This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

LOW 60FE ABUNDANCE IN SEMARKONA AND SAHARA 99555

and

Published 2015 March 17 © 2015. The American Astronomical Society. All rights reserved.
, , Citation Haolan Tang and Nicolas Dauphas 2015 ApJ 802 22 DOI 10.1088/0004-637X/802/1/22

0004-637X/802/1/22

ABSTRACT

Iron-60 (t1/2 = 2.62 Myr) is a short-lived nuclide that can help constrain the astrophysical context of Solar System formation and date early Solar System events. A high abundance of 60Fe(60Fe/56Fe ≈ 4 × 10−7) was reported by in situ techniques in some chondrules from the LL3.00 Semarkona meteorite, which was taken as evidence that a supernova exploded in the vicinity of the birthplace of the Sun. However, our previous multi-collector inductively coupled plasma mass spectrometry (MC-ICPMS) measurements of a wide range of meteoritic materials, including chondrules, showed that 60Fe was present in the early Solar System at a much lower level (60Fe/56Fe ≈ 10−8). The reason for the discrepancy is unknown but only two Semarkona chondrules were measured by MC-ICPMS and these had Fe/Ni ratios below ∼2× chondritic. Here, we show that the initial 60Fe/56Fe ratio in Semarkona chondrules with Fe/Ni ratios up to ∼24× chondritic is (5.39 ± 3.27) × 10−9. We also establish the initial 60Fe/56Fe ratio at the time of crystallization of the Sahara 99555 angrite, a chronological anchor, to be (1.97 ± 0.77) × 10−9. These results demonstrate that the initial abundance of 60Fe at Solar System birth was low, corresponding to an initial 60Fe/56Fe ratio of (1.01 ± 0.27) × 10−8.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Chondrules are quenched spherical beads that were once molten in space and are found in large abundance in primitive meteorites known as chondrites (Scott & Krot 2014). Although the mechanism responsible for their melting is uncertain (shock waves, planetary collisions, or lightning), they have been precisely dated using several radioactive chronometers. Their crystallization ages span a few million years with a peak at ∼2 Myr after formation of calcium-aluminum-rich inclusions (CAIs; Hutcheon & Hutchison 1989; Kita et al. 2000; Rudraswami et al. 2008; Villeneuve et al. 2009; Kita & Ushikubo 2012), which is taken to represent time zero in early Solar System chronology (Dauphas & Chaussidon 2011). Chondrules formed early and contain some phases that have high Fe/Ni ratios, which makes them particularly well suited to establish the abundance of 60Fe in the early Solar System.

Using multi-collector inductively coupled plasma mass spectrometry (MC-ICPMS), Tang & Dauphas (2012) found a low and uniform initial 60Fe/56Fe ratio in early Solar System materials, including in chondrites and their constituents However, other in situ measurements by secondary ionization mass spectrometry (SIMS) have yielded much higher 60Fe/56Fe ratios (Telus et al. 2013 and references therein; Mishra & Goswami 2014; Mishra & Chaussidon 2014). Several explanations to this discrepancy are possible: (1) SIMS measurements suffer from an unidentified isobaric interference on 60Ni, (2) 60Fe was heterogeneously distributed and the 60Fe/56Fe ratio is highly variable from chondrule-to-chondrule, or (3) parent-body alteration and metamorphism have disturbed 60Fe-60Ni systematics in chondrules. Semarkona is one of the significant candidates to study 60Fe-decay because this LL3.00 ordinary chondrite has experienced little thermal metamorphism (Quirico et al. 2003; Grossman & Brearley 2005) with a peak temperature probably not exceeding ∼200°C (Huss et al. 2006), so disturbance to the 60Fe-60Ni system should be minimal. A limitation to studying this sample is its availability, as its total mass is only 691 g and less than of third of this is available in meteorite collections for scientific studies. In Tang & Dauphas (2012), we analyzed 2 bulk chondrules from Semarkona, 8 chondrules and magnetic/size separates from NWA 5717 (an ungrouped ordinary chondrite with a petrologic type of 3.05, i.e., slightly more metamorphosed than Semarkona). The spread in Fe/Ni ratio was limited (up to ∼5× chondritic) but this was sufficient to set an upper-limit on the initial 60Fe/56Fe ratio of <3 × 10−8, clearly lower than the values found by SIMS of ∼4 × 10−7 More recently, Telus et al. (2013) measured Fe and Ni distribution in chondrules by synchrotron X-ray fluorescence and found that late-stage fluids had mobilized these elements in most chondrites. However, only 5/16 Semarkona chondrules showed evidence of mobilization, meaning that most chondrules in that meteorite (∼69%) should be relatively pristine. The chondrules affected by parent-body disturbance should have relatively low Fe/Ni ratios.

In an effort to better constrain the abundance of 60Fe in the early Solar System, we have analyzed 6 new chondrules from Semarkona, some of which have high Fe/Ni ratios (up to ∼24× chondritic) We have also studied the initial 60Fe/56Fe ratio at the time of crystallization of the quenched angrite Sahara 99555, a sample that has been dated by several techniques (Connelly et al. 2008b; Spivak-Birndorf et al. 2009; Schiller et al. 2010; Larsen et al. 2011; Kleine et al. 2012), and can serve as a chronological anchor to back-calculate the initial 60Fe/56Fe ratio at the time of CAI formation. We confirm our earlier conclusion that the initial 60Fe/56Fe ratio was uniform across the inner protoplanetary disk at a level of 10−8.

In Section 2, the samples and their processing are presented. This includes retrieval of chondrules from the Semarkona meteorite; magnetic, density, and size separations of grains from the Sahara 99555 angrite; characterization of the samples, digestion, purification of Ni by chromatography, and isotopic analysis by mass spectrometry. In Section 3, the Ni isotopic results are presented, and the implications of those measurements for the abundance of 60Fe in the chondrule-forming region and quenched angrites are discussed in Section 4. Section 5 concludes that the abundance of 60Fe was low; implying that 26Al in meteorites came from the winds of one or several massive stars while 60Fe was inherited from galactic background.

