Articles

RAPIDLY ACCRETING SUPERGIANT PROTOSTARS: EMBRYOS OF SUPERMASSIVE BLACK HOLES?

, , and

Published 2012 August 20 © 2012. The American Astronomical Society. All rights reserved.
, , Citation Takashi Hosokawa et al 2012 ApJ 756 93 DOI 10.1088/0004-637X/756/1/93

0004-637X/756/1/93

ABSTRACT

Direct collapse of supermassive stars (SMSs) is a possible pathway for generating supermassive black holes in the early universe. It is expected that an SMS could form via very rapid mass accretion with $\dot{M}_*\sim 0.1\hbox{--}1 \,M_\odot \,{\rm yr}^{-1}$ during the gravitational collapse of an atomic-cooling primordial gas cloud. In this paper, we study how stars would evolve under such extreme rapid mass accretion, focusing on the early evolution until the stellar mass reaches 103M. To this end, we numerically calculate the detailed interior structure of accreting stars with primordial element abundances. Our results show that for accretion rates higher than 10−2M yr−1, stellar evolution is qualitatively different from that expected at lower rates. While accreting at these high rates, the star always has a radius exceeding 100 R, which increases monotonically with the stellar mass. The mass–radius relation for stellar masses exceeding ∼100 M follows the same track with R*M1/2* in all cases with accretion rates ≳ 10−2M yr−1; at a stellar mass of 103M, the radius is ≃ 7000 R (≃ 30 AU). With higher accretion rates, the onset of hydrogen burning is shifted toward higher stellar masses. In particular, for accretion rates exceeding $\dot{M}_*\gtrsim 0.1 \,M_\odot \,{\rm yr}^{-1}$, there is no significant hydrogen burning even after 103M have accreted onto the protostar. Such "supergiant" protostars have effective temperatures as low as Teff ≃ 5000 K throughout their evolution and because they hardly emit ionizing photons, they do not create an H ii region or significantly heat their immediate surroundings. Thus, radiative feedback is unable to hinder the growth of rapidly accreting stars to masses in excess of 103M as long as material is accreted at rates $\dot{M}_*\gtrsim 10^{-2} \,M_\odot \,{\rm yr}^{-1}$.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Recent observations reveal that supermassive black holes (SMBHs) exceeding 109M already existed in the universe less than 1 Gyr after the big bang (e.g., Fan 2006; Mortlock et al. 2011; Treister et al. 2011). The origins of such SMBHs must be intimately related to structure formation in the early universe. Some scenarios on the birth and growth of SMBHs postulate the existence of remnant BHs from Population III (Pop III) stars as their seeds (e.g., Madau & Rees 2001; Schneider et al. 2002). For several decades, theoretical studies have predicted that the majority of Pop III stars were very massive, exceeding 100 M (e.g., Bromm & Larson 2004). Pop III stars more massive than 300 M end their lives by directly collapsing to form BHs (e.g., Heger & Woosley 2002). If such a ∼100 M BH grows via continuous mass accretion at the Eddington limited rate, then its mass barely attains 109M in 1 Gyr.

This scenario, however, has recently been challenged. First, it is suspected that most Pop III stars were much less massive than previously thought. A circumstellar disk forming after the cloud's collapse easily fragments due to gravitational instability and could produce multiple protostars (e.g., Machida et al. 2008; Stacy et al. 2010; Clark et al. 2011). The final stellar masses would be reduced as the accreting gas is shared by multiple stars. Moreover, strong stellar UV light creates an H ii region around the protostar when the stellar mass exceeds a few × 10 M. The resulting feedback terminates the growth of Pop III protostars via mass accretion at a few × 10 M (e.g., McKee & Tan 2008; Hosokawa et al. 2011b; Stacy et al. 2012). A large amount of gas would be expelled from the dark halo due to the expansion of H ii regions and the onset of core-collapse supernovae (e.g., Whalen et al. 2004; Kitayama et al. 2004; Kitayama & Yoshida 2005), which quenches the supply of gas to any remnant BH. Even if a BH gets some gas supply, radiative feedback from the BH accretion disk could regulate mass accretion onto the BH-disk system (e.g., Alvarez et al. 2009; Milosavljević et al. 2009; Jeon et al. 2011).

Another pathway for generating SMBHs is BH binary mergers. However, this process could also be limited due to the strong recoil resulting from gravitational wave emission (e.g., Campanelli et al. 2007; Herrmann et al. 2007). It is not straightforward that a seed BH  ≲  100 M can grow to a ∼109M SMBH within 1 Gyr of its birth.

An alternative possibility is that massive BHs exceeding 105M form directly in some rare occasions in the primordial gas (e.g., Bromm & Loeb 2003). The primary cooling process in the primordial gas is line emission of molecular hydrogen. However, the thermal evolution of a gravitationally collapsing cloud can change significantly, if this cooling process is suppressed, for example, due to photodissociation of molecules by strong background radiation (Omukai 2001; Oh & Haiman 2002; Shang et al. 2010; Inayoshi & Omukai 2011) or collisional dissociation in dense shocks (Inayoshi & Omukai 2012). If the dark halo is sufficiently massive (≳ 108M), then the baryonic gas can contract with atomic hydrogen cooling even without molecular hydrogen. The collapse proceeds nearly isothermally at ≃ 8000 K. Without efficient molecular cooling, fragmentation is suppressed and single or binary protostars form within dense cloud cores of ≳ 105M (Bromm & Loeb 2003; Regan & Haehnelt 2009). The protostar's mass is initially ∼10−2M but quickly increases via mass accretion. The expected accretion rates are 0.1–1 M yr−1, more than 100 times higher than the standard value ≃ 10−3M yr−1 expected for Pop III star formation. The stellar mass could reach 105–106M in ∼1 Myr with this very rapid mass accretion. General relativity predicts that such a supermassive star (SMS) becomes unstable (e.g., Chandrasekhar 1964) and collapses to form a BH, which subsequently swallows most of the surrounding stellar material (e.g., Shibata & Shapiro 2002). Some authors are exploring a different picture, whereby only a central part of the SMS collapses to form a ∼100 M BH and heat input from the accreting BH inflates the outer envelope of the SMS ("quasi-star"; Begelman et al. 2006, 2008; Begelman 2010; Ball et al. 2011; Dotan et al. 2011).

