Articles

ASTROPHYSICAL SHRAPNEL: DISCRIMINATING AMONG NEAR-EARTH STELLAR EXPLOSION SOURCES OF LIVE RADIOACTIVE ISOTOPES

, , and

Published 2015 February 10 © 2015. The American Astronomical Society. All rights reserved.
, , Citation Brian J. Fry et al 2015 ApJ 800 71 DOI 10.1088/0004-637X/800/1/71

0004-637X/800/1/71

ABSTRACT

We consider the production and deposition on Earth of isotopes with half-lives in the range 105–108 yr that might provide signatures of nearby stellar explosions, extending previous analyses of Core-Collapse Supernovae (CCSNe) to include Electron-Capture Supernovae (ECSNe), Super-Asymptotic Giant Branch (SAGB) stars, Thermonuclear/Type Ia Supernovae (TNSNe), and Kilonovae/Neutron Star Mergers (KNe). We revisit previous estimates of the 60Fe and 26Al signatures, and extend these estimates to include 244Pu and 53Mn. We discuss interpretations of the 60Fe signals in terrestrial and lunar reservoirs in terms of a nearby stellar ejection ∼2.2 Myr ago, showing that (1) the 60Fe yield rules out the TNSN and KN interpretations, (2) the 60Fe signals highly constrain SAGB interpretations but do not completely them rule out, (3) are consistent with a CCSN origin, and (4) are highly compatible with an ECSN interpretation. Future measurements could resolve the radioisotope deposition over time, and we use the Sedov blast wave solution to illustrate possible time-resolved profiles. Measuring such profiles would independently probe the blast properties including distance, and would provide additional constraints for the nature of the explosion.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The most violent stellar explosions are the sources of most of the heavy elements on Earth, and supernovae (SNe) in particular are estimated to occur at a rate of ∼1–3 per century in our Galaxy (e.g., Adams et al. 2013, and references therein). It is inevitable that, over the course of geological time, some such explosions will have occurred within ∼100 pc of the Earth, close enough to have deposited some ejecta on the Earth and Moon (e.g., Shklovskij 1969; Fields 2004, and references therein). Indeed, the Geminga pulsar located ∼250 pc away (Faherty et al. 2007) is the remnant of an SN explosion estimated to have occurred ∼300 kyr ago, and may be partly responsible for the low density of the interstellar medium (ISM) around the solar system (Bignami & Caraveo 1996). Similarly, 26Al gamma-ray line emission and large-angle Hα filaments suggest an SN toward the Antlia constellation 60–240 pc away (McCullough et al. 2002); if this event created a neutron star associated with the high-proper-motion pulsar PSR J0630-2834, then the explosion occurred about 1.2 Myr ago at about 140 pc (Tetzlaff et al. 2013). The question then arises whether some closer astrophysical explosion might have left detectable traces on the Earth itself in the form of geological isotope anomalies. Moreover, with a closer astrophysical explosion, the possibility for biological damage, even a mass extinction arises (for recent references, see, e.g., Melott & Thomas 2011; Beech 2011; Dartnell 2011; Atri & Melott 2014).

Discussions of this possibility date back to the pioneering study of Alvarez et al. (1980). These authors discovered an iridium anomaly associated with the Cretaceous–Tertiary transition that they argued could not, in fact, be associated with an SN explosion, but instead with a giant impact. Subsequently, Ellis et al. (1996) surveyed possible isotope signatures of a nearby SN explosion, including 26Al, 53Mn, 60Fe, and 244Pu. Motivated by this study, Knie et al. (1999) searched for an anomaly in the 60Fe abundance in a deep-ocean ferro–manganese (Fe-Mn) crust, and found one that appeared ∼2.2 Myr ago. Although primordial solar system composition shows enrichment from extra-solar origins, to our knowledge, this is the first such specific extra-solar event to be identified. Following the Knie et al. (1999) discovery, its interpretation was discussed in Fields & Ellis (1999) and possible corroborating isotope signatures were discussed in Fields et al. (2005). Benítez et al. (2002) proposed that the event arose in the Sco-Cen OB association, which was ∼130 pc away at the time of the 60Fe-producing event. Fields et al. (2008) presented hydrodynamic models for the SN blast impact with the solar wind, and Athanassiadou & Fields (2011) highlighted the importance of the ejecta condensation into dust grains.