2. SAMPLES AND METHODS

2.1. Semarkona Chondrules and Sahara 99555 Angrite

Chondrules are quenched droplets of magma that formed within a few million years of the formation of the Solar System (Kita & Ushikubo 2012; Scott & Krot 2014). After incorporation in chondrite parent-bodies, they were subjected to thermal metamorphism and aqueous alteration, processes that could disturb 60Fe-60Ni systematics, so care must be exercised in selecting the most pristine samples for study. Semarkona is a LL3.00 chondrite (Quirico et al. 2003; Grossman & Brearley 2005), meaning that it was minimally modified by parent-body aqueous alteration or metamorphism (the degree of aqueous alteration increases from type 3–1 while the degree of metamorphism increases from type 3–6) The Smithsonian Institution provided a fragment of Semarkona of approximately ∼370 mg, from which 14 chondrules were hand-picked under a binocular microscope In order to identify the chondrules with relatively high Fe/Ni ratios, small areas were polished using 1200 grit abrasive paper. The internal areas thus exposed allowed us to measure the chemical compositions of the chondrules using energy dispersive spectroscopy on a JEOL JSM-5800LV scanning electron microscope operated in low vacuum mode to prevent charging. The surfaces were not well polished and the samples were not carbon coated, so the chemical analyses were of limited quality but sufficient for our purposes. Out of the 14 starting chondrules, 6 had relatively high Fe contents and were selected for Ni isotopic analysis. The bulk chondrules (∼2–14 mg) were first rinsed with acetone to get rid of possible surface contamination. To avoid the risk of accidental sample loss, recover the maximum amount of Ni for high-precision isotopic analysis (a low 60Fe/56Fe ratio implies that 60Ni-excess should be barely resolvable in chondrules), and preserve the bulk nature of the measurements, no further characterization or fragmentation was done on these samples, which were directly digested.

Angrites are a group of basaltic achondrites that record some early igneous activity (Mittlefehldt et al. 2002; Mittlefehldt 2003) According to their textural characteristics, angrites are divided into two subgroups. Fine-grained angrites, such as D'Orbigny and Sahara 99555, are characterized by quenched textures indicative of rapid cooling (Mittlefehldt et al. 2002). Coarser-grained angrites, such as Angra dos Reis and NWA 4801, experienced more protracted cooling histories (e.g., Nyquist et al. 2009; Kleine et al. 2012). High-resolution chronometers including 26Al-26Mg, 53Mn-53Cr, 182Hf-182W, and Pb-Pb have been applied to date angrites (Lugmair & Galer 1992; Nyquist et al. 1994, 2003; Lugmair & Shukolyukov 1998; Baker et al. 2005; Amelin 2007, 2008; Markowski et al. 2007; Connelly et al. 2008a, 2008b; Spivak-Birndorf et al. 2009), providing a means of testing the concordance between different extant and extinct radiochronometers. Due to its quenched texture and rapid cooling, Sahara 99555 is a well-suited anchor for early Solar System chronology. It is mainly composed of Ca-rich olivine (∼31–42%), Al-Ti rich pyroxene (∼24–28%), and anorthitic plagioclase (∼33–39%; Mikouchi et al. 2000; Mikouchi & McKay 2001). A 500 mg sample of Sahara 99555 purchased from L. Labenne was crushed into powder in an agate mortar. The fragmented samples were separated into two parts: one was processed with a hand magnet followed by sieving to separate the grains into three silicate size fractions (100–166 μm, 166–200 μm, and >200 μm); the other was processed with a hand magnet and split according to density (below or above 3.10 g cm−3) using sodium polytungstate solution. The mineral fractions were not characterized but Tang & Dauphas (2012) applied the same procedure to D'Orbigny angrite and the low-density fraction was relatively rich in anorthite while the high-density fraction was rich in olivine and pyroxene. For density separation, the fragmented sample was placed in the separatory funnel and heavy liquid was then added. The funnel was left for 10 minutes until dense grains (>3.10 g cm−3) sank to the bottom while light grains(<3.10 g cm−3) remained suspended in the liquid. Silicate fractions with different densities were collected onto pieces of weighing paper, rinsed with Millipore Milli-Q water, and dried in an oven.

To assess data quality and make sure that no analytical artifact was present, terrestrial standards were processed and analyzed together with the meteorite samples.

2.2. Sample Preparation, Digestion, and Chemical Separation

The chemistry was performed under clean laboratory conditions at the Origins Lab of the University of Chicago. Optima grade HF, reagent grade acetone, and double distilled HCl and HNO3 were used for digestion and column chromatography. Millipore Milli-Q water was used for acid dilution.

Tang & Dauphas (2012, 2014) provide details on the procedures of sample digestion and chemical separation. Semarkona chondrules and Sahara 99555 fractions weighing 2–130 mg were digested in 5–20 ml HF-HNO3 (in a 2:1 volume ratio) in a Teflon beaker placed on a hot plate at ∼90°C for 2 days. The solution was subsequently evaporated to dryness and re-dissolved in a 5–20 mL mixture of concentrated HCl-HNO3 (in a 2:1 volume ratio) until all the sample powder was completely digested. The solutions were dried down and the residues taken back in solution with a minimum amount of concentrated HCl (∼11 M) for loading on the first column. In order to obtain sufficiently clean Ni for isotopic measurements, chemical separation of Ni from matrix elements and isobars was done in three steps of chromatography that are described below.

