00 −1.00 0.00 −1.00 KW CHNGE
15.00 0.52 0.00 11 15 0.00 −1.00 0.00 −1.00 MAX TIME

Notes. The BSE results for our entire set of simulations for all Kepler systems are published in their entirety in an online, machine-readable format.

aTime in the online supplement is in Myr. bNTCP. cTCP.

Only a portion of this table is shown here to demonstrate its form and content. A machine-readable version of the full table is available.

Download table as:  DataTypeset image

In the following section we describe how we use the initial and BSE-generated final binary masses and orbital separations for each Kepler CBP system to explore the dynamical response of the planets during the respective CE stages.

3. DYNAMICAL SIMULATIONS

While the concept of CE evolution for binary stars was proposed 40 years ago (Paczynski 1976), detailed understanding of this important stage is a challenging task that requires computationally intensive hydrodynamical calculations. The outcome of this complex phase has been studied both numerically and analytically (see Table 1 of Iaconi et al. 2016 for a list of references). For example, hydrodynamical simulations by Ricker & Taam (2012) for a 1.05 ${M}_{\odot }$+0.36 ${M}_{\odot }$ binary star with an initial orbital period of 44 days (similar to Kepler CBP hosts) show that in the later stages of the CE the system loses mass at a rate of 2 ${M}_{\odot }$ yr−1, the binary orbital separation decreases by a factor of 7 after ∼56 days, and ∼90% of the outflow is contained within 30 degrees of the binary's orbital plane.

To numerically account for the dynamical impact on Kepler CBPs from such drastic orbital reconfigurations and tremendous, short-timescale mass-loss phases, we use REBOUNDx, a library for incorporating additional effects beyond point-mass gravity in REBOUND simulations (https://github.com/dtamayo/reboundx). In particular, we shrink the binary orbits with the REBOUNDx modify_orbits_forces routine, which adds to the equations of motion a simple drag force for opposite particles' velocity vectors (Papaloizou & Larwood 2000). When orbit-averaged over a Keplerian orbit, this approach results in an exponential damping of the semimajor axis where the e-folding timescale is set through the particle parameter τa. Hydrodynamical simulations by Passy et al. (2012) and Iaconi et al. (2016) exhibit similar orbital decay during the CE stage (e.g., see their respective Figures 5 and 14), validating our approach.

Given the complexities and uncertainties of the mass-loss mechanism during the CE phases, we adopt a simple analytical parameterization similar to the method of Portegies Zwart (2013), who tackled the problem by using a constant mass-loss rate. Here we choose to instead use the same functional forms for the CE-driven changes in Mstar as we do for abin (utilizing the REBOUNDx modify_mass routine), that is, an exponential mass loss for particles on their assigned e-folding timescales τM. Namely, for a given CE stage where one of the stars transitions from an initial mass M0 to a final mass Mf (both provided by BSE), the mass loss occurs over a time ${T}_{\mathrm{CE}}$  = (MfM0)/${\alpha }_{M}$. We note that an exponential mass loss has been explored by Adams et al. (2013) and Adams & Bloch (2013) as well, but for the case of single-star systems.

To ensure a smooth mass evolution (and avoid possible numerical artifacts) that asymptotes at Mf, we linearly evolved REBOUNDx's exponential mass-loss rate so that it vanished at time ${T}_{\mathrm{CE}}$. In particular, in terms of the e-folding timescale τM,

Equation (4)

 This differential equation can be solved analytically, yielding

Equation (5)

as long as τM,0 in Equation (4) is chosen as

Equation (6)

As the star's mass decreases from M0 to Mf during the respective CE stage, the binary orbit must shrink from abin,0 to abin,PCE (as provided by BSE) as well, and both must occur over a time ${T}_{\mathrm{CE}}$. We apply the same prescription for evolving abin as that for the mass loss (i.e., Equations (4)–(6) with τa instead of τM), but we caution that while this would evolve abin from abin,0 to abin,PCE in isolation, the mass loss is simultaneously increasing abin. To correct for this and achieve the desired abin,PCE, we manually adjust the corresponding τa,0 parameter.

While tides are not directly accounted for in our dynamical simulations as the particles involved are treated as point masses, we capture their effects in a simple parameterized way. Specifically, we impose the same exponential decay on the evolution of ebin as we do for the binary's semimajor axis shrinking and stellar mass loss (i.e., Equations (4)–(6) but with τe). As a result, ebin undergoes dampened oscillations during the CE phase, at the end of which it settles within a few percent of zero, as required by BSE.

As mentioned in the previous section, while the BSE code yields initial and final values for the stellar masses and abin at the beginning and end of the CE phases, it does not resolve the extremely rapid mass-loss phase. We bridge this gap by exploring two regimes of mass loss, namely ${\alpha }_{1}$ = 1.0 M yr−1 and ${\alpha }_{0.1}$ = 0.1 M yr−1, denoted by their respective subscripts. The former is based on the results of Ricker & Taam (2008, 2012) and Portegies Zwart (2013) for systems similar to the Kepler  CBP hosts. Values of ${\alpha }_{0.1}$ (and smaller) typically guarantee that the orbital evolution of these CBPs is in the adiabatic regime (see Equation (1)).

To test the applicability of our numerical simulations to the stated problem, we examine the evolution of a mock Kepler-38-like system where the binary is replaced with a single star of the same total initial and final mass according to the BSE prescription for Kepler-38 (see Table 4) and using the same initial orbit of the planet. Here we test two mass-loss regimes: runaway (for ${\alpha }_{M}$  = 100.0 ${M}_{\odot }$ yr−1) and adiabatic (for ${\alpha }_{M}$  = 0.1 ${M}_{\odot }$ yr−1${\alpha }_{0.1}$). The results are shown in Figure 1, where the three panels represent (from upper to lower, respectively) the evolution of the stellar mass, of the planet's semimajor axis, and of the planet's eccentricity as a function of time (such that mass loss starts at t = 0 and ends at t = Tmass-loss). According to the analytic prescription, the planet should remain bound even in the runaway mass-loss regime as the mass lost is smaller than the critical mass limit for ejection, as indicated by the dotted line in the upper panel of the figure.16 The horizontal dashed lines in the middle and upper panels represent the theoretical adiabatic (green) and runaway (red) approximations. As seen from these two panels, the numerical simulations are fully consistent with the theoretical approximations, demonstrating the validity and applicability of our simulations.

Figure 1.

Figure 1. Evolution of, from top to bottom, the stellar mass, the CBP semimajor axis, and eccentricity for a Kepler-38-like system, where the binary is replaced with a single star that loses mass according to the αCE  = 0.5 BSE simulation (for the primary CE phase) for two mass-loss regimes: runaway (red line) and adiabatic (green). Here, Tmass-loss is the same as the corresponding ${T}_{\mathrm{CE}}$ for Kepler-38, αCE = 0.5. The black dotted line in the upper panel represents the critical mass loss that would result in the ejection of the planet. The dashed lines in the middle and lower panels indicate the runaway (red) and adiabatic (green) approximations. The numerical evolution of the planet's semimajor axis and eccentricity is fully consistent with the theoretical expectation, validating the applicability of our simulations.

Standard image High-resolution image

In order to interpret the results from numerical integrations, it is useful to define a critical mass-loss rate, αcrit, corresponding to Ψ = 1 (see Equation (1)). CBPs in systems with αM ≫ αcrit  will experience "instantaneous" mass loss, while the orbits of those in systems with αM ≪ αcrit  should evolve adiabatically. A CBP in a system where αM becomes comparable to or larger than αcrit  at any point during the binary evolution risks becoming unbound. Secondary CE stages will have two different αcrit  values as the respective CBPs will have two different values for aCBP  and eCBP (but the same μ) at the end of the preceding CE stages (see Equation (1)): one for ${\alpha }_{1}$ and another for ${\alpha }_{0.1}$. We denote these critical values as αcrit (α1) and αcrit (α0.1). In the subsections below, for each CBP Kepler  system, we compare αcrit  to α1 and α0.1 for each CE phase. Where the adopted ${\alpha }_{1}$ is within a factor of 10 of the respective αcrit, we pay special attention to the CBP orbit and test its evolution for 0.1 ${\alpha }_{1}$, 0.2 ${\alpha }_{1}$, and 0.5 ${\alpha }_{1}$ as well. Unless indicated otherwise, the final fate of the CBP for these cases is similar to the case for ${\alpha }_{1}$.

The description of a three-body system containing an evolving binary star is inherently multidimensional, and minor changes in one parameter can cascade into major changes in the rest. In addition to the importance of αCE, tides, and αM, another key issue—as shown in the next section—is the phase offset between the binary star and the CBP at the onset of the CE phase, Δθ0 ≡ θCBP,0 − θbin,0 − ωbin (i.e., the difference between their true anomalies, taking into account the binary's argument of the pericenter).

Thus for each system we test four initial binary phases corresponding to the binary orbit turning points, that is, eastern and western elongations (EE and WE), and superior and inferior conjunctions (SC and IC), and 50 initial CBP phases (between 0 and 1, with a step of 0.02) for a total of 200 initial conditions. We explore each of these initial conditions for αCE = (0.5, 1.0, 3.0, 5.0, 10.0), for TCP/NTCP, and for ${\alpha }_{1}$ and ${\alpha }_{0.1}$.

Finally, we note that the post-CE semimajor axis of a CBP (aCBP,PCE) that was initially on an eccentric orbit (i.e., eCBP,0 > 0) depends on its orbital phase at the onset of the CE stage. For those Kepler  systems that undergo a secondary CE phase with mass loss, we start the respective integrations at the modes of aCBP,PCE and eCBP,PCE from the preceding CE stage.

4. RESULTS

Here we describe the results of our simulations for those Kepler  CBP systems that experience at least one CE phase. Unless specified otherwise, the evolution of each system is for the default CE parameters listed in Table 1. The initial orbital parameters for each system are listed in Table 5. We note that our N-body integrations only cover the CE phases.

Table 5.  Kepler System Default Initial Parameters

Name Kepler ID M1 M2 Pbin ebin  Z PCBP eCBP
    (${M}_{\odot }$) (${M}_{\odot }$) (days)     (days)  
Kepler-16 12644769 0.69 0.20 41.08 0.16 0.010 228.78 0.01
Kepler-34 8572936 1.05 1.0208 27.79 0.52 0.016 288.82 0.18
Kepler-35 9837578 0.89 0.81 20.73 0.14 0.009 131.46 0.04
Kepler-38 6762829 0.95 0.25 18.79 0.1 0.015 105.59 0.03
Kepler-47a 10020423 0.96 0.34 7.45 0.02 0.011 49.5 0.03 (P1)
              187.4 0.02 (P2)
              303.16 0.04 (P3)
Kepler-64 4862625 1.53 0.41 20 0.22 0.031b 138.5 0.1
Kepler-413 12351927 0.82 0.54 10.12 0.04 0.012 66.26 0.12
Kepler-453 9632895 0.94 0.2 27.34 0.05 0.023 240.5 0.04
Kepler-1647 5473556 1.22 0.97 11.259 0.15 0.014 1107.59 0.06

Notes.

aJ. Orosz (2016, private communication). bBSE allows maximum metallicity of Z = 0.03.