The 60Fe signal has subsequently been confirmed in another Fe-Mn crust sample (Knie et al. 2004; Fitoussi et al. 2008) and in lunar regolith (Cook et al. 2009; Fimiani et al. 2012, 2014), but no other accompanying isotope anomaly has been found in studies of 26Al abundances (Feige et al. 2013). Searches for 244Pu have produced just a single count, albeit with no stable isobar background (Wallner et al. 2000, 2004).

In this paper, we broaden our previous analyses in four ways. In a first step, we provide yields for isotopes from the Core-Collapse Supernovae (CCSNe) considered previously and extend our analysis to include the cases of Electron-Capture Supernovae (ECSNe), Thermonuclear/Type Ia Supernovae (TNSNe, also known as a Type Ia SN), KNe (also known as Neutron Star Mergers), and Super-Asymptotic Giant Branches (SAGBs), which have not been considered previously in this context. For this paper, we distinguish between ECSNe and the more massive CCSNe since there are qualitative differences in the collapse and explosion mechanism as well as nucleosynthesis of these two classes. Secondly, we revisit the formalism surrounding the deposition calculations, including the impact and some geology of the uptake factor, and also the possibility of using sediments to get time-resolved signals and give predictions for these profiles. We also discuss the filtering processes impacting the transport of the signal via dust. Next, we discuss the compatibility between the terrestrial and lunar evidence for a 60Fe anomaly, and we also analyze the existing limits on the 26Al abundance from samples bracketing the 60Fe anomaly. In combination with these previous steps, as a fourth and final step, we survey the possible interpretations of the 60Fe anomaly and make predictions for upcoming measurements.

We find that a TNSN and a KN would yield too little 60Fe, and can be ruled out as possible sources for the Knie et al. (2004) 60Fe signal. Additionally, we find an SAGB source constrained, but not eliminated due to uncertainty in the Local Bubble's magnetic field and the location of a possible SAGB source. CCSNe from our set of masses and ECSNe cannot be ruled out based on the available measurements.

2. PROGENITORS AND DELIVERY TO THE SOLAR SYSTEM

Previous papers have focused on CCSNe as the likeliest progenitor for the 60Fe signal. However, there are other astrophysical ejections that are thought to produce 60Fe but have not been considered previously. These include TNSNe, ECSNe, KNe, and SAGBs. Table 1 summarizes the yields for possible CCSNe, ECSNe, and SAGBs progenitors as used in our model calculations; yields are expressed in units of M.

Table 1. Ejected Masses for Various Radioactive Isotopes, in M

Progenitor 15-M CCSNa 19-M CCSNa 20-M CCSNa 21-M CCSNa 25-M CCSNa 8–10-M ECSNb
26Al 2.6 × 10−5 3.2 × 10−5 3.0 × 10−5 4.6 × 10−5 7.0 × 10−5 4.4 × 10−8
53Mn 1.8 × 10−4 2.1 × 10−4 1.3 × 10−4 2.3 × 10−4 3.6 × 10−4 1.1 × 10−6
60Fe 6.6 × 10−5 1.1 × 10−4 3.6 × 10−5 2.5 × 10−5 1.5 × 10−4 3.6 × 10−5
41Ca 4.3 × 10−6 2.7 × 10−5 4.3 × 10−4 6.9 × 10−6 3.2 × 10−5 2.0 × 10−7
93Zr 1.3 × 10−8 4.7 × 10−8 9.8 × 10−9 5.9 × 10−8 1.5 × 10−7 N/Ad
97Tc 4.8 × 10−11 4.2 × 10−11 1.9 × 10−10 1.3 × 10−10 8.3 × 10−11 N/Ad
107Pd 4.1 × 10−10 8.4 × 10−10 4.6 × 10−10 1.4 × 10−9 1.4 × 10−9 N/Ad
146Sm 3.9 × 10−10 6.3 × 10−12 3.4 × 10−10 8.5 × 10−10 1.2 × 10−9 N/Ad
182Hf 1.4 × 10−10 1.5 × 10−9 2.5 × 10−10 5.5 × 10−10 4.3 × 10−10 N/Ad
244Puc 2.0 × 10−11 2.2 × 10−10 3.7 × 10−11 8.1 × 10−11 6.3 × 10−11 N/Ad
Progenitor 6.5-M SAGBe 7.0-M SAGBe 7.5-M SAGBe 8.0-M SAGBe 8.5-M SAGBe 9.0-M SAGBe
26Al 5.0 × 10−6 5.0 × 10−6 5.0 × 10−6 8.0 × 10−6 1.0 × 10−5 1.1 × 10−5
53Mn 0 0 0 0 0 0
60Fe 5.0 × 10−6 3.0 × 10−6 4.0 × 10−6 9.0 × 10−6 1.4 × 10−5 1.4 × 10−5