U/TEVA cartridge (Horwitz et al. 1992) was used for the first step of chemistry to get rid of Ti, Co, Zr and Fe. The column (2 mL volume, 2.5 cm length, 1 cm diameter) was pre-cleaned with 10 mL water, 15 mL 0.4 M HCl, 15 mL 4 M HCl and was then conditioned with 10 mL of concentrated HCl. The sample solution was loaded onto the column in 5–10 mL 10 M HCl. The load solution was collected in clean Teflon beakers and an additional 10 mL of concentrated HCl was passed through the resin and collected in the same beaker. This eluate contained Ni together with Na, Mg, Ca, and other matrix elements. After drying down, the Ni elution cut from the first column chemistry was re-dissolved in 5 mL of a mixture of 20% 10 M HCl-80% acetone (by volume) and loaded onto a Teflon column containing 5 mL (40 cm length, 0.4 cm diameter) of pre-cleaned Bio-Rad AG50-X12 200–400 mesh hydrogen-form resin, previously conditioned with 10 mL 20% 10 M HCl-80% acetone. After loading the sample solution and rinsing with 30 mL 20% 10 M HCl-80% acetone mixture to eliminate Cr and any remaining Fe, Ni was collected by eluting 150 mL of the HCl-acetone mixture into a jar containing 30 mL of water to dilute HCl and stabilize Ni in the eluate. In those conditions, Mg, Na, Ca, and other matrix elements were retained on the resin (Strelow et al. 1971) The collected Ni solution was evaporated at moderate temperature (<90°C) under a flow of N2 to avoid the formation of organic complexes with acetone and accelerate evaporation. After evaporation, the Ni fraction was dissolved in 1 mL of aqua regia (1:3 HNO3:HCl) to remove any organic residue formed during evaporation. This HCl-acetone column was repeated five times to ensure thorough separation of major rock forming element Mg from Ni, two elements that are notoriously difficult to separate. Zinc is a significant interference on low abundance isotope 64Ni. It was removed using a third column filled with 1 mL (2 cm length, 0.8 cm diameter) Bio-Rad AG1W-X8 anionic ion exchange resin in 8 M HBr medium (Moynier et al. 2006). Nickel was eluted in 8 mL 8 M HBr, whereas Zn was retained on the resin.

The entire procedural blank was ∼35 ng, which is negligible compared to the amounts of Ni in the samples. The nickel yield of the entire procedure was 90–100%.

2.3. Mass Spectrometry

All measurements were performed at the Origins Laboratory of the University of Chicago using a Neptune MC-ICPMS equipped with an OnTool Booster 150 (Pfeiffer) interface jet pump. Jet sampler and X skimmer cones were used. The samples were re-dissolved in 0.3 M HNO3 and introduced into the mass spectrometer with Ar + N2 using an Aridus II desolvating nebulizer at an uptake rate of ∼100 μL minutes−1. The instrument sensitivity for 58Ni was 100 V/ppm. One analysis consisted of 25 cycles, each acquisition lasting for 8.4 s. During a session, each sample solution was measured 13 times bracketed by SRM 986. A small isobaric interference from the least abundant isotope of iron, 58Fe, on the most abundant isotope of nickel, 58Ni, was corrected by monitoring 57Fe (the correction is always smaller than 0.5 $\varepsilon $ ). All isotopes were measured using Faraday cups with 1011 Ω resistance amplifiers. Background was subtracted using an on-peak zero procedure. Internal normalization was used to correct mass-dependent isotopic fractionation by fixing 61Ni/58Ni to 0.016720 or 62Ni/58Ni to 0.053389 (Gramlich et al. 1989) using the exponential law (Maréchal et al. 1999). Nickel-64 is reported only for samples that gave ∼15 μg Ni because below this level, an isobaric interference from $^{48}{\rm T}{{{\rm i}}^{16}}{{{\rm O}}^{+}}$ can affect the results (Tang & Dauphas 2012).

Approximately 20% of the original sample solutions were kept as safety aliquots and for Fe/Ni ratio measurements by MC-ICPMS using both standard bracketing and standard addition techniques. The Fe/Ni ratios measured by standard addition agree well with Fe/Ni ratios measured by simple sample-standard bracketing. Fe/Ni ratios in terrestrial standards were all within 3% of their reference values, demonstrating the accuracy of our measurements.

3. RESULTS

Table 1 shows the Ni isotopic compositions and Fe/Ni ratios measured in Semarkona chondrules, Sahara 99555 fractions and terrestrial standards. Terrestrial standards passed through the same column chemistry as meteoritic samples have normal Ni isotopic ratios, attesting to the accuracy of the measurements. The intercepts and slopes of the $\varepsilon $60Ni versus 56Fe/58Ni correlations were calculated using Isoplot (Ludwig 2012) to estimate the initial 60Fe/56Fe ratios and $\varepsilon $60Ni values based on the following isochron equation (see Figure 2 of Dauphas & Chaussidon 2011 for an explanation),

Equation (1)

where 2.596 is a constant that corresponds to the 58Ni/60Ni ratio in Solar System material. No significant variations were detected for 61Ni, 62Ni, and 64Ni isotopes relative to terrestrial standards.

