Brought to you by:

26Al AND THE FORMATION OF THE SOLAR SYSTEM FROM A MOLECULAR CLOUD CONTAMINATED BY WOLF–RAYET WINDS

, , , and

Published 2009 April 27 © 2009. The American Astronomical Society. All rights reserved.
, , Citation Eric Gaidos et al 2009 ApJ 696 1854 DOI 10.1088/0004-637X/696/2/1854

0004-637X/696/2/1854

ABSTRACT

In agreement with previous work, we show that the presence of the short-lived radionuclide (SLR) 26Al in the early solar system was unlikely (less than 2% a priori probability) to be the result of direct introduction of supernova (SN) ejecta into the gaseous disk during the Class II stage of protosolar evolution. We also show that Bondi–Hoyle accretion of any contaminated residual gas from the Sun's natal star cluster contributed negligible 26Al to the primordial solar system. Our calculations are consistent with the absence of the oxygen isotopic signature expected with any late introduction of SN ejecta into the protoplanetary disk. Instead, the presence of 26Al in the oldest solar system solids (calcium–aluminum-rich inclusions (CAIs)) and its apparent uniform distribution with the inferred canonical 26Al/27Al ratio of (4.5–5) × 10−5 support the inheritance of 26Al from the Sun's parent giant molecular cloud. We propose that this radionuclide originated in a prior generation of massive stars that formed in the same molecular cloud and contaminated that cloud by Wolf–Rayet winds. We calculated the Galactic distribution of 26Al/27Al ratios that arise from such contamination using the established embedded cluster mass and stellar initial mass functions, published nucleosynthetic yields from the winds of massive stars, and by assuming rapid and uniform mixing into the cloud. Although our model predicts that the majority of stellar systems contain no 26Al from massive stars, and that the a priori probability that the 26Al/27Al ratio will reach or exceed the canonical solar system value is only ∼6%, the maximum in the distribution of nonzero values is close to the canonical 26Al/27Al ratio. We find that the Sun most likely formed 4–5 million years (Myr) after the massive stars that were the source of 26Al. Furthermore, our model can explain the initial solar system abundance of a second, co-occurring SLR, 41Ca, if ∼5 × 105 yr elapsed between ejection of the radionuclides and the formation of CAIs. The presence of a third radionuclide, 60Fe, can be quantitatively explained if (1) the Sun formed immediately after the first SNe from the earlier generation of stars; (2) only 5% of SN ejecta was incorporated into the molecular cloud, or (3) the radionuclide originated in an even earlier generation of stars whose contributions to other radionuclides with a shorter half-life had completely decayed.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

1.1. Short-Lived Radionuclides in the Early Solar System

Primitive meteoritic materials contain compelling evidence for the short-lived radionuclides (SLRs) 10Be, 26Al, 36Cl, 41Ca, 53Mn, 60Fe, 107Pd, and 182Hf in the early solar system (Goswami et al. 2005). These radionuclides have a half-life τ1/2 < 10 Myr and are potential high-resolution chronometers of events during the epoch of planet formation (Kita et al. 2005; Krot et al. 2008b), but only if they were introduced at discrete times and were uniformly distributed in the solar system. The potential sources of these radionuclides inform us about the young Sun's stellar neighborhood (Hester et al. 2004), or its magnetic interaction with an accretion disk (Shu et al. 1997). The decay of one SLR, 26Al, might have been the principal heat source in planetesimals and responsible for the differentiation of the parent bodies of magmatic iron meteorites in the first 1–2 Myr of the solar system (Greenwood et al. 2005; Scherstén et al. 2006; Markowski et al. 2006).

The origin of SLRs is controversial. The half-life of each is much shorter than the ∼100 Myr mixing time of the interstellar medium (de Avillez & Mac Low 2002) and the excess abundances of at least five radionuclides (10Be, 36Cl, 26Al, 41Ca, and 60Fe) require one or more "local" sources in addition to the average Galactic background (Jacobsen 2005). Two principal scenarios emerged soon after the first reports of fossil SLRs in meteorites: (1) an origin in one or more neighboring massive stars, either Type II supernova (SN) progenitors (Cameron & Truran 1977) or Wolf–Rayet (W-R) stars (Arnould & Prantzos 1986) and (2) production by irradiation of gas or dust with energetic particles from the active young Sun (Heymann & Dziczkaniec 1976). An alternative scenario invoking an origin in a nearby intermediate-mass asymptotic giant branch (AGB) star (Wasserburg et al. 1994) is generally discounted because of the very low probability of such an encounter (Kastner & Myers 1994). Each of the two schools of thought has developed elaborate models (Lee et al. 1998; Gounelle et al. 2006; Ouellette et al. 2007) but neither has produced a comprehensive explanation of the origin, abundance, and distribution of all the SLRs (Goswami et al. 2005; Gounelle et al. 2006; Duprat & Tatischeff 2007).

The origins of three SLRs seem unambiguous: (1) 10Be (τ1/2 ∼ 1.5 Myr) inferred from excess 10B (McKeegan et al. 2000) could have been produced by energetic particle irradiation but not by stellar nucleosynthesis. Although magnetic trapping of Galactic cosmic rays in the protosolar molecular cloud has been advanced as an alternative explanation (Desch et al. 2004), it is inconsistent with variations in the inferred initial 10Be/9Be ratio (McKeegan et al. 2000; Sugiura et al. 2001; Marhas et al. 2002; MacPherson et al. 2003). (2) Excess 36S correlated with the ratio 35Cl/34S and attributable to the decay of 36Cl (τ1/2 ∼ 0.3 Myr) was reported in sodalite, an alteration phase that replaced anorthite in calcium–aluminum-rich inclusions (CAIs) and chondrules from CV chondrites (Lin et al. 2005; Hsu et al. 2006). The inferred 36Cl/35Cl ratio at the time of sodalite formation is 5 × 10−6. If the sodalite formed late (greater than 1.5 Myr after CAI crystallization, based on absence of 26Al), the initial 36Cl/35Cl ratio was >1.6 × 10−4. This is inconsistent with a massive stellar source and requires a late episode of irradiation. (3) 60Ni excess correlated with the 56Fe/58Ni ratio is evidence for live 60Fe, a radionuclide that cannot be produced by irradiation and must have originated in one or more massive stars (Tachibana & Huss 2003). High-precision nickel isotope measurements in several groups of magmatic iron meteorites indicate that 60Fe was uniformly distributed in the solar nebula (Dauphas et al. 2008) but its initial abundance, (60Fe/56Fe)0 ∼ (0.5–1) × 10−6, is uncertain (Tachibana et al. 2006; Bizzarro et al. 2007; Quitté et al. 2007; Guan et al. 2007). The lower end of the range of estimates is consistent with the expected abundance in star-forming regions if star formation rates were approximately twice as high at the epoch of solar system formation (Williams 2008, Gounelle & Meibom 2008a) and/or the half-life of 60Fe is actually longer than the published value (τ1/2 ≈ 1.5 Myr).