Notes. In addition to the cited CCSN yields from Rauscher et al. (2002), 26Al and 60Fe yields from Limongi & Chieffi (2006) (11–120 M) were investigated as well. These did not show any additional features beyond those shown with the Rauscher et al. (2002) yields. aS15, S19, S20, S21, and S25 Models respectively, Rauscher et al. (2002). b"Unchanged" configuration, Wanajo et al. (2013). c244Pu yields calculated as outlined in Fields et al. (2005) but using Rauscher et al. (2002) 182Hf yields. dr-process yields for ECSN are not available at present although Wanajo et al. (2013) stated that ECSN may produce some weak r-process elements. eDoherty et al. (2013).

Download table as:  ASCIITypeset image

2.1. Supernovae

SNe include both CCSNe and TNSNe; CCSNe are the results of massive stars completing Fe/Ni fusion in their cores and collapsing under the influence of gravity whereas TNSNe result from runaway nuclear fusion in a C-O white dwarf near its Chandrasekhar limit. Both types have similar explosive energies and modes of transporting ejecta. However, although known to be sources of stable iron isotopes, TNSNe are calculated to produce relatively little 60Fe, namely ∼2.3 × 10−9M, according to the W7 Type Ia Model of Nomoto et al. (1984). ECSNe form a special subcategory of CCSNe with significantly different radioactive isotope yields. While the 60Fe yields for CCSNe and ECSNe are similar (∼10−6–10−3M, Rauscher et al. 2002; Limongi & Chieffi 2006), their yields for other isotopes (e.g., 26Al and 53Mn) are vastly different (Wanajo et al. 2013). For the purposes of this paper, ECSNe will refer to SNe from 8–10 M stars, CCSNe will refer to SNe with progenitor masses > 10 M, and SNe will refer to CCSNe, ECSNe, and TNSNe. It should be noted that the yields for 244Pu were calculated using the same method from Fields et al. (2005); the proportions of r-process elements are generally consistent to that found in metal-poor globular clusters and in the Sun. In this paper, however, the yields for 244Pu were based on the yields for 182Hf from Rauscher et al. (2002) using the ratios given in Fields et al. (2005).

SNe can show large variations in their isotope yields. TNSN 60Fe yields show variations over several orders of magnitude (∼10−18–10−7M) due to variations in the number and location of ignition points and the transition from deflagration to detonation (Seitenzahl et al. 2013).4 CCSNe/ECSNe yields are highly dependent on a number of factors including when different layers are mixed. This can be seen in the variations of yields from one mass to another (Rauscher et al. 2002). The yields within a given mass are also subject to uncertainties in nuclear reaction rates (3-α and 12C(α,γ)16O) which can lead to an almost order of magnitude shift in the production of 26Al and 60Fe (Tur et al. 2010). In Section 6, Figure 3, we show the calculated distances with uncertainties indicated by dashed lines for a factor of five variation in the 60Fe yield for each CCSN/ECSN type.

Fiducial parameters for the explosions and the ISM are chosen as follows. We assume CCSNe and TNSNe deposit $E_{\rm CCSN \ \& \ TNSN} = 10^{51}$ erg into their ejecta, while ECSNe deposit EECSN = 1050 erg (Wanajo et al. 2009). Because the Local Bubble shows evidence of multiple SN explosions, we will assume that if an SN were the source of the 60Fe signal, it would be the most recent SN, meaning the SN occurred in an already depleted ISM, but not as depleted as the current density of the Local Bubble (i.e., nAverage ISM = 1.0 cm−3 > nISM > nLocal Bubble = 0.005 cm−3). Therefore, we estimated an SN would have occurred in an ISM of density, nISM = 0.1 cm−3, temperature, T = 8000 K, and sound speed, cs = 10 km s−1 (i.e., approximate values for the Local Cloud, Fields et al. 2008).

