Articles

DEUTERIUM BURNING IN MASSIVE GIANT PLANETS AND LOW-MASS BROWN DWARFS FORMED BY CORE-NUCLEATED ACCRETION

, , , , and

Published 2013 June 3 © 2013. The American Astronomical Society. All rights reserved.
, , Citation Peter Bodenheimer et al 2013 ApJ 770 120 DOI 10.1088/0004-637X/770/2/120

0004-637X/770/2/120

ABSTRACT

Using detailed numerical simulations, we study the formation of bodies near the deuterium-burning limit according to the core-nucleated giant planet accretion scenario. The objects, with heavy-element cores in the range 5–30 M, are assumed to accrete gas up to final masses of 10–15 Jupiter masses (MJup). After the formation process, which lasts 1–5 Myr and which ends with a "cold-start," low-entropy configuration, the bodies evolve at constant mass up to an age of several Gyr. Deuterium burning via proton capture is included in the calculation, and we determined the mass, M50, above which more than 50% of the initial deuterium is burned. This often-quoted borderline between giant planets and brown dwarfs is found to depend only slightly on parameters, such as core mass, stellar mass, formation location, solid surface density in the protoplanetary disk, disk viscosity, and dust opacity. The values for M50 fall in the range 11.6–13.6 MJup, in agreement with previous determinations that do not take the formation process into account. For a given opacity law during the formation process, objects with higher core masses form more quickly. The result is higher entropy in the envelope at the completion of accretion, yielding lower values of M50. For masses above M50, during the deuterium-burning phase, objects expand and increase in luminosity by one to three orders of magnitude. Evolutionary tracks in the luminosity versus time diagram are compared with the observed position of the companion to Beta Pictoris.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

There is considerable debate over the question of defining a precise boundary between the class of objects known as "planets" and those known as "brown dwarfs." It has been suggested that the two types of objects could be distinguished by their formation mechanism; however, it is generally difficult to deduce this property from observations of specific objects. Nevertheless, there is a well-defined minimum in the mass distribution (actually Msin i), for substellar companions to G and K main-sequence stars, in the range 20–30 MJup (Lovis et al. 2006; Sahlmann et al. 2011; Schneider et al. 2011), suggesting that objects near the deuterium-burning limit can "form like planets." These authors suggest that this minimum does correspond to a (somewhat imprecise) dividing line between formation mechanisms and that the upper limit to planet masses should be set at about Msin i ≈ 25 MJup. However, no break is seen near this mass in the distribution of free-floating objects observed in the Sigma Orionis young cluster (Peña Ramírez et al. 2012) down to 4 MJup. Together, the observations imply that formation mechanisms do not define a unique mass boundary between planets and brown dwarfs.

Another commonly used criterion to classify planets, brown dwarfs, and stars is based on nuclear fusion that does or does not occur within the object. Brown dwarfs are defined to be those objects that at some point in their evolution become hot enough in their interiors to burn a majority of the deuterium that was initially present in the object; however, they never become hot enough to burn 1H by the proton-plus-proton reaction in a self-sustaining manner as true stars do. On the other hand, the term planet is applied only to objects that will not burn much deuterium. This criterion was used by Burrows et al. (1997) to separate the two types of objects, and the dividing line was stated to be ≈13 MJup, where MJup = 1.898 × 1030 g. This dividing line depends on the helium mass fraction, the deuterium abundance, and the metallicity, and Spiegel et al. (2011) found that for a reasonable range of parameters, 50% of the initial D is burned in the mass range 12–14 MJup. The evolutionary models used to establish the criterion have a uniform chemical composition, a defined total mass in the vicinity of the 13 MJup limit, constant in time, an initial radius of about 2–3 RJup where RJup = 7.15 × 109 cm, and an initial photospheric temperature (Teff) of about 2500 K. The corresponding initial luminosities are (2–3) × 10−3L. A starting model of this type has become known as a "hot-start" model, characterized by a relatively high initial entropy (Marley et al. 2007).

The question of whether, either for brown dwarfs or planets, the formation mechanism actually leads to such hot-start initial conditions is still under investigation. For objects formed either by collapse of interstellar clouds or by fragmentation in a protostellar disk by gravitational instability, it is plausible that the hot-start initial condition could be reached (Baraffe et al. 2002). In the case of gravitational instability, Galvagni et al. (2012) have found, from three-dimensional (3D) numerical simulations, that the entropy of newly formed clumps, near the point where molecular dissociation sets in at the center, is high, possibly consistent with a hot start. However, the dominant process for giant planet formation is most likely the core-nucleated accretion mechanism, in which solid particles first accumulate to form a heavy-element core, then later when the core has attained roughly 5–10 M, gas is captured from the disk. A particular set of evolutionary calculations based on this theory (Marley et al. 2007) shows that once the planet has become fully formed, its entropy is relatively low, with luminosities on the order of 10−5 to 10−6L. The low entropy is a direct consequence of the assumption made in these calculations that, during the phase of rapid gas accretion, all of the accretion energy is radiated away at the accretion shock at the planet's surface. Thus, the core accretion process can lead to a "cold start." However, the shock treatment is approximate, and the accretion flow cannot actually be modeled correctly with one-dimensional spherically symmetric calculations. Thus, other possibilities can arise. Mordasini et al. (2012) show that core accretion formation calculations in which none of the energy is radiated at the shock lead to hot-start conditions very similar to those assumed by Baraffe et al. (2003) and Burrows et al. (1997). Furthermore, intermediate "warm" states are also possible outcomes (Spiegel & Burrows 2012). In the core-accretion picture, also, the chemical composition is not uniform because of the presence of the core, which turns out for the case of a Jupiter mass planet to fall in the range 4–20 M (Movshovitz et al. 2010).

A massive object of 25 MJup formed by core accretion (Baraffe et al. 2008) has been shown to burn all of its initial deuterium despite the presence of a heavy-element core of 100 or a few hundred M. Cold-start models, including the core and calculations of the formation phase, have been investigated to determine the D-burning mass limit (Mollière & Mordasini 2012). The results show that the limit still falls within the range 12–14 MJup. The purpose of the present paper is to present further formation calculations for bodies formed by core-nucleated accretion that end up with a total mass in the 10–15 MJup range in the low-entropy state, and to investigate the effect of various possible initial conditions, as well as physical parameters during the formation stage, on the corresponding deuterium-burning limit.

2. COMPUTATIONAL METHOD

The evolutionary calculations for giant planets are started at the point where the heavy-element core has a mass of about 1 M, and are carried through the entire formation process as well as the subsequent contraction/cooling phase at constant mass, up to a final age of several Gyr. The assumptions and computational procedures were described in detail in previous publications (Pollack et al. 1996; Bodenheimer et al. 2000; Hubickyj et al. 2005; Lissauer et al. 2009; Movshovitz et al. 2010). The early phase of the formation process is dominated by the accretion of planetesimals onto the core; during this phase the gaseous envelope has low mass, ≪1 M, and a low accretion rate compared to that of the core. The latter is given by

Equation (1)

where $\pi R^2_{\rm capt}$ is the effective geometrical capture cross section, σ is the surface density of solid particles (planetesimals) in the protoplanetary disk, Ωp is the planet's orbital frequency, and Fg is the gravitational enhancement factor, which is obtained from the calculations of Greenzweig & Lissauer (1992). The planetesimal radius is taken to be 50 km for the cases with a central star of 2 M and 100 km for the cases with a star of 1 M (see Table 1). The smaller size, or a reasonable distribution of planetesimal sizes, tends to reduce the formation time but has little effect on the basic results of this paper.