1.2. The Origin of 26Al

Ironically, the origin of 26Al (τ1/2 ∼ 0.73 Myr), the first SLR to be discovered (Lee et al. 1976) and the one best studied, remains an enigma. A significant contribution by irradiation is disputed on the grounds that (1) production models adjusted to achieve the required 26Al/27Al ratio overpredict the observed initial abundance of 41Ca (τ1/2 ∼ 0.1 Myr) whose co-occurrence indicates a common origin (Sahijpal & Goswami 1998; Goswami et al. 2005), (2) the flux of energetic particle inferred from X-ray observations of young solar-type stars would have been insufficient to produce the 26Al (Duprat & Tatischeff 2007), and (3) a lack of correlation between the presence of 26Al and 10Be, which was formed by irradiation (MacPherson et al. 2003; Marhas & Goswami 2004). Finally, the canonical value [(4.5–5) × 10−5] of the initial 26Al/27Al ratio in the majority of CAIs from primitive chondrites (MacPherson et al. 1995; Jacobsen et al. 2008; Makide et al. 2008), the consistent chronology of CAI and chondrule formation between 207Pb–206Pb and 26Al–26Mg systematics (Amelin et al. 2002; Halliday & Kleine 2006; Connelly et al. 2008), and the apparently uniform Mg–isotope compositions of bulk chondrites, Mars, Moon, and the Earth (Thrane et al. 2006) are evidence for a uniform distribution thought inconsistent with a central (solar) irradiation source.

Although these observations favor an origin of 26Al in massive stars, the mechanism and timing of its delivery to the early solar system remains unclear. Two models have been proposed: instability-induced mixing of gaseous SN ejecta during the molecular cloud core phase (Cameron & Truran 1977) and injection of SLR-bearing dust grains into the later protoplanetary disk (Ouellette et al. 2005). The relatively small cross section of the solar nebula dictated that the progenitor was within ∼1 pc of the solar system (Looney et al. 2006). Such a circumstance is possible only in the dense environment of a large stellar cluster (Hester et al. 2004). Incorporation of hot, low-density gas from SN ejecta into the denser, cooler disk gas is inefficient (Vanhala & Boss 2002; Ouellette et al. 2007; Boss et al. 2008). Instead, Ouellette et al. (2005) proposed that several SLRs, including 26Al, were delivered to the early solar system as grains that condensed from SN ejecta and vaporized upon entering the relatively high-density gas in the disk (Ouellette et al. 2007). Disks around low-mass cluster stars persist for up to 6 Myr (Haisch et al. 2005; Jayawardhana et al. 2006), longer than the main-sequence lifetime of the most massive SN progenitors (Schaller et al. 1992).

However, any explanation for solar system 26Al invoking a late introduction of SN ejecta has four significant shortcomings : (1) not all stars form in large clusters and most clusters are dynamically unbound and disperse in ∼10 Myr (Lada & Lada 2003). It is statistically unlikely that the Sun would have been sufficiently close to a massive star at the end of the latter's main-sequence life (Williams & Gaidos 2007; Gounelle & Meibom 2008b). (2) Even the most massive stars have a main-sequence lifetime of at least 3 Myr and, if they and the Sun formed simultaneously, the former would have ended in SNe late in the evolution of the solar nebula and probably long after CAIs containing the canonical 26Al/27Al ratio had formed. These CAIs, the oldest dated solids from the solar system (Amelin et al. 2002), have a narrow range of inferred initial 26Al/27Al ratios suggesting that they formed in ≪1 Myr (Thrane et al. 2006; Jacobsen et al. 2008), consistent with the duration of Class 0–I stages of protostars (Smith 2004; Ward-Thompson et al. 2007). A short interval of CAI formation is also consistent with the narrow range of their oxygen isotope compositions (Δ17O = −24 ±  2% MacPherson et al. 2008; Makide et al. 2008), which are similar to the inferred oxygen isotopic composition of the Sun (Hashizume & Chaussidon 2005). (Later CAI formation would presumably have reflected the rapid oxygen isotopic evolution of the solar nebula along the slope-one carbonaceous chondrite anhydrous mineral (CCAM) line toward the terrestrial value (Δ17O = 0%), a trend attributed to CO photochemical self-shielding and radial mixing in the disk (Yurimoto et al. 2007; Aléon et al. 2007).) (3) Late injection of 26Al into the protoplanetary disk would likely have disturbed the oxygen isotopic composition of the disk from the CCAM line, and this is not observed (Gounelle & Meibom 2007). (4) SN models that produce the canonical 26Al/27Al ratio invariably overpredict the abundance of 53Mn and must impose fallback of the innermost layers, e.g., Meyer (2005). (These models also overpredict the abundance of 60Fe, see Section 5.3.)

A scenario in which SLRs from coeval SN progenitors were injected into the protoplanetary disk can be evaluated for its statistical plausibility. Gounelle & Meibom (2008b) estimated that the probability that any given disk is contaminated by a SN with enough 26Al to reach the canonical 26Al/27Al ratio is ⩽0.3%. However, there are three limitations to their model: (1) it underestimated the probability by assuming a maximum disk radius of 50 AU based on observations and theoretical arguments that disks within 0.1 pc of massive (O) stars suffer from photoevaporation (Johnstone et al. 1998; Chevalier 2000). But disks further from O stars are invariably larger (Vicente & Alves 2005; Andrews & Williams 2007; Balog et al. 2007) and this effect compensates the greater distance from the source of radionuclides. (2) Their model overestimated the probability by neglecting the expansion of the host star cluster and using the ⩽1 Myr old Orion Nebular Cluster as a template, rather than a cluster at the minimum age (3 Myr) when SNe occur. In older clusters, stars are more widely separated. (3) It only included the contribution of SNe to 26Al. 26Al is also produced and ejected from stars with initial masses ⩾40 M in their luminous blue variable (LBV) and W-R phases (Arnould & Prantzos 1986; Palacios et al. 2005; Sahijpal & Soni 2006).