However, we only have limited knowledge on how stars evolve under such extreme conditions of rapid mass accretion. Begelman (2010) predicts that, based on simple analytic arguments, such stars have a very different structure from their main-sequence counterparts. Stellar evolution at lower accretion rates $\dot{M}_*\lesssim 10^{-2} \,M_\odot \,{\rm yr}^{-1}$ has been studied in detail by numerically solving the stellar interior structure (e.g., Stahler et al. 1986; Omukai & Palla 2001, 2003; Ohkubo et al. 2009; Hosokawa & Omukai 2009; Hosokawa et al. 2010). Omukai & Palla (2001, 2003) showed that rapid mass accretion with $\dot{M}_*> 4 \times 10^{-3} \,M_\odot \,{\rm yr}^{-1}$ causes the protostar's abrupt expansion before its arrival to the zero-age main sequence (ZAMS). Further comprehensive studies on stellar evolution with rapid mass accretion are indispensable for considering their radiative feedback and observational signatures (e.g., Johnson et al. 2011).

We present here our first results of this sort, whereby as a first step, we study the early evolution up to a stellar mass of 103M. Our results show that rapid accretion with $\dot{M}_*\gtrsim 10^{-2} \,M_\odot \,{\rm yr}^{-1}$ causes the star to bloat up like a red giant. The stellar radius increases monotonically with stellar mass and reaches ≃ 7000 R(≃ 30 AU) at a mass of 103M. Unlike the cases with lower accretion rates previously studied, the mass–radius relation in this phase is almost independent of the assumed accretion rate. Such massive "super-giant" protostars could be the progenitors that eventually evolve to the observed SMBHs in the early universe.

The organization of this paper is as follows. First, we briefly review our numerical method and summarize the calculated cases in Section 2. We describe our results in Section 3; we first focus on the fiducial case with $\dot{M} = 0.1 \,M_\odot \,{\rm yr}^{-1}$ and then examine effects of varying accretion rates and boundary conditions. Finally, summary and discussions are described in Section 4.

2. NUMERICAL MODELING OF ACCRETING STARS

2.1. Method

We calculate stellar evolution with mass accretion using the numerical codes developed in our previous work (see Omukai & Palla 2003; Hosokawa & Omukai 2009; Hosokawa et al. 2010, for details). The four stellar structure equations, i.e., equations of continuity, hydrostatic equilibrium, energy conservation, and energy transfer, including effects of mass accretion, are solved assuming spherical symmetry. We focus on the early evolution until slightly after the ignition of hydrogen fusion in this paper. To this end, the appropriate nuclear network for the thermo-nuclear burning of deuterium, hydrogen, and helium is considered.

The codes are designed to handle two different outer boundary conditions for stellar models: shock and photospheric boundaries. The shock boundary condition presupposes spherically symmetric accretion onto a protostar, whereby the inflow directly hits the stellar surface and forms a shock front (e.g., Stahler et al. 1980; Hosokawa & Omukai 2009). We solve for the structure of both the stellar interior and outer accretion flow. With this boundary condition, the photosphere is located outside the stellar surface where the accretion flow is optically thick to the stellar radiation. The photospheric boundary condition, on the other hand, presupposes a limiting case of mass accretion via a circumstellar disk, whereby accretion columns connecting the star and disk are geometrically compact and most of the stellar surface radiates freely (e.g., Palla & Stahler 1992; Hosokawa et al. 2010). In this case, we only solve the stellar interior structure without considering details of the accretion flow; the location of the photosphere always coincides with the stellar surface.

The different outer boundary conditions correspond to the two extremes of accretion flow geometries, or more specifically to different thermal efficiencies of mass accretion, which determine the specific entropy of accreting materials (e.g., Hosokawa et al. 2011a). The accretion thermal efficiency controls the entropy content of the star, which determines the stellar structure. With the shock boundary condition, the accreting gas obtains a fraction of the entropy generated behind the shock front at the stellar surface. The resulting thermal efficiency is relatively high ("hot" accretion).

For the photospheric boundary condition, on the other hand, the accreting gas is assumed to have the same entropy as in the stellar atmosphere. The underlying idea is that when the accreting gas slowly approaches the star via angular momentum transport in the disk, its entropy should be regulated to the atmospheric value. This is a limiting case of thermally inefficient accretion ("cold" accretion). In general, with even a small amount of angular momentum, mass accretion onto the protostar would be via a circumstellar disk, perhaps with geometrically narrow accretion columns connecting the disk with the star. For extremely rapid mass accretion, however, the innermost part of the disk becomes hot and entropy generated within the disk is advected into the stellar interior (e.g., Popham et al. 1993). Thus, cold accretion as envisioned for the photospheric boundary condition is not appropriate for the case of rapid mass accretion (also see discussions in Hartmann et al. 1997; Smith et al. 2011). We therefore expect that the shock boundary condition is a good approximation for the extremely high accretion rates considered here and mostly focus on stellar evolution with the shock boundary condition. We also consider a few cases with the photospheric boundary condition for comparison to test potential effects of reducing the accretion thermal efficiency (also see Section 2.2 below).