Table 1.  Nickel Isotopic Compositions and Fe/Ni Ratios of Semarkona Chondrules and Whole Rock Sahara 99555

Sample Name Type Sample Mass (mg) Fe/Ni 56Fe/58Ni (at.) Norm. 61Ni/58Ni Norm. 62Ni/58Ni n Replicates
          $\varepsilon $60Ni $\varepsilon $62Ni $\varepsilon $64Ni $\varepsilon $60Ni $\varepsilon $61Ni $\varepsilon $64Ni  
Terrestrial Standards                    
SRM986         −0.036 ± 0.042 0.044 ± 0.170 0.443 ± 0.488 −0.058 ± 0.085 −0.033 ± 0.129 0.370 ± 0.453 12
BHVO-02 195 727 1010 ± 30 −0.033 ± 0.118 −0.040 ± 0.086 −0.013 ± 0.098 0.030 ± 0.065 10
BHVO-02 (2) 154   −0.001 ± 0.074 0.072 ± 0.147 −0.037 ± 0.054 −0.054 ± 0.111 12
DNC-1 105 258 359 ± 23 0.013 ± 0.118 0.141 ± 0.286 −0.059 ± 0.091 −0.107 ± 0.202 7
Semarkona Chondrules                    
SC-10–2a Type I 0.3 38 50 ± 2.9 −0.03 ± 0.12 −0.11 ± 0.17 0.03 ± 0.12 0.09 ± 0.13 10
SC-30–6a Type I 0.2 13 17 ± 1.0 −0.12 ± 0.25 −0.38 ± 0.42 0.08 ± 0.14 0.28 ± 0.32 10
SC-13–1 Type II 7.4 69.7 96.9 ± 7.9 −0.004 ± 0.045 0.045 ± 0.133 −0.027 ± 0.065 −0.034 ± 0.101 12
SC-13–2 Type I 3.6 13.2 18.3 ± 2.0 −0.045 ± 0.046 −0.011 ± 0.136 0.256 ± 0.643 −0.040 ± 0.075 0.008 ± 0.102 0.272 ± 0.525 12
SC-13–3 Type I 11.1 23.1 32.1 ± 2.8 −0.026 ± 0.057 0.050 ± 0.096 0.223 ± 0.295 −0.051 ± 0.057 −0.038 ± 0.073 0.149 ± 0.326 12
SC-13–4 Type I 13.5 20.7 28.8 ± 2.2 −0.026 ± 0.045 −0.047 ± 0.141 0.026 ± 0.406 −0.003 ± 0.057 0.035 ± 0.106 0.095 ± 0.269 12
SC-13–5 Type I 7.5 31.2 43.4 ± 5.2 −0.016 ± 0.064 0.009 ± 0.120 −0.045 ± 0.046 −0.011 ± 0.136 12
SC-13–6 Type II 1.7 435.0 605 ± 56.6 0.051 ± 0.043 0.009 ± 0.056 0.047 ± 0.050 −0.006 ± 0.042 12
Sahara 99555                      
WRa 368 3376 4680 ± 311 0.31 ± 0.16 0.07 ± 0.22 0.28 ± 0.16 −0.05 ± 0.17 6
<100 μm 121 3078 4267 ± 322 0.306 ± 0.090 −0.075 ± 0.184 0.344 ± 0.064 0.057 ± 0.140 7
100–166 μm 98 5613 7780 ± 1354 0.443 ± 0.134 0.085 ± 0.302 0.400 ± 0.065 −0.064 ± 0.228 7
166–200 μm 101 4512 6255 ± 934 0.310 ± 0.076 −0.156 ± 0.177 0.389 ± 0.072 0.118 ± 0.134 10
>200 μm 131 5155 7145 ± 762 0.408 ± 0.048 0.112 ± 0.129 0.351 ± 0.032 −0.084 ± 0.097 10
<3.10 g cm−3 53 2538 3517 ± 346 0.214 ± 0.088 −0.045 ± 0.173 0.237 ± 0.146 0.034 ± 0.131 7
>3.10 g cm−3 150 6872 9524 ± 903 0.540 ± 0.106 0.087 ± 0.175 0.496 ± 0.076 −0.066 ± 0.133 7

Note. $\varepsilon $iNi = ([iNi/58Ni]sample/[iNi/58Ni]SRM986−1) × 104. The uncertainties are 95% confidence intervals.

aNickel isotopic composition and Fe/Ni ratio in Semarkona chondrules and whole rock Sahara 99555 are from Tang & Dauphas (2012).

Download table as:  ASCIITypeset image

A total of eight Semarkona chondrules (6 out of 14 chondrules surveyed from this study, Table 1; 2 chondrules from Tang & Dauphas 2012) have been measured and the results are shown in Figure 1(a). Fe/Ni ratios range from 13 to 435 (for reference, the CI chondrite value is 17). No significant 60Ni excess was detected in the chondrules with relatively low Fe/Ni ratios (Fe/Ni < 100). One Type II chondrule (FeO ∼ 16.7 wt%), SC-13-6, has a high Fe/Ni ratio (∼435) as well as barely resolvable $\varepsilon $60Ni excess of +0.051± 0.043. Combining data from all Semarkona chondrules, a single isochron can be defined corresponding to initial 60Fe/56Fe = (5.39 ± 3.27) × 10−9 and $\varepsilon $60Ni = −0.032 ± 0.023 (MSWD = 0.26), at the time of equilibration (Figure 1). This initial 60Fe/56Fe ratio is much lower than the value inferred by SIMS of ∼4 × 10−7 at the time of chondrule formation (Mishra & Goswami 2014; Mishra & Chaussidon 2014). The value of the slope is heavily leveraged by SC-13-6. Even if this data point is excluded from the regression, the initial 60Fe/56Fe ratio calculated based on MC-ICPMS data remains low 60Fe/56Fe = (1.77 ± 2.77) × 10–8, and is inconsistent with SIMS results.

Figure 1.