1.3. Was Solar System 26Al Inherited?

The uniform distribution of 26Al, the existence of the canonical 26Al/27Al ratio in CAIs, the adherence of primitive oxygen isotope compositions to the CCAM line, and the low likelihood of an SN injection event all point to the introduction of 26Al before the collapse of the protosolar cloud. Indeed, the oxygen isotopic composition of the entire solar system appears to be displaced from the locus of mean Galactic evolution (Young et al. 2008), suggesting primordial contamination. The source of 26Al must also have introduced 41Ca, which co-occurs with an inferred initial abundance of 41Ca/40Ca ∼1.5 × 10−8 (Sahijpal & Goswami 1998). The half-life of this radionuclide is only 0.1 Myr, limiting any time delay between production at the source and its incorporation into CAIs. The source of 26Al cannot have been accompanied by substantial 53Mn, as is the case of SN ejecta without fallback onto the remnant (Meyer 2005). (The predicted 53Mn abundance in the interstellar medium is consistent with its inferred initial abundance in the solar system and a "local" source is not required.) Any relationship between the source of 26Al and that of 60Fe remains to be determined. There is as yet no evidence that 60Fe and 26Al are correlated and had the same origin. Indeed, there may be a deficit of 60Fe relative to 26Al compared with SN ejecta that cannot be explained by free decay of the two radionuclides: the initial ratio of 60Fe to 26Al in the solar system was 0.1–0.2 (Tachibana et al. 2006), lower than the 0.3 deduced from the Galactic average γ-ray emission (Wang et al. 2007) and theoretical predictions (Limongi & Chieffi 2006; Woosley & Heger 2007).

Very massive (⩾40 M) stars eject 26Al (and other SLRs) during the W-R phase of mass loss near the end of hydrogen core burning, as well as in SN (Arnould & Prantzos 1986). W-R winds might account for a large fraction of the total fluence and Galactic distribution of γ-rays from the decay of 26Al (Palacios et al. 2005; Diehl 2006; Voss et al. 2008; Martin et al. 2008). (The nondetection of γ-ray emission in the decay line of 26Al from the nearest W-R star γ2-Velorum can be explained by the dispersal of most of the radionuclide to large angular separation Mowlavi & Meynet 2006). We propose that most or all of the 26Al in the early solar system originated in W-R winds from one or more massive stars that contaminated the molecular cloud from which the Sun formed. (Sahijpal & Soni 2006 also considered the contribution of W-R winds to solar system inventories of SLRs.) These stars could have been members of the same embedded cluster as the Sun, or, more likely, members of another cluster that formed in the same giant molecular cloud (GMC; Figure 1). An analogous "self-contamination" scenario has been invoked to explain anomalous abundance patterns in some globular clusters (Smith 2006).

Figure 1.

Figure 1. Cartoon of the molecular cloud contamination scenario to explain the presence of 26Al in the early solar system: (a) A young cluster of stars forms an H ii region in a GMC; (b) massive stars eject 26Al into the cloud; (c) a clump collapses from the contaminated cloud; (d) a second generation of stars (including the Sun) forms from the clump.

Standard image High-resolution image

Our proposal is based upon the following: (1) the amount of 26Al ejected in the winds of a single 60 M star (Limongi & Chieffi 2006) is sufficient to contaminate 2 × 104 M of solar-metallicity gas to the canonical 26Al/27Al ratio of the solar system. (2) The W-R phase can occur as soon as 1–2 Myr after star formation and more than 1 Myr before the SN (Palacios et al. 2005), making it more likely to contaminate residual star-forming molecular gas. (3) W-R winds have speeds of up to 2000 km s−1 (Niedzielski & Skórzynński 2002) and can traverse star-forming regions (10–100 pc) in 104–105 yr. (4) In contrast to single clusters where star-formation may be coeval, star formation in a molecular cloud can occur over an interval of at least a few Myr (Hartmann et al. 2001), and possibly longer (Williams et al. 2000) as clumps of gas with a mass spectrum M−1.7±0.1 (Pudritz 2002) form multiple embedded stellar clusters (Williams et al. 2000). For example, the Orion star-forming complex contains several subgroups that are several Myr older than the Orion Nebula Cluster (Bally 2008). (5) W-R winds contain multiple SLRs, including 26Al, 41Ca, and 36Cl, but little or no 53Mn or 60Fe (Arnould et al. 2006). We propose that the collapse of the protosolar cloud homogenized the distribution of these isotopes (but see Section 5.2). Our scenario does not preclude an SN-triggered collapse (Cameron & Truran 1977; Boss et al. 2008), which would have occurred after the W-R phase.

In Section 2, we revisit the scenario of 26Al introduction by SN into the protoplanetary disk with a Monte Carlo approach that used more realistic disk sizes, accounted for the expansion of clusters, and included the contribution from both W-R winds and SN. While our results differ quantitatively from those of Gounelle & Meibom (2008b), the calculated probability of a disk having the canonical solar system 26Al/27Al ratio is nevertheless small (less than 2%). We also consider Bondi–Hoyle accretion of contaminated intracluster gas, as proposed by Throop & Bally (2008) and find its inclusion does not significantly alter this result (Section 3). We then use a similar Monte Carlo model to investigate a scenario where 26Al is introduced by W-R winds into the parent molecular cloud of the Sun (Section 4). This model readily reproduces the canonical 26Al/27Al ratio of the solar system. We discuss the delivery of 26Al from W-R winds into the molecular cloud, interpret CAIs that lack the canonical 26Al/27Al ratio, and, in the context of the cloud contamination scenario, present additional calculations for three other SLRs (41Ca, 60Fe, and 36Cl) in Section 5.

2. DISK CONTAMINATION SCENARIO

We calculated the Galactic distribution of 26Al/27Al ratio in the disk injection scenario (Ouellette et al. 2005) by extending the model of Williams & Gaidos (2007) to include 26Al from both winds and SN, a dynamically realistic description of cluster expansion, and the effect of the UV field of massive stars on disk size. Unlike Williams & Gaidos (2007) we do not explicitly consider disk evolution and disappearance in our model because this occurs on a timescale of ∼6 Myr (Haisch et al. 2005), much longer than the likely epoch of CAI formation.