Table 1. Input Parameters

Run M/M Distance σ ρneb Tneb Opacity αt Miso
(AU) (g cm−2) (g cm−3) (K) (M)
1A 1 5.2 10 9 × 10−11 115 gs 1.0 × 10−2 11.6
1B 1 5.2 10 9 × 10−11 115 ngs 1.0 × 10−2 11.6
1C 1 5.2 4 3.7 × 10−11 115 gs 1.0 × 10−2 2.9
2A 2 9.5 4 1.8 × 10−11 125 ngs 1.0 × 10−2 12.6
2B 2 9.5 4 1.8 × 10−11 125 ngs 4.0 × 10−3 12.6
2C 2 9.5 6 2.8 × 10−11 125 ngs 1.0 × 10−2 23.2

Download table as:  ASCIITypeset image

If no gaseous envelope is present, then Rcapt = Rcore, the radius of the heavy-element core. However, even if the envelope mass is relatively small compared with the core mass, the planetesimals interact with the envelope gas, are slowed down by gas drag, and are subject to ablation and fragmentation. The trajectories of planetesimals through the envelope are calculated (Podolak et al. 1988), and the effective Rcapt is determined. The material that is deposited in the envelope is then allowed to sink to the core, as discussed by Pollack et al. (1996). Calculations by Iaroslavitz & Podolak (2007) show that this assumption is valid at least for the organic and rock components of the planetesimals. The ices, however, can dissolve in the envelope, so that our "core mass" is somewhat overestimated; the quoted value actually refers to the total excess of heavy-element material, above the solar abundance, in the entire planet. Erosion of the core and possible mixing of some core material into the envelope is not considered. This process has been shown to be unlikely for the case of Jupiter (Lissauer & Stevenson 2007), but such estimates have not been extended to the case of planets in the 10 MJup range.

The structure of the hydrogen–helium envelope is calculated according to the differential equations of stellar structure (Kippenhahn & Weigert 1990), which assume hydrostatic equilibrium, a spherically symmetric mass distribution, radiative or convective energy transport, and energy conservation. The energy sources are provided by planetesimal accretion, contraction of the gaseous envelope, and cooling. The additional energy source provided by deuterium burning is included in the later phases of accretion and during the constant-mass final cooling phase, once the mass has exceeded 10 MJup and internal temperatures exceed ≈105 K. The full set of equations, supplemented by calculation of the mass accretion rates onto the core and the envelope, and of the planetesimal trajectories, is solved by the Henyey method (Henyey et al. 1964).

At the inner boundary of the envelope the radius is set to Rcore, which is determined from its mean density. During the earlier phases of the evolution, when the envelope mass is less than about 0.1 MJup, the core is assumed to be a mixture of rock and ice with a mean density of 3.0 g cm−3. During the later phases, when the pressure at the base of the envelope increases to values above ∼1011 dyne cm−2, an ANEOS equation of state with 50% rock and 50% ice (Thompson 1990) for the core is used to determine its mean density, which can increase to 60 g cm−3 or higher. In the hydrogen–helium envelope, the equation of state is taken to be given by the tables of Saumon et al. (1995), which take into account the partial degeneracy of the electrons as well as non-ideal effects. The chemical composition is taken to be near-solar, with X = 0.70, Y = 0.283, and Z = 0.017, where X, Y, Z are, respectively, the mass fractions of hydrogen, helium, and heavy elements. The tables of course do not include a Z component, so the Y component was adjusted upward to partially compensate.

The Rosseland mean opacity during the formation phase combines the low-temperature atomic/molecular calculation of Alexander & Ferguson (1994) with the interstellar grain opacities of Pollack et al. (1985). The opacity values of the grain component are reduced by a factor 50 to approximately represent the reduction caused by grain growth and settling in the protoplanet (Podolak 2003; Movshovitz & Podolak 2008). However, in two of the runs the grain growth and settling are calculated in detail in the temperature range 100–1800 K as described in Movshovitz et al. (2010). The grain size distributions and the opacities are recalculated in every layer at every time step in that temperature range. These opacities are important in regulating the rate at which the envelope can contract, and therefore the rate at which it accretes gas. However, once the envelope is well into the rapid gas accretion phase, at about 0.25 MJup, the gas accretion rate is limited by the physical properties of the protoplanetary disk near the planet, and the precise values of envelope opacity assume a less important role. Once the planet reaches its final mass, say 12 MJup, the grains are assumed to settle rapidly and to evaporate in the interior. For the final contraction/cooling phase at constant mass, the molecular opacities of Freedman et al. (2008) are used, with solar composition, up to a temperature of 3500 K. At and above that temperature, with any reasonable opacity, the interior is convective.

At the outer surface of the envelope, the mass addition rate of gas, during the earlier phases of accretion, is determined by the requirement that the planet radius Rp match the effective accretion radius, which is given by (Lissauer et al. 2009)

Equation (2)

where cs is the sound speed in the disk, RH is the Hill sphere radius, and Mp is the total mass of the planet. The constant K ≈ 0.25 is determined by 3D numerical simulations which calculate the accretion rate of gas from the protoplanetary disk onto the planet (Lissauer et al. 2009). As a result, in the limit where RH is small compared with the Bondi accretion radius $GM_p/c_s^2$, Reff = 0.25 RH.

Additional boundary conditions at the surface depend on the evolutionary phase. During the early phases when Mp < 0.25 MJup, the density and temperature are set to constant values appropriate for the protoplanetary disk, ρneb and Tneb, respectively. The density ρneb is determined from the assumed value of σ using a standard gas-to-solid ratio of 70 and Hp/ap = 0.05, where Hp is the (Gaussian) disk scale height and ap is the distance of the planet from the star. However, at some point during the rapid gas accretion phase, the mass addition rate required by condition (2) exceeds the rate at which matter can be supplied by the disk. The disk-limited rates, based on 3D hydrodynamic simulations, are described in the next section. During that phase, the boundary conditions at the actual surface of the planet, whose radius falls well below Reff, are determined through the properties of the accretion shock at this surface, as described in detail by Bodenheimer et al. (2000). The basic assumption is that practically all of the gravitational energy released by the infalling gas is radiated away at the shock; this energy escapes through the infalling envelope ahead of the shock. This assumption defines the "cold start" for planetary evolution.

During the final phase of cooling at constant mass, the planet becomes isolated from the disk and the surface boundary conditions change again, to those of a blackbody in hydrostatic equilibrium

Equation (3)

where σB is the Stefan–Boltzmann constant, Teff is the surface temperature, L is the total luminosity, and κ, P, and g are, respectively, the photospheric values of Rosseland mean opacity, pressure, and acceleration of gravity. Insolation from the star is not included.

Significant deuterium burning in the mass range considered begins near the end of the phase of rapid gas accretion. The burning occurs via the reaction

Equation (4)

with an energy release Qdp = 5.494 MeV per reaction. The initial deuterium abundance by mass fraction is set to 4 × 10−5, consistent with the value derived from the local interstellar medium (Prodanović et al. 2010). The reaction rate (reactions per second per gram) is taken from the Nuclear Astrophysics Compilation of Reaction Rates (Angulo et al. 1999):

Equation (5)

where T6 is the temperature in 106 K, ρ is the density in cgs, X1H is the mass fraction of 1H, and X2H is the mass fraction of 2H (deuterium). This rate is then multiplied by the screening factor, which takes into account ion–ion and ion–electron screening in partially degenerate dense plasmas (Potekhin & Chabrier 2012). The energy generation epsilon, per gram per second, is then obtained, zone by zone, from the rate multiplied by Qdp in the appropriate units. To get the change in the deuterium abundance during one time step, it is assumed that the planet interior is fully convective and therefore fully mixed. This assumption is valid for the planets considered during the phase of contraction and cooling, even if no deuterium is burned. The convective velocities of order 10–100 cm s−1, calculated according to the mixing-length approximation, give a mixing timescale far shorter than the D-burning timescale. The reaction rate multiplied by zone mass is integrated over the entire envelope and used to calculate the abundance change.

Given the central stellar mass M*, the solid surface density σ, and the distance of the planet from the star ap, the isolation mass for the solid material is

Equation (6)

where C is the number of Hill-sphere radii on each side of the planetary core from which it is able to capture planetesimals; C = 4 in our simulations. Once the core mass approaches Miso, the dMcore/dt slows down drastically, and beyond that point, gas accretion continues and surpasses the core accretion rate. The core mass increases to a value of about $\sqrt{2} M_\mathrm{iso}$ at crossover, when Mcore = Menv (Pollack et al. 1996). This phase of relatively slow accretion rates onto both core and envelope is known as "Phase 2."

3. DISK-LIMITED GAS ACCRETION RATES

The epoch of rapid gas accretion in the core-nucleated accretion model generally begins soon after the envelope mass, Menv, exceeds the core mass, Mcore, as can also be shown by means of simple thermodynamical arguments (D'Angelo et al. 2011). In a proto-solar nebula at ∼5 AU, this condition typically occurs when the planet mass Mp = Mcore + Menv is between ∼10 and a few tens of Earth masses. After this point, the planet's envelope tends to contract very rapidly, limited only by the rate of energy escape at the surface, and a high rate of gas accretion is required to maintain the condition Rp = Reff. At or about 0.25 MJup this condition can no longer be met, the rate is set by the ability of the protoplanetary disk to deliver gas to the planet, and Rp contracts well within Reff.