2.2. Cases Considered

The cases considered are summarized in Table 1. In this paper, we only consider the evolution with constant accretion rates for simplicity. The adopted accretion rates range from 10−3M yr−1 to 1 M yr−1. Stellar evolution with accretion rates less than 10−2M yr−1 (cases MD1e3 and MD6e3) has been studied in detail in our previous work (e.g., Omukai & Palla 2003; Hosokawa & Omukai 2009). As described in Section 2.1 above, we calculate the protostellar evolution assuming shock outer boundary conditions for most of the cases. Cases MD3e1-HC10 and MD3e1-HC50 are the only exceptions, whereby we switch to the photospheric boundary condition after the stellar mass reaches 10 M and 50 M, respectively, at a constant accretion rate of 0.3 M yr−1. The underlying idea for switching the boundary at some moment is that the specific angular momentum of the inflow, and thus the circumstellar disk, grow with time, which reduces entropy of the accreted matter. The higher mass at switching point corresponds to the higher angular momentum of the parental core. As in our previous work, we start the calculations with initial stellar models constructed assuming that the stellar interior is in radiative equilibrium (e.g., Hosokawa & Omukai 2009). We adopt a slightly higher initial stellar mass of ∼1 M for stability reasons. The calculated initial stellar radii are also summarized in Table 1.

Table 1. Cases Considered

Case $\dot{M}_*$ M*, 0 R*, 0 Notes References
  (M yr−1) (M) (R)    
MD1e0 1.0 2.5 298.4   Section 3.2
MD3e1 0.3 2.0 238.3   Sections 3.23.3
MD3e1-HC10 0.3 2.0 (10) 238.3 (437.1) shock → photo. BC at M* = 10 M Section 3.3
MD3e1-HC50 0.3 2.0 (50) 238.3 (826.8) shock → photo. BC at M* = 50 M Section 3.3
MD1e1 0.1 2.0 177.8 fiducial case Sections 3.13.2
MD6e2 0.06 1.0 118.6   Section 3.2
MD3e2 0.03 1.0 95.2   Section 3.2
MD6e3 0.006 1.0 52.3 also see Omukai & Palla (2003) Section 3.2
MD1e3 0.001 0.05 12.5 also see Omukai & Palla (2003) Sections 3.13.2

Notes. Column 2: mass accretion rate, Column 3: initial stellar mass, Column 4: initial radius. For cases MD3e1-HC10 and MD3e1-HD50, the values when the boundary condition is switched are given in parentheses.

Download table as:  ASCIITypeset image

3. RESULTS

3.1. Evolution in the Fiducial Case ($\dot{M}_*= {\it 0.1} \,M_\odot \,{\rm yr}^{\it -1}$)

We first consider the fiducial case (MD1e1), whereby the stellar mass increases with the constant accretion rate of $\dot{M}_*= 0.1 \,M_\odot \,{\rm yr}^{-1}$. The calculated evolution of the stellar interior structure is presented in Figure 1. We see that the stellar radius is very large and increases monotonically with the stellar mass. The stellar radius exceeds 103R when the stellar mass is M* ≃ 45 M and reaches ≃ 6500 R at M* ≃ 103M. This evolution differs qualitatively from that calculated assuming a lower accretion rate 10−3M yr−1 as depicted in Figure 1 by the blue line (see also, e.g., Omukai & Palla 2003). At the lower accretion rate, the stellar radius initially increases with stellar mass but begins to decrease at M* ≳ 6 M.

Figure 1.

Figure 1. Evolution of the stellar interior structure for the fiducial case, whereby the stellar mass increases at a rate of 0.1 M yr−1 (case MD1e1). The thick solid line depicts the stellar surface, which is the position of the accretion shock front. The dotted lines show the radial positions of the mass coordinates of M = 3, 10, 30, 100, and 300 M. The dot-solid line indicates the radial position within which 70% of the stellar mass is enclosed. The gray-shaded areas represent the convective layers. The hatched areas indicate layers of active nuclear burning, where the energy production rate exceeds 10% of the steady rate 0.1LD, st/M* for deuterium burning (see Equation (9)), and 0.1L*/M* for hydrogen burning. The blue solid line shows the evolution of the radius of a star accreting material at 10−3M yr−1 for comparison.

Standard image High-resolution image

A key quantity for understanding the contrast between the two cases is the balance between the two characteristic timescales: the accretion timescale

Equation (1)

and the Kelvin–Helmholtz (KH) timescale

Equation (2)

where R* and L* are the stellar radius and luminosity, and G is the gravitational constant (e.g., Stahler et al. 1986; Omukai & Palla 2003; Hosokawa & Omukai 2009). In the early stage, during which the stellar radius increases with mass, the timescale balance is tacctKH and radiative loss of the stellar energy is negligible (adiabatic accretion stage). However, the radiative energy loss becomes more efficient as the stellar mass increases. This is because opacity in the stellar interior, which is due to the free–free absorption following Kramers' law κ∝ρT−3.5, decreases and the stellar luminosity L* increases with the stellar interior temperature (and thus with its mass M*). The upper panel of Figure 2 indeed shows that the maximum luminosity within the star Lmax increases as a power-law function of M*.

Figure 2.

Figure 2. Evolution of the stellar surface luminosity L* (dashed line) and maximum luminosity within the star Lmax (solid line) for the cases with 10−3M yr−1 (MD1e3; upper panel) and 0.1 M yr−1 (fiducial case MD1e1; lower panel). The mass–luminosity relations given by Equations (3), (4), and (5) are shown with the thin green, blue, and red lines, respectively. In each panel, the vertical dot-dashed line (magenta) indicates the epoch when the accretion time is equal to the KH time.