For simplicity, we assume that every young low-mass star has a disk and that the properties of the disk are independent of stellar mass. According to Lada & Lada (2003), 72% of stars form in clusters with N*>100 members distributed in size according to dNc/dN*N−2*. The remaining 28% form in isolation or in clusters with fewer than N* = 100 stars. The most massive members of such small clusters will be ∼5 M B stars that do not produce winds or Type II core-collapse SN. Disks around members of such clusters will not receive any exogenous 26Al. The other 72% of disks were represented by 105 Monte Carlo calculations. The size of the host cluster of each disk was drawn from an N−1* distribution (N* ⩾ 100). The number of massive stars (SN progenitors with M*>8 M) in the host cluster was selected from a Poisson distribution with an average of 3 × 10−3N* (Williams & Gaidos 2007). The masses of these stars were drawn from a power-law initial mass function dN*/dM*M−2.5* (Kroupa 2002). Clusters form over ∼1 Myr (Hillenbrand et al. 2007) and most star formation in a single cluster occurs within ∼3 Myr (Hartmann 2001; Hartmann 2003; Huff & Stahler 2006; Jeffries 2007; Hillenbrand et al. 2007). We assumed the instantaneous formation of all massive stars and an exponentially decaying rate of low-mass star formation after massive star formation. We used age statistics for members of the Orion Nebula Cluster (Palla & Stahler 1999) to infer a decay time of 2.7 Myr, which we adopted for clusters of all sizes. We assumed that the rate of star formation does not depend on stellar mass.

The 26Al/27Al ratio in the disk around the ith low-mass star at an interval T after the star's formation at time t*i was calculated by summing over the product of the yield mj of the radionuclide from the jth wind or SN, the solid angle subtended by the disk at the time of ejection tj, and the factor of free decay between that time and t*i + T:

Equation (1)

where rd and md are the (constant) radius and mass of the disk, fAl is the mass fraction of 27Al, and dij is the distance of the disk from the source star at the time of ejection. We considered injection as instantaneous because the speeds of SN ejecta and winds are >1000 km s−1, and the 26Al-producing W-R phase of an individual massive star is brief (∼1 Myr) compared with the dispersal timescale of the cluster (∼10 Myr). To account for the changing perspective of each star as it orbits inside the cluster, we used the isotropic average of the projected cross section of each disk, πr2d/2. mj and tj were estimated by spline interpolation in a grid of yield calculations by Limongi & Chieffi (2006). We used a default disk radius of 200 AU (Vicente & Alves 2005; Andrews & Williams 2007) but to account for photoevaporation by the UV radiation from massive stars (Johnstone et al. 1998) we reduced this to 30 AU for disks within 0.2 pc of the cluster center at 3 Myr, and to zero for disks within 0.1 pc. This was a conservative assumption, since the evidence for significant disk truncation is weak (Balog et al. 2007), but it only has a minor effect on our results. Like Ouellette et al. (2005), we adopted a minimum mass solar nebula disk mass of 0.013 M and the fractional abundance of 27Al given by Lodders (2003).

As in Williams & Gaidos (2007), we placed the massive stars at the cluster center (Grebel 2007) and assumed that, at any time, low-mass stars were distributed with an inverse-square density profile, such that an equal number of stars reside in shells of constant thickness out to the edge of the cluster at rc. To model cluster expansion, and thereby determine the distance of a disk to each source of 26Al, we developed an empirical relationship for the time dependence of rc based on a series of numerical simulations of clusters containing between 100 and 15,000 stars. The dynamical simulations were preformed using the NBODY4 code running on the Cambridge University GRAPE-6a card (Aarseth 2003). In each case, the stars were initially (3 Myr) distributed in a Plummer sphere (Binney & Tremaine 1988) with a radius set by the requirement that the initial surface density Σ3 = 100 pc−2 (Williams & Gaidos 2007). The virial parameter Ω (ratio of kinetic to gravitational potential energy) was set to 1.5. This condition is brought about by a cluster local star formation efficiency of 33% in the embedded cluster and the instantaneous removal of the remaining gas at 3 Myr (Bastian & Goodwin 2006). The size of the cluster was explicitly determined at regular intervals until an age of 10 Myr. We found that the expansion of the cluster from its size at 3 Myr was closely approximated by

Equation (2)

where G is the gravitational constant and $\bar{M_*}$ is the average stellar mass. The expression multiplying 0.45 is the cluster's virial speed (Binney & Tremaine 1988). In each of our simulations the surface density of the cluster fell to the background level of field stars (2–3 pc−2) by 10 Myr, in agreement with observations (Lada & Lada 2003).

Equation (2) specifies the radius of the cluster at the epoch tj of a massive stellar wind or SN. The uniform distribution of low-mass stars with distance from the cluster center 0 < dij < rc(tj) was then used in Equation (1) to produce a distribution of 26Al/27Al. These distributions were summed over all events in a cluster, corrected for free decay, averaged over 105 realizations, and multiplied by 0.78 to produce a Galactic distribution.

The calculated distributions of 26Al/27Al are plotted in Figure 2 and the fractions of systems that have any initial 26Al or 26Al abundances exceeding the canonical value are given in Table 1. We also report the 95 percentile values of 26Al/27Al. We considered two values for T, which in the solar system represents the epoch of CAI formation. The probability is 1.1% for T = 0.5 Myr, rising to 1.9% by 1 Myr. The probability is higher at still later, but unlikely CAI formation times (not shown). Our probabilities are several times higher than that reported by Gounelle & Meibom (2008b), mostly due to the larger disk size we used, but are still small. Our estimates are nonetheless optimistic because we assume all high-mass stars form prior to low-mass stars. If all stars form instantaneously, then nearly 3 Myr must elapse before 26Al is produced, and no disks contain 26Al by the time of CAI formation.

Figure 2.

Figure 2. Distribution of nonzero values of the 26Al/27Al ratio predicted by a Monte Carlo model of its formation in massive stars and incorporation injection into the protoplanetary disk (see the text for details). The units of the ordinate are fractional number of Monte Carlo systems per unit common logarithm. The vertical bar demarks the canonical solar system value of 5 × 10−5. The curves do not integrate to unity because a large majority of systems are not contaminated with 26Al (Table 1). The solid curves are for the disk injection scenario (see the text), and the dashed curves include Bondi–Hoyle accretion of nebular gas. Two values for the elapsed time between the formation of the Sun and 26Al-containing CAIs are considered.