2.2. Kilonovae

KNe are thought to result from Neutron Star–Neutron Star (NS–NS) or in some cases Neutron Star–Black Hole (NS-BH) mergers (Li & Paczyński 1998; Metzger et al. 2010; Tanvir et al. 2013). For this paper, we only consider the KN explosion's lower-energy, spherical/torical ejection and not its highly beamed gamma-ray burst (GRB) jet. The rapid decompression and ejection of neutron-rich NS matter makes these events a natural site for the r-process (Lattimer & Schramm 1974; Symbalisty & Schramm 1982). While KNe are less energetic than SNe (EKN = 1049 erg, Goriely et al. 2011), we will consider a possible KN source of the 60Fe signal as occurring in the same ISM conditions as an SN. However, given the axisymmetric nature of NS–NS mergers, we will not apply the same constraints to KNe as SNe, but will instead evaluate KNe with respect to isotope yields and frequency.

While there has been some modeling of KN yields, none we are aware of have specifically stated a yield for 60Fe. However, it is possible to determine an upper limit on the range for a KN. In Goriely et al. (2011), they list mass fractions for every atomic number up to ∼200 for a NS–NS merger with a total merger mass of 2.7 M. If we assume all of the isotopes with A = 60 are in the form of 60Fe (Mej, total = 10−3–10−2, $X_{{60}_{\rm Fe}} = 10^{-5}$) then the upper limit to the mass of ejecta in 60Fe is 10−7M.

2.3. Super-asymptotic Giant Branch Stars

SAGBs (6.5–9 M) are post-main-sequence stars that produce large amounts of dust (see, e.g., Ventura et al. 2012) and have strong winds (∼30 km s−1) capable of carrying dust great distances. SAGBs produce 10−6–10−5M of 60Fe (Doherty et al. 2013), but are distinguishable from SNe in that they produce practically no 53Mn (Wasserburg et al. 2006; Fimiani et al. 2014). We note that SAGB yields are subject to an uncertainty in the onset of the super-wind phase (Doherty et al. 2013); a delayed onset results in generally increased yields. The implication for distance is shown with dashed error bars on SAGB results in Section 6, Figure 3. In contrast to SNe, we do not expect an SAGB wind to affect the density of the Local Bubble appreciably, and we assume that an SAGB source for the 60Fe signal would have occurred in an ISM like that found in the Local Bubble today (i.e., nISM = 0.005 cm−3, temperature, T = 106 K, and sound speed, cs = 100 km s−1). Finally, we assume the initial velocity of the SAGB grains to be: vgrain, 0 = 30 km s−1.

2.4. Dust Transport to the Solar System

Regardless of the source, any 60Fe arriving in the solar system will need to be in the form of dust. Fields et al. (2008) showed that the solar wind will keep any gaseous isotopes from reaching the Earth (unless an SN is sufficiently close, but this will be used as a constraint later in Sections 3.1 and 6). We assume that the dust grains are spherical and select as our fiducial values for dust grains: density, ρgrain = 3.5 g cm−3 (an average value for silicates), radius, a = 0.2 μm (this selection is based on discussion in Section 4.2), and voltage, ${\cal V} = 0.5 \ {\rm V}$. Departures from these values will be specifically stated.

3. FORMALISM

In order to identify the most likely progenitor, we will attempt to constrain the source and its allowable distances using the measured 60Fe fluence and calculated yields. In the future, with additional measurements of other isotopes, we can use the other isotope yields to constrain the source using the observed isotope to 60Fe ratio.