Download table as:  ASCIITypeset image

Below we outline the major binary evolution stages (as produced by BSE), along with the dynamical responses of the CBPs; the main results are presented in Tables 610. All reported times are in Gyr since ZAMS. For simplicity, we assume that the current Kepler  CBP systems are at ZAMS and that all CBPs start on initially circular, coplanar orbits.

Table 6.  Binary Evolution from BSE

αCE Tides M1 M2 abin  ebin  Time RLOFa Notes
    (MSun) (MSun) (R)   (Gyr)  
Kepler-34              
1.05 1.02 49.2 0.52 0
0.5/1/3/5/10 NTCP 9.64(1.7) DD SN
0.5/1/3/5/10 TCP 9.66(1.7) DD SNb
Kepler-38              
0.95 0.25 31.6 0.1 0
0.5 NTCP 0.76 13.72(0.6) CE + Merger
0.5 TCP 0.91 13.70(0.6) CE + Merger
1/3/5/10 NTCP 0.27 0.25 1.0/2.8/4.5/7.8 13.72(1.2) CE
1/3/5/10 TCP 0.25 0.25 0.7/2.0/3.2/5.7 13.70(1.1) CEd
Kepler-47              
0.96 0.34 17.5 0.02 0
0.5 NTCP 1.08 12.03(1.3) CE + Merger
0.5 TCP 1.14 11.99(1.3) CE + Merger
1.0 NTCP 0.78 12.03(1.5) CE + Merger
1.0 TCP 0.96 11.99(1.5) CE + Merger
3/5/10 NTCP 0.23 0.34 1.7/2.6/4.5 12.03(1.7) CE
3/5/10 TCP 0.21 0.34 1.2/1.9/3.4 11.99(1.6) CE
Kepler-64              
1.53 0.41 38.7 0.22 0
0.5 NTCP 1.53 3.02(0.3) CE + Merger
0.5 TCP 1.68 3.02(0.3)e CE + Merger
1.0 NTCP 0.8 3.02(0.3) CE + Merger
1.0 TCP 1.35 3.02(0.3) CE + Merger
3/5/10 NTCP 0.28 0.41 2.4/3.9/7.0 3.02(0.3) CE
3/5/10 TCP 0.26 0.41 1.6/2.6/4.7 3.02(0.3) CE
3 TCP 0.67 6.20(0.4) CEc + Merger
Kepler-1647              
1.22 0.97 27.4 0.15 0
0.5 NTCP 1.81 5.45(0.45) CE + Merger
0.5 TCP 1.87 5.44(0.46) CE + Merger
1.0 NTCP 1.28 5.45(0.44) CE + Merger
1.0 TCP 1.45 5.44(0.43) CE + Merger
3/5/10 NTCP 0.25 0.97 4.0/6.2/10.1 5.45(0.45) CE
3/5/10 TCP 0.24 0.97 3.6/5.4/8.9 5.44(0.36) CE
3 TCP 1.16 6.07(0.35) CEc + Merger
5 NTCP 0.25 0.17 0.6 12.06(0.9) CEc
5 NTCP 12.93(1.0) DD SN
5 TCP 1.15 9.85(0.5) CEc + Merger
10 NTCP 0.25 0.2 1.9 12.27(1.1) CEc
10 TCP 0.24 0.16 0.7 11.86(0.9) CEc
10 TCP 13.99(0.5) DD SN
Kepler-35             (Z − 1σZ)
0.88 0.81 37.9 0.14 0
0.5 NTCP 1.1 12.99(1.44) CE + Merger
0.5 TCP 1.09 12.99(1.44) CE + Merger
1/3/5/10 NTCP 0.29 0.81 3.5/8.8/12.7/19.1 12.99(1.44) CE
1/3/5/10 TCP 0.28 0.81 3.1/7.9/11.5/17.4 12.99(1.44) CE

Notes. The numbers quoted in parentheses for the time represent the corresponding range obtained from BSE for Z ± 3σZ (see text for details). The typical range for M1 and M2 is ∼0.01–0.1 ${M}_{\odot }$, and ∼0.1–1 R for abin.

aRoche-lobe overflow. bDouble-degenerate supernova. cSecondary CE. dSecond Roche lobe fill and merger shortly after first CE for αCE = 1. eOnly lower range because BSE does not accept Z > 0.3 (the metallicity of Kepler-64).

Download table as:  ASCIITypeset image

Table 7.  PCE CBP Semimajor Axis and Eccentricity for ${\alpha }_{M}$  = 1.0 MSun yr−1 and for Primary CE

αCE Tides   aCBP,PCE (au)     eCBP,PCE   Notes
    Mode Median Range Mode Median Range  
Kepler-38               [aCBP,0 = 0.46]
0.5 NTCP ${0.6}_{-0.01}^{+0.2}$ 0.8 0.58–1.7 ${0.15}_{-0.07}^{+0.12}$ 0.26 0.05–0.67
0.5a TCP ${0.5}_{-0.01}^{+0.4}$ 0.6 0.49–0.9 ${0.4}_{-0.35}^{+0.01}$ 0.2 0.03–0.42
1/3/5/10 NTCP ${1.3}_{-0.2}^{+0.1}$ 1.3 0.9–3.8 ${0.38}_{-0.01}^{+0.16}$ 0.42 0.02–0.82
1/3/5/10 TCP ${1.4}_{-0.2}^{+0.01}$ 1.4 1.1–1.7 ${0.46}_{-0.08}^{+0.01}$ 0.42 0.28–0.56
Kepler-47                
0.5, P1b NTCP ${0.31}_{-0.01}^{+0.09}$ 0.34 0.3–0.44 ${0.12}_{-0.03}^{+0.04}$ 0.14 0.06–0.27 [aCBP,0 = 0.29]
0.5, P2b NTCP ${0.66}_{-0.01}^{+0.42}$ 0.88 0.61–1.67 ${0.14}_{-0.05}^{+0.35}$ 0.22 0.06–0.56 [aCBP,0 = 0.70]
0.5, P3b NTCP ${0.9}_{-0.01}^{+0.61}$ 1.2 0.8–2.87 ${0.17}_{-0.05}^{+0.29}$ 0.23 0.09–0.66 [aCBP,0 = 0.96]
0.5a, P1b TCP ${0.32}_{-0.01}^{+0.07}$ 0.34 0.31–0.39 ${0.1}_{-0.02}^{+0.08}$ 0.1 0.07–0.18
0.5a, P2b TCP ${0.7}_{-0.01}^{+0.5}$ 0.87 0.67–1.24 ${0.08}_{-0.01}^{+0.32}$ 0.21 0.06–0.42
0.5a, P3b TCP ${0.93}_{-0.01}^{+0.91}$ 1.18 0.88–1.89 ${0.11}_{-0.01}^{+0.35}$ 0.22 0.09–0.48
1.0, P1b NTCP ${0.49}_{-0.02}^{+0.01}$ 0.49 0.47–0.5 ${0.08}_{-0.04}^{+0.05}$ 0.08 0.04–0.13
1.0, P2b NTCP ${1.33}_{-0.17}^{+0.17}$ 1.41 1.12–1.97 ${0.42}_{-0.14}^{+0.11}$ 0.4 0.22–0.58
1.0, P3b NTCP ${1.65}_{-0.01}^{+0.53}$ 2.13 1.52–4.15 ${0.5}_{-0.14}^{+0.18}$ 0.51 0.29–0.75
1.0a, P1 TCP ${0.38}_{-0.01}^{+0.03}$ 0.39 0.37–0.41 ${0.11}_{-0.1}^{+0.01}$ 0.07 0.01–0.12
1.0a, P2 TCP ${0.91}_{-0.01}^{+0.29}$ 1.03 0.89–1.22 ${0.36}_{-0.21}^{+0.01}$ 0.26 0.14–0.38
1.0a, P3 TCP ${1.23}_{-0.01}^{+0.47}$ 1.46 1.2–1.83 ${0.44}_{-0.26}^{+0.01}$ 0.32 0.17–0.45
3/5/10, P1d NTCP ${0.67}_{-0.01}^{+0.01}$ 0.67 0.66–0.68 ${0.11}_{-0.01}^{+0.03}$ 0.11 0.04–0.17
3/5/10, P2d NTCP ${3.23}_{-0.01}^{+0.57}$ 3.4 2.23–7.95 ${0.71}_{-0.06}^{+0.1}$ 0.73 0.57–0.89
3/5/10, P3d NTCP ${7.64}_{-2.71}^{+7.21}$ 9.97 4.48–22.52 ${0.92}_{-0.02}^{+0.02}$ 0.89 0.74–0.95c 46% ejected
3/5/10, P1d TCP ${0.69}_{-0.01}^{+0.01}$ 0.69 0.68–0.7 ${0.1}_{-0.01}^{+0.03}$ 0.11 0.08–0.14
3/5/10, P2d TCP ${3.68}_{-0.23}^{+0.91}$ 4.18 3.34–5.62 ${0.75}_{-0.01}^{+0.07}$ 0.78 0.72–0.83
3/5/10, P3d TCP ${15.69}_{-0.01}^{+6.16}$ 19.7 15.3–23.01 ${0.93}_{-0.01}^{+0.02}$ 0.94 0.93–0.95c 81% ejected

Notes. The modes are shown with their respective 68% range. We caution that most of the probability distribution functions (PDFs) are notably nonnormal, often with two dominant peaks at the range limits.

aSinusoidal distributions in phase versus final aCBP/eCBP (see, e.g., Figure 5), typically with double-peaked PDFs. bP1/P2/P3 have 40%/15%/75% probability to become dynamically unstable within 1 Myr after the CE phase. cGiven the large parameter space and the associated uncertainties, we consider eCBP > 0.95 as ejection. dP1/P2/P3 have 60%/15%/85% probability to become dynamically unstable within 1 Myr after the CE phase.