Standard image High-resolution image

Table 1. Statistics of 26Al Abundance in Different Scenarios

Scenario >0 % ⩾5 × 10−5 26Al/27Al 95%
Disk injection
T = 0.5 Myr 16 1.2 5 × 10−6
T = 1 Myr 20 1.9 1.3 × 10−5
Disk injection with Bondi–Hoyle accretion
T = 0.5 Myr 16 1.2 8 × 10−6
T = 1 Myr 21 1.9 1.6 × 10−5
Molecular gas contamination (T = 0)
2.7 Myr exp. SF 16 6.2 9 × 10−5
1.7 Myr exp. SF 8 4.3 6 × 10−5
3.7 Myr exp. SF 21 6.5 9 × 10−5

Notes. Only simulations which produced 26Al-contaminated systems are reported.

Download table as:  ASCIITypeset image

3. BONDI–HOYLE ACCRETION SCENARIO

Disks might also accrete gas from their natal cluster before it is removed by winds and SN explosions. Throop & Bally (2008) proposed that Bondi–Hoyle accretion of residual, SN-contaminated cluster gas onto the protosolar nebula produced isotopic anomalies in the solar system, including the presence of SLRs. They estimated that disks around solar-mass stars in a 3000-star cluster could accrete an additional ∼0.01 M of gas, an amount comparable to the original mass of a disk, over the 2–4 Myr that gas remained in the cluster. We estimated the amount of 26Al that could be introduced by this process by specifying the fraction of disk mass f acquired by Bondi–Hoyle accretion by the time of CAI formation, and assuming that accretion is constant as long as intracluster gas is present. We assumed that the maximum amount of mass that can be accreted by a disk is equal to the initial disk mass (Throop & Bally 2008), i.e., f ⩽ 0.5. To account for the absence of cluster gas during the history of later forming stars we adjusted the accreted mass by the ratio of the time interval between low-mass star formation and the disappearance of cluster gas, and the lifetime of the cluster gas. Like Throop & Bally (2008), we assumed a gas lifetime of 3 Myr. We calculated the average 26Al/27Al ratio of cluster gas during the period of Bondi–Hoyle accretion (see Section 4 for details) and then determined the final 26Al/27Al of the disk as 26Al/27Al = (1 − f)(26Al/27Al)0 + f(26Al/27Al)acc, where the subscripts refer to the initial value and the average value during accretion.

We calculated distributions of the 26Al/27Al ratio for combined Bondi–Hoyle accretion and disk injection and found that the former has a negligible effect (Figure 2 and Table 1). This is because low-mass stars that form late enough (∼3 Myr) to acquire significant 26Al will accrete little gas because the intracluster gas disappears soon thereafter. This is of course entirely a result of our (reasonable) assumption that intracluster gas is evacuated by the time the massive stars leave the main sequence, if not earlier. Nevertheless, the same cosmochemical timing arguments that apply to the disk injection scenario also apply to the Bondi–Hoyle scenario; CAIs probably formed by the time a disk had formed, or very soon thereafter, and thus later accretion of contaminated gas cannot be responsible for the presence of 26Al in them.

4. MOLECULAR CLOUD CONTAMINATION SCENARIO

In this scenario the Sun's natal GMC spawned an earlier generation of massive stars (Figure 1(a)) whose W-R winds contaminated the rest of the cloud (Figure 1(b)), from which the Sun subsequently formed, perhaps in a second cluster (Figure 1(c and d)). We calculated the distribution of 26Al/27Al ratios in this scenario by 105 Monte Carlo simulations, each corresponding to a disk formed from a GMC contaminated by an immediately previous generation of massive cluster stars. The number and masses of those stars were drawn from the distributions described in Section 2. The mass of gas in the GMC was calculated using an average stellar mass derived from the initial mass function of Kroupa (2002), and a total star-formation efficiency of 10% (Williams et al. 2000). The amount of 26Al added to the molecular cloud by massive stellar winds was calculated as a function of time using the yields and times of Limongi & Chieffi (2006) and by assuming 100% delivery efficiency and instantaneous mixing into the cloud (we discuss this assumption further in Section 5.1). We did not include SN ejecta in these calculations, but include it when we consider 60Fe in Section 5.3 (but see the footnote on SN ejecta delivery in Section 5.1). We assumed that the Sun and CAIs formed simultaneously.

We first carried out a series of calculations for different intervals of elapsed time (3–6 Myr) between the formation of the earlier generation of massive stars and the Sun (dashed lines in Figure 3). If the interval is less than 3 Myr contamination has yet to take place in our model and newly formed stellar systems lack 26Al. An interval of 4–5 Myr is most likely to produce the CAI value. The history of low-mass star formation in molecular cloud complexes is poorly known but clearly nonmonotonic (Hartmann et al. 2001; Hartmann 2003). We calculated more realistic distributions by again adopting the exponential rate with a 2.7 Myr decay time based on the data of Palla & Stahler (1999). This is plotted as the heavy solid line in Figure 3. That curve can be understood as a convolution of the star formation history with the 26Al/27Al distributions for "starburst" scenarios. The probability of 26Al/27Al exceeding the canonical value is 6.2%. This figure depends on the assumed star formation history: for example, varying the decay time constant by ±1 Myr changes the fraction between 4.3% and 6.5%. However, the peaks in all three distributions are near the CAI value (solid lines in Figure 3). This robustness is a result of negligible 26Al production at times earlier than 3 Myr and negligible low-mass star formation at times much later than 5 Myr. We found that simulations which produce a 26Al/27Al ratio within 2σ (σ = 0.1 × 10−5 (Goswami et al. 2005)) of the canonical value were most likely to involve contamination of a GMC having ∼3 × 105 M of gas by the massive members of a cluster with ∼105 stars. Such a situation is exemplified in our galaxy by NGC 3603, which contains multiple W-R stars within an H ii region (Melena et al. 2008).

Figure 3.

Figure 3. Distribution of nonzero values of the 26Al/27Al ratio predicted by an alternative model in which 26Al was produced in an earlier generation of massive stars and introduced by W-R winds into the molecular cloud that formed the Sun (see the text for details). The units of the ordinate are fractional number of Monte Carlo systems per unit common logarithm. The vertical bar is the canonical solar system value of 5 × 10−5. The integral of the curves is not unity because the majority of systems are not contaminated with 26Al (Table 1). The broken curves are for a monotonic elapsed time between the formation of the earlier generation of massive, 26Al-producing stars, and the Sun. The solid curves (shown one-tenth scale) are for an exponentially decaying rate of star formation with a decay time of 2.7 Myr (heavy curve) or 1.7 and 3.7 Myr (light curves). The area under the solid curves changes with assume star formation history, but the location of the peak does not.