Several previous works have presented the formalism for calculating deposited material from an SN (Ellis et al. 1996; Fields & Ellis 1999; Fields et al. 2005). These works focused primarily on short ranges (SN distances, D ∼ 10 pc) and on the isotope 60Fe. For such short distances, the losses due to decay of live radionuclides en route from the SN to Earth amount to ≲ 1% and can be ignored. At greater ranges (D ∼ 100 pc) and for shorter-lived isotopes (in particular, 26Al with $\tau _{1/2, {}^{26}{\rm Al}}$ = 0.717 Myr) decays en route become a significant issue. Accounting for this, the observed fluence today, ${\cal F}_{{\rm obs}, i}$, for each isotope, i, in atoms per area on the surface of the Earth within a given substance (e.g., crust, sediment, etc.) becomes:

Equation (1)

where Mej, i is the mass of the ejecta by isotope, D is the distance from the progenitor to Earth, Ai is the atomic number of the isotope, mu is the atomic mass unit, tarr is the time from today since the ejecta arrived at Earth, and ttravel is the time the ejecta traveled from the source to Earth. Also, τi is the mean lifetime of the isotope (τi = τ1/2, i/ln 2).

The uptake, Ui ≡ (Amount Collected)/(Amount Deposited), is the fraction of the isotope deposited on a surface that is collected by that material. The quantity is dimensionless, and ranges from one (the material collects 100% of deposited element) to zero (the material collects 0% of deposited element).5 It is further discussed in Section 4.1. Additionally, the dust fraction, fi, is the amount of the isotope in the form of dust that arrives at Earth (Section 4.2). It is similar to uptake in that it is also dimensionless and ranges from one (all of the isotope is in the form of dust and reaches Earth) to zero (none of the isotope is dust and/or reaches Earth). There is a factor of one-fourth from the ratio of the Earth's cross sectional to surface areas, because it is assumed material is distributed evenly over Earth's entire surface through collisional accretion only. Equation (1) also assumes an isotropic dispersal of material from the source (4πD2 factor for spherical distribution), that no additional isotopes are created after the ejection from the progenitor, and that the ejected material passes through a homogeneous ISM.

There are three other fluence quantities that appear in the literature: "decay-corrected" fluence, "surface" fluence, and "interstellar" fluence. The decay-corrected or arrival fluence, ${\cal F}_{{\rm arr}, i}$, is the total number of atoms per area that would have been measured at the time the signal arrived. It is calculated by correcting our previous description of fluence (see Equation (1)) for radioactive decay since the isotope was deposited ${\cal F}_{{\rm arr}, i} = e^{t_{\rm arr}/\tau _{i}} {\cal F}_{{\rm obs}, i}$. The surface fluence or global mean fluence, ${\cal F}_{{\rm surface}, i}$, is the total number of atoms per area that arrive at the surface of the Earth regardless of what substance they might be incorporated into, and is found by dividing the decay-corrected fluence by the uptake (i.e., ${\cal F}_{{\rm surface}, i} = {\cal F}_{{\rm arr}, i}/U_i$). It will be used in Sections 5 and 6 and will be specifically stated when used. The interstellar fluence, ${\cal F}_{{\rm interstellar}, i}$, also appears in the literature (e.g., Fitoussi et al. 2008; Cook et al. 2009), namely the number of atoms per area on the surface of the spherical shock front. It is related to the surface fluence by a factor of four, the ratio of Earth's cross section to surface area (i.e., ${\cal F}_{{\rm interstellar}, i} = 4{\cal F}_{{\rm surface}, i}$). Interstellar fluence will not be used in this paper, but the reader should be aware of the distinction when reviewing the literature.

In order to find the time delay, ttravel, from ejection to deposition on Earth, we must account for the propagation of the ejection through the intervening ISM, which depends on the progenitor. SNe transmit material via an explosive shock, whereas SAGBs use a wind-driven ejection. In the case of SNe, we will assume that the shock has transitioned from the free-expansion phase into the adiabatic/energy-conserving phase. For SAGBs, we will assume dust has been blown by winds from the star and experiences drag as it travels to Earth.