Standard image High-resolution image

This increase of L* is consistent with the analytic scaling relation for radiative stars with Kramers' opacity, LM11/2*R*−1/2 (e.g., Hayashi et al. 1962). Our numerical results are well fitted by the analytic relation

Equation (3)

The increase of L* causes an inversion of the timescale balance to tacc > tKH at low accretion rates. The star contracts by losing its energy (KH contraction stage), which is seen for M* ≳ 6 M. The opacity in the stellar interior has fallen down to the constant value of electron scattering. Figure 2 shows that, in this stage, luminosity takes its maximum value at the stellar surface and increases as L* = LmaxM3*, which is valid for the constant opacity cases (e.g., Hayashi et al. 1962). The relation

Equation (4)

roughly agrees with our results. Temperature at the stellar center increases during the KH contraction stage. Hydrogen burning finally begins and the stellar radius begins to increase following the mass–radius relation of ZAMS stars for M* ≳ 50 M. Figure 2 shows that the stellar luminosity gradually approaches the Eddington limit

Equation (5)

By contrast, there is no contraction stage for the case with a much higher accretion rate $\dot{M}_*= 0.1 \,M_\odot \,{\rm yr}^{-1}$ (MD1e1). Nevertheless, the evolution of the timescales still follows the picture described above (Figure 3(a)). We see that the timescale balance changes from tacc < tKH to tacc > tKH at M* ≃ 40 M. The protostar is in the adiabatic accretion stage for M* ≲ 40 M. The accretion luminosity $L_{\rm acc} \equiv G M_* \dot{M}_*/ R_*$ at this stage is much higher than the stellar luminosity L*, since the luminosity ratio Lacc/L* is equal to the timescale ratio tKH/tacc by definition. Stahler et al. (1986) derived the approximate analytic formulae describing radial positions of the stellar surface R* and photosphere Rph (located within the accretion flow) during the adiabatic stage:

Equation (6)

Equation (7)
Figure 3.

Figure 3. Evolution of several physical quantities for the fiducial case with 0.1 M yr−1 (MD1e1). Top panel: comparison between the accretion timescale tacc (dashed line) and KH timescale tKH (dotted line). The vertical magenta dot-dashed line indicates the epoch when tKH is equal to tacc. The thin solid line represents 100 times of the stellar free-fall timescale $t_{{\rm ff}} \equiv \sqrt{3 \pi / 32 G \bar{\rho }}$, where $\bar{\rho }$ is the average mass density of the star. The fact that tff is much shorter than tKH and tacc verifies the hydrostatic balance assumption implicit in the stellar structure equations. Middle panel: evolution of the accretion luminosity Lacc (dashed line), stellar luminosity L* (dotted line), and total luminosity LtotLacc + L* (solid line). The red line indicates the Eddington luminosity at each stellar mass. Bottom panel: evolution of the radial positions of the photosphere Rph(R) (dashed line) and stellar surface R*(R) (black solid line). The red solid and dashed lines denote the analytic formulae for these radii by Stahler et al. (1986) (Equations (6) and (7)). The evolution of the stellar effective temperature Teff(K) is also shown with the blue line.

Standard image High-resolution image

The bottom panel of Figure 3 shows that these formulae still agree with our numerical results with 0.1 M yr−1. Equation (6) shows that the stellar radius is larger for the higher accretion rate at the same mass. The larger radius is due to the higher specific entropy of accreting material and the resulting higher entropy content of the star (e.g., Hosokawa & Omukai 2009). Comparing two stars of the same mass, the one with a larger radius has a lower interior temperature, which then implies a higher opacity due to the strong T dependence of Kramers' law κ∝ρT−3.5. For this reason, the adiabatic accretion stage is prolonged up to higher stellar masses for higher accretion rates. (See the Appendix for an analytic expression describing this dependence.) Figure 2(b) shows that the evolution of the stellar maximum luminosity Lmax still obeys Equations (3)–(5). Unlike the case for 10−3M yr−1, however, it is only after Lmax approaches the relation of LmaxM3* (Equation (4)) that tKH becomes equal to tacc. The rapid heat input via mass accretion prevents the star from losing internal energy until the stellar luminosity becomes sufficiently high.

The fact that tKH is shorter than tacc for M* ≳ 40 M indicates that most of the stellar interior is contracting, as shown by the trajectories of the mass coordinates (dashed lines in Figure 1). The figure also shows that the bloated surface layer occupies only a small fraction of the total stellar mass. When the stellar mass is ≃ 300 M, for example, the layer which has 30% of the total mass measured from the surface covers more than 98% of the radial extent. The star has a radiative core surrounded by an outer convective layer. Although the convective layer covers a large fraction of the stellar radius, even the central radiative core is much larger than a ZAMS star with the same mass (compare with the blue curve for M* ≳ 40 M in Figure 1). Figure 4 shows the radial distributions of physical quantities, i.e., specific entropy, luminosity, temperature, and density, in the stellar interior. We see that the specific entropy is at its maximum value near the boundary between the radiative core and convective layer. The stellar entropy distribution is controlled by the energy equation,

Equation (8)

where s is the specific entropy and epsilon is the energy production rate by nuclear fusion. For M* ≳ 40 M, most of the stellar luminosity comes from the release of gravitational energy. In fact, as seen in Figure 4(b), (∂L/∂M)t > 0 in the radiative core, which means that the internal energy of the gas is decreasing. The local luminosity in the radiative core is close to the Eddington value given by Equation (5), using the mass coordinate M rather than the stellar mass M* (Figure 4(b)). Note that the opacity in the radiative core is only slightly higher than that expected from electron scattering alone.

Figure 4.

Figure 4. Radial profiles of the (a) specific entropy, (b) luminosity, (c) temperature, and (d) gas mass density in the stellar interior for the fiducial case with 0.1 M yr−1 (MD1e1). The profiles when the stellar mass is 200 M, 600 M, and 103M are shown in each panel. The magenta parts indicate the layers are convective. The thin red line in panel (b) represents the Eddington-limit luminosity as a function of the mass coordinate M.