Standard image High-resolution image

5. DISCUSSION

5.1. Delivery of 26Al into the Molecular Cloud

Our scenario requires a plausible mechanism for the efficient introduction of the 26Al carrier in W-R winds into the surrounding molecular cloud. In the calculations above, we assumed a 100% delivery efficiency but this will clearly not be the case. In general, mixing between the hot, tenuous gas from massive stars and cooler, denser molecular gas is thought to be very inefficient (de Avillez & Mac Low 2002). SN-enriched gas from H ii regions is thought to find its way back into the interstellar medium (if at all) through a circuitous route taking ∼100 Myr (Tenorio-Tagle 2000). Impact of SN ejecta onto a protostellar cloud core may induce mixing via Rayleigh–Taylor and Kelvin–Helmholtz instabilities (Boss & Foster 1998; Boss et al. 2008) but the injection efficiency is too low to explain the solar system's canonical 26Al/27Al ratio.6 Gas in W-R winds will be less dense than SN ejecta by several orders of magnitude and efficient mixing in the gas phase is even less likely. Instead, the high-velocity (500–2000 km s−1) winds will develop a reverse shock upon encountering the much denser molecular cloud (Weaver et al. 1977).

We propose that refractory dust grains are the principal carrier of 26Al and 41Ca in W-R winds and that these can dynamically decouple from the shocked wind and imbed themselves into the surrounding molecular cloud, analogous to the "aerogel" model described by Ouellette et al. (2005). Presolar grains of Al2O3 (corundum or other forms) are found in primitive meteorites but their oxygen isotopes indicate a source in red giant or AGB stars, not in very massive stars (Hutcheon et al. 1994; Ott 2007; Nittler et al. 2008). Although many of these grains have large 26Mg excesses produced by the decay of 26Al, the abundance of these grains is insufficient to account for the canonical 26Al/27Al ratio (Hutcheon et al. 1994). In fact, the 26Al in CAIs must have been processed through the gas phase and subsequently recondensed into the refractory inclusions. The presolar grains are also much larger (∼1 μm) than the silicate grains predicted to form from winds and ejecta. Their size may be why they survived the incorporation process and the latter did not.

Simple models of grain nucleation and growth predict oxide grain growth to sizes of 0.01–0.1 μm in SN ejecta (Nozawa et al. 2007). SN are predicted to be copious sources of dust, but observations have so far produced evidence only for a few times 10−5 M of dust in individual events (Meikle et al. 2007). Dust production in W-R stars is poorly investigated, although such stars appear to be minor contributors to the overall interstellar dust budget (Tielens et al. 2005). Copious amorphous carbon dust is observed around carbon-rich W-R CO stars and is thought to be the result of colliding stellar winds in binary systems (Crowther 2003). WCO stars do not produce 26Al, but dust production by predecessor LBV and WN phases predicted to contain 26Al in their winds has recently been established (Rajagopal et al. 2007; Barniske et al. 2008). The η Carinae LBV star ejected ∼10 M of dust-rich material during its 1843 eruption (Smith et al. 2003), including aluminum oxide (de Koter et al. 2005), although its 26Al content has yet to be definitely established by γ-ray observations (Knödlseder et al. 1996).

To reach the molecular cloud, grains must survive sputtering after they pass the reverse shock and move at high speed (∼103 km s−1) with respect to the shocked gas in the H ii region. They also must not be completely decelerated within the reverse shock zone. These conditions place a lower limit on grain size. Grains too small are sputtered to destruction or eventually vaporized in the hot, shocked gas (Nozawa et al. 2007). The density of the wind 1 pc from a W-R star-losing mass at a rate of 10−5 M yr−1 is ∼10−2 cm−3 and the stopping distance of grains even as small as 0.01 μm grains is 10 pc, comparable or larger to the size of H ii regions. The deceleration across the scale of the shocked region (∼1 pc; Weaver et al. 1977) will be low and the fraction of material sputtered from the grains, which is related to the deceleration (Equation (1) in Nozawa et al. 2007) will be likewise small.7

Grains that escape the W-R wind will not penetrate far into a GMC. Typical hydrogen number densities in clouds are 102–103 cm−3 and the stopping distance of 0.01–0.1 μm grains will be only of order ∼103 AU. The gas densities in portions of the molecular cloud that are shocked and swept up by the expanding wind or SN ejecta will be higher and the stopping distances proportionally shorter. Thus, only the surfaces of clouds will be initially contaminated by 26Al. Further transport of dust grains into cloud depends on their kinematics, which are poorly understood. The smallest scale on which turbulence in clouds can affect the mass distribution and cause mixing is the sonic transition where the turbulent velocity is equal to the sound speed; this is roughly 1 pc in solar-metallicity clouds (Padoan 1995). Thus, mixing might be inefficient at the surfaces of clouds. The degree to which this controls the incorporation of 26Al into new low-mass stars depends on the extent of large-scale mixing and whether star formation is triggered or at least assisted, by the interaction of winds or SN ejecta with cloud gas (Zavagno et al. 2007). In that case, star formation is spatially correlated with 26Al abundance.

W-R progenitors may themselves migrate into and contaminate regions of a molecular cloud where low-mass star formation has yet to occur. 10%–30% of O stars have large peculiar velocities (up to 200 km s−1) relative to most early-type stars as a result either of dissolution of binary systems by SN explosions, or encounters between two binary systems (Hoogerwerf et al. 2001). One of these so-called runaway O stars moving at a typical speed of 30 km s−1 would traverse a molecular cloud in ∼1 Myr. Mass loss from this star, if occurring, could contaminate a larger region of the molecular cloud with SLRs.