There are various regimes of disk-limited gas accretion (see D'Angelo & Lubow 2008). For the purpose of this study, we are mainly interested in the high-mass limit RHHp, where $R_{H}=a_{p} \sqrt [3] {M_{p}/\left(3M_{\star }\right)}$ is the Hill radius of the planet and Hp is the disk thickness at the planet's orbital radius, ap. In this regime, disk-limited accretion rates can be affected by disk–planet gravitational interactions if tidal torques overcome viscous torques. Assume that the turbulent (kinematic) viscosity of the disk at the orbital distance of the planet is given by $\nu _{t}=\alpha _{t} H^{2}_{p} \Omega _{p}$, where Ωp is the local Keplerian rotation frequency of the disk and αt is the viscosity parameter. Then tidal torques exerted by the planet on the disk exceed viscous torques exerted by adjacent disk rings on each other if

Equation (7)

where Δ = max (Hp, RH) and f is a factor of order unity (see, e.g., D'Angelo et al. 2011, and references therein). When the left-hand side of Equation (7) is much greater than the right-hand side, a gap forms in the disk surface density along the planet's orbital radius.

We estimated disk-limited accretion rates, $\dot{M}_{p}$, using high-resolution 3D hydrodynamical simulations of a planet embedded in a protoplanetary disk. We used an approach along the lines of D'Angelo et al. (2003). We considered a disk with a constant aspect ratio of Hp/ap = 0.05 and with the parameter αt ranging from 4 × 10−4 to 2 × 10−2. The unperturbed surface density of the disk is taken to be a power law of the distance from the star with exponent −1/2. The planet is kept on a fixed circular orbit, and the continuity and Navier–Stokes equations (written in terms of linear and angular momenta) for the gas are solved in a reference frame co-rotating with the planet.

The disk is assumed to be vertically isothermal and, radially, the temperature drops as the inverse of the distance from the star. At the radial distance ap, the temperature is $T_{p}=(\mu _d m_{\mathrm{H}}/k_{B}) H^{2}_{p} \Omega ^2_{p}$, which is equal to 53.8μd K at 5 AU from a solar-mass star (μd indicates the mean molecular weight of the disk's gas). The relatively simple equation of state adopted here (the pressure pTρ, where ρ is the mass density and the temperature T is a given function of the orbital distance) allows us to write the fluid equations in a non-dimensional form so that the gas accretion rates can be expressed in terms of $a^{2}_{p}\Sigma _{p}\Omega _{p}$, where Σp is the unperturbed gas surface density of the disk at the planet location (i.e., that the disk would have in the absence of the planet). Furthermore, the planet's mass enters the calculations only via its ratio to the mass of the star. We considered values of the ratio Mp/M up to 0.02. The planet's gas accretion rate starts to decline for planet masses greater than the value for which the inequality in Equation (7) is satisfied. This critical mass is larger within more viscous disks. The equation also suggests that there is a dependence on the disk thickness, which, however, was not explored here. We notice that reasonable values of Hp/ap for evolved disks, between 1 and 10 AU, range from ≈0.03 to ≈0.05 (e.g., D'Angelo & Marzari 2012), affecting the right-hand side of Equation (7) by a factor of less than three (for RH > Hp), whereas uncertainties on αt are much larger, spanning two orders of magnitude or more. An unperturbed surface density with a power index different from that adopted here (−1/2) may also affect the accretion rates. We expect these effects to be small, especially when tidal torques exerted by the planet drastically modify the surface density, which is typically the case in the models discussed here.

In the calculations, the disk domain extends in radius as close to the star as 0.1 ap (and 0.05 ap, in some calculations) and as far as 9.4 ap. More vigorous perturbations exercised by larger mass planets cause the inner/outer disk radius to decrease/increase with increasing planet-to-star mass ratio. Figure 1 shows the surface density, averaged in azimuth, for a case in which Mp = 10−2M and αt = 10−2. Notice that the low densities in the disk inside the orbit of the planet are a consequence of tidal torques and planet accretion (see Lubow & D'Angelo 2006), with possibly some impact from the finite radius of the grid inner boundary (0.05ap in the calculation shown in the figure). The analysis of Lubow & D'Angelo (2006), where applicable, suggests that the effects of the finite inner grid radius are small.

Figure 1.

Figure 1. Average surface density of gas in a disk, as a function of distance from the star, that tidally interacts with a planet with Mp = 10−2M. The orbital radius of the planet is ap. The surface density is plotted in scaled units of $M_{\star }/a^{2}_{p}$. The aspect ratio of the disk is Hp/ap = 0.05, and the turbulent viscosity parameter is αt = 10−2.

Standard image High-resolution image

High resolution in and around the planet's Hill sphere is achieved by means of multiple nested grids (D'Angelo et al. 2002, 2003) centered on the planet's position. This methodology allows us to solve the fluid equations (locally around the planet), and hence to resolve the accretion flow, on length scales of order 0.01 RH, or ≈7 RJup at 5 AU.

The gas that orbits the planet deep within its gravitational potential is eventually accreted. We assume that gas can be accreted within a spherical region of radius 0.1 RH (or 0.05 RH in some models), centered on the planet. The amount of accreted gas is proportional to the amount of gas available in the region (see D'Angelo & Lubow 2008, and references therein). In these calculations, accreted gas is removed from the computational domain but not added to the mass of the planet in order to achieve a stationary accretion flow (see Lissauer et al. 2009).

We determined an interpolation procedure for the disk-limited gas accretion rates obtained from calculations, by performing piece-wise parabolic fits (in a logarithmic plane) to each $(\dot{M}_{p},M_{p}/M_{\star })$ data set, relative to a given value of the turbulent viscosity. Two such fitting curves are shown in Figure 2 (explicit expressions are provided in the Appendix). Linear interpolations among these curves provide the accretion rate at the desired viscosity parameter, αt. In doing so, we derived a function $\dot{M}_{p}=\dot{M}_{p}(M_{p},M_{\star },a_{p},\Sigma _{p},\alpha _{t})$, which we employ in our planet formation calculations. We recall that Σp here represents the disk gas surface density at the planet's orbital radius, in the absence of the planet. An analytic formula is available for the accretion rate at the low-mass end (D'Angelo & Lubow 2008). However, in the formation calculations the use of these curves is not required until Mp exceeds ≈0.25 MJup.

Figure 2.

Figure 2. Disk-limited gas accretion rates as a function of the planet-to-star mass ratio, Mp/M, for Hp/ap = 0.05 and for two values of the disk turbulent viscosity parameter: αt = 4 × 10−3 and 10−2. Symbols are data obtained from 3D hydrodynamics simulations: filled pentagons/open circles refer to the lower/higher viscosity case. The solid and dashed lines represent results from the fitting procedure outlined in the text. In the units of $\dot{M}_{p}$, Σp represents the unperturbed disk gas surface density and Ωp the Keplerian rotation rate at the planet's orbital radius, ap.

Standard image High-resolution image

During the disk-limited gas accretion phase, the solid accretion rate is arbitrarily limited to a fraction of the value at crossover; the precise value has practically no effect on the results. We do not expect the core-accretion prescription to be valid at this stage, because most solids in the disk will not be in the form of planetesimals, and we do not have the capability to model giant impacts. As the planet reaches within 2% of the desired final mass (e.g., 12 MJup), the gas accretion rate, already quite low, is smoothly reduced to zero.

4. CALCULATIONS AND RESULTS

A recent paper on deuterium burning in objects formed through the core-accretion scenario (Mollière & Mordasini 2012) considered the basic case of a body forming at 5.2 AU in a disk around a 1 M star with a solid surface density of σ = 10 g cm−2 and Tneb = 150 K. Their study compared results obtained by varying the following parameters: initial entropy of the object after formation (hot start versus cold start), helium abundance, metal abundance, initial deuterium mass fraction, σ, which determines the final planet core mass, and maximum gas accretion rate. Their calculations differ from ours in the phase of rapid gas accretion, when disk-limited rates apply. They take that rate to be an arbitrary parameter, while we use the 3D simulations mentioned above (Section 3) to determine it. Here we concentrate on cold-start models and consider a somewhat different set of parameters: stellar mass, formation position of the planet in the disk, solid surface density σ, method of computation of the opacity in the planetary envelope during the formation phase, and protoplanetary disk viscosity parameter αt. The planet's core mass is determined through the calculation itself, and it depends on the first three of these quantities. Note that the final core masses found in our calculations fall in the range 4.8–31 M, while those of Mollière & Mordasini (2012) are higher (30–100 M). The formation and evolution are assumed to take place at a fixed orbital radius.