Download table as:  ASCIITypeset image

Table 8.  Same as Table 7

αCE Tides   aCBP,PCE (au)     eCBP,PCE    Notes
    Mode Median Range Mode Median Range  
Kepler-64               [aCBP,0 = 0.65]
0.5 NTCP ${0.8}_{-0.2}^{+0.6}$ 0.9 0.6–4.8 ${0.16}_{-0.}^{+0.24}$ 0.35 0.03–0.85
0.5a TCP ${0.6}_{-0.0}^{+0.6}$ 0.8 0.6–1.2 ${0.19}_{-0.07}^{+0.23}$ 0.19 0.11–0.43
1.0 NTCP ${1.4}_{-0.0}^{+1.1}$ 2.0 1.3–7.5 ${0.62}_{-0.31}^{+0.15}$ 0.53 0.12–0.8
1.0a TCP ${0.9}_{-0.0}^{+0.1}$ 0.9 0.9–1.0 ${0.2}_{-0.18}^{+0.0}$ 0.13 0.01–0.21
3/5/10 NTCP ${1.6}_{-0.1}^{+2.6}$ 3.1 1.4–19.6 ${0.8}_{-0.26}^{+0.11}$ 0.66 0.04–0.95b 4% ejected
3/5/10 TCP ${2.4}_{-0.1}^{+0.}$ 2.4 2.2–2.6 ${0.49}_{-0.09}^{+0.01}$ 0.44 0.38–0.5
Kepler-1647c               [aCBP,0 = 2.71]
0.5 NTCP ${1.6}_{-0.0}^{+2.3}$ 2.4 1.4–39.5 ${0.72}_{-0.43}^{+0.06}$ 0.69 0.3–0.95 47% ejected
0.5a TCP ${2.3}_{-0.0}^{+10.6}$ 3.6 2.0–13.2 ${0.26}_{-0.01}^{+0.5}$ 0.33 0.23–0.79
1.0 NTCP ${1.9}_{-0.0}^{+2.1}$ 2.6 1.7–37.5 ${0.53}_{-0.11}^{+0.11}$ 0.57 0.4–0.95 47% ejected
1.0a TCP ${3.9}_{-0.0}^{+5.6}$ 5.5 3.8–9.7 ${0.29}_{-0.0}^{+0.4}$ 0.5 0.27–0.71
3/5/10 NTCP ${2.3}_{-0.3}^{+1.5}$ 3.1 1.9–47.1 ${0.4}_{-0.12}^{+0.12}$ 0.44 0.16–0.95 54% ejected
3/5/10 TCP ${9.6}_{-1.6}^{+4.6}$ 11.5 7.6–23.2 ${0.7}_{-0.01}^{+0.14}$ 0.76 0.63–0.88
Kepler-1647d                
5.0a NTCP ${11.3}_{-1.1}^{+5.3}$ 15.3 10.2–53.3 ${0.76}_{-0.03}^{+0.03}$ 0.79 0.68–0.95 86% ejected
10.0a NTCP ${8.2}_{-0.6}^{+6.4}$ 10.5 7.6–36.9 ${0.64}_{-0.07}^{+0.07}$ 0.71 0.57–0.95 83% ejected
10.0 TCP ${14.5}_{-1.0}^{+11.1}$ 18.2 13.5–96.7 ${0.52}_{-0.4}^{+0.28}$ 0.5 0.01–0.95 85%ejected
Kepler-35e               [aCBP,0 = 0.60]
0.5 NTCP ${0.87}_{-0.01}^{+0.23}$ 1.04 0.75–10.13 ${0.44}_{-0.18}^{+0.27}$ 0.41 0.03–0.93 4% ejected
0.5a TCP ${0.87}_{-0.01}^{+0.47}$ 1.01 0.84–1.37 ${0.43}_{-0.37}^{+0.01}$ 0.25 0.04–0.45
1/3/5/10 NTCP ${0.84}_{-0.01}^{+0.16}$ 1.03 0.77–6.93 ${0.43}_{-0.26}^{+0.01}$ 0.36 0.04–0.91
1/3/5/10 TCP ${1.0}_{-0.04}^{+0.03}$ 0.98 0.86–1.24 ${0.18}_{-0.04}^{+0.04}$ 0.19 0.01–0.39

Notes.

aSinusoidal-like distributions in aCBP,PCE/eCBP,PCE as a function of Δθ0; typically with double-peaked PDFs. bGiven the large parameter space and the associated uncertainties, we consider $e\gt 0.95$ as ejection. cPrimary RLOF and CE. dSecondary RLOF and CE. eExperiences CE for Z − 1σZ.

Download table as:  ASCIITypeset image

Table 9.  Same as Table 7 but for ${\alpha }_{M}$  = 0.1 MSun yr−1

αCE Tides   aCBP,PCE (au)     eCBP,PCE    Notes
    Mode Median Range Mode Median Range  
Kepler-38               [aCBP,0 = 0.46]
0.5 NTCP 0.73 0.73 0.12 0.12
0.5a TCP 0.6 0.61 0.03 0.03
1/3/5/10 NTCP 1.06 1.06 0.13 0.13
1/3/5/10 TCP 1.1 1.11 0.02 0.02
Kepler-47                
0.5, P1 NTCP 0.34 0.34 0.04 0.04 [aCBP,0 = 0.29]
0.5, P2 NTCP 0.83 0.83 0.04 0.04 [aCBP,0 = 0.70]
0.5, P3 NTCP 1.13 1.14 0.06 0.06 [aCBP,0 = 0.96]
0.5a, P1 TCP 0.34 0.34 0.01 0.01
0.5a, P2 TCP 0.8 0.83 0.02 0.02
0.5a, P3 TCP 1.13 1.14 0.03 0.03
1.0, P1 NTCP 0.48 0.48 0.03 0.03
1.0, P2 NTCP 1.18 1.18 0.05 0.05
1.0, P3 NTCP 1.6 1.61 0.06 0.06
1.0a, P1 TCP 0.39 0.39 0.01 0.01
1.0a, P2 TCP 0.96 0.96 0.02 0.02
1.0a, P3 TCP 1.3 1.31 0.03 0.03
3/5/10, P1 NTCP 0.66 0.66 0.04 0.04
3/5/10, P2 NTCP 1.61 1.61 0.07 0.07
3/5/10, P3 NTCP 2.2 2.2 0.08 0.08
3/5/10, P1 TCP 0.69 0.69 0.01 0.01
3/5/10, P2 TCP 1.67 1.67 0.03 0.03
3/5/10, P3 TCP 2.28 2.28 0.04 0.04

Notes. Kepler-38b, Kepler-47bcd, and Kepler-64b evolve adiabatically. Unless indicated otherwise, the spread of their semimajor axis and eccentricity range is practically zero and for simplicity omitted.

aSinusoidal distributions in phase versus final aCBP/eCBP (see Figure 5), typically with double-peaked PDFs. bGiven the large parameter space and the associated uncertainties, we consider e > 0.95 as ejection.

Download table as:  ASCIITypeset image

Table 10.  Same as Table 9

αCE Tides aCBP,PCE (au) eCBP,PCE  Notes
    Mode Median Range Mode Median Range  
Kepler-64               [aCBP,0 = 0.65]
0.5 NTCP 0.82 0.82 0.25 0.25 0.22–0.27
0.5${}^{{\rm{f}}}$ TCP 0.72 0.72 0.02 0.02
1.0 NTCP 1.58 1.58 0.31 0.33 0.31–0.35
1.0${}^{{\rm{f}}}$ TCP 0.91 0.91 0.01 0.01
3/5/10 NTCP 1.83 1.83 0.3 0.35 0.31–0.39
3/5/10 TCP 1.9 1.9 0.02 0.02
Kepler-1647b               [aCBP,0 = 2.71]
0.5 NTCP ${2.52}_{-0.01}^{+1.53}$ 3.55 2.33–40.61 ${0.7}_{-0.08}^{+0.02}$ 0.71 0.62–0.95a 27% ejected
0.5${}^{{\rm{f}}}$ TCP ${3.1}_{-0.01}^{+0.19}$ 3.19 3.08–3.3 ${0.11}_{-0.08}^{+0.01}$ 0.07 0.02–0.12
1.0 NTCP ${3.91}_{-0.01}^{+0.67}$ 4.45 3.83–20.27 ${0.72}_{-0.01}^{+0.22}$ 0.81 0.71–0.95 27% ejected
1.0${}^{{\rm{f}}}$ TCP 4.1 4.1 0.07 0.07
3/5/10 NTCP ${4.7}_{-0.26}^{+0.63}$ 4.93 4.29–19.14 ${0.7}_{-0.01}^{+0.19}$ 0.76 0.57–0.95 4% ejected
3/5/10 TCP 4.93 4.93 0.08 0.08
Kepler-1647c                
5.0e NTCP ${7.9}_{-0.7}^{+6.3}$ 9.9 7.2–55.9 ${0.48}_{-0.1}^{+0.9}$ 0.55 0.2–0.95 82% ejected
10.0e NTCP ${7.7}_{-0.8}^{+4.0}$ 9.2 6.9–31.1 ${0.32}_{-0.08}^{+0.25}$ 0.54 0.24–0.95 81% ejected
10.0 TCP 100% ejected
Kepler-35d               [aCBP,0 = 0.60]
0.5 NTCP 0.94 0.94 0.37 0.37 0.33–0.41
0.5 TCP 0.95 0.95 0.03 0.03
1/3/5/10 NTCP 0.94 0.94 0.27 0.27 0.18–0.39
1/3/5/10 TCP 0.95 0.95 0.02 0.02

Notes.

aGiven the large parameter space and the associated uncertainties, we consider e > 0.95 as ejection. bPrimary RLOF and CE. cSecondary RLOF and CE. dExperiences CE for Z − 1σZ. eNonnormal PDF with multiple peaks of comparable strength, with mode not well defined.

Download table as:  ASCIITypeset image

Overall, five of the nine Kepler CBP systems experience at least one CE phase for their default metallicities (see Table 1) over the duration of the BSE simulations (15 Gyr): Kepler-34, -38, -47, -64, and -1647. Kepler-35 experiences a CE for Z − 1σZ, and Kepler-453 for Z − 2σZ.

4.1. Kepler-34

The binary consists of 1.05 ${M}_{\odot }$ and 1.02 ${M}_{\odot }$ stars on a highly eccentric (ebin,0 = 0.52), 28 day orbit that changes little in the first ∼9 Gyr. The CBP has an initial orbit of 1.09 au.

The primary star fills its Roche lobe at ∼9.6 Gyr, and the binary enters a CE stage and triggers a double-degenerate supernova event, disrupting the system.

4.2. Kepler-38

The binary consists of 0.95 ${M}_{\odot }$ and 0.25 ${M}_{\odot }$ stars on a slightly eccentric (ebin,0 = 0.1), 19 day orbit that changes little in the first ∼13 Gyr. The CBP has an initial orbit of 0.46 au. The main evolution stages of the binary star and their effect on the CBP are as follows.

4.2.1. The Binary

The primary star fills its Roche lobe at t ∼ 13.7 Gyr, and the binary enters a primary CE stage. During this stage, Ψ(${\alpha }_{1}$) ≈ 0.04 and αcrit  ≈ 26 ${M}_{\odot }$ yr−1, much larger than ${\alpha }_{1}$ and ${\alpha }_{0.1}$. As a result of the CE, the binary merges as a first giant branch star and loses ∼25–40% of its mass for αCE = 0.5; here β > βeject (see Equation (3)), and the planet should remain bound in a runaway regime (i.e., Ψ ≫ 1) even at pericenter (see Equations (37)–(44), Veras et al. 2011). The system continues to slowly lose mass by the end of the BSE simulations (15 Gyr), so aCBP,PCE adiabatically expands by up to a factor of 2.

For αCE = 1.0, 3.0, 5.0, and 10.0, the binary orbit shrinks by a factor of 4–30 and its mass decreases by ∼60%; here β < βeject, that is, the planet should be ejected in a runaway regime. For αCE  = 1 and TCP, the system experiences a secondary RLOF at t ∼ 13.74 Gyr and coalesces into a first giant branch star without mass loss (thus with no changes in aCBP,PCE or eCBP,PCE). The binary does not lose additional mass by the end of the BSE simulations.

Overall, by 15 Gyr the binary evolves into either a merged helium or carbon/oxygen white dwarf (HeWD, COWD) or a very close HeWD–MS star binary.