Standard image High-resolution image

In the outer parts of the star, where temperature and density are lower, however, opacity is higher than in the core because of bound–free absorption of H, He atoms, and the H ion. Energy transport via radiation is inefficient there, and most of the energy coming from the core is carried outward via convection. Figure 4 shows that the surface convective layer lies in the temperature and density range where T ≲ 105 K and ρ ≲ 10−8 g cm−3, which is almost independent of the stellar mass. We also see that the specific entropy is not constant over the convective layer, decreasing toward the stellar surface (a so-called super-adiabatic layer). This is because convective heat transport is inefficient near the stellar surface. A part of the outflowing energy is absorbed there, as indicated by the fact that the surface layer has a negative luminosity gradient (∂L/∂M)t < 0. This explains the high specific entropy in the outer convective layer.

The energy flux coming from the radiative core increases as the star grows in mass. As a result, the energy absorbed in the outer convective layer also increases with stellar mass, which raises the entropy peak at the bottom of the convective layer (Figure 4(a)). As the entropy increases, the outer layers expand, while the central part of the core contracts further. Therefore, the star assumes a more centrally condensed structure for the higher stellar mass.

The outermost part of the star has a density inversion, i.e., the density increases toward the stellar surface. Here, opacity assumes very high values because of H absorption. Radiation pressure is so strong that the hydrostatic balance is not achieved only with gravity; the additional inward force by the negative gas pressure gradient helps maintain the hydrostatic structure. Note, however, that this density inversion could be unstable in realistic multi-dimensions (see, e.g., Begelman et al. 2008).

Although deuterium burning is ignited when the stellar mass is ≃ 50 M, its influence on the subsequent evolution is negligible (Figure 1). The total energy production rate by deuterium burning is approximately

Equation (9)

where δD is the energy available from deuterium burning per unit gas mass. Since this is much lower than the Eddington luminosity LEdd in the mass range considered, energy production by deuterium burning contributes only slightly to the luminosity in the stellar interior. Hosokawa & Omukai (2009) showed that deuterium burning influences the stellar evolution only when the accretion rate is low $\dot{M}_*\lesssim 10^{-4} \,M_\odot \,{\rm yr}^{-1}$.

Figure 4(c) shows that temperature in the stellar interior increases with total mass. The central temperature reaches 108 K when the stellar mass is ≃ 600 M. Soon after that, hydrogen burning begins and a central convective core develops (Figure 1). This convective core can also be seen in the radial profiles for the 103M model in Figure 4 (indicated by magenta). The luminosity profile tells that most of the energy produced by hydrogen burning is absorbed within the convective core. The star still shines largely by releasing its gravitational energy even after hydrogen ignition.

When the stellar radius is sufficiently large, the accretion flow reaches the stellar surface before becoming opaque to the outgoing stellar light. In fact, soon after the end of the adiabatic accretion stage, the accreting envelope remains optically thin throughout (Figure 3(c)). We also see that the stellar effective temperature is almost constant at Teff ≃ 5000 K during this period. In general, the stellar effective temperature never assumes a lower value due to the strong temperature dependence of H absorption opacity (e.g., Hayashi 1961). Stars that have a compact core and bloated envelope (e.g., red giants) commonly have an almost constant effective temperature, regardless of their stellar masses.

3.2. Cases with Different Accretion Rates

We now investigate how stellar evolution changes with the accretion rate. Figure 5 shows the evolution of the stellar radius for several cases, including $\dot{M}_*= 10^{-3} \,M_\odot \,{\rm yr}^{-1}$ (case MD1e3) and 0.1 M yr−1 (case MD1e1) explained in Section 3.1. We see that, with accretion rates higher than 6 × 10−2M yr−1, the evolution becomes similar to that of the fiducial case for 0.1 M yr−1 (MD1e1); the stellar radius monotonically increases with mass. The protostars undergo adiabatic accretion in the early stage. As Equation (6) shows, the stellar radius is larger for higher accretion rates at a given stellar mass, say, at M* = 10 M.

Figure 5.

Figure 5. Evolution of the protostellar radius for various accretion rates. Upper panel: the different curves represent the cases with $\dot{M}_*= 10^{-3} \,M_\odot \,{\rm yr}^{-1}$ (case MD1e3, black), 6 × 10−3M yr−1 (MD6e3, blue), 3 × 10−2M yr−1 (MD3e2, red), and 6 × 10−2M yr−1 (MD6e2, magenta). The open and filled circles on each curve denote the epoch when tKH = tacc and when the central hydrogen burning begins, respectively. Lower panel: same as the upper panel, but for higher accretion rates of 6 × 10−2M yr−1 (MD6e2, magenta), 0.1 M yr−1 (MD1e1, red), 0.3 M yr−1 (MD3e1, blue), and 1 M yr−1 (MD1e0, black). The case MDe2 is illustrated in both panels as a reference. For the cases with 0.3 M yr−1 and 1 M yr−1 (MD3e1 and MD1e0), hydrogen fusion has not ignited by the time the stellar mass reaches 103M. In both panels, the thin green line represents the mass–radius relation given by Equation (11).

Standard image High-resolution image

Even after tKH becomes shorter than tacc, the stellar radius continues to increase for accretion rates ≳ 10−2M yr−1. The variations of radii among cases with different accretion rates gradually disappear. The stellar radii finally converge to a unique mass–radius relation with R*M1/2* in all the cases. We can derive the approximate mass–radius relation from the following simple argument. First, the stellar luminosity is generally written as

Equation (10)

where σ is the Stefan–Boltzmann constant. As discussed in Section 3.1 for $\dot{M}_*= 0.1 \,M_\odot \,{\rm yr}^{-1}$ (case MD1e1), the stellar luminosity generally approaches the Eddington value LEdd(M*) for M* ≳ 100 M. Figure 6(b) and Figure 7 show the evolution of the stellar effective temperature as a function of the stellar mass and on the H-R diagram, respectively. We see that the effective temperature stays constant at Teff ≃ 5000 K for the high accretion rates considered. Substituting these relations into Equation (10), we obtain

Equation (11)

Figure 5 shows that our numerical results approximately follow this relation. The stellar luminosity is actually a bit lower than the Eddington value, which is consistent with the fact that the numerical results show a slightly smaller stellar radius than the analytic relation (11).