The parameters for the runs are given in Table 1. The columns in the table give, respectively, the run identifier, the mass of the central star in M, the distance of the planet from the star, the solid surface density σ, the density ρneb at the surface of the planet during the earlier phases when this surface connects with the disk, the temperature Tneb at the surface during the same phases, the method of opacity calculation during the formation phase—that is, whether it includes the calculation of grain settling and coagulation (gs) or not (ngs)—the value of the viscosity parameter αt in the disk during the phases of disk-limited gas accretion, and the isolation mass (Equation (6)).

Some results for the six runs are presented in Table 2. Each run is given two lines, the first for a final planet mass that burns less than half of its deuterium, the second for a nearby mass that burns more than half. The columns give, respectively, the run identification, the final planet mass in MJup, the total time to reach the final mass (the formation time), the final core mass, the central temperature (Tc, f, at the core/envelope interface) just after formation, the maximum central temperature during D-burning, the central density ρc, f just after formation, the planet's radiated luminosity Lc, f just after formation, and the mass fraction of deuterium that remains after 4 Gyr of evolution, in units of the initial D mass fraction of 4 × 10−5.

Table 2. Selected Results

Run Mfinal/MJup tform Mcore/M Tc, f Tmax ρc, f log Dfinal/Dinit
(yr) (K) (K) (g cm−3) (Lf/L)
1A 12.26 1.19 × 106 16.8 2.60 × 105 2.60 × 105 49.7 −6.22 0.895
1A 12.48 1.20 × 106 16.8 2.62 × 105 3.54 × 105 51.4 −6.23 0.164
1B 12.14 2.67 × 106 16.8 2.69 × 105 2.69 × 105 47.0 −5.84 0.860
1B 12.26 2.68 × 106 16.8 2.72 × 105 3.18 × 105 47.7 −5.91 0.328
1C 13.5 4.37 × 106 4.83 2.31 × 105 2.31 × 105 65.1 −6.57 0.938
1C 13.6 4.39 × 106 4.83 2.33 × 105 4.27 × 105 65.9 −6.57 0.292
2A 11.9 2.14 × 106 18.7 2.73 × 105 2.81 × 105 41.4 −5.64 0.767
2A 12.0 2.14 × 106 18.7 2.76 × 105 3.06 × 105 41.8 −5.63 0.435
2B 12.0 3.23 × 106 18.8 2.78 × 105 2.87 × 105 44.6 −5.73 0.582
2B 12.1 3.26 × 106 18.8 2.80 × 105 3.28 × 105 45.1 −5.72 0.320
2C 11.6 8.75 × 105 31.0 3.37 × 105 3.46 × 105 27.9 −4.80 0.670
2C 11.7 8.75 × 105 31.0 3.39 × 105 3.56 × 105 28.3 −4.80 0.390

Download table as:  ASCIITypeset image

4.1. Results for 1 M

Run 1A, with standard parameters of 1 M, 5.2 AU, and σ = 10 g cm−2, was originally calculated by Movshovitz et al. (2010) through most of the formation phase, including the detailed calculation of grain opacity (their run σ10). Their run, whose characteristics are listed in that paper, ended at the beginning of disk-limited gas accretion, with a core mass of 16.8 M and an envelope mass of 56.8 M at a total elapsed time of 1 Myr. In this work, the run was continued through the disk-limited phase with αt = 10−2 (Section 3) up to the mass range required for deuterium burning. The maximum gas accretion rate was 2.5 × 10−1 M yr−1 at a total planet mass of 96 M, declining to 10−2 M yr−1 at 10 MJup. For several different masses in that range, the accretion was terminated, the opacity was reset in the surface layers to the values given by Freedman et al. (2008), and the evolution was followed at constant mass up to Gyr times. The runs were terminated when deuterium burning ceased, and the mass M50, where 50% of the original deuterium had been burned, was determined. In the Run 1A, the total formation time at M50, up to termination of accretion, was 1.2 Myr, well within the lifetime of protoplanetary disks.

The planetary luminosity as a function of time for three different final masses in Run 1A is shown in Figure 3, where it is compared with that typically obtained in a "hot-start" model. In the case of 16 MJup, just after formation the central temperature Tc, f = 2.8 × 105 K, too low for substantial burning on a short timescale, even though ρc, f = 80 g cm−3. Under these conditions the screening correction factor to the nuclear reaction rate is high, about 88. Consequently, deuterium burning can take place at relatively low temperatures compared to those (≈106 K) where deuterium burns in solar-mass stars. The central temperature Tc as a function of time (at the core/envelope interface) gradually increases as a result of slow deuterium burning and is accompanied by a slight increase in radius. When Tc reaches 3.2 × 105 K, a rapid increase in burning occurs, leading to a peak in luminosity at about 108 yr. At the peak about 60% of the deuterium has burned, and Tc is near its maximum of 5.1 × 105 K. At the same time the radius has increased from 7.4 × 109 cm to 1.25 × 1010 cm; then it contracts again after the luminosity peak. At the end of the evolution essentially all the deuterium has burned. A similar process, involving a rapid increase in deuterium burning in the context of a slowly accreting brown dwarf, was studied by Salpeter (1992); he denotes the event a "deuterium flash." The radii for the three masses, as well as for the hot-start case, are shown in Figure 4. The general result that cold-start models result in a radius increase during deuterium burning agrees with the previous results of Mollière & Mordasini (2012).

Figure 3.

Figure 3. Luminosity (in solar units) as a function of time for Run 1A during the post-formation deuterium-burning phase for three different planet masses. Solid curve: 16 MJup; dashed curve: 12.5 MJup; short-dash dot curve: 12 MJup. The long-dash dot curve shows the results for a hot-start model of 10 MJup (Baraffe et al. 2003).

Standard image High-resolution image
Figure 4.

Figure 4. Radius (in 1010 cm) as a function of time for Run 1A during the post-formation deuterium-burning phase for three different planet masses. Solid curve: 16 MJup; dashed curve: 12.5 MJup; short-dash dot curve: 12 MJup. The long-dash dot curve shows the results for a hot-start model of 10 MJup (Baraffe et al. 2003).

Standard image High-resolution image

In the case of 12.5 MJup, right after formation the central temperature is lower, only 2.6 × 105 K, with a central density of 52 g cm−3 and a screening factor of 70. It takes almost 109 yr for rapid deuterium burning to start, at Tc = 2.8 × 105 K, with the burning occurring on a much longer timescale than in the case of 16 MJup. Eventually about 98% of the deuterium is burned, and the luminosity peak, which is somewhat lower, is shifted to later times. At the peak, about half of the D has burned, and this point is also close to the maxima in radius and Tc. In the case of 12 MJup only 6% of the deuterium is burned, and no peak in luminosity appears. At M50 itself, the peak involves only a factor two increase in luminosity. In the peaks, the total energy ∫Ldt is found to agree closely with the total energy available from D-burning, given by the quantity Qdp/md × Mp × Xdfd, where Qdp, the energy production per reaction, is expressed in erg, md is the mass of a deuterium atom, Mp is the planet mass, Xd is the initial mass fraction of deuterium, and fd is the fraction of the initial D that burned. The total energy is about 2.5 × 1045 erg for the 12.5 MJup case.

Results for two runs whose final masses closely bracket M50 are shown in Table 2. The main result of this case is that M50 = 12.37 MJup with a heavy-element core mass of 16.8 M. By way of comparison, a cold-start model with a core, calculated by Mollière & Mordasini (2012) with about the same basic parameters (1 M, σ = 10 g cm−2, ap = 5.2 AU), with a similar helium mass fraction of 28%, but with some differences in assumptions and computational procedure, gives M50 = 12.6 MJup.