The orbital reconfiguration of the binary system during the primary CE stage is shown on the left panels of Figure 2 for αCE = 0.5 (upper panels) and αCE = 1.0, 3.0, 5.0, and 10.0 (lower panels). The observer is looking along the axis, from below the figure. The maximum (red dashed line; binary initially at EE, CBP initially at phase 0.73) and minimum (red solid line; binary at EE, CBP at phase 0.08) orbital expansion of the CBP is shown on the right panels of the figure for ${\alpha }_{1}$ (red), as well as the maximum expansion for ${\alpha }_{0.1}$ (green solid line). To test the dynamical stability of the CBP at maximum expansion (dashed red line), we integrated the system for a thousand planetary orbits. The orbit precesses but does not exhibit chaotic behavior for the duration of the integration. We leave examination of the long-term behavior and stability of the system for future work.

Figure 2.

Figure 2. Orbital reconfiguration of the Kepler-38 system during the primary CE stage. The binary orbits are shown on the left panels (blue lines for the primary star, magenta lines for the secondary) and the CBP orbits on the right (green lines for ${\alpha }_{0.1}$, red lines for ${\alpha }_{1}$, where dashed red corresponds to the largest and the solid red to the smallest achieved orbit). The binary merges for αCE  = 0.5. The square symbols indicate the initial positions of the two stars and the planet; the circle symbols in the left panels indicate the final positions of the primary (blue) and secondary (black) star. The right panels show the evolution of aCBP  from different initial conditions, as indicated by the respective squares.

Standard image High-resolution image

4.2.2. The CBP

In Figure 3 we show the αCE  = 0.5 CE evolution of M1, of abin (blue lines, upper panel), and of aCBP  (middle panels) and eCBP (lower panel) for ${\alpha }_{1}$ (red) and ${\alpha }_{0.1}$ (green); on Figure 4 we show the corresponding evolution for αCE  = 1, 3, 5, and 10. Unlike the case of a planet orbiting a single star, where the evolution of the planet's orbit follows a single path for each ${\alpha }_{M}$ (see Figure 1), the dynamical response of a CBP to a binary undergoing a CE stage depends on the initial configuration of the system, that is, the phase difference Δθ0 ≡ θCBP,0 − θbin,0 − ωbin, where θCBP,0and θbin,0 are the initial true anomaly of the CBP and the binary, respectively, and ωbin is the binary's argument of pericenter. This is reminiscent of the importance of the initial true anomaly of a planet orbiting a single star on an eccentric orbit (see Veras et al. 2011). The dependence is shown in the second panel of Figure 3, where each curve represents the evolution of aCBP  for Δθ0 varying between 0 and 1 (for clarity, the curves here represent every tenth phase where in the simulations the phase difference varies with a step of 0.02).

Figure 3.

Figure 3. Evolution of, from top to bottom, the mass of the Kepler-38 primary star and the binary semimajor axis (blue), aCBP, and eCBP (red for ${\alpha }_{1}$, green for ${\alpha }_{0.1}$) during the primary RLOF and CE for αCE = 0.5. The binary merges at the end of the CE, at t = TCE. The second panel shows the dependence of the evolution of aCBP on Δθ0, the initial true anomaly difference between the binary star and the CBP at the onset of the CE (Δθ0θCBP,0θbin,0ωbin). Unlike the case of a planet orbiting a single star losing mass, where the planet's orbit follows a single evolutionary track, a CBP orbiting a binary star can experience different evolutionary tracks during a CE phase for different Δθ0. The individual evolutionary tracks for aCBP  and eCBP are condensed into hatched regions on the third and fourth panels (and in the subsequent figures) for clarity. The dotted black line in the upper panel represents the critical mass for ejection of the CBP, and the dashed lines in the middle and lower panels represent the respective adiabatic and runaway expansions.

Standard image High-resolution image
Figure 4.

Figure 4. Same as Figure 3 but for αCE  = 1, 3, 5, and 10.

Standard image High-resolution image

Since our simulations encompass a discrete set of initial conditions for αCE, tides (NTCP or TCP for minimum to maximum strength, respectively), and Δθ0, the true evolution of aCBP  and eCBP would be continuous. We represent this by the hatched regions on the lower panels of Figures 3 and 4, where the different hatches correspond to TCP or NTCP. For clarity, we show individual aCBP  evolutionary tracks only in Figure 3; these are replaced with hatched regions in all subsequent figures. We note that the distribution of aCBP,PCE and eCBP,PCE  (i.e., aCBP  and eCBP at the end of the CE phase) is neither uniform nor necessarily normal, as seen from the second panel on Figure 3, but instead depends on Δθ0. This dependence is demonstrated on Figure 5 for αCE  = 0.5 and on Figure 6 for αCE  = 1, 3, 5, and 10, and the corresponding probability distribution functions (PDFs) are shown in Figures 7 and 8. The PDFs can have a prominent peak with a long one-sided tail (e.g., upper left panel on the first figure), or be double-peaked with the peaks near the edges (e.g., lower left panel on the first figure); this is a natural consequence of the vaguely sinusoidal dependencies seen in Figures 5 and 6. Given the prominent diversity of the PDFs, the distributions are better represented by their modes than their medians. Thus for completeness we list the mode, median, and min/max range for each aCBP,PCE and eCBP,PCE  in Tables 7 through 10, but we cite the 68%-range upper and lower bounds on the modes only because the medians may not be appropriate in some cases. Unless otherwise noted, we refer to the modes of aCBP,PCE and eCBP,PCE  as the default results.

Figure 5.

Figure 5. Dependence of aCBP,PCE (upper) and eCBP,PCE (lower) on the initial phase difference between the binary and the CBP Δθ0, and on ${\alpha }_{M}$  for Kepler-38, and for αCE  = 0.5. The different colors for NTCP (blue, red, green, cyan) correspond to different initial binary phases: eastern and western elongation (EE and WE) and superior and inferior conjunction (SC, IC). The case of TCP is represented by a single color (magenta) as the binary is circular at the onset of the RLOF and thus has a single initial condition (set to inferior conjunction for simplicity). The dashed black lines in the upper panels represent the corresponding adiabatic expansion approximation. As seen from the upper right panel, the results from our numerical simulations for aCBP,PCE and for ${\alpha }_{0.1}$ are fully consistent with the adiabatic approximation. For ${\alpha }_{1}$, the modes of the aCBP,PCE and eCBP,PCE  distributions are 0.5/0.6 au and 0.4/0.15 for TCP/NTCP, respectively, and the planet can reach aCBP,PCE of up to 1.7 au and eCBP,PCE  of up to 0.67.

Standard image High-resolution image
Figure 6.

Figure 6. Same as Figure 5 but for αCE  = 1, 3, 5, and 10. For ${\alpha }_{1}$, mode(aCBP,PCE) = 1.4/1.3 au and mode(eCBP,PCE ) = 0.46/0.38 for TCP/NTCP, respectively; aCBP,PCE can reach 3.8 au, and eCBP,PCE  can be as high as 0.82.

Standard image High-resolution image
Figure 7.

Figure 7. Probability distribution functions for aCBP,PCE (left) and eCBP,PCE (right) for Kepler-38 for αCE = 0.5. Red histograms represent αM = 1∼Msun/yr, green histograms represent αM = 0.1∼Msun/yr (/hatch for NTCP, \∼hatch for TCP).

Standard image High-resolution image
Figure 8.

Figure 8. Same as Figure 7 but for αCE  = 1, 3, 5, and 10.

Standard image High-resolution image

It is important to note that the osculating orbital elements of a CBP do not describe a closed ellipse during the CE phase, and its orbital motion is not Keplerian. Instead, the orbit spirals outward with time and, as seen from Figures 3 and 4, while the binary is losing mass, aCBP  and eCBP oscillate together with the oscillations in ebin (see inset figure, lower panel). Similar behavior was observed in numerical simulations of planets around single stars experiencing mass loss (Veras et al. 2011; Adams et al. 2013), where the planet's orbital elements oscillate on the planet's period due to the different effects of mass loss at pericenter and apocenter. This is reproduced in our simulations for the binary star itself; the orbital elements of the secondary star oscillate on the binary period as seen from the inset figure in the lower panel of Figure 3. The CBP, initially on a circular orbit, quickly gains eccentricity and responds to the binary's oscillations.

Overall, the orbital expansion of the Kepler-38 CBP for ${\alpha }_{0.1}$ is fully consistent with the adiabatic approximation, and the planet gains slight eccentricity (see Table 9). The outcome for the case of ${\alpha }_{1}$ is diverse: aCBP,PCE ranges from below the corresponding adiabatic approximation up to 3.8 au, and eCBP,PCE  can reach up to 0.8 (see Table 7).

4.3. Kepler-47

The binary consists of 0.96 ${M}_{\odot }$ and 0.34 ${M}_{\odot }$ stars on a nearly circular (ebin,0 = 0.02), 7.45 day orbit that changes little in the first ∼12 Gyr. The three CBPs have initial orbits of 0.3, 0.72, and 0.99 au and masses of 2 M, 19 M, and 3 M (J. Orosz 2016, private communication), from inner to outer, respectively. For completeness, we integrate the system for the appropriate CBP masses to take into account planet–planet interactions. The main evolutionary stages are as follows.

4.3.1. The Binary

The primary star experiences an RLOF at t ∼ 12 Gyr, and the binary enters a CE stage. During this stage, Ψ(${\alpha }_{1}$) ≈ 0.02/0.06/0.1 and αcrit  ≈ 60/16/10 ${M}_{\odot }$ yr−1 for planets 1, 2, and 3, respectively. For ${\alpha }_{1}$ the time of CE mass loss (${T}_{\mathrm{CE}}$ ) is comparable to the orbital periods of planets 2 and 3, that is, TCE/PCBP ∼ 0.5–2, indicating that these two CBPs evolve in the transition regime.

As a result of the primary CE, for αCE  = 0.5/1 the binary merges as a first giant branch star and loses ∼15–40% of its mass; β < βeject, and the CBPs should remain bound even in a runaway regime. From the end of the CE to the end of the BSE simulations, the system slowly loses 50% of its mass, and aCBP,PCE of those planets that remain long-term stable after the CE (discussed below) expand adiabatically by a factor of 2.

For αCE  = 3/5/10, the binary evolves into a HeWD–MS star pair and loses 56% of its mass, and abin decreases by a factor of 4–10; here β > βeject, and the CBPs should be ejected in a runaway regime. The system does not experience further mass loss by the end of the BSE simulations.

4.3.2. The CBPs

In Figures 9 and 10 we show the evolution of aCBP  and eCBP for the three CBPs (and for ${\alpha }_{1}$ and ${\alpha }_{0.1}$) caused by the primary RLOF and CE. For simplicity here and for the rest of the Kepler  systems, we do not show the evolution of M1 and abin since it is qualitatively very similar to the case of Kepler-38. We note that, unlike the corresponding figures for Kepler-38, here the colors represent the three planets (magenta, green, and red for planets 1, 2, and 3, respectively), and the different panels represent the evolution of aCBP  and eCBP for ${\alpha }_{1}$ (upper and middle panels) and for ${\alpha }_{0.1}$ (lower panels).

Figure 9.