Figure 6.

Figure 6. Upper panel: evolution of radial positions of the stellar surface (solid) and photosphere (dotted). The black, blue, and magenta curves denote cases with $\dot{M}_*= 10^{-3} \,M_\odot \,{\rm yr}^{-1}$ (case MD1e3), 6 × 10−3M yr−1 (MD6e3), and 1 M yr−1 (MD1e0), respectively. The mass–radius relation given by Equation (11) is plotted with the thin green line. Lower panel: the stellar effective temperature for the same cases as in the upper panel.

Standard image High-resolution image
Figure 7.

Figure 7. Evolutionary tracks in the H-R diagram. The different colors denote the evolution for different accretion rates, 10−3M yr−1 (case MD1e3, black), 6 × 10−3M yr−1 (MD6e3, blue), 0.1 M yr−1 (MD1e1, red), and 1 M yr−1 (MD1e0, magenta). For each case, the values at the stellar photosphere (Ltot and Teff) and at the stellar surface (L* and Teff, *L*/4πσR2*) are plotted with solid and dashed lines, respectively. The thick dashed line represents the loci of non-accreting ZAMS stars taken from Marigo et al. (2001) (M* ⩽ 100 M) and Bromm et al. (2001) (M* ⩾ 100 M). The filled circles and open squares on the lines mark the positions for M* = 30 M, 100 M, 300 M, and 103M in ascending order.

Standard image High-resolution image

Begelman (2010) also considered stellar evolution with very rapid mass accretion using simple analytic arguments. His model predicts that the stellar radius is proportional to the mass accretion rate (his Equation (24)), which does not agree with our numerical results. Begelman (2010), however, does not take into account the detailed structure of the outermost layer of the star, where H opacity is important. The fact that the strong T dependence of H opacity keeps the stellar effective temperature almost constant is essential for our results.

Cases MD3e2 and MD6e3 for intermediate accretion rates (3 × 10−2 and 6 × 10−3M yr−1, respectively) exhibit a different behavior than the higher accretion-rate cases described above (see Figure 5). Figure 6(a) shows the evolution of radial positions of the stellar surface and the photosphere. In this case, the photospheric radius is located outside the stellar surface, i.e., the accreting envelope remains optically thick after the onset of KH contraction. For an accretion rate of 3 × 10−2M yr−1 (MD3e2), the protostar initially contracts after the adiabatic accretion stage.

At M* ≃ 70 M, however, the stellar radius sharply increases and eventually converges to the mass–radius relation given by Equation (11). The case with 6 × 10−3M yr−1 exhibits an oscillatory behavior of the stellar radius for M* ≳ 70 M, but its photospheric radius still follows the mass–radius relation RphM1/2*. Figures 6(b) and 7 show that the effective temperature also assumes the constant value Teff ≃ 6000 K. These evolutionary features for $\dot{M}_*\lesssim 10^{-2} \,M_\odot \,{\rm yr}^{-1}$ have also been found in previous studies (e.g., Omukai & Palla 2001, 2003).

We have seen that the stellar radius at M* ≃ 103M is almost independent of accretion rate as long as $\dot{M}_*\gtrsim 3 \times 10^{-2} \,M_\odot \,{\rm yr}^{-1}$. However, the stellar interior structure at this moment is not identical among these cases. This is seen in Figure 8, which displays the radial distributions of physical quantities within the 103M models for the different accretion rates. Although each of these stars has a radiative core and a convective envelope, the mass is more centrally concentrated for the lower accretion rates; the less massive envelopes have a higher entropy and inflate more to achieve the same stellar radius.

Figure 8.

Figure 8. Comparisons of the interior structure of 103M stars produced by different accretion rates. Radial profiles of mass (a), specific entropy (b), and luminosity (c) are shown. In each panel, the solid and dashed curves represent the cases for 0.3 M yr−1 (MD3e1), 0.1 M yr−1 (MD1e1), 3 × 10−2M yr−1 (MD3e2), and 10−3M yr−1 (MD1e3). The magenta lines mark convective layers in the stellar interior. The thin red line in panel (c) represents the Eddington luminosity as a function of mass M.

Standard image High-resolution image

This can be understood by the following consideration. Recall that after the KH timescale tKH falls below the accretion timescale tacc, the star becomes more and more centrally condensed with increasing stellar mass as explained in Section 3.1. Since this timescale inversion occurs earlier, i.e., at a lower stellar mass for the lower accretion rate (also see the Appendix), the prolonged tKH < tacc phase until the stellar mass reaches 103M causes a more centrally condensed structure.

Figure 9(a) shows the evolution of the maximum temperature in the stellar interior, which is helpful for understanding the variation of the stellar interior structure with accretion rates. As Equation (A1) shows, the central part of the star begins to contract and release gravitational energy at a lower stellar mass for the lower accretion rates. The central temperature quickly increases with stellar mass once the KH time becomes shorter than the accretion time. The maximum temperature Tmax reaches 108 K at a lower stellar mass for the lower accretion rate. After that, Tmax assumes an almost constant value due to the strong T dependence of the energy production rate of hydrogen burning. For the case with 3 × 10−2M yr−1 (MD3e2), hydrogen burning begins at M* ≃ 200 M; the resulting central convective core is seen in the profiles in Figure 8. This feature is not seen for the case with 0.3 M yr−1, because hydrogen has not yet ignited by the time M* = 103M for $\dot{M}_*\gtrsim 0.3 \,M_\odot \,{\rm yr}^{-1}$.

Figure 9.

Figure 9. Evolution of the maximum temperature in the stellar interior (upper panel) and Eddington ratio Ltot/LEdd (lower panel) with increasing stellar mass. The solid and dashed curves alternately represent the cases with different accretion rates, 10−3M yr−1 (case MD1e3), 6 × 10−3M yr−1 (MD6e3), 3 × 10−2M yr−1 (MD3e2), 0.1 M yr−1 (MD1e1), and 0.3 M yr−1 (MD3e1).