The maximum Tc at the core/envelope interface for M50 in this case is close to 3.2 × 105 K, a very sensitive function of mass. Whether significant D-burning occurs depends sensitively on this temperature. If it reaches, say, 2.5 × 105 K, practically no D is burned for the corresponding mass of 12.0 MJup. If it reaches 4.0 × 105 K, practically all (98%) of the D is burned for the corresponding mass of 12.5 MJup. Once the threshold is reached, energy deposition from burning increases the temperature, which increases the reaction rate, as it is proportional to T12. The resulting expansion leads to a near thermal equilibrium, with the energy produced from D-burning matched closely by the total radiated luminosity.

Run 1B differs from 1A only with respect to the calculation of the opacity resulting from grains in the protoplanetary envelope during the formation phase. As mentioned above, in Run 1A this opacity is obtained through detailed consideration of grain settling and coagulation (Movshovitz et al. 2010). In 1B a table of interstellar grain opacities is used, reduced by a factor of about 50. The characteristics of this run, up to a mass of about 1 MJup, are very similar to those listed for Run 1sG in Lissauer et al. (2009). The crossover mass is 16.16 M, the crossover time is 2.31 Myr, and the onset of disk-limited rapid gas accretion occurs at a core mass of 16.8 M and a time of 2.41 Myr. Note that the evolution time up to this point is 2.4 times longer than in Run 1A. Note also that the core mass is the same as in Run 1A; the substantial difference in opacity, which can be up to two orders of magnitude in certain (ρ, T) regions, has practically no effect on the core mass.

Here, the disk-limited accretion rates are used to continue the evolution up to the D-burning mass range. The luminosity as a function of time up to the end of accretion is shown in Figure 5. The results for D-burning after that time show that M50 = 12.20 MJup, not significantly different from the results of Run 1A. As Table 2 shows, Tc, f in Run 1B, at the same final mass, is slightly higher than that in Run 1A, just after formation. Correspondingly, ρc, f is slightly lower. These small differences indicate a slightly higher entropy for 1B after formation, as indicated by the slightly higher luminosity at this point. The increased envelope opacity in 1B as compared with 1A results in slower heat loss and tends to keep internal temperatures higher. However, this effect is almost compensated by the fact that the formation time is more than twice as long in 1B. Even the slight increase in Tc, f in 1B as compared with 1A allows M50 to be pushed to a slightly lower mass.

Figure 5.

Figure 5. Luminosity (in solar units) as a function of time for Run 1B (long-dashed curve), Run 1C (short-dash long-dash curve), Run 2A (solid curve), and Run 2C (short-dashed curve) during the formation phase.

Standard image High-resolution image

Run 1C differs from Run 1A in that σ is reduced to 4 g cm−2, a value only slightly greater than that in a minimum-mass solar nebula (Weidenschilling 1977). Grain settling and coagulation are included in the opacity calculation. The earlier portions of this run, up to the onset of disk-limited gas accretion, are described in Movshovitz et al. (2010), their run σ4. The time to reach this point, 3.5 Myr, is considerably longer than in Run 1A, first, because the core accretion rate is considerably lower, and second, because the lower isolation mass results in reduced luminosity and reduced gas accretion rate during Phase 2 (Pollack et al. 1996). The crossover mass is 4.09 M, and the core mass at the time of onset of disk-limited accretion is 4.74 M.

The calculations were continued up to the point where gas accretion terminated, at which point the core mass was 4.8 M. The total time to reach 12 MJup was about 4.1 Myr, and to 14 MJup, about 4.5 Myr. The peak disk-limited accretion rate was 1.0 × 10−1 M yr−1 at 0.3 MJup, a factor of 2.5 lower than in Run 1A because of the reduction in Σp by the same factor. By the time the total mass was 5 MJup the rate was down to 1.8 × 10−2M yr−1, and at 10 MJup it had further declined to 4.3 × 10−3 M yr−1. Much of the time during the disk-limited accretion phase was spent in accreting the last 1–2 MJup to reach the D-burning point. The luminosity as a function of time for this run, up to the end of accretion, is shown in Figure 5.

The luminosity versus time plots for Run 1C during the D-burning phase look similar to those for 1A, except in this case M50 noticeably increases to 13.55 MJup. The reduced core mass in 1C (4.8 M) as compared to that in 1A (16.8 M) is clearly associated with the difference, in agreement with the results of Mollière & Mordasini (2012). In our calculations, the core equation of state gives a core radius of 3.8 × 108 cm for the Run 1C core of mass 4.8 M when the total mass is 12 MJup. For the core of 16.8 M in Run 1A, at the same total mass, the radius is 6.0 × 108 cm. Thus, at the core boundary, the gravitational potential is more negative, and the gravity is about 40% greater in 1A than in 1C. The calculated values of Tc, f are ≈2.6 × 105 K and ≈2.1 × 105 K in Runs 1A and 1C, respectively.

It follows from the equation of hydrostatic equilibrium (Mollière & Mordasini 2012) that in a convective envelope the adiabatic temperature gradient at the interface should be proportional to the core gravity, so a higher gravity most probably gives a higher temperature. However, this statement is inconclusive. We calculated static models for a planet of 12 MJup, all with the envelope entropy of Run 1C, with core masses ranging from 0 to 15 M. We found practically no difference in Tc as a function of Mcore, with Tc decreasing by less than 1% when the core mass increases from 0 to 15 M.

The real source of the difference in Tc, f between Runs 1A and 1C is the entropy in the envelope. The lower Tc, f and higher ρc, f for 1C as compared with 1A indicate a lower entropy, which is consistent with the fact that the luminosity just after formation is lower by more than a factor two in 1C (Table 2). The values of entropy just after formation for a planet of 12 MJup in Runs 1A and 1C are, respectively, 8.02 and 7.52 kB per baryon. The entropy is determined through the physical processes that occur during the entire formation phase; for example, the formation time for Run 1C is almost four times longer than that for 1A, and the same opacities were used, which suggests a lower entropy. Thus, there exists a qualitative understanding of the relation between core mass and Tc, f, but a quantitative theory, apart from the numerical simulations, is quite difficult.

In Run 1C, Tc, f = 2.1 × 105 K is the maximum reached for a final mass of 12 MJup, and it is insufficient for D-burning. In the case of Run 1A, the corresponding Tc, f is much closer to the threshold required for burning. Thus, the planet with the higher Mcore is able to produce significant D-burning at a lower total mass. As Table 2 shows, in the mass range for Run 1C where D-burning begins, just above 13.5 MJup, Tc, f is somewhat less (2.3 × 105 K) than in the corresponding mass range for Run 1A. However, to compensate, ρc, f is higher, about 65 g cm−3, and the screening factor at the center has increased to 160. Again, the lower entropy at formation for 1C, as compared with 1A, a result of various processes associated with the accretion of core and envelope, leads to a higher M50.

4.2. Results for 2 M

The formation phases of Runs 2A and 2C, for a central star of 2 M, are illustrated in Figure 5, which gives the luminosity as a function of time, and Figure 6, which gives the core mass, envelope mass, and total mass as a function of time. Run 2A differs from 1A in that the planet is placed 9.5 AU away from a star of 2 M, in a disk with σ = 4 g cm−2. In a minimum mass solar nebula, scaled to the mass of this star, the corresponding value would be 2 g cm−2.

Figure 6.

Figure 6. Planet mass (in M) as a function of time for Run 2A and Run 2C during the formation phase. The final masses are 15 MJup and 12 MJup, respectively. For Run 2A, the solid curve gives the core mass, the dotted curve the envelope mass, and the short-dash dot curve the total mass. For Run 2C, the short-dashed curve gives the core mass, the long-dashed curve the envelope mass, and the long-dash dot curve the total mass.

Standard image High-resolution image

The isolation mass, however, is quite similar to that in 1A, 12.6 rather than 11.6 M. The opacity during the formation phase of 2A is taken from a table of interstellar grain opacities, reduced by a factor of 50, as in Run 1B. However, the comparison between 1A and 1B showed that these opacities have little effect on M50. The formation time is longer in 2A than in 1A because of the longer dynamical time at the larger distance, the reduced solid surface density, and the somewhat higher envelope opacity. However, these effects are partially compensated for by the smaller planetesimal size (50 km in 2A; 100 km in 1A), which increases the capture cross section $\pi R^2_\mathrm{capt}$, and by the increased gravitational focusing factor Fg at the larger distance.