Figure 1. 60Fe-60Ni isochron diagrams of Semarkona chondrules and silicate minerals in Sahara 99555. Ni isotopic ratios are reported using the $\varepsilon $-notation; $\varepsilon $60Ni = [(60Ni/58Ni)sample/(60Ni/58Ni)standard−1] × 104, where 60Ni/58Ni ratios have been corrected for natural and laboratory-introduced mass fractionation by internal normalization to a constant 61Ni/58Ni ratio. The error bars represent 95% confidence intervals. In $\varepsilon $60Ni vs. 56Fe/58Ni isochron diagrams, the intercept gives the initial Ni isotopic composition $\varepsilon $60Ni while the slope is proportional to the initial 60Fe/56Fe ratio; slope = 25,961 × (60Fe/56Fe). Live 60Fe was detected in (A) Semarkona bulk chondrules [60Fe/56Fei = (529 ± 3.27) × 10−9], and (B) mineral separates of Sahara 99555 [WR, whole rock, the other labels represent fractions with different grain sizes or densities; 60Fe/56Fei = (1.96 ± 0.77) × 109]. The green dashed line in panel A shows the expected isochron if the initial 60Fe/56Fe ratio was 4 × 10−7 as suggested by SIMS (Mishra & Goswami 2014; Mishra & Chaussidon 2014).

Standard image High-resolution image

The Fe-Ni results for mineral separates in the quenched angrite Sahara 99555are given in Table 1. The mineral separates display high Fe/Ni ratios ranging from ∼2,500 (for the low-density fraction) to ∼6,900 (for the high-density fraction) and radiogenic $\varepsilon $60Ni values between +0.21 and +0.54. The initial 60Fe/56Fe ratio and $\varepsilon $60Ni value inferred from mineral separates from the Sahara 99555 angrite are shown in Figure 1(b). The data points define an internal isochron (MSWD = 0.49) of slope 60Fe/56Fe = (1.96 ± 0.77) × 10−9 and intercept $\varepsilon $60Ni = +0.43 ± 0.115 at the time of closure to isotope exchange of the minerals investigated. This value agrees well with independent results reported for this meteorite, which gave an initial 60Fe/56Fe ratio of (1.8 ± 0.5) × 10−9 (Quitté et al. 2010).

4. DISCUSSION

4.1. Abundance of 60Fe in the Chondrule-forming Region

The nickel isotopic compositions of Semarkona chondrules (LL3.00) were analyzed and give an initial 60Fe/56Fe ratio of (539 ± 3.27) × 10−9 at the time of chondrule formation. Telus et al. (2013, 2014) measured Fe and Ni distribution in chondrules by synchrotron X-ray fluorescence and found that late-stage fluids had mobilized these elements from the matrix to deposit them as iron and nickel oxide/hydroxide in chondrule fractures. Such mobilization has little bearing on the inferred low 60Fe abundance in Semarkona chondrules measured in this study for the following three reasons.

  • (i)  
    In chondrites of metamorphic grade as low as 3.10, Telus et al. (2014) found that 100% of the chondrules were affected by Fe-Ni mobilization. However, in Semarkona (type LL3.00, one of the most pristine meteorite samples available), they report that approximately 70% of the chondrules analyzed did not display any evidence for mobilization (the probability to have a random Semarkona chondrule free of Fe-Ni mobilization is $p=0.7)$. The probability that $k$ out of $n=8$ random Semarkona chondrules be free of Fe-Ni mobilization is,
    Equation (2)
    The calculated probabilities are $P\left( 0 \right)\sim 0\%$, $P\left( 1 \right)\sim 0\%$, $P\left( 2 \right)\sim 1\%$, $P\left( 3 \right)\sim 5\%$, $P\left( 4 \right)\sim 13\%$, $P\left( 5 \right)\sim 25\%$, $P\left( 6 \right)\sim 30\%$, $P\left( 7 \right)\sim 20\%$, and $P\left( 8 \right)\sim 6\%$. At 95% confidence level, the majority (k > 4) of the chondrules analyzed here were not affected by Fe-Ni mobilization. Given that the chondrules were pre-screened to select those with high Fe/Ni ratios, the sample set is biased (p is probably higher than 0.7) and k > 4 is a conservative estimate.
  • (ii)  
    In bulk measurements, the effect of Fe-Ni mobilization would be to lower the Fe/Ni ratios and $\varepsilon $60Ni values toward chondritic values Such physical admixture of Fe-Ni in chondrules through fractures could have added some scatter to the data points and could have brought the samples toward the chondritic value but overall, the points should have moved along the isochron line. Indeed, mixing between two components in the $\varepsilon $60Ni versus Fe/Ni diagram is a straight line, so the isochronous behavior of bulk chondrule measurements should have been preserved at some level.
  • (iii)  
    The Semarkona chondrules most likely to have been affected by Fe-Ni addition should be among those that display low Fe/Ni ratios in bulk. Chondrules with high Fe/Ni ratios such as SC-13-6 (Fe/Ni ratio ∼24× chondritic) provide the most leverage to define the slope of the isochron and are least likely to have been affected by Fe/Ni mobilization. For reference, an initial 60Fe/56Fe ratio of 4 × 10−7 in Semarkona chondrules (Mishra & Goswami 2014; Mishra & Chaussidon 2014) should have been associated with excess epsilon60Ni of +6.3 in chondrule SC-13-6 (Equation (1)) while a value of +0.051± 0.043 was measured.

The results presented here thus reaffirm our earlier conclusion (Tang & Dauphas 2012) that 60Fe was present at a low level in the chondrule-forming region, 60Fe/56Fe = (5.39 ± 3.27) × 10−9.

4.2. Abundance of 60Fe in the Sahara 99555 Angrite

The mineral separates in the Sahara 99555 angrite give an initial 60Fe/56Fe ratio of (1.96 ± 0.77) × 10−9 (Figure 1(b)), which is in excellent agreement with the initial value of (1.8 ± 0.5) × 10−9 reported by Quitté et al. (2010) in the same meteorite. The weighted average of those two values is (1.85 ± 0.42) × 10−9.

The Ni isotopic compositions of mineral separates from the D'Orbigny meteorite, a quenched angrite like Sahara 99555, were measured independently by Quitté et al. (2010), Spivak-Birndorf et al. (2011) and Tang & Dauphas (2012). The initial 60Fe/56Fe ratios reported by these three studies are (4.1 ± 2.6) × 10−9, (2.81 ± 0.86) × 10−9, and (3.42 ± 0.58) × 10−9, respectively. The three values agree and the weighted average is (3.26 ± 0.47) × 10−9.