5.2. CAIs with Low Initial 26Al/ 27Al Ratios

Any scenario that explains the canonical solar system 26Al/27Al ratio must also accommodate the exceptions. Several classes of CAIs show either no excess of 26Mg produced by the decay of 26Al or have an inferred 26Al/27Al ratio ≪1 × 10−5, much lower than the canonical value of (4.5–5)× 10−5. These include (1) igneous CAIs associated with chondrule-like materials (relict CAIs inside chondrules and CAIs surrounded by chondrule-like, ferromagnesian silicate rims; Krot et al. 2005a, 2005b); (2) some igneous CAIs in metal-rich (CB and CH) carbonaceous chondrites (Gounelle et al. 2007, Krot et al. 2008a); (3) fractionation and unidentified nuclear (FUN) effects CAIs (Lee 1988); (4) isolated platy hibonite crystals (PLACs; Ireland & Fegley 2000); (5) pyroxene–hibonite spherules (Ireland et al. 1991; Russell et al. 1998); (6) some corundum-bearing CAIs (Simon et al. 2002); and (7) most of the grossite- and hibonite-rich inclusions in CH chondrites (Kimura et al. 1993; Weber et al. 1995, Krot et al. 2008a). (1 and 2) CAIs associated with chondrule materials and some igneous CAIs in CB and CH carbonaceous chondrites (Krot et al. 2001, 2005a, 2005b, 2008a) are 16O-depleted to varying degrees (Δ17O ranges from −25 to −5) relative to typical CAIs in primitive chondrites which uniformly have 16O-rich compositions (Δ17O ∼ −25; Itoh et al. 2004; Makide et al. 2008). We infer that the 26Al-poor and 16O-depleted CAIs experienced late-stage melting and oxygen isotope exchange, probably during chondrule formation, which could have reset their 26Al–26Mg systematics. (3–7) The lower than the canonical 26Al/27Al ratio in FUN CAIs, PLACs, pyroxene–hibonite spherules, some corundum-bearing CAIs, and most of the grossite- and hibonite-rich inclusions in CH chondrites can be explained by (1) their late formation, after decay of 26Al; (2) their early formation, prior to introduction of 26Al; or (3) the lack of the canonical budget of 26Al in their precursors. Most of these CAIs have 16O-rich compositions Goswami et al. 2001; Simon et al. 2002; Krot et al. 2008a, 2008c), indistinguishable from those of typical CAIs with the canonical 26Al/27Al ratio. The rapid evolution of the oxygen isotopic composition of the inner solar system (Krot et al. 2005c; Aléon et al. 2007) and the short (less than 105 yr) duration of CAI formation (Thrane et al. 2006; Jacobsen et al. 2008) thus make (1) unlikely. Similar arguments can be used against (2). Although an early formation is possible if complete melting and exchange of oxygen isotopes occurred, this seems unlikely considering evidence for incomplete melting of CAIs (MacPherson et al. 2005).

We infer that the CAIs of categories 3–7 formed contemporaneously with 26Al-rich inclusions. The absence of canonical 26Al in their precursors suggests either they formed prior to homogenization of 26Al in the solar system Sahijpal & Goswami 1998; Krot et al. 2008a, 2008c), or preferential loss of the (uniformly distributed) 26Al carrier during thermal processing of the CAI precursors (Wood 1996). Both explanations can be reconciled with the presence of nucleosynthetic anomalies in some of these CAIs (Lee et al. 1998) if the 26Al carrier contributed a distinct component to the solar system's stable isotope composition (Lee et al. 1998). The second explanation is more speculative, however. It hypothesizes that (1) the precursors of these CAIs was isotopically heterogeneous and, contrary to typical refractory inclusions, escaped a cycle of complete evaporation-condensation; (2) the carrier of 26Al was relatively volatile; and (3) it was lost to varying degrees by sublimation of these CAI precursors prior to their melting. Although these inclusions can be used as an evidence for heterogeneous distribution of 26Al among CAI precursors, the scale of any heterogeneity was probably limited because such inclusions are rare relative to 26Al-rich CAIs.

5.3. Other SLRs

41Ca: A further test of the wind model is whether it can also reproduce the inferred initial 41Ca/40Ca ratio of 1.5 × 10−8 (Sahijpal & Goswami 1998). Published calculations of 41Ca yields in winds from massive stars are limited. We considered a 60 M progenitor for which 41Ca and 26Al yields from the W-R winds were separately published (Arnould et al. 2006; Limongi & Chieffi 2006). We accounted for the additional free decay of 26Al, which is ejected during the hydrogen-burning WN W-R phase, while 41Ca is produced in the later core He-burning WCO phase 1.9 × 105 yr later. The corrected ratio of 41Ca to 26Al in the ejecta is ∼25 times higher than in the solar system, but would be consistent if an additional time Δ ≈ 0.5 Myr elapsed before CAI formation. It is interesting that this is approximately the same duration as the W-R phase itself before the final SN Ib/c.

60Fe: W-R winds contain negligible amounts of 60Fe (Arnould et al. 2006). Live 60Fe in the early solar system could have originated in SN from the same early generation of massive stars that produced the 26Al, or in an even earlier generation of stars (Gounelle & Meibom 2008a). We repeated the calculations described in Section 4 but included SN contributions and calculated 60Fe/56Fe ratios in the same manner as 26Al/27Al, using the yields of Limongi & Chieffi (2006) and the solar iron abundance of Lodders (2003). We assume Δ = 0.5 Myr based on the 41Ca abundance. In Figure 4, we plot Monte Carlo realizations for different epochs (3–8 Myr before the Sun) for the earlier generation of massive stars. If all SN ejecta is incorporated, the 60Fe/56Fe is overpredicted by an order of magnitude relative to 26Al/27Al. A comparison with the ratio of the two SLRs (black line) inferred from γ-ray measurements (Wang et al. 2007) suggests that this discrepancy may be in part the result of an overprediction of 60Fe yield—or underprediction of 26Al yield—by the nucleosynthesis models (see Section 6). There are two other explanations suggested by Figure 4: (1) the Sun formed 3 Myr after the massive stars, when many massive stars were in the W-R phase but few SN had occurred (a scenario represented by the red dots extending below the primary locus); or (2) the SN contribution was attenuated by an effect such as described in Section 5.1.

Figure 4.

Figure 4. Calculated abundances of 26Al and 60Fe relative to stable comparison isotopes in star-forming regions contaminated by W-R winds and SN ejecta (see the text for details). Each point represents a Monte Carlo calculation of the composition of the gas in a well-mixed molecular cloud 3–8 Myr after massive star formation. The large black point is the inferred composition of the solar system. The line represents the Galactic average abundance ratio from γ-ray observations (Wang et al. 2007).