The first luminosity peak for Run 2A (Figure 5) occurs at t = 3.54 × 105 yr, with log L/L = −5.14, with Mcore = 9.3 M, with Menv = 0.024 M, and with $\dot{M}_\mathrm{core} = 6.67 \times 10^{-5}$M yr−1. This peak corresponds to the maximum in the accretion rate of solids onto the core. The crossover mass (Figure 6) of 17.6 M is reached in 2.0 × 106 yr. The second, much higher luminosity peak at 2.07 × 106 yr corresponds to the phase of rapid gas accretion up to a final mass of 15 MJup. At that time the maximum gas accretion rate is 2.2 × 10−1 M yr−1 and Mp = 0.47 MJup. Formation is complete, up to 15 MJup, in 2.2 × 106 yr. The final Mcore = 18.7 M is slightly higher than in Run 1A.

The luminosity as a function of time during the later deuterium-burning phase is shown for five different final masses in Run 2A in Figure 7. In the cases of 16 and 15 MJup, practically all (>99%) of the deuterium is burned; in the case of 13.5 MJup, about 92% is burned; for 13.0 MJup, 75% is burned; and for 12.0 MJup, just over 50% is burned. Thus, the value for M50 ≈ 11.95 MJup is very close to the values obtained for Runs 1A/1B despite substantial differences in assumptions and initial conditions. As discussed in the comparison between Runs 1A and 1C, the somewhat larger Mcore in 2A as compared to 1A is the main reason for the slightly lower M50 in 2A. After formation, 2A has a slightly higher Tc, f than 1A and slightly lower ρc, f, leading to a slightly higher entropy.

Figure 7.

Figure 7. Luminosity (in solar units) as a function of time for Run 2A during the post-formation deuterium-burning phase for five different planet masses. Long-dash dot curve: 16 MJup; solid curve: 15 MJup; short-dashed curve: 13.5 MJup; long-dashed curve: 13 MJup; short-dash dot curve: 12 MJup.

Standard image High-resolution image

Comparing the luminosity curves for a mass of 16 MJup in Figures 3 and 7, they look very different but in fact are consistent. In Run 2A (Figure 7) the higher Tc, f (because of the somewhat higher Mcore) allows D-burning to start earlier than in Run 1A, and the value of L at the starting point is a factor of four higher. In fact, the full widths of the two curves are quite similar, the peak values agree to better than a factor two, and the integrated luminosities over time of the two curves agree to within 10%.

Run 2B has exactly the same parameters as 2A except that αt is reduced by a factor 2.5, which affects the gas accretion rates during the disk-limited phase. Thus, the formation time in 2B turns out to be a factor of 1.5 longer at 3.2 × 106 yr, but still within the range of observed disk lifetimes. Table 2 shows that for final mass 12 MJup the fractions of deuterium burned are in agreement for runs 2A and 2B, within the uncertainties of the calculations. Thus, αt has practically no effect on M50 in this case. Run 2B has a slightly lower entropy than 2A at 12 MJup, 8.2 kB per baryon versus 8.25, and therefore a slightly higher M50. Thus, it appears that the longer time during disk-limited accretion in Run 2B has only a weak effect on both M50 and the entropy, at the same Mcore.

Run 2C has the same parameters as 2A except that the solid surface density σ is increased by a factor 1.5 to 6 g cm−2. The first luminosity peak (Figure 5) occurs at t = 3.07 × 105 yr with log L/L = −4.45 at Mcore = 15.6 M. The crossover mass (Figure 6) is reached at t = 7.88 × 105 yr with a value of 30.7 M. The maximum luminosity in the second peak is above log L/L = −1, at t = 7.915 × 105 yr and a total mass of 0.62 MJup. The higher σ with respect to Run 2A results in a markedly higher Mcore = 31 M and a markedly shorter formation time (8.75 × 105 yr at M50). Despite these relatively large differences, the value for M50 in 2C is only 2.5% smaller than in 2A. At the end of the formation phase, central temperatures are higher and central densities are lower in 2C as compared with 2A. Also, the screening factor is only 14 in 2C compared with 41 for 2A. The entropy at formation, for a final mass of 12 MJup, is higher in 2C, 9.08 kB/baryon as compared with 8.25, corresponding to a higher luminosity at that point. The slope in the (Mcore, M50) diagram between Mcore= 18.7 and 31 M is −0.024, a result which differs somewhat from that of Mollière & Mordasini (2012). They obtain a slope (in the same units) of −0.01, although for a different core mass range, 30 to 100 M.

Plots of luminosity versus time are shown in Figure 8 for several different masses in Run 2C. As in Figure 7 the higher masses give higher peak luminosity at earlier times than the lower masses, and at M50 there is only a very small peak. The L(t) curve for 15 MJup starts at a higher value and reaches a maximum sooner than for the same mass in Run 2A, because of the higher internal temperature, but the value of log L at the peak is about the same. At 13.7 and 15 MJup practically all of the initial D is burned. At 12 MJup, 72% is burned, while at 11.7 MJup, which is very close to M50, 61% is burned.

Figure 8.

Figure 8. Luminosity (in solar units) as a function of time for Run 2C during the post-formation deuterium-burning phase for four different planet masses. Solid curve: 15 MJup; long-dashed curve: 13.7 MJup; short-dash dot curve: 12 MJup; short-dash curve: 11.7 MJup. The long-dash dot curve shows a hot-start model for 10 MJup (Baraffe et al. 2003). The cross gives the position and error bars for the companion to Beta Pic (Bonnefoy et al. 2013).

Standard image High-resolution image

Plots of Tc versus time, during the deuterium-burning phases, are shown for masses 12 and 15 MJup in Figure 9, where they are compared with the results from Run 2A. The plot shows the effect of varying the core mass at fixed total mass, and of varying the total mass at fixed core mass. For example, for Run 2A at 15 MJup the maximum Tc is 4.6 × 105 K, while for 12 MJup it is only 3.06 × 105 K and is reached at a much later time. The vertical portions of these curves show the effect of rapid gas accretion from about 1 MJup to the final mass of either 12 or 15 MJup. The two nearly horizontal curves are for a total mass of 12 MJup and core masses of 18.7 (lower; Run 2A) and 31 M (upper; Run 2C). The higher core mass results in a higher temperature by a factor of about 1.3. In the case of the 31 M core, about 75% of the deuterium is burned; in the 18.7 M core, a little over 50%. Note that the D-burning occurs late in the evolution, where small peaks in the temperature are seen. The remaining two curves correspond to a total mass of 15 MJup, with the same two core masses just mentioned. The D-burning occurs earlier than in the case of the lower total mass, and the higher core mass again gives a higher maximum Tc. In both cases for 15 MJup practically all the D is burned, and the residual mass fraction is smaller (8.9 × 10−11) for the higher core mass as compared to the lower (2.6 × 10−9).

Figure 9.

Figure 9. Central temperature Tc (at the core/envelope interface) as a function of time for four cases. Solid curve: Run 2A (15 MJup); long-dashed curve: Run 2A (12 MJup); dash-dot curve: Run 2C (15 MJup); short-dashed curve: Run 2C (12 MJup). The phases of rapid gas accretion and final phases of constant mass with deuterium burning are shown. Core masses for Runs 2A and 2C are 18.7 and 31 M, respectively.

Standard image High-resolution image

In Run 2C with 15 MJup, Tc goes up to about 5 × 105 K and there are actually two minor peaks. Figure 10 illustrates in more detail how various quantities vary during this phase. In this case, with a high Tc, f, nuclear burning starts very early. During most of the phase, the object is not in thermal equilibrium. The first maximum in Tc occurs when about 25% of the D has burned, close to the time of the maximum in the nuclear burning luminosity Lnuc. Here Lnuc is well above the radiated luminosity L, and the extra power goes into expansion, resulting in slight cooling of the interior. When half the deuterium has burned (1.3 × 107 yr), there is a maximum in luminosity and radius, corresponding to the slight minimum in Tc. Then contraction along with a slow decrease in nuclear burning leads to slight heating, and the second maximum occurs when 98% of the D has burned. This maximum corresponds to the time when Lnuc starts to drop rapidly and to fall well below L. Beyond that point, even though contraction is occurring, there is insufficient burning to maintain the high temperature, and the object enters its final cooling phase. In contrast, in the case of 12 MJup, the main D-burning in Run 2C takes place at practically constant Tc, radius, and L, with a slight maximum in Tc of 3.67 × 105 K at about 108 yr. In this case the configuration is close to thermal equilibrium through most of the D-burning phase.

Figure 10.