The calculated initial 60Fe/56Fe ratio of D'Orbigny is significantly higher than the value measured in Sahara 99555. The age difference between these two angrites is given by Δt2−1 = In(r1/r2)/λ with an associated error of $\sigma ({\rm \Delta }{{t}_{2-1}})=\sqrt{\sigma _{{{r}_{1}}}^{2}/r_{1}^{2}+\sigma _{{{r}_{2}}}^{2}/r_{2}^{2}}/\lambda ,$ where r denote the initial 60Fe/56Fe ratio in either Sahara 99555 or D'Orbigny, and λ is the half-life of 60Fe (2.62 Myr). The calculated age difference between D'Orbigny and Sahara 99555 is +2.1 ± 1. Myr.

The relative chronology of formation of D'Orbigny and Sahara 99555 can be compared with independent estimates obtained using various dating techniques. Internal 26Al-26Mg isochrons in Sahara 99555 and D'Orbigny give initial 26Al/27Al ratios of (4.50 ± 0.54) × 10−7 and (3.97 ± 0.26) × 10−7, respectively (Spivak-Birndorf et al. 2009; Schiller et al. 2010). These two values are almost indistinguishable, meaning that the two objects crystallized within ∼0.2Myr of each other Similarly, 182Hf-182W internal isochrons in Sahara 99555 and D'Orbigny give initial ratios of (6.83 ± 0.14) × 10−5 and (7.15 ± 0.17) × 10−5 (Kleine et al. 2012), corresponding to a time difference between D'Orbigny and Sahara 99555 of +0.6 ± 0.4 Myr. Pb-Pb ages have also been reported for D'Orbigny (4563.37 ± 0.25 Myr; Brennecka and Wadhwa, 2012) and Sahara 99555 (4563.64 ± 0.14 Myr, Connelly et al. 2008b; Larsen et al. 2011). These absolute Pb-Pb ages should be regarded with caution because inter-laboratory calibration for this dating system is lacking, yet the ages are indistinguishable. All available evidence thus suggests that D'Orbigny and Sahara 99555 crystallized at approximately the same time.

The difference between 60Fe-56Fe and other decay systems cannot be explained by a difference in closure age because the cooling rates of quenched angrites estimated from diffusion profiles and petrographic textures are rapid, 7–50 °C/hr (Mikouchi & McKay 2001). This means that it would take 1–6 days for the samples to cool by 1000 °C, which covers more than the span of closure temperatures for the systems discussed above. This is instantaneous in regard to planetary timescales, so one would expect all extant and extinct chronometers to be closed to isotopic exchange at approximately the same time.

Most likely, the low 60Fe/56Fe ratio measured in Sahara 99555 versus D'Orbigny reflects terrestrial contamination in the former. Sahara 99555 was found in the Sahara desert. As pointed out by Crozaz et al. (2003), Floss et al. (2003), and Amelin (2008), Sahara 99555 shows more evidence of terrestrial alteration than other angrites, including mobilization of rare earth elements. The sample that we studied was partly covered with some desert-weathering product that we physically removed but Crozaz et al. (2003) and Floss et al. (2003) showed that chemical alteration also affected Sahara 99555 samples with fresh appearances. The 60Fe/56Fe ratio of (3.26 ± 0.47) × 10−9 from D'Orbigny internal isochrons may thus provide the best estimate of the 60Fe/56Fe ratio at the time of crystallization of the quenched angrites (Quitté et al. 2010; Tang & Dauphas 2012).

4.3. Abundance of 60Fe at Solar System Birth and Astrophysical Implications

Tang & Dauphas (2012) were able to constrain the 60Fe/56Fe ratio at Solar System birth to (115 ± 0.26) × 10−8 using internal isochrons in Gujba CB chondrite, D'Orbigny angrite, and unequilibrated ordinary chondrites as well as bulk rock isochrons for angrites and HED meteorites. The new results presented here for Semarkona chondrules and Sahara 99555 allow us to refine the initial 60Fe/56Fe ratio. Table 2 gives the ages relative to CAIs and initial 60Fe/56Fe ratios for all MC-ICPMS measurements for which resolvable 60Ni-excess could be resolved, that is to say Semarkona chondrules, bulk HEDs, bulk angrites, and mineral separates of D'Orbigny and Sahara 99555. Those data relate to the 60Fe/56Fe ratio at the time of CAI formation using the free decay equation,

Equation (3)

where t and tCAI are the formation ages of the samples considered and CAIs, respectively. The 60Fe/56Fe ratios inferred from 60Ni-56Fe/58Ni isochrons are plotted in Figure 2 and all the results obtained thus far point to an initial 60Fe/56Fe ratio of (1.01 ± 0.27) × 10−8 and a homogeneous distribution of 60Fe (see Table 2).

Figure 2.

Figure 2. Initial 60Fe/56Fe ratios as a function of time after CAI formation for various meteoritic objects. The values measured by MC-ICPMS are in blue (Tang & Dauphas 2012; this study; also see Quitté et al. 2010 for angrites). Those measured by SIMS in Semarkona chondrules are in yellow (Telus et al. 2013; Mishra & Goswami 2014; Mishra & Chaussidon 2014). Note the two-orders of magnitude discrepancy between in situ SIMS data (60Fe/56Fe = 70 × 10−8 at CAI formation) and those measured by MC-ICPMS (60Fe/56Fe = 1.01 × 10−8 at CAI formation). The isochrons used to infer initial 60Fe/56Fe ratio by SIMS are not very well defined (Figure 3).