Figure 9. Evolution of Kepler-47 CBP's aCBP  and eCBP during the primary RLOF and CE for αCE = 0.5/1.0 (binary merger), for ${\alpha }_{1}$ (upper two panels), and for ${\alpha }_{0.1}$ (lower panel), where the colors represent planets 1 (magenta), 2 (green), and 3 (red). The dashed lines represent the corresponding adiabatic and runaway expansions. The orbital periods of planets 2 and 3 are comparable to the mass-loss timescale of the CE (TCE) and evolve in the transition regime for ${\alpha }_{1}$. All three planets evolve adiabatically for ${\alpha }_{0.1}$. PCE planet–planet interactions result in 40%/15%/75% ejection probability for the modes of aCBP,1/2/3 after 1 Myr. For ${\alpha }_{0.1}$ the three aCBP  evolve adiabatically.

Standard image High-resolution image
Figure 10.

Figure 10. Same as Figure 9 but for αCE  = 3/5/10. The orbital periods of planets 2 and 3 are comparable to the CE mass-loss timescale TCE and evolve in the transition regime for ${\alpha }_{1}$; planet 3 is ejected during the CE in 46%/81% of the simulations for NTCP/TCP, respectively. For ${\alpha }_{1}$, planets 1/2/3 are ejected in 60%/15%/85% after 1 Myr. For ${\alpha }_{0.1}$ their orbits evolve adiabatically and do not experience ejections.

Standard image High-resolution image

As seen from the lower panels in Figures 9 and 10, the orbital expansion of all three of Kepler-47's CBPs is fully consistent with the adiabatic approximation for ${\alpha }_{0.1}$, and the planets gain slight eccentricity (not shown in the figures, but listed in Table 9). The orbital evolution of planets 2 and 3 for the primary RLOF is much richer for ${\alpha }_{1}$, as shown on the upper and middle panels of Figures 9 and 10 (see also Table 7). The planets gain high eccentricities and for αCE  = 3/5/10 may even become ejected during the CE stage, where, given the complexity, uncertainties, and large parameter space of the evolution of the binary and of the CBP, here and throughout the paper we define ejection as eCBP > 0.95. Specifically, planet 3 becomes unbound in 46% of the NTCP simulations and in 81% of the TCP simulations.

We note that, during the primary CE stage, for the case of ${\alpha }_{1}$, the orbit of the inner CBP grows adiabatically, while those of the outer two do not (see upper panel of Figure 10). This is because the orbital period of the inner planet is much shorter than the duration of ${T}_{\mathrm{CE}}$, while those of the outer two are comparable. Thus a multiplanet CB system can experience two regimes of planetary orbital evolution during the same CE stage.

The dependence of aCBP,PCE and eCBP,PCE of planet 2 on Δθ0 is shown in Figures 11 and 12. The planet's orbit evolves adiabatically for ${\alpha }_{0.1}$ (upper right panels), nonadiabatically for ${\alpha }_{1}$, and remains bound during the CE in both cases, regardless of Δθ0.

Figure 11.

Figure 11. Same as Figure 5 but for Kepler-47, planet 2, and for αCE  = 0.5/1. For ${\alpha }_{1}$, mode(aCBP,PCE) = 0.6–1.33 au and mode(eCBP,PCE ) = 0.08–0.42, depending on αCE  and tidal evolution, and can reach 1.97 au and 0.58 eccentricity. The results from our numerical simulations for aCBP,PCE and for ${\alpha }_{0.1}$ are fully consistent with the adiabatic approximation. The planet remains bound regardless of Δθ0.

Standard image High-resolution image
Figure 12.

Figure 12. Same as Figure 11 but for αCE  = 3, 5, and 10. For ${\alpha }_{1}$, mode(aCBP,PCE) = 3.68/3.23 au, mode(eCBP,PCE ) = 0.71/0.75 for TCP/NTCP, aCBP,PCE can reach 7.95 au, and eCBP,PCE  can be as high as 0.89. The orbit of the planet evolves adiabatically for ${\alpha }_{0.1}$ and nonadiabatically for ${\alpha }_{1}$. The planet remains bound in both cases, regardless of Δθ0.

Standard image High-resolution image

Even if an ejection does not occur during the primary CE phase, aCBP,PCE and eCBP,PCE of all three CBPs for ${\alpha }_{1}$ are such that the planets will continue to interact with each other after the end of the CE in such a way that further ejections or collisions are possible. To study the long-term evolution of the system, we integrated the equations of motion for all five bodies (two stars and three planets) for 1 Myr, using the planets' mode(aCBP,PCE) and mode(eCBP,PCE), and randomizing their initial arguments of pericenter and true anomalies. Any planet that achieves eCBP > 0.95 is marked as ejected and removed from the simulations, and collision is defined as a planet–planet separation smaller than dmin = RP,i + RP,j.

Overall, planet 2 dominates the 1 Myr dynamical evolution and has the highest probability of remaining bound to the system. For αCE  = 0.5/1 the modes of aCBP,1/2/3 have 40%/15%/75% ejection probability in 1 Myr. Where planet 2 remains bound, its orbital elements remain mostly within the respective 68% range of mode(aCBP,PCE,2) and mode(eCBP,PCE,2).17 Here the modes of aCBP,1/2/3 have 60%/15%/85% ejection probability in 1 Myr, and the orbital elements of planet 2 again remain mostly within the 68% range of mode(aCBP,PCE,2) and mode(eCBP,PCE,2).

4.4. Kepler-64

The binary consists of 1.53 ${M}_{\odot }$ and 0.41 ${M}_{\odot }$ stars on a slightly eccentric (ebin,0 = 0.22), 20 day orbit that changes little in the first ∼3 Gyr; the CBP has an initial orbit of 0.65 au. The system evolves as follows.

4.4.1. The Binary

The primary star fills its Roche lobe at t ∼ 3.02 Gyr, and the binary enters a primary CE stage. During this stage, Ψ(${\alpha }_{1}$) ≈ 0.03 and αcrit  ≈ 32 ${M}_{\odot }$ yr−1, much larger than ${\alpha }_{1}$ and ${\alpha }_{0.1}$. However, for ${\alpha }_{1}$ the time of mass loss (TCE) is comparable to the orbital period of the CBP, that is, TCE/PCBP ∼ 1.1–3.4, depending on αCE. Thus while the CBP is theoretically in the adiabatic regime, the comparable timescales result in a nonadiabatic dynamical response.

The binary merges as a first giant branch star and loses ∼20–58% of its mass for αCE  = 0.5/1. In this regime, β > βeject for αCE  = 0.5, and for αCE  = 1, TCP and the planet should remain bound in a runaway regime even at pericenter; β < βeject for αCE  = 1.0 and NTCP, indicating a potential ejection of the CBP in a runaway regime. For αCE  = 0.5 and 1, the system continues to slowly lose mass by the end of the BSE simulations (15 Gyr), so aCBP,PCE expands adiabatically by up to a factor of 3.

For αCE = 3.0, 5.0, and 10.0 the binary shrinks by a factor of 5–25 and its mass decreases by ∼65%; here, β < βeject, and the CBP should be ejected in a runaway regime. For αCE  = 3 and TCP, the system experiences a secondary RLOF at t ∼ 6.2 Gyr and coalesces without mass loss (no changes in aCBP  or eCBP). Except for αCE  = 3 and TCP, the binary does not lose mass after the primary CE.

The binary reaches 15 Gyr as either a merged COWD or a very close HeWD–MS star binary.

4.4.2. The CBP

In Figure 13 we show the evolution of aCBP  and eCBP for ${\alpha }_{1}$ (red) and ${\alpha }_{0.1}$ (green), respectively, for αCE  = 0.5/1 (upper two panels), which result in a binary merger, and for αCE  = 3/5/10 (lower two panels). The dependence of aCBP,PCE and eCBP,PCE  on the initial phase difference between the binary and the CBP (Δθ0) is shown in Figure 14 for αCE  = 0.5/1.0 and in Figure 15 for αCE  = 3/5/10.

Figure 13.

Figure 13. Evolution of Kepler-64 CBP's aCBP  and eCBP during the primary RLOF and CE for αCE = 0.5,1.0 (upper two panels) and for αCE = 3, 5, and 10 (lower two panels), where red indicates ${\alpha }_{1}$ and green indicates ${\alpha }_{0.1}$. The binary merges at the end of the CE for αCE = 0.5, 1.0. The dashed lines represent the corresponding adiabatic and runaway expansions. For the case of ${\alpha }_{1}$ and NTCP, ∼4% of the simulations result in ejection of the CBP (defined as eCBP > 0.95).

Standard image High-resolution image
Figure 14.

Figure 14. Same as Figure 5 but for Kepler-64 and αCE  = 0.5/1. For ${\alpha }_{1}$, mode(aCBP,PCE) = 0.6–1.4 au and mode(eCBP,PCE ) = 0.16–0.62 depending on αCE  and tidal evolution; aCBP,PCE can reach 7.5 au, and eCBP,PCE  can reach 0.85. The results from our numerical simulations for aCBP,PCE and for ${\alpha }_{0.1}$ are fully consistent with the adiabatic approximation.

Standard image High-resolution image
Figure 15.

Figure 15. Same as Figure 14 but for αCE  = 3/5/10. Here mode(aCBP,PCE) = 2.4/1.6 au and mode(eCBP,PCE ) = 0.49/0.8 for NTCP/TCP, max(aCBP,PCE) = 19.6 au and max(eCBP,PCE ) =  0.95. The planet is ejected (eCBP > 0.95) in 4% of the simulations, indicated by the missing Δθ0 coverage for the green and red colors in the left-hand panels.

Standard image High-resolution image

Similar to Kepler-38, the orbital expansion of Kepler-64's CBP is fully consistent with the adiabatic approximation for ${\alpha }_{0.1}$, and again the planet gains slight eccentricity (Table 10). The outcome for the case of ${\alpha }_{1}$ is quite diverse. Depending on Δθ0, the planet can gain very high eccentricity and even become ejected (eCBP > 0.95). This is shown as missing points in Figures 14 and 15. The Kepler-64 CBP can reach such eccentricities in ∼4% of our simulations for ${\alpha }_{1}$ and αCE  = 3/5/10. Interestingly, for αCE  = 0.5 and TCP, aCBP  can decrease during the CE, and mode(aCBP,PCE) = ${0.6}_{-0.0}^{+0.6}\,[\mathrm{au}]$ is smaller than the initial orbit aCBP,0 = 0.65 au, though the mode's 68% range is significant. On the other side of the spectrum, aCBP  can reach up to ∼20 au for αCE  = 3/5/10 and NTCP.

4.5. Kepler-1647

The initial binary consists of 1.22 ${M}_{\odot }$ and 0.97 ${M}_{\odot }$ stars on a slightly eccentric (ebin,0 = 0.15), 11 day orbit that changes little in the first ∼5 Gyr. The CBP has an initial orbit of 2.7 au. The main stages of the evolution of the system are as follows.