Figure 10. Detail of the deuterium-burning phase for Run 2C, 15 MJup. Solid curve: outer radius Rp as a function of time, in units of 5 × 109 cm; long-dashed curve: central temperature Tc, at the core/envelope interface, as a function of time, in units of 105 K; dash-dot curve: nuclear luminosity as a function of time in units of log (Lnuc/L) +7; short-dashed curve: radiated luminosity L as a function of time, in the same units as Lnuc.

Standard image High-resolution image

4.3. Comparison with Beta Pictoris b

The cross in Figure 8 gives the approximate location of the directly imaged companion (Lagrange et al. 2009) to the well-known star Beta Pictoris. That star, according to http://exoplanet.eu, has a mass of about 1.8 M and an age of 12 (+8,−4) Myr. The planet is located between 8 and 15 AU from the star (Lagrange et al. 2010); thus, an approximate comparison can be made with these calculations. Beta Pic b's position in the (log L, t) diagram is plotted in Marois et al. (2010) where it is shown to fall on a theoretical track with mass 10 MJup as calculated from a "hot start" by Baraffe et al. (2003). In http://exoplanet.eu that mass is given as 8 (+5,−2) MJup. The surface temperature Teff has been estimated from observed near-infrared colors (Bonnefoy et al. 2011; Quanz et al. 2010) at 1700 K, with considerable uncertainty (≈300 K). Further infrared and astrometric observations (Bonnefoy et al. 2013) are essentially in agreement, giving log (L/L) = −3.87 ± 0.08, Teff = 1700 ± 100 K, ap = 8–10 AU, and "hot-start" masses in the range 7–13 MJup. The bolometric luminosity found by Marleau & Cumming (2013) is in agreement with the above value, and they find "hot-start" masses in the range 7–12 MJup.

In our "cold-start" calculations the track for Run 2C, 13.7 MJup, passes close to the object in the (log L, t) diagram, and the calculations give Teff = 1627 K at an age of 12 Myr. Our mass 10 MJup cannot possibly provide a fit. The "hot-start" models thus would show that the object is a planet, as defined by an object with mass not high enough to burn deuterium. However, this particular "cold-start" model indicates that beta Pictoris b is currently burning deuterium, which, according to the same definition, would classify it as a brown dwarf. As mentioned in Section 1, this definition is not universally agreed upon; an alternative definition, based on the minimum in the mass distribution of low-mass companions, observed within several AU of sunlike stars, places the limit at ≈25MJup. In this case Beta Pic b would still be a planet. Note that in the "cold-start" calculations, the fit at 13.7 MJup with an assumed σ = 6 g cm−2 is not unique; the companion could also be fit at σ = 4 g cm−2 at a slightly higher mass, about 15.6 MJup. Furthermore, these masses are uncertain and will probably change when the calculations are redone in the future with more detailed model atmospheres. Nevertheless, as such they are marginally consistent with the upper limits to the mass of Beta Pic b derived from radial velocity measurements (Lagrange et al. 2012). For a planet at 9 AU the limit is 12 MJup; at 10 AU it is 15.4 MJup.

We note also that the luminosity curve for 11.7 MJup in Figure 8 agrees well with the observed luminosity of the directly imaged planet HR 8799 c at the stellar age (≈6 × 107 yr). The observed value is given by Marley et al. (2012) as log L/L = −4.9 ± 0.1. The agreement of course requires a core mass of ≈30 M. A hot-start model of about 10 MJup without a core also agrees. However we do not make a detailed comparison with HR 8799 c, because the metallicity of the star is low ([Fe/H] =−0.47) and the planet orbits at 43 AU, making it highly debatable whether it could have formed by core-nucleated accretion.

5. SUMMARY AND CONCLUSIONS

We investigate the boundary between brown dwarfs and giant planets, according to the definition that brown dwarfs can burn the deuterium that is present when they form, and giant planets cannot. The main parameters and the results for M50, the boundary mass at which half of the original deuterium is burned after 4 Gyr, are summarized in Table 3. The columns give, respectively, the run identification, the stellar mass in M, ap, the initial disk solid surface density σ at ap, the resulting Mcore, and M50. The main cases considered involve a planet/brown dwarf at 5.2 AU around a solar-mass star, and a planet/brown dwarf at 9.5 AU around a star of 2 M. The table shows that there is only a small variation in the values of M50, which, however, correlate with the core mass in the sense that the smaller the core mass, the higher the value of M50.

Table 3. Summary

Run M/M Distance σ Mcore M50
(AU) (g cm−2) (M) (MJup)
1A 1 5.2 10 16.8 12.37
1B 1 5.2 10 16.8 12.20
1C 1 5.2 4 4.83 13.55
2A 2 9.5 4 18.7 11.95
2B 2 9.5 4 18.8 12.05
2C 2 9.5 6 31.0 11.65

Download table as:  ASCIITypeset image

The calculations, taken as a whole, indicate that the envelope entropy, which is a function of initial conditions and which is closely related to the core mass through the accretion processes during the formation phase, is an important factor in determining M50. However, certain physical processes during formation are shown to have only a small effect. Run 1B has the same parameters as Run 1A except that the dust opacity during the formation phase is higher by a factor that ranges from 2 to 100, depending on the depth in the envelope. This difference has a negligible effect on M50. Run 2B has the same parameters as 2A except that the disk viscosity during the phase of disk-limited gas accretion is lower by a factor 2.5. This difference also has a negligible effect on M50. However, the disk viscosity is important in another respect. If it is significantly lower than the range presented here (αt ≈ 10−2), then there will not be time to accrete a planet with mass necessary to burn deuterium during the lifetime of the disk. The gas accretion rate onto a planet of 4 MJup around a star of 2 M, in a disk with αt = 4 × 10−4, is reduced by a factor 400 compared with a disk with αt = 4 × 10−3 (Lissauer et al. 2009), corresponding to less than a Jupiter mass in a million years for the initial conditions of Run 2B (formation time about 3 Myr). Of course, the minimum viscosity required to build a planet up to about 12 MJup will depend on parameters such as σ and ap. This question is discussed in more detail in the Appendix.

Core accretion models, in the cold-start case, are known to have low entropy compared with hot-start models. In Marley et al. (2007) the entropy just after formation for 10 MJup was found to be 8.2 kB per baryon for Mcore = 16.8 M. The corresponding luminosity at ages of 107 to 108 yr was about 2 × 10−6L, certainly fainter than observed values for directly imaged planets. In this mass range, for the given core mass, the entropy is very insensitive to the planet's total mass, as shown in that paper and confirmed by the present results. However, our calculations show that the entropy is quite sensitive to the core mass, as illustrated in Figure 11 (a similar effect has been found independently by Mordasini (2013) for Mcore > 20 M). The points shown are all calculated with the same total mass and the same disk viscosity. All used the reduced interstellar grain opacity, except for the point at Mcore = 4.8 M, for which the grain-settling opacities were used (if the interstellar opacities had been used, the formation time would have been considerably longer). However, the comparison between Runs 1A and 1B, which looked at the effect of changing the opacities, showed that the difference in entropy was less than 0.1 kB per baryon at the same total mass. The effect on the entropy of changing the viscosity (Runs 2A and 2B) was even smaller. Physical effects that do affect the entropy include the planetesimal accretion rate and the rate of contraction of the envelope, both of which affect the internal heating of the envelope. Thus, the luminosities of newly formed massive planets, depending on formation conditions, can vary by up to two orders of magnitude. Information on the runs whose entropies are plotted in the figure is given in Table 4. The table is in the same format as Table 2 and gives the runs in order of decreasing entropy. Clearly, for this set of models, a lower entropy is associated with a longer formation time. The luminosity plots for these four cases in Figure 5 illustrate the same effect.

Figure 11.

Figure 11. The entropy in the interior of planets of total mass 12 MJup, immediately after formation, is plotted against their core masses, in M. The points plotted, from top to bottom, are from Runs 2C, 2A, 1B, and 1C.