Standard image High-resolution image

Table 2.  Relative Ages and 60Fe/56Fe Ratios in Different Meteoritic Samples and Back-calculated 60Fe/56Fe Initial Ratio in the Solar Protoplanetary Disk

Sample Age (Myr) Method Reference 60Fe/56Fei Reference 60Fe/56Fe CAI
Bulk HED meteorites 3.7 ± 0.5 53Mn-53Cr anchored to D'Orbigny Trinquier et al. (2008) (3.45 ± 0.32) × 10−9 Tang & Dauphas (2012) (9.26 ± 1.58) × 10−9
Bulk angrites 4.9 ± 0.2 53Mn-53Cr anchored to D'Orbigny Shukolyukov & Lugmair (2007) (2.20 ± 1.16) × 10−9 Tang & Dauphas (2012) (8.00 ± 4.25) × 10−9
D'Orbigny minerals 5.1 ± 0.1 26Al-26Mg Spivak-Birndorf et al. (2009); Schiller et al. (2010) (3.26 ± 0.47) × 10−9 Quitté et al. (2010), Spivak-Birndorf et al. (2011), Tang & Dauphas (2012) (1.27 ± 0.19) × 10−8
Sahara 99555 minerals 5.0 ± 0.2 26Al-26Mg Spivak-Birndorf et al. (2009); Schiller et al. (2010) (1.85 ± 0.42) × 10−9 This study, Quitté et al. (2010) (6.96 ± 1.60) × 10−9
Semarkona chondrules 2.0 ± 0.8 26Al-26Mg Kita & Ushikubo (2012) and references therein (5.29 ± 3.27) × 10−9 This study (8.93 ± 5.85) × 10−9
Gujba chondrules 5.2 ± 0.9 53Mn-53Cr anchored to D'Orbigny Yamashita et al. (2010) <3.49 × 10−9 Tang & Dauphas (2012)
NWA5717 chondrules and fragments 2 ± 1 Chondrule ages in UOC Kita et al., (2005) <2.14 × 10−8 Tang & Dauphas (2012)
60Fe/56FeCAI weighted averagea   (1.01 ± 0.27) × 10−8

aCalculated using initial 60Fe/56Fe ratios at CAI formation excluding bulk angrites, which show disturbed 60Fe-60Ni systematics.

Download table as:  ASCIITypeset image

The results from the present study reaffirm and strengthen our earlier conclusion that the abundance of 60Fe in the early Solar System was low. This contrasts with recent in situ studies of chondrules from unequilibrated ordinary chondrites that report high initial 60Fe/56Fe ratios. Telus et al. (2014) made the case that most chondrites other than Semarkona have been affected by Fe-Ni mobilization. For this reason, we focus our comparison on the results obtained by SIMS on chondrules from the Semarkona meteorite. The initial 60Fe/56Fe ratios reported by Mishra & Goswami (2014) and Mishra & Chaussidon (2014) correspond to an initial 60Fe/56Fe ratio at Solar System birth of ∼70 × 10−8. This is almost two orders of magnitude higher than the initial 60Fe/56Fe ratio obtained by MC-ICPMS (∼10−8, Tang & Dauphas 2012; this study). This cannot be due to 60Fe heterogeneity because Tang & Dauphas (2012) did not detect isotopic anomalies in 58Fe, which is produced in stars together with 60Fe. Furthermore, the same sample types, Semarkona chondrules, were measured by both SIMS and MC-ICPMS. Telus et al. (2013) also reported SIMS measurements of 3 Semarkona chondrules and found barely detectable 60Fe in only one chondrule 60Fe/56Fe ∼(1.4± 1.2) × 10−7, while the other two chondrules only provided upper-limits. One such chondrule has an upper-limit of 60Fe/56Fe <5.1 × 10−8, which is significantly lower than the values reported by Mishra & Goswami (2014) and Mishra & Chaussidon (2014; Figure 3).

We have no explanation for the discrepancy between the MC-ICPMS and SIMS studies but we note that all SIMS estimates are based on measurements of a single sample type (i.e., pyroxene in chondrules from unequilibrated ordinary chondrites) and that most of the SIMS isochrons are defined by points that have error bars that largely overlap with zero, the significance of the isochron arising from a few data points with small errors (Figure 3) In contrast, the estimate from MC-ICPMS is derived from isochrons measured on a variety of samples formed at different times (bulk HEDs—Vesta, SNCs—Mars, bulk angrites, D'Orbigny and Sahara 99555 angrites, Gujba, NWA 5717, and Semarkona chondrules) and the significances of the regressions are in most cases very well established (Figure 1(b); Tang & Dauphas 2012). The isochron for Semarkona chondrules presented here (Figure 1(a)) is barely resolvable from zero but the corresponding initial 60Fe/56Fe ratio is much lower than the initial ratio expected based on SIMS data. Care was put in the SIMS studies to avoid analytical artifacts and it is not clear how much improvement can be made on that front but other instruments exist or are being developed that may provide a direct comparison with SIMS measurements, such as resonant ionization mass spectrometry, atom probe, or megaSIMS. At present, the weight of evidence supports a low and uniform abundance of 60Fe in the early Solar System.

Figure 3.

Figure 3. Compilation of epsilon60Ni-56Fe/58Ni values measured by SIMS (Mishra & Goswami 2014; Mishra & Chaussidon 2014) for Semarkona chondrules in which evidence for live 60Fe was reported as significant. The reported initial 60Fe/56Fe values for each chondrule are given in the legend (each set of colored symbols corresponds to different data points measured on the same chondrule). Panel A shows the data on a linear x-axis scale while panel B shows the data on logarithmic scale. The two black lines are internal isochrons for initial 60Fe/56Fe ratios of 4× 10−7 (the initial value at the time of Semarkona chondrule formation reported by Mishra & Goswami 2014; Mishra & Chaussidon 2014) and 5× 10−9 (inferred value by MC-ICPMS, Tang & Dauphas 2012; this study). As shown, most of the data points measured by SIMS that are taken as evidence for a high initial 60Fe/56Fe initial ratio have uncertainties that overlap with terrestrial isotopic composition.