4.5.1. The Binary

The primary star fills its Roche lobe at t ∼ 5.4 Gyr, and the binary enters a primary CE stage. During this stage, Ψ(${\alpha }_{1}$) ≈ 0.2 and αcrit  ≈ 4.6 ${M}_{\odot }$ yr−1, comparable to ${\alpha }_{1}$. Additionally, for NTCP and ${\alpha }_{0.1}$, the time of mass loss (TCE) is comparable to the orbital period of the CBP, that is, TCE/PCBP ∼ 1.3/3.0 for αCE  = 0.5/1.0, respectively. Thus the orbit of the CBP evolves in the transition regime.

Compared to the other Kepler  CBP systems, the binary star of Kepler-1647 experiences the richest evolution. For αCE  = 0.5/1 the binary merges as a first giant branch star and loses ∼15–40% of its mass. In this regime, β > βeject, and the planet should remain bound in a runaway regime even at pericenter. The system continues to slowly lose mass by the end of the BSE simulations (15 Gyr), so aCBP,PCE further expands (adiabatically) by up to a factor of 3.

For αCE = 3/5/10, the binary shrinks by a factor of 2–7 into a HeWD–MS star binary, and its mass decreases by ∼45%, close to the critical mass loss for runaway ejection (0.5). Here, β < βeject and the CBP should remain bound even in a runaway regime.

The binary experiences a secondary RLOF and CE for αCE  = 3/5/10 (see Table 6 for details), and its subsequent evolution can follow several paths. By the end of the BSE simulations, the binary (1) merges without mass loss (i.e., no changes in aCBP,PCE and eCBP,PCE ) into a first giant branch for αCE  = 3/5 and TCP, which by 15 Gyr evolves into a COWD; and (2) evolves into a PCE WD–WD binary (with mass loss, thus aCBP,PCE and eCBP,PCE  change) for αCE  = 5 (and for NTCP) and for αCE  = 10 (for both TCP and NTCP), which can experience a third and final RLOF, triggering an SN explosion (for αCE  = 5/10 and TCP).

4.5.2. The CBP

The corresponding evolution of aCBP  and eCBP for the primary RLOF and for ${\alpha }_{1}$ and ${\alpha }_{0.1}$ is shown in Figure 16. Unlike the previous systems, the ${\alpha }_{0.1}$ case for Kepler-1647 represents an adiabatic evolution only for the TCP. For ${\alpha }_{0.1}$ and NTCP, where ${T}_{\mathrm{CE}}$  and PCBP are comparable for all αCE, the CBP's orbit evolves in the transition regime (as it does for ${\alpha }_{1}$).

Figure 16.

Figure 16. Evolution of Kepler-1647 CBP's aCBP  and eCBP during the primary RLOF and CE for αCE = 0.5/1 (upper two panels) and for αCE = 3/5/10 (lower two panels), where red indicates ${\alpha }_{1}$, and green indicates ${\alpha }_{0.1}$. The binary merges at the end of the CE (at t = TCE) for αCE = 0.5/1. The dashed lines represent the corresponding adiabatic and runaway expansions. The CBP is ejected (eCBP > 0.95) in ∼5–50% of the simulations depending on αCE and the tidal evolution, for both ${\alpha }_{1}$ and ${\alpha }_{0.1}$. The latter case is nonadiabatic as TCE is comparable to the period of the CBP at the start of the simulations (PCBP,0) and the CBP can expand to large orbits, gaining very high eccentricities.

Standard image High-resolution image

Specifically, for ${\alpha }_{0.1}$ and NTCP, the CBP reaches sufficiently high eccentricities (eCBP > 0.95) in ∼5–25% of the simulations to be ejected from the system; for ${\alpha }_{1}$ the planet is ejected in ∼45–55% of the simulations. In addition, for αCE  = 0.5/1 and NTCP, and for αCE  = 3/5/10 and NTCP, aCBP  can decrease during the primary CE phase, and the corresponding mode of aCBP,PCE (1.6–2.3 au, depending on αCE, with a wide 68% range) is smaller than the initial semimajor axis of the CBP (2.71 au). Some of the simulations produce the opposite result, namely orbital expansion by up to a factor of ∼20 such that aCBP,PCE reaches ∼50 au. The corresponding dependence of aCBP,PCE and eCBP,PCE  on Δθ0 is shown in Figures 17 and 18, where the missing Δθ0 phase coverage in all panels represents planet ejection.

Figure 17.

Figure 17. Same as Figure 5 but for the primary CE stage of Kepler-1647 and αCE  = 0.5/1. Here mode(aCBP,PCE) = 1.6–3.9 au, mode(eCBP,PCE ) = 0.26–0.72 depending on αCE  and tidal evolution, max(aCBP,PCE) = 37.5 au, and max(eCBP,PCE) = 0.95. The planet is ejected in 47%/23% of the ${\alpha }_{1}$/${\alpha }_{0.1}$ simulations, indicated by the missing Δθ0 phase coverage for all but the magenta curves in the left panels and points in the right panels.

Standard image High-resolution image
Figure 18.

Figure 18. Same as Figure 17 but for αCE = 3/5/10. Mode(aCBP,PCE) = 9.6/2.3 au, mode(eCBP,PCE ) = 0.7/0.4 for NTCP/TCP, max(aCBP,PCE) = 47.1 au, and max(eCBP,PCE) = 0.95. The planet is ejected in 54%/4% of the ${\alpha }_{1}$/${\alpha }_{0.1}$ simulations.

Standard image High-resolution image

Using the respective abin,PCE and ebin,PCE and the modes of aCBP,PCE and eCBP,PCE  from the end of the preceding primary RLOF as initial conditions, we further explore the evolution of the CBP for the secondary RLOF and CE stage for the three cases where the binary evolves into a WD–WD pair with mass loss: (1) αCE  = 5, NTCP; (2) αCE  = 10, TCP; and (3) αCE  = 10, NTCP. The results are as follows.

During this secondary CE phase, for (1) and (2) Ψ(${\alpha }_{1}$) ≈ 0.4 and αcrit  ≈ 2.4 ${M}_{\odot }$ yr−1, comparable to ${\alpha }_{1}$, indicating that the CBP evolves in the transition regime. For (3) Ψ(${\alpha }_{1}$) ≈ 3.6, αcrit  ≈ 0.3 ${M}_{\odot }$ yr−1, and the evolution of the planet's orbit is in the runaway regime. The time of mass loss (TCE) is close to the orbital period of the CBP, that is, TCE/PCBP ∼ 0.3–2.5. The binary shrinks by a factor of ∼5–10 into a WD–WD pair and loses ∼60–70% of its mass. As a result, β < βeject and the CBP should become unbound in a runaway regime.

The evolution of aCBP  and eCBP during the secondary CE phase is shown in Figure 19, where the upper panel represents case (1) (αCE  = 5, NTCP), the middle panel case (2) (αCE  = 10, NTCP), and the lower panel case (3) (αCE  = 10, TCP). The CBP is ejected in more than ∼80% of the simulations. For ${\alpha }_{1}$, the CBP's aCBP  expands to a mode value of ∼8–15 au (depending on the scenario described above), with a maximum of up to 50–100 au; the mode of eCBP is in the range of 0.5–0.75, reaching a maximum of 0.95. For ${\alpha }_{0.1}$, the CBP's orbit can expand up to ∼30–50 au.

Figure 19.

Figure 19. Evolution of Kepler-1674 CBP's aCBP  and eCBP during the secondary RLOF and CE for αCE  = 5, NTCP (upper panel); αCE  = 10, NTCP (middle panel); and αCE  = 10, TCP (lower panel) for ${\alpha }_{1}$ (red) and ${\alpha }_{0.1}$ (green). The planet is ejected in ∼80–100% of the simulations.

Standard image High-resolution image

By the end of the BSE simulations, the CBP can remain bound only for case (2) as the other two scenarios trigger an SN shortly after the secondary CE.

4.6. Kepler-35

The system does not experience a CE evolution for Znominal, but the primary undergoes an RLOF and CE for Z − 1σZ. Given the sensitivity of the binary evolution to the uncertainties in Z, here we outline the results for this system as well.

4.6.1. The Binary

The binary consists of 0.88 ${M}_{\odot }$ and 0.81 ${M}_{\odot }$ stars on a slightly eccentric (ebin,0 = 0.15), 21 day orbit that changes little in the first ∼13 Gyr. The CBP has an initial orbit of 0.6 au.

The primary star fills its Roche lobe at t ≈ 12.99 Gyr, and the binary enters a primary CE stage. During this stage, Ψ(${\alpha }_{1}$) ≈ 0.03 and αcrit  ≈ 29.6 ${M}_{\odot }$ yr−1, much larger than ${\alpha }_{1}$. As a result of the CE, the binary merges as a first giant branch star and loses ∼35% of its mass for αCE  = 0.5; here β > βeject, and the planet should remain bound in a runaway regime. The system does not experience significant mass loss by the end of the BSE simulations (15 Gyr), so there are no major changes in aCBP.

For αCE  = 1/3/5/10 the binary orbit shrinks by a factor of ∼2–10, loses ∼35% of its initial mass, and evolves into a He–WD pair; here β > βeject, and the planet should remain bound in a runaway regime.

4.6.2. The CBP

Overall, for ${\alpha }_{0.1}$ the orbital evolution of the CBP is consistent with the adiabatic approximation. For ${\alpha }_{1}$, mode(aCBP,PCE) ∼ 0.8–1 au, and the planet gains an eccentricity of 0.2–0.4; the min/max ranges can be significant. In particular, for αCE  = 0.5, ${\alpha }_{1}$, and NTCP, the CBP reaches sufficiently high eccentricities (eCBP > 0.95) in ∼4% of the simulations to be ejected from the system; the CBP remains bound in all other cases. For ${\alpha }_{1}$ and NTCP, aCBP,PCE can reach up to 7–10 au, with eCBP,PCE  > 0.9.

5. DISCUSSION AND CONCLUSIONS

Several key results emerge from our study:

(1) Five of the nine Kepler  CBP hosts experience at least one RLOF and a CE phase for their default parameters (masses, metallicities) by the end of our BSE simulations (15 Gyr); Kepler-35 experiences a primary RLOF for Z − 1σZ. Depending on the treatment of the CE stage, the binaries either merge after a primary or secondary CE (typically for αCE  = 0.5/1) or evolve into very short-period WD–MS pairs (for αCE  = 1/3/5/10)18 ; Kepler-1647 evolves into a WD–WD pair after a secondary RLOF and CE for αCE  = 5/10. Two systems trigger a double-degenerate supernova explosion: Kepler-34 for primary RLOF, and Kepler-1647 for a third RLOF and αCE  = 5, TCP, and αCE  = 10, NTCP. The binary systems lose a tremendous amount of mass—up to ∼60–70%—during the CE stages and can shrink to sub-R⊙ orbital separations.

(2) From the dynamical perspective presented here, Kepler CBPs predominantly remain bound to their host binaries after the respective CE phases even for mass-loss rates as high as 1.0 ${M}_{\odot }$ yr−1. There are only four scenarios where the respective CBP has a 100% probability of becoming unbound: (a) for Kepler-34 where the SN explosion occurs with the first CE; (b) for the secondary CE phase of Kepler-1647 for ${\alpha }_{0.1}$, αCE  = 10, TCP; (c) for the SN caused by the third RLOF of Kepler-1647, αCE  = 5, NTCP; and (d) SN caused by the third RLOF of Kepler-1647, αCE  = 10, TCP. In all other scenarios (106 total, not accounting for the initial phase differences Δθ0), the CBPs remain bound in the majority of cases (only Kepler-1647 has nonnegligible ejection probability, again highly dependent on Δθ0).