Standard image High-resolution image

Figure 5 shows that for $\dot{M}_*\gtrsim 6 \times 10^{-3} \,M_\odot \,{\rm yr}^{-1}$, the protostar does not reach the ZAMS stage by KH contraction. Omukai & Palla (2003) explained this by the fact that the total luminosity LtotL* + Lacc becomes close to the Eddington limit during the contraction to the ZAMS. Figure 9(b) shows the evolution of the Eddington ratio Ltot/LEdd for several cases. We see that for the cases with 6 × 10−3 and 3 × 10−2M yr−1 (cases MD6e3 and MD3e2), for example, the abrupt expansion terminates the KH contraction when the total luminosity approaches the Eddington limit. Omukai & Palla (2003) analytically derived the maximum accretion rate ≃ 4 × 10−3M yr−1 with which the protostar can reach the ZAMS following KH contraction. Figure 5 indicates that there is another critical accretion rate ≃ 6 × 10−2M yr−1, above which the stellar evolution changes qualitatively; the KH contraction stage disappears entirely at higher rates. This critical rate can also be derived from a similar argument as the one above.

Note that the increase of stellar mass during the KH contraction stage is smaller for the case with 3 × 10−2M yr−1 than with 6 × 10−3M yr−1. Extending this fact for our critical case, the total luminosity would nearly reach the Eddington limit just at the end of the adiabatic accretion stage, i.e., when tKHtacc. Since the opacity in the surface layer is higher than from Thomson scattering during this epoch, having the total luminosity only slightly lower than the Eddington value causes the star to expand. Thus, the condition for the critical case is

Equation (12)

where CEdd is a factor less than the unity and we have used the fact that the total luminosity is written as Ltot ≃ 2Lmax when tKHtacc. Using CEdd = 0.25 as a fiducial value (Figure 9(b)) and Equation (4) for Lmax, the stellar mass which satisfies the condition (12) is

Equation (13)

On the other hand, Equation (A1) also gives the stellar mass when tKHtacc for a given accretion rate. Equating M*, teq and M*.Edd, teq with Equations (A1) and (13), we obtain the critical mass accretion rate

Equation (14)

which agrees with our numerical results.

3.3. Effects of Lower-entropy Accretion

We have used the shock outer boundary condition for the stellar models presented and discussed above. As discussed in Section 2.1, the shock boundary condition, which implies that the accreting gas joins the star with relatively high entropy, would be valid for cases with the very rapid mass accretion considered in this paper. If the accreted gas had lower entropy, however, then the stellar radius would be reduced because of the resulting lower entropy throughout the stellar interior. Here, we examine potential effects of the colder mass accretion by adopting the photospheric boundary conditions (e.g., Hosokawa et al. 2010, 2011a).

Figure 10(a) shows the evolution of the stellar radius for three cases with 0.3 M yr−1, whereby the shock boundary condition is used throughout in one case (MD3e1), whereas the outer boundary condition is changed to the photospheric one for >10 M (MD3e1-HCm10) and >50 M (MD3e1-HCm50), respectively. The stars are still in the adiabatic accretion stage when the boundary condition is changed in both cases. The different outer boundary conditions do affect the stellar evolution. For the case where the photospheric boundary condition is adopted at M* = 10 M (MD3e1-HCm10), for example, the star initially contracts after the boundary condition is switched at M* = 10 M, and then abruptly inflates at M* ≃ 45 M. The stellar radius exceeds 103R and gradually increases with the stellar mass thereafter. In spite of the different behaviors in the early stages, the subsequent evolution for M* ≳ 100 M is quite similar to that in the case with the shock boundary condition throughout (MD3e1). The evolution when the boundary condition switching occurs at M* = 50 M (MD3e1-HCm50) is much closer to that in the shock-boundary case.

Figure 10.

Figure 10. Effect of reducing the thermal efficiency of mass accretion (upper panel: stellar radius, lower panel: maximum temperature within the star). The same accretion rate of 0.3 M yr−1 is adopted for all three cases presented. The solid line represents the evolution with the shock boundary condition, i.e., thermally efficient or "hot" accretion, throughout (MD3e1). The dashed and dot-dashed lines show the evolution in cases MD3e1-HC10 and MD3e1-HC50, where the photospheric boundary condition (i.e., thermally inefficient or "cold" accretion) is adopted after the stellar mass exceeds 10 M and 50 M, respectively. In the lower panel, the dot-dashed line is indistinguishable from the solid line.

Standard image High-resolution image

The uniqueness of the mass–radius relation for M* ≳ 100 M can be explained by the fact that the argument leading to the analytic expression, Equation (11), does not assume a specific boundary condition. When the boundary condition is switched at M* = 10 M (MD3e1-HCm10), the stellar interior temperature is higher and thus the opacity in the stellar interior (∝T−3.5 according to Kramers' law) is lower than for the shock-boundary case (MD3e1) at the same stellar mass. As a result, the star begins to release its internal energy earlier than for the shock-boundary case. Indeed, the timescale equality between tKH and tacc occurs at M* ≃ 40 M, earlier than for the shock-boundary case, which occurs at the time of abrupt expansion of the stellar radius.