Standard image High-resolution image

The low initial 60Fe/56Fe ratio is consistent with background abundances in the Galaxy with no compelling need to invoke late injection from a nearby star (Tang & Dauphas 2012). Indeed, the average 60Fe/56Fe ratio in the ISM at Solar System birth inferred from γ-ray astronomy (Wang et al. 2007) is (2.8± 1.4) × 10−7, which is 30 times higher than the initial ratio in the early Solar System (Tang & Dauphas 2012). For comparison, the average 26Al/27Al ratio in the ISM at Solar System birth inferred from γ-ray astronomy (Diehl et al. 2006, 2010) is (3.0± 0.8) × 10−6, which is 17 times lower than the early Solar System ratio (Lee et al. 1976). It thus appears that 26Al and 60Fe in meteorites have different origins (Tang & Dauphas 2012).

Several scenarios can be considered to incorporate freshly made 26Al without adding too much 60Fe. Adjusting the timescale between nucleosynthesis and Solar System formation does not help because 26Al has a shorter half-life than 60Fe, so any delay would cause the 26Al/60Fe ratio to decrease, making the problem worse. The possible scenarios to explain the high 26Al/60Fe ratio of the early Solar System include (1) supernova (SN) explosion with fallback of the inner layers, so that only 26Al is efficiently ejected while 60Fe is trapped in the stellar remnant (Meyer & Clayton 2000) but this is unlikely because one would need to have a lot of fallback (a cutoff in the C/O-burning layer) to prevent 60Fe from escaping (Takigawa et al. 2008), (2) interaction of a SN with an already formed cloud core, so that only the outer layers are efficiently injected while the inner layers are deflected (Gritschneder et al. 2012), and (3) ejection of 26Al as winds from one or several massive stars (Arnould et al. 1997, 2006; Gaidos et al. 2009; Tatischeff et al. 2010; Gounelle et al. 2012; Young 2014). The last scenario is appealing because it could be a natural outcome of the presence in the Solar System forming region of one or several Wolf–Rayet (W–R) stars, as such stars shed their mass through winds rich in 26Al whereas 60Fe is ejected at a later time following the SN explosion (Tang & Dauphas 2012). W–R winds would have carved 26Al-rich bubbles in molecular cloud material, which could have subsequently been incorporated in the molecular cloud core that formed the Solar System. Semi-analytic approaches have been used recently to assess the feasibility of such a scenario (Gounelle & Meynet 2012; Young 2014) but it remains to be seen whether high 26Al/low 60Fe regions can exist because SN explosion is the main factor that drives the mixing between W–R product and surrounding cloud material, so that the 26Al-rich W–R material may be contaminated with 60Fe-rich SN ejecta. Existing semi-analytic models aimed at explaining the high 26Al/60Fe ratio in meteorites have not addressed this critical aspect of the process, which will require high-resolution modeling of the interactions of stellar winds and ejecta with the surrounding medium.

5. CONCLUSION

We report Ni isotope measurements by MC-ICPMS of mineral separates from the Sahara 99555 quenched angrite (formed ∼5 Myr after Solar System formation) and bulk chondrules from the Semarkona LL3.00 ordinary chondrite (formed ∼2 Myr after Solar System formation). These two objects are important anchors in early Solar System chronology. In both cases, resolvable excess 60Ni from 60Fe-decay is found. The Semarkona chondrule isochron defines an initial 60Fe/56Fe ratio of (5.29 ± 3.27) × 10−9 at the time of chondrule formation (Figure 1(a)). The Sahara 99555 mineral separate isochron defines and initial 60Fe/56Fe ratio of (1.96 ± 0.77) × 10−9 (Figure 1(b)). These two 60Fe/56Fe add to an already long list of meteoritic materials for which the 60Fe/56Fe abundances are constrained namely bulk HEDs (Vesta), SNCs (Mars), bulk angrites, D'Orbigny angrite, Gujba (CB) chondrules, and NWA 5717 chondrules. The initial 60Fe/56Fe ratio in Sahara 99555 may have been disturbed by terrestrial alteration but all samples measured by MC-ICPMS give a consistent uniform initial 60Fe/56Fe of (1.01 ± 0.27) × 10−8 (Figure 2, Table 2). This is almost two orders of magnitude lower than the estimated value from chondrule pyroxene measurements by SIMS, which give an initial 60Fe/56Fe ratio of ∼7 × 10−7. We have no explanation for the discrepancy but note that in many cases, the SIMS 60Fe/56Fe isochrons are defined by points that largely overlap with zero (Figure 3) and were only measured on one sample type (pyroxene in chondrules from unequilibrated ordinary chondrites). In contrast, the MC-ICPMS results were measured on a variety of planetary materials formed at different times and the significances of the isochrons are in most cases very high. Until unambiguous internal isochrons are measured in situ, the weight of evidence favors a low 60Fe/56Fe ratio at Solar System birth. Such low ratio contrasts with the high 26Al/27Al ratio, which can be explained if 60Fe was derived from the long-term chemical evolution of the Galaxy while 26Al was derived from the wind of a Wolf-Rayet star. To test this idea, detailed modeling of the interaction of SN ejecta with W–R bubbles and surrounding medium is needed.

We thank V. Dwarkadas and B. S. Meyer for discussions. G.J. MacPherson (Smithsonian Institution) generously provided the Semarkona chondrules analyzed in this study. This work was supported by grants NNX12AH60G and NNX14AK09G from NASA to ND.

Please wait… references are loading.
10.1088/0004-637X/802/1/22