For mass-loss rates of 0.1 ${M}_{\odot }$ yr−1, the orbits of the CBPs evolve adiabatically—well reproduced by REBOUNDx—except for Kepler-1647. The orbital expansion of some of the Kepler  CBPs is consistent with the adiabatic approximation even for mass-loss rates of 1.0 ${M}_{\odot }$ yr−1 (e.g., inner planet of Kepler-47, Kepler-64); in other cases, the mode of aCBP,PCE is smaller than the adiabatic approximation (e.g., Kepler-1647).

According to the analytical prescription, some systems should retain their CBPs even in a runaway mass-loss regime (Ψ ≫ 1). Our code fully reproduces the runaway approximation for single-star systems.

(3) The transition regime (Ψ ∼ 0.1–1)—where the evolution of some of Kepler  CBPs falls for ${\alpha }_{1}$—is complex, and the final eccentricities and semimajor axes of the planets depend both on the treatment of the CE stage (i.e., αCE, treatment of tides) and on the initial configuration of the system (i.e., Δθ0). We find that the orbits of Kepler  CBPs can expand by more than an order of magnitude over the course of a single CBP year, reaching aCBP,PCE of tens of astronomical units (during the secondary CE of Kepler-1647, aCBP  can expand from ∼10 au to ∼100 au; see Table 8).

Multiplanet CBP systems add yet another level of complexity because they can experience both regimes simultaneously. For example, where the orbit of the inner Kepler-47 CBP expands adiabatically, the middle and outer CBP orbits grow nonadiabatically—all during the primary CE phase.

Overall, the CE-induced orbital evolution of CBPs is a dynamically rich and complex process in which a planet can migrate adiabatically during one CE phase and nonadiabatically during another. As a result, a CBP cannot experience the same CE twice.

(4) If CE mass-loss rates are indeed high (e.g., ∼1.0 ${M}_{\odot }$ yr−1)—as suggested by theoretical work—and the orbits of Kepler-like CBPs evolve (and survive) according to our simulations, we should expect to detect potentially highly eccentric planets orbiting PCE systems on very large orbits. Alternatively, if Ψ ≪ 1, then we should find PCE CBPs on low-eccentricity orbits. Interestingly, recent observational efforts have produced a number of PCE eclipse-time variations (ETV) CBP candidates on wide, notably eccentric orbits (e.g., Zorotovic & Schreiber 2013 and references therein), suggesting a possible connection with PCE Kepler-like CBPs.

In Figure 20 we show aCBP,PCE versus eCBP,PCE of Kepler's CBPs19 and compare them to those of the PCE ETV CBP candidates.20 We caution that the comparison is not direct as the CE stages of the different Kepler systems do not occur all at the same time, and the PCE ETV CBPs do not have identical ages. Instead, given the handful of known targets—Kepler's nine compared to estimated millions of similar CBPs (Welsh et al. 2012) and a dozen PCE ETV CBPs—the lower two panels on the figure portray a potential distribution of an underlying PCE CBP population with a range of ages and evolutionary stages.

Figure 20.

Figure 20. Upper six panels: aCBP,PCE vs. eCBP,PCE  for Kepler's CBPs for various CE treatments (green, blue, and red symbols) and initial conditions (${\rm{\Delta }}{\theta }_{0}=0\div1$), along with the respective ejection probabilities. Lower two panels: comparison between Kepler's aCBP,PCE and eCBP,PCE  and those of the currently known PCE ETV CBP candidates (black symbols) at the end of the respective Kepler  CEs (left) and at the end of our BSE simulations (15 Gyr, right). Given the associated observational and numerical uncertainties, our forward-evolution results are qualitatively consistent with the observed aCBP,PCE vs. eCBP,PCE  distributions. See text for details.

Standard image High-resolution image

Overall, the PCE orbital configurations of Kepler's CBPs are qualitatively consistent with those of the observed population of the currently known PCE ETV CBP candidates.21 Thus our results assist in both interpreting the nature of these candidates and in guiding future observational efforts to discover new systems. Such discoveries will provide a deeper understanding of the evolution of planets in a binary star system and also much-needed observational constraints on the stellar astrophysics of the complex CE phase, as "CE is one of the most important unsolved problems in stellar evolution," according to Ivanova et al. (2013), "and is arguably the most significant and least-well-constrained major process in binary evolution" (but also see Taam & Sandquist 2000; Taam & Ricker 2010 and Webbink 2008 for alternative reviews).

(5) A CB body gravitationally perturbs its host binary star, and if the latter is eclipsing, these perturbations can be manifested as deviations from linear ephemeris in the measured stellar eclipse times. The cause of these variations can be either dynamical, where the third body's influence changes the orbital elements of the binary star, or a light travel time effect, where the tertiary object and the binary revolve around a common center of mass. ETVs are indeed a powerful and highly productive method of discovering and studying stellar triple and higher order systems (e.g., Borkovits et al. 2016; Orosz 2015, and references therein), and they have recently been used to study CBPs as well. In particular, the former effect has been measured for several of the Kepler CBP systems and used to constrain the respective planetary masses (Doyle et al. 2011; Welsh et al. 2012; Kostov et al. 2016). The latter effect has been suggested as the cause for measured ETVs for a number of PCE EB systems with proposed CB companions (as mentioned above; also see Volschow et al. (2016) and references therein).

Assuming that the orbits of the Kepler  CBPs studied here evolve only dynamically and their masses remain constant (i.e., no changes due to, for example, interactions with a CE-triggered CB disk), it is informative to evaluate the respective planets' ETV-based detectability after the CE stages of the relevant systems. To calculate the expected amplitudes of the two effects discussed above, Adyn and ALTTE, we use the formalism of Borkovits et al. (2012, 2015).

The respective amplitudes and periods for Kepler-35 (Mp = 0.13 MJup) and Kepler-1647 (Mp = 1.5 MJup)—the only two systems with well-constrained CBP masses—are listed in Table 11 (for the respective range of binary and CBP parameters listed in Tables 8) and compared to the detected signals for the two PCE CBP candidates NN Ser and HW Vir. As seen from Table 11, detection of such ETVs is feasible in terms of both the amplitudes and periods of the expected signal, which are qualitatively similar to the ETV signals of the two PCE CBP candidates.

Table 11.  Expected Dynamical and Light Travel Time ETV Amplitudes and Periods after the Respective CE Stages for Kepler CBPs with Known Masses (Kepler-1647 and Kepler-35)

αM Adyn (min/max) ALTTE(min/max) P(min/max) Notes
(${M}_{\odot }$ yr−1) (s) (s) (year)  
Kepler-1647ba        
1.0 0.2/13.4b 68/1682 2.4/292
0.1 0.1/13 154/680 8.1/75
Kepler-1647bc        
1.0 0./0.2 736/9680 31/1490
0.1 0./0.2 668/5420 27/624
Kepler-35ba        
1.0 0.1/56.5 4.3/36.9 0.7/17
0.1 0.2/44.2 5.0/5.1 0.9/0.9
NN Ser bc        
  ∼5/30d ∼5/30 ∼7/15
HW Vir bc        
  ∼55/550d ∼50/550 ∼13/55

Notes. For comparison, we also list the detected ETV amplitudes and periods for the CBP candidates in NN Ser and HW Vir.

aAfter primary CE. bMin/max for the range of parameters from Table 8. cAfter secondary CE. dShowing the combined effect of inner/outer CBPs (instead of min/max).

Download table as:  ASCIITypeset image

Another option for the detection of a PCE CBP is to directly image the planet. The benefits for such detection are twofold: (a) the contrast ratio between the binary and the CBP decreases after each CE phase as the primary or secondary evolves from a luminous MS star to a much fainter WD; and (b) the angular separation between the binary and the CBP increases22 (e.g., Hardy et al. 2015). There is potentially a third benefit, where the CBP can accrete mass after the CE stage (e.g., Zorotovic & Schreiber 2013; Bear & Soker 2014) and becomes brighter, thus further decreasing the contrast ratio.

Recently, Hardy et al. (2015) used VLT/SPHERE to observe the V471 Tau system—an eclipsing binary star with measured ETVs—with the goal to directly image a substellar-mass CB candidate suspected to be the cause of the ETVs. Given the known age of the system and the masses of the binary and the CB candidate, direct detection of such a tertiary body should have been well within the capabilities of the instrument. Hardy et al. (2015), however, report a null detection, casting doubt on the interpretation of the measured ETV signal (but see Vaccaro et al. 2015 for an alternative explanation). Nevertheless, based on our results for the dynamical survivability of CBPs around evolving binary stars, and on estimates of the occurrence rates of planets orbiting short-period MS binary stars (e.g., Welsh et al. 2012), we encourage the continuation of such direct imaging efforts, aimed at both PCE, very short period binaries and PCE, WD-coalesced binaries where the contrast ratio is even more favorable.

(6) The closer a planet's orbit is to a single star during the MS stage, the worse its prospects are for survival when the star ascends the RGB and AGB. For example, Mercury, with an orbit of 0.46 au, will be destroyed as the Sun evolves off the MS (e.g., Schroder & Connon Smith 2008). In contrast, our results suggest that, despite their dramatic evolution, close binary stars may in fact be better for the survival of planets compared to single stars. For example, the Kepler-38 CBP has an initial orbital separation similar to Mercury's, and yet it survives the evolution of a binary star with a total mass of ∼1 ${M}_{\odot }$. A CBP with an even smaller orbit—as small as 0.3–0.35 au—can survive the evolution of a 1.3–1.4 ${M}_{\odot }$ (total mass) central binary star, as is the case for the inner planet in Kepler-47. A planet with the same orbit around a single star of the same mass will have no chance for survival (Villaver & Livio 2007). Thus, despite being a disruptive stage for the binary star itself, the CE phase may in fact promote planet survivability.

(7) With the exception of Kepler-1647, the Kepler CBPs currently orbit their binary star hosts within a factor of 2 of the critical limit for dynamical stability (Holman & Wiegert 1999, HW hereafter), that is, ${a}_{\mathrm{CBP},0}(1-{e}_{\mathrm{CBP},0})/{a}_{\mathrm{crit},0}\sim 2$ (the factor for Kepler-1647 is ∼7). However, as the binaries shrink and lose mass during the CE and the planets migrate to larger, eccentric orbits, this ratio will change. To calculate acrit,PCE—the PCE critical limits (for those scenarios where the binaries do not coalesce, i.e., αCE  = 3/5/10)—we use Equation (3) from HW using a binary mass ratio μ = MA,PCE/(MA,PCE + MB,PCE).23 As the CE phases result in a variety of abin, aCBP,PCE, and eCBP,PCE, here we quote only the most constraining critical limits, where the respective separation between abin and aCBP,PCE (1 − eCBP,PCE ) is at a minimum. For ${\alpha }_{1}$, the limits for Kepler-38, Kepler-47, Kepler-64, Kepler-1647 (NTCP), and Kepler-1647 (TCP) in terms of aCBP,PCE (1 − eCBP,PCE)/acrit,PCE are 3.3, 7.9, 5.3, 15.9, 14.7, and 29.5, respectively, indicating that the CBPs remain in a dynamically stable regime.