Standard image High-resolution image

Explanation (1) is statistically unlikely if star formation is uncorrelated and demands precise timing between the formation of massive stars and the Sun (∼3 Myr later), but could be demanded in a scenario where the Sun's formation was triggered by an SN (Cameron & Truran 1977; Boss et al. 2008). Our simulations indicate that the initial solar 26Al/27Al and 60Fe/56Fe ratios can be reproduced in this manner only in star clusters with N*>105 whose most massive members have ∼100 M. As such large clusters are relatively rare, this scenario is a priori less likely. Explanation (2) requires a reduction in the SN contribution to 60Fe (and 26Al) by a factor of 20. This could be due to a combination of effects; retention or fallback of the central region of the progenitor (Meyer 2005), inefficient delivery of SN ejecta into the cloud (Section 5.1) or the collapse of the protosolar cloud and a decrease in its cross section by the time the SN ejecta arrived. An alternative scenario (3) is that 60Fe in the solar system is the relict of an even earlier episodes of massive star-formation and contamination (Gounelle & Meibom 2008a) of which the 41Ca and most 26Al has decayed. 60Fe will decay to 5% of its initial abundance in 6.5 Myr, during which 26Al decays to 0.2%, and 41Ca essentially vanishes. This last explanation is viable only if this earliest generation of stars ceased to contribute SLRs to the molecular cloud after ∼6 Myr.

36Cl: We estimated the abundance of 36Cl, which is also ejected by W-R stars during the WCO phase (Arnould et al. 2006), and calculated the 36Cl:37Cl ratio in the same manner as 41Ca/40Ca. We find that our model underpredicts the ratio by at least 3 orders of magnitude. It has already been recognized that stellar nucleosynthetic models cannot account for this isotope, especially if it was introduced at the epoch of CAI formation 1–2 Myr before the host sodalite alteration phases were formed. At the present time, the only viable explanation appears to be a late episode of irradiation by energetic particles (Hsu et al. 2006)

6. SUMMARY AND OUTLOOK

The canonical abundance of 26Al in the solar system cannot be explained in terms of a late injection of debris from a nearby SN into the gaseous protoplanetary disk because (1) the dispersal of the natal cluster and the finite time window for injection make it a priori an unlikely event (less than 2%), (2) 26Al was already present in CAIs, which formed within ∼105 yr of the initial collapse of the protosolar nebula and the formation of the protoplanetary disk, and (3) the oxygen isotope systematics of primitive solar system materials show no sign of a late introduction of SN ejecta. The apparently uniform distribution of 26Al in meteorites and samples of the Earth, Moon, and Mars suggests homogenization during the collapse of the protosolar cloud.

We showed that the canonical 26Al/27Al ratio can be explained if the solar system formed from a molecular cloud contaminated by W-R winds from massive stars that formed 4–5 Myr earlier. An SN contribution is not required to explain the abundance of 26Al, although it is not necessarily excluded. The a priori probability that such a level of contamination occurred depends on the poorly understood star formation histories in GMCs; we estimate that it is ∼6%. However, our model predicts that the canonical value is close to the most likely nonzero value in the Galactic distribution.

The initial 41Ca/40Ca ratio can also be explained by W-R wind contamination if ∼0.5 Myr elapsed between its introduction by winds and the formation of CAIs. If this scenario is also to explain primordial 60Fe in the solar system, the cloud must have been contaminated with SN ejecta as well. If SN ejecta is included, our model overpredicts the abundance of 60Fe by an order of magnitude. This discrepancy could be rectified by some combination of the following: (1) most 60Fe falls back onto SN remnants rather than be ejected (Meyer 2005; Takigawa et al. 2008); (2) most dust grains in SN ejecta are retained and destroyed in the shocked ejecta and never enter the molecular cloud; (3) the protosolar cloud was already collapsing and presented a smaller cross section when the SN ejecta arrived; and (4) the 60Fe is a relict of an even earlier episode of massive star formation and contamination for which all the other SLRs have decayed away. The absence of a significant excess of 53Mn seems to require (1), but not to the exclusion of the other explanations. Our model does not explain the inferred abundance of 36Cl, and another mechanism such as irradiation much be invoked.

A key uncertainty in our model is the efficiency with which 26Al is introduced into the host molecular cloud and the degree to which it becomes uniformly mixed. We propose that the carrier of 26Al was dust grains and that these dynamically decoupled from the wind and embedded themselves (intact) into the cloud, but this hypothesis needs further investigation. Furthermore, our model does not account for the inhomogeneities in SN ejecta and W-R winds that could produce spatial variation in the contamination of a molecular cloud. There are also uncertainties in calculations of the evolution and nucleosynthesis of massive stars that could quantitatively alter our results. Production of 26Al by neon burning during the Type Ib/c SN that follows the W-R phase is sensitive to the progenitor mass (Higdon et al. 2004) and for a 60 M progenitor could be as large as the yield from the wind (Woosley & Heger 2007). New models that include stellar rotation predict higher yields of 26Al, an earlier appearance of 26Al in the W-R wind (as early as 1 Myr), and a smaller minimum initial mass for entry into the W-R phase (Palacios et al. 2005). Larger 26Al yields would relieve the requirement for high delivery efficiency to the molecular cloud and may resolve the discrepancy between the predicted amount of concomitant 60Fe from SN ejecta and the inferred initial abundance of the radionuclide in the solar system. Future tests of this model could compare predicted W-R stellar contamination with short-lived isotopes (e.g., 107Pd) whose abundances seem consistent with models of the interstellar medium (G. Huss, personal communication), as well as the solar system's oxygen isotopic composition.

This material is based upon work supported by the National Aeronautics and Space Administration through the NASA Astrobiology Institute under Cooperative Agreement No. NNA04CC08A issued through the Office of Space Science. S.R. is a NASA Postdoctoral Program Fellow. Some of this work was performed while E.G. was a Visiting Scholar at the University of California Berkeley. We thank Gary Huss and Kazuhide Nagashima for enlightening discussions and John Bally and Marcel Arnould for helpful comments and corrections.

Footnotes

  • Adopting an 27Al mass fraction of 5.8 × 10−5 (Lodders 2003), the protosolar cloud core initially contained 3 × 10−9 M of 26Al. The mass fraction of 26Al in the convective hydrogen shell of a 25 M star at the end of its main-sequence life is ∼1 × 10−6 (Meyer 2005). Therefore at least 3 × 10−3M of SN ejecta must have been introduced (assuming no free decay). Boss et al. (2008) report that only 5 × 10−5 M is injected in their model.

  • In contrast, the density of shocked SN ejecta is ∼105 cm−3, the stopping distance of a 0.1 μm grain is only ∼2 AU, whereas the ejecta scale length can be as large as 1 pc. Thus, grains in SN ejecta are more likely to be trapped in the ejecta and never introduced into star-forming molecular gas. This is another argument for W-R winds as the source of 26Al in the solar system.

Please wait… references are loading.
10.1088/0004-637X/696/2/1854