Standard image High-resolution image

Table 4. Data for Figure 11

Run Mfinal/MJup tform Mcore/M Tc, f Tmax ρc, f log Dfinal/Dinit
(yr) (K) (K) (g cm−3) (Lf/L)
2C 12.0 8.83 × 105 31.0 3.48 × 105 3.67 × 105 29.6 −4.80 0.278
2A 12.0 2.14 × 106 18.7 2.76 × 105 3.06 × 105 41.8 −5.63 0.435
1B 12.0 2.66 × 106 16.8 2.66 × 105 2.66 × 105 46.1 −5.91 0.930
1C 12.0 4.10 × 106 4.80 2.08 × 105 2.08 × 105 54.0 −6.56 1.000

Download table as:  ASCIITypeset image

The combination of M*, ap, and σ determines the isolation mass, and thereby the ultimate core mass, which turns out to be a key factor in determining the entropy of the planet at formation. Higher entropy, in particular the higher temperature, favors more rapid nuclear burning, so the higher entropy runs result in lower values of M50. Nevertheless, the range of initial conditions explored here, which is considerable, produces only a small range in M50, about 11.6–13.6 MJup, in agreement with previous independent calculations. We can further conclude that for cold-start core-accretion models that do burn deuterium, the tracks in the luminosity versus time diagram can potentially provide agreement with the properties of directly imaged low-mass stellar companions.

Primary funding for this project was provided by the NASA Origins of Solar Systems Program grant NNX11AK54G (P.B., G.D., J.L.). G.D. acknowledges additional support from NASA grant NNX11AD20G. P.B. acknowledges additional support from NSF grant AST0908807. D.S. is supported in part by NASA grants NNH11AQ54I and NNH12AT89I. The authors are indebted to Gilles Chabrier for the use of his nuclear screening factors. The 3D hydrodynamical simulations reported in this work were performed using resources provided by the NASA High-End Computing (HEC) Program through the NASA Advanced Supercomputing (NAS) Division at Ames Research Center. G.D. thanks Los Alamos National Laboratory for its hospitality. The authors thank the referee Dr. Christoph Mordasini for a detailed and constructive review.

APPENDIX: ANALYTIC APPROXIMATIONS OF THE DISK-LIMITED GAS ACCRETION RATES

In this section, we provide analytic approximations for the gas accretion rate in the regime where this rate is limited by the ability of the disk to transfer gas to the planet. In the calculations, we used piece-wise functions obtained by fitting the data from the 3D hydrodynamical calculations (see Section 3), for various values of the turbulent viscosity parameter, αt, which quantifies the kinematic viscosity of the disk at the radial location of the planet, $\nu _{t}=\alpha _{t} H^{2}_{p} \Omega _{p}$. We recall that the hydrodynamical calculations used an aspect ratio Hp/ap = 0.05, which is a reasonable value in evolved disks between 5 and 10 AU (e.g., D'Angelo & Marzari 2012, and references therein).

We fitted $(\log {\dot{M}_{p}},\log {M_{p}})$ data using multiple second-order polynomials, which were then smoothly joined in overlapping regions. Since this procedure is somewhat cumbersome, here we provide simpler analytic approximations derived from data in the range of Mp/M from 10−4 to 10−2. In the calculations, as explained in Section 3, disk-limited accretion sets in when Mp ≳ 0.25 MJup.

Let us introduce the four functions

Equation (A1)

Equation (A2)

Equation (A3)

Equation (A4)

where q = Mp/M and all logarithms are in base 10. The coefficients ai, bi, ci, and di are given in Table 5. For αt = 10−2, the following analytic approximation for the disk-limited gas accretion rate, in units of $a^{2}_{p}\Sigma _{p}\Omega _{p}$, may be used:

Equation (A5)

For αt = 4 × 10−3, the following analytic approximation may be applied:

Equation (A6)

The fitting functions are shown in the upper panel of Figure 12, along with the data obtained from the 3D hydrodynamical calculations.

Figure 12.

Figure 12. Upper panel: disk-limited gas accretion rates vs. the planet-to-star mass ratio (see also Figure 2). The dashed line is Equation (A5), the solid line is Equation (A6), and the symbols represent the simulations' data for the disk turbulent parameter αt = 10−2 (open circles) and 4 × 10−3 (filled pentagons). Lower panel: final mass of a planet accreting at a disk-limited gas accretion rate in various situations: αt = 4 × 10−3 (diamonds), αt = 10−2 (circles), ap = 5.2 AU (thin lines), ap = 9.5 AU (thick lines), M = 1 M (open symbols), and M = 2 M (filled symbols). See text for further details.

Standard image High-resolution image

Table 5. Coefficients in Equations (A1)–(A4)

  i = 0 i = 1 i = 2
ai −6.8345179 −2.5152600 −0.3542960
bi −12.471200 −6.3732500 −1.0144400
ci −11.429400 −5.0171300 −0.697484
di −17.819292 −8.5200885 −1.172181

Download table as:  ASCIITypeset image

In the range of the turbulent parameter αt that we explored (4 × 10−4 ⩽ αt ⩽ 0.02), the maximum of $\dot{M}_{p}$ occurs at a ratio Mp/M similar (within a factor of ≈2) to the square root of the right-hand side of Equation (7), i.e., before gas begins to be depleted significantly because of the formation of the density gap. The maximum of $\dot{M}_{p}$ can be compared to the (steady state) accretion rate through the disk in absence of the planet, 3πνtΣp (Lynden-Bell & Pringle 1974), at the radial location of the planet. In units of $a^{2}_{p}\Sigma _{p}\Omega _{p}$, this accretion rate can be written as 3παt(Hp/ap)2, giving ≈10−4 and 2.4 × 10−4 for αt = 0.004 and 0.01, respectively. As can be seen in Figure 12, these disk accretion rates are smaller than the maximum of $\dot{M}_{p}$. However, it should be noted that the tidal perturbation of the planet can modify the accretion rate through the disk (Lubow & D'Angelo 2006).

Equations (A5) and (A6) can be integrated to find the final (asymptotic) mass of a planet, Mfinal, that accretes gas at a disk-limited gas accretion rate. We solved numerically the differential equation

Equation (A7)

for Mp, using an adaptive Adams–Bashforth–Moulton method of variable order with adaptive step-size and error control (available through the SLATEC Common Mathematical Library). Notice from Equation (A7) that, although the dependence of $\dot{M}_{p}$ on M, ap, and Σp is trivial, the dependence of Mp(t) on those three quantities is not!

During the integration of Equation (A7), we assumed that M, ap, and αt are constants. In order to mimic the viscous evolution of the (unperturbed) gas surface density at the radial position of the planet, Σp, we applied the solution of Lynden-Bell & Pringle (1974) for a disk with no central couple. Using the same notations and indicating with R1 the initial standard deviation of the (Gaussian) surface density distribution and with M1 the initial disk mass, Lynden-Bell & Pringle (1974) found that $\nu _{t}/R^{2}_{1}=(2/3)\dot{M}_{*}/M_{1}$ (where $\dot{M}_{*}$ is the initial accretion rate on the star). Introducing the non-dimensional "viscous" time7$t_{\mathrm{vis}}=6 (\nu _{t}/R^{2}_{1}) t +1$, which can be written as $t_{\mathrm{vis}}=4 (\dot{M}_{*}/M_{1}) t +1$, the surface density evolution can be approximated by

Equation (A8)

where Σref is a parameter and which represents the behavior of the Lynden-Bell & Pringle solution for tvis ≫ 1. We assumed that $\dot{M}_{*}/M_{1}$ is ≈10−6 yr−1. According to the equation above, the surface density ratio Σpref decreases by more than two orders of magnitude over 10 Myr.

We integrated Equation (A7) for the values of M, ap, and αt used in the calculations, applying Equation (A8), and determined Mfinal as a function of Σref. The results are shown in the lower panel of Figure 12 (see figure caption for a description of the different curves). The final mass is reached within about 5.5 Myr, when typically $M_{p}/\dot{M}_{p}\sim 100$ Myr. The effect of disk viscosity is evident in this figure. In fact, around a solar mass star, the mass threshold for deuterium burning can only be achieved for αt ≳ 10−2. Among the varied parameters, αt produces the largest differences in Mfinal, whereas ap produces the smallest. Notice that the values of Mfinal shown in the lower panel of Figure 12 should not necessarily agree with those in the D-burning calculations because of the different assumptions made for the nebula evolution. In particular, Σp in those calculations was taken as a constant.

Footnotes

  • Notice that the power of R1, in the definition of Lynden-Bell & Pringle (1974), should be −2. Also, the subscript "1" in R1 and M1 refers to the viscous time tvis = 1, when the physical time t = 0.

Please wait… references are loading.
10.1088/0004-637X/770/2/120