While it is beyond the scope of this study, we note that tidal evolution of the binary orbit prior to the CE phase may affect the CBP orbits. Specifically, as most of the planets are currently orbiting close to their host binaries, a pre-CE decay in abin means that strong mean motion resonances (MMR) may sweep over the planet, leading to eccentricity excitation and potential destabilization. As an example, abin of Kepler-38 decreases from 31.6 R to 24.6 R prior to the CE phase for TCP (see Table 4), indicating that the CBP's orbit crosses 6:1, 7:1, and 8:1 MMR. The MMR crossings are (a) 7:1, 8:1, and 9:1 for Kepler-47 (planet 1); (b) 7:1 through 11:1 for Kepler-64; and (c) 99:1 through 116:1 for Kepler-1647.

(8) "...It is interesting to consider the situation at a much earlier time in the past when the primary was near the zero-age main sequence...," note Orosz et al. (2012a) for the case of Kepler-38b: "The primary's luminosity would have a factor ≈3 smaller...." While Kepler-38b is currently closer to its binary host than the inner edge of the habitable zone (HZ) for the system, four of the Kepler  CBPs (Kepler-16, Kepler-47c, Kepler-453, Kepler-1647) reside in the HZ. As their host binaries evolve, however, the location of the HZ will change.

Unlike the case for single stars where both the HZ and a (surviving) planet's orbit will expand during the RGB and AGB stages (Villaver & Livio 2007), as a close binary star evolves through a CE stage the HZ can shrink while its CBP migrates outward. The shrinking is caused by the luminosity drop of the central binary as one of its stars rapidly transitions from the MS to a compact object. Thus a CBP residing in the HZ during the pre-CE binary will leave the zone after the CE phase (e.g., Kepler-1647), whereas a CBP that is initially interior to the HZ may migrate to a PCE orbit coinciding with the PCE HZ.

Consequently, it would be equally interesting to consider the configuration of a CBP system at a much later time in the binary evolution. For example, the current orbital configuration of Kepler-35b is such that the planet's orbit is internal to the HZ (Welsh et al. 2012). However, as its host binary evolves through the CE phase, for αCE  = 1/3/5/10 the primary star evolves into a HeWD (while the secondary remains on the MS as a 0.81 ${M}_{\odot }$ star), and the CBP migrates to mode(aCBP,PCE) = 0.84–1.0 au, mode(eCBP,PCE) = 0.0–0.4 (see Tables 8). This migration places the planet near the conservative HZ of 0.98–1.77 au (for 1 M; Kopparapu et al. 2014), using the luminosity, 0.4 LSun, and temperature of the secondary star, 5200 K (Welsh et al. 2012), and ignoring the flux contribution of the WD.

5.1. Limitations

The results presented here are based on the assumptions that the CBPs do not interact with the material ejected from their binary star. However, numerical studies have indicated that this material is neither lost isotropically from the binary during the CE phase, nor does it all become unbound. Instead, 1–10% of the ejecta may fall back into a CB disk according to Kashi & Soker (2011), and Passy et al. (2012) suggest that ∼80% of the ejected material may remain gravitationally bound to the binary (also see Kuiper 1941; Shu et al. 1979; and Pejcha et al. 2016 for mass loss outflows through the L2 Lagrange point24 ). Either of these scenarios would significantly complicate the dynamical evolution of the system as the CBP could accrete material and gain mass and also experience migration similar to that during planetary formation.25 Such accretion of material of a different specific angular momentum will change the orbital evolution of the planet, as will gravitational interaction with the bulk gas in an accretion-favorable environment such as a CB disk. Additionally, interactions between CBPs and the CE ejecta of close binary stars (including ejection or destruction of the planets) may also play an important role in the elusive mechanisms responsible for the shaping, morphology, and chemistry of planetary nebulae (see Bear & Soker 2016 for triple star origin of asymmetric planetary nebulae; for recent reviews see Zijlstra 2014 and Jones 2015, and references therein). We note that the CE-triggered dynamical disturbances discussed here occur on timescales of ∼1 year and are thus "instantaneous" compared to any subsequent planet migration, which occurs on much longer timescales.

PCE CB disks are also suspected to trigger a new episode of planet formation, a second-generation scenario (e.g., Verhoelst et al. 2007; Perets 2010; Tutukov & Fedorova 2012 and references therein), and have also been suggested to be the birthplace of several PCE CBP candidates (Schleicher & Dreizler 2014; Volschow et al. 2014).26 Indeed, a recent discovery of dust around NN Ser supports the presence of a remnant PCE disk (Hardy et al. 2016). Alternatively, based on angular momentum constraints, Bear & Soker (2014) argue that these CBP candidates formed together with their host binary stars (i.e., first generation). Mixed scenarios, where the CBP forms with the binary, survives the CE stage and subsequently acts as a seed for the accretion of more mass, are also possible (Armitage & Hansen 1999; Perets 2010; Schleicher et al. 2015), showcasing the complexity and diversity of the formation and evolution mechanisms of planets around close binary stars.

These complications are, however, beyond the scope of our study, and we defer their implementation to future work.

5.2. Conclusions

We have presented numerical studies of the dynamical evolution of the nine Kepler  CBP systems as their host binary stars undergo CE phases. Five of the systems undergo at least one RLOF and CE phases; Kepler-1647 experiences three RLOFs. Two systems trigger a double-degenerate SN explosion. Depending on the treatment of the CE phase (i.e., efficiency and tidal evolution), the binaries either coalesce or shrink into very close WD–MS or WD–WD pairs of stars. Despite the dramatic reconfiguration of the binary stars during these violent evolutionary phases, the planets predominantly survive the CE stage even for mass-loss rates of 1 ${M}_{\odot }$ yr−1. The PCE orbital configurations of these CBPs depend on the rate at which the binary loses mass. For mass-loss rates of 1 ${M}_{\odot }$ yr−1, the CBPs migrate nonadiabatically outward to highly eccentric orbits, while for slower mass-loss rates (0.1 ${M}_{\odot }$ yr−1) our simulations reproduce the expected adiabatic orbital expansion. Overall, the PCE semimajor axes and eccentricities of Kepler  CBPs are qualitatively consistent with those of the currently known PCE CBP candidates. The mode of orbital expansion depends on the particular configuration of the system, such that a CBP can migrate adiabatically during the primary CE stage but nonadiabatically during the secondary CE stage. Multiplanet CBP systems can experience both modes simultaneously: an inner planet can migrate adiabatically while an outer does so nonadiabatically, both occurring during the same CE stage. Our results also indicate that planets are more likely to survive around evolving close binary stars than around evolving single stars, thus improving the discovery prospects for CBPs in PCE systems.

We thank the anonymous referee for the insightful comments that helped us improve this paper. The authors are grateful to Jarrod Hurley and Marten van Kerkwijk for valuable discussions. VBK gratefully acknowledges support by an appointment to the NASA Postdoctoral Program at the Goddard Space Flight Center. DT was supported by a postdoctoral fellowship from the Centre for Planetary Sciences at the University of Toronto at Scarborough and is grateful for additional support from the Jeffrey L. Bishop Fellowship. This work was supported in part by NSERC grants to RJ. We acknowledge conversations with Daniel Fabrycky, Nader Haghighipour, Kaitlin Kratter, Boyana Lilian, Jerome Orosz, and William Welsh.

APPENDIX: LIST OF ABBREVIATIONS AND PARAMETERS

  • BSE: Binary star evolution code
  • CE: Common envelope
  • CBP: Circumbinary planet
  • EB: Eclipsing binary
  • NTCP: No tidal circularization path (tides "OFF" in BSE)
  • PCE: Postcommon envelope
  • RLOF: Roche-lobe overflow
  • TCP: Tidal circularization path (tides "ON" in BSE)

  • ${\alpha }_{M}$ : Common-envelope mass-loss rate
  • ${\alpha }_{1}$: 1 ${M}_{\odot }$ yr−1 mass-loss rate
  • ${\alpha }_{0.1}$: 0.1 ${M}_{\odot }$ yr−1 mass-loss rate
  • αcrit: Critical common-envelope mass-loss rate
  • αCE: Common-envelope efficiency parameter
  • aCBP,0: Initial semimajor axis of the circumbinary planets
  • aCBP,PCE: Semimajor axis of the circumbinary planets at the end of the common envelope
  • β: Ratio between initial and final mass of the system
  • βeject: Runaway ejection ratio between initial and final mass of the system
  • eCBP,0: Initial eccentricity of the circumbinary planets
  • eCBP,PCE: Eccentricity of the circumbinary planets at the end of the common-envelope phase
  • ψ: Common-envelope mass-loss index
  • ${T}_{\mathrm{CE}}$: Common-envelope mass-loss timescale

Footnotes

  • In line with thermodynamics nomenclature, this is an isoeccentric process.

  • We note that both the total amount of mass lost and the rate at which it is lost will affect the evolution of a CBP's orbit during a CE phase.

  • 10 

    Throughout this paper, we use subscripts "bin" and "p" or "CBP" to denote the binary star and the CBP, and M1 and M2 to denote the mass of the primary and the secondary star, respectively.

  • 11 

    Also see Equation (48) from Veras et al. (2011) for a comprehensive treatment.

  • 12 

    A collision between a dense core and a star can also trigger a CE phase.

  • 13 

    BSE also accounts for collisions that do not proceed through CE evolution, and for the possibility of core sinking depending on the stellar types (which may lead to rejuvenation).

  • 14 

    Interestingly, their metallicities are typically subsolar.

  • 15 

    The BSE upper limit of Z = 0.03 restricts some systems with greater Z, such as Kepler-64, to this extreme.

  • 16 

    For the planet to be ejected, the final stellar mass must be smaller than M1(runaway, eject).

  • 17 

    Except in one simulations where planet 1 is ejected, planet 2 migrates to a 5.67 au, 0.71-eccentricity orbit, and planet 3 migrates to a 0.55 au, 0.66-eccentricity orbit.

  • 18 

    Which can also merge for a secondary-triggered CE phase.

  • 19 

    For Kepler-47 we only show the middle planet (planet 2) as it has the highest probability to survive both the CE phase and subsequent planet–planet interactions.

  • 20 

    For those candidates that do not have published eccentricities, we set the respective eCBP to zero.

  • 21 

    With the caveat that the latter are much too massive.

  • 22 

    Such that the minimum increase is typically for adiabatic orbital expansion.

  • 23 

    Since the primary star becomes the less massive component of the binary after the CE phase.

  • 24 

    Mass loss through L2 results in several possible outcomes, e.g., isotropic or equatorial wind, CB disk; for details, see Table 1 and Figures 12 and 13 of Pejcha et al. (2016).

  • 25 

    There are, however, two potential benefits of the former in terms of detection: a more luminous planet would be more amenable to direct imaging efforts, and a more massive planet would cause stronger ETVs.

  • 26 

    Second generation has also been proposed for the origin of the first detected exoplanet (Wolszczan & Frail 1992).

Please wait… references are loading.
10.3847/0004-637X/832/2/183