4. SUMMARY AND DISCUSSIONS

We have studied the evolution of stars growing via very rapid mass accretion with $10^{-2} \,M_\odot \,{\rm yr}^{-1}\lesssim \dot{M}_*\lesssim 1 \,M_\odot \,{\rm yr}^{-1}$, which potentially leads to formation of SMBHs in the early universe. In contrast to previous attempts to address this problem, we study the stars' evolution by numerically solving the stellar structure equations including mass accretion. Our calculations show that stellar evolution in such cases is qualitatively different from that expected for the normal Pop III star formation, which proceeds at much lower accretion rates ∼10−3 to 10−2M yr−1. Rapid mass accretion causes the star to inflate; the stellar radius further increases monotonically with stellar mass at least up to M* ≃ 103M. For masses exceeding ∼100 M, the star consists of a contracting radiative core and a bloated surface convective layer. The surface layer, which contains only a small fraction of the total stellar mass, fills out most of the stellar radius. The evolution of the stellar radius in this stage follows a unique mass–radius relation R*M1/2*, which reaches ≃ 7000 R(≃ 30 AU) at M* = 103M, in all the cases with ≳ 10−2M yr−1. Hydrogen burning begins only after the star becomes very massive (M* ≳ 100 M); its onset is shifted toward higher masses for higher accretion rates. With very high accretion rates $\dot{M}_*\gtrsim 0.1 \,M_\odot \,{\rm yr}^{-1}$, hydrogen is ignited after the stellar mass exceeds 103M. The stellar radius continues to grow as R*M1/2* even after hydrogen ignition.

In this paper, we have focused on the early evolution until the stellar mass reaches 103M. The subsequent evolution remains unexplored because of convergence difficulties with the current numerical codes. If the star continues to expand following the same mass–radius relation (11) also for M* > 103M, then the stellar radius at 105M would be ≃ 400 AU. Since the stellar effective temperature remains ≃ 5000 K, the stellar ionizing photon flux is estimated as

Equation (15)

where h is Planck's constant, Bν(Teff) is the Planck function, ν is the frequency, and Equation (11) is used for the mass–radius relation R*(M*). This value is about 105 times lower than that of ordinary Pop III ZAMS stars (e.g., Bromm et al. 2001). Therefore, it is unlikely that stellar growth is limited by the radiative feedback via formation of an H ii region as discussed by Hosokawa et al. (2011b).

Johnson et al. (2011) also reached an analogous conclusion that UV feedback does not hinder SMS formation. In their argument, however, the star is assumed to reach the ZAMS and to emit a copious amount of ionizing photons, but the expansion of the H ii region is squelched by rapid spherical inflow. They also expected that, as a result of confinement of the H ii region, strong emission lines reprocessed from the ionizing photons (e.g., H α and He ii 1640) would escape from the accretion envelope to be an observational signature of these objects. By contrast, our calculations show that the stellar UV luminosity and thus the luminosities in those lines should be much weaker than supposed. Note that the argument by Johnson et al. (2011) assumes perfect spherical symmetry, which allows the H ii region to be confined within the accretion envelope. Given that mass accretion will likely occur through a circumstellar disk, the H ii region should grow toward the polar region where the gas density is much lower than the spherical accretion flow (e.g., Hosokawa et al. 2011b). This should be the case with the high stellar UV luminosity assumed in Johnson et al. (2011).

Even without stellar radiative feedback, stellar growth via mass accretion might be hindered by some other process, e.g., rapid mass loss. Indeed, evolved massive stars in the Galaxy (M* ∼ 10–100 M), which have large radii (R* ≳ 100 R) and high luminosities close to the Eddington limit (L* ≃ 106L), generally have strong stellar winds with mass losses ∼10−4M yr−1 (e.g., Humphreys & Davidson 1994). Although the line-driven winds of primordial stars are predicted to be weak or non-existent (Krtička & Kubát 2006), pulsational instability of massive stars has also been found to drive mass loss (Baraffe et al. 2001; Sonoi & Umeda 2011). Further work is necessary to address how massive SMSs could form via mass accretion in spite of such disruptive effects.

Stellar evolution under conditions of very rapid mass accretion as presented and discussed here is mostly relevant to the formation of stars in the atomic-cooling halos. However, our results could be also important for normal Pop III star formation where H2 molecular cooling operates. The typical mass accretion rate for this case is around 10−3M yr−1, but in some exceptional situations, e.g., when a progenitor cloud core is extremely slow rotating, higher accretion rates $\dot{M}_*\sim 10^{-2} \,M_\odot \,{\rm yr}^{-1}$ can be realized (e.g., Hosokawa et al. 2011b). Since the stellar effective temperature is low at ≃ 5000 K with rapid mass accretion, formation of the H ii region would be postponed until the mass accretion rate falls below 10−2M yr−1. This would help the primordial star to grow to more than 100 M in the molecular-cooling halos (see also Omukai & Palla 2003).

The authors thank Francesco Palla, Mitchell Begelman, Zoltan Haiman, Milos Milosavljevic, Jarrett Johnson, Neal Turner, Rolf Kuiper, Kohei Inayoshi, and Naoki Yoshida for fruitful discussions and comments. T.H. appreciates the support by Fellowship of the Japan Society for the Promotion of Science for Research Abroad. K.O. is supported by the Grants-in-Aid by the Ministry of Education, Science and Culture of Japan (2168407 and 21244021). Portions of this work were conducted at the Jet Propulsion Laboratory, California Institute of Technology, operating under a contract with the National Aeronautics and Space Administration (NASA).

APPENDIX: EPOCH OF THE END OF THE ADIABATIC ACCRETION STAGE

Here, we derive an analytic expression to explain how the length of the adiabatic accretion phase depends on the accretion rate. As long as the KH timescale tKH is longer than the accretion timescale tacc, we have adiabatic accretion (e.g., Hosokawa & Omukai 2009). Figure 2 shows that Lmax, whose evolution is well described by simple analytic expressions (Equations (3) and (4)), converges to L* at the end of the adiabatic accretion stage. Thus, the stellar mass at which adiabatic accretion terminates M*, teq can be estimated using the relation tacctKH = GM2*/R*Lmax. Eliminating Lmax and R* with Equations (4) and (6), we obtain

Equation (A1)

We have confirmed that the epoch of the timescale equality in our numerical calculation is approximately described by this equation.

Please wait… references are loading.
10.1088/0004-637X/756/1/93