3.1. SN Expansion Profile and Constraints

For SNe in the adiabatic/energy-conserving phase, the shock follows a self-similar or Sedov–Taylor expansion profile. With the explosion at time t = 0, a shock is launched with radius, RSN, at elapsed time, t, given by

Equation (2)

for an SN explosion depositing energy ESN into the ejecta, propagating into a local ISM of density ρISM = munISM. The quantity ξ0 is a dimensionless constant that is of the order of unity for γ = 5/3 using the derivation in Zel'dovich & Raizer (1967). Thus the time interval to traverse distance, D, is

Equation (3)

where we have assumed a uniform ISM density. While we know this is a crude approximation (see, e.g., Abt 2011), deviations from the uniform case would be encoded into the signal and could be determined if the signal is time-resolved (for examples of non-uniform media, see Book 1994). We use the uniform ISM case as a baseline.

The density versus radius profile for an SN signal may be approximated by a "saw-tooth" profile. As will be described in greater detail in Section 3.3, Section 3.4, and Appendix B, measurements in sediment open the possibility to making time-resolved fluence measurements. A saw-tooth pattern gives a better approximation of the more exact Sedov solution than a uniform shell profile. The saw-tooth pattern reaches its maximum density value at the outer edge of the shock, then decreases linearly to a fraction of the total shock radius, epsilon, where the density is zero from that point to the center of the remnant. See Appendix B for a comparison of the exact Sedov, saw-tooth, and uniform shell profiles.

Possible SNe will be constrained by an inner "kill" distance and an outer "fadeaway" distance. The kill distance is the range at which an SN can occur and create extinction-level disruptions to the Earth's biosphere. The primary mechanism for an SN to accomplish this is for ionizing radiation (i.e., gamma-rays, hard X-rays, and cosmic-rays) to destroy O3 and N2 in the atmosphere producing nitric oxides (NOy) and leaving the biosphere vulnerable to UVB rays from the Sun (first described by Shklovskii & Sagan 1966, described in detail by Ruderman 1974, and updated by Ellis & Schramm 1995). This can be accomplished either by direct exposure (i.e., an X-ray flash from the SN) or by a "descreening" boost in cosmic-ray flux.6 Gehrels et al. (2003) calculated a kill distance Rkill ≲ 8 pc for the direct exposure case using the galactic gamma-ray background and SN rates, although, as pointed out by Melott & Thomas (2011), this is probably an underestimation based on more recent rate estimations. This work was expanded upon by Ejzak et al. (2007) and Thomas et al. (2008) to include X-rays and showed that for exposure durations up to 108 s, the effects on the biosphere were the same and that the critical value for an extinction-level SN event was an energy fluence of 108 erg cm−2 (not to be confused with the description of fluence used throughout the rest of this paper). As noted in Melott & Thomas (2011), these direct exposure calculations use SN rate and photon and cosmic-ray emission information that are improving, but still subject to large uncertainties.

Here we calculate the kill distance using the descreening case described in Fields et al. (2008); it yields the same range as the direct exposure calculations, and is scalable to the energy of the SN, ESN. The descreening kill distance is the range at which an SN can occur and its shock will penetrate the solar system to within 1 AU of the Sun. It is determined by setting the solar wind pressure, PSW, equal to the pressure of the SN shock (see, e.g., Fields et al. 2008). In this case, the pressure from the SN has little effect on the Earth, but by pushing back the solar wind, the Earth is inside the SN remnant and now directly exposed to the SN cosmic-rays that would normally diffuse out over 104 yr (Fujita et al. 2010) in addition to an increased galactic cosmic-ray background. In-turn, these destroy O3 and N2 in the atmosphere, just as in the direct exposure case, in addition to increased radionuclide deposition (Melott & Thomas 2011). Using our fiducial SN values, we find

Equation (4)

The fadeaway distance is the range at which the SN shock dissipates and slows to the sound speed of the ISM. Because of uncertainty in when SN dust decouples from the rest of the shock, the fadeaway distance is not an absolute limitation like the kill distance, but can serve as a guide to the likelihood of a progenitor. Using the derivation from Draine (2011, Equation (39.31)), we find

Equation (5)

3.2. SAGB Expansion Profile and Constraints

In the case of SAGBs, we assume the dust is ejected radially and that the distance traveled by SAGB dust is determined only by a drag force, Fdrag (magnetic forces will be considered later in this section as a constraint). Using the description in Draine & Salpeter (1979), the drag force due to only collisional forces (in cgs units) is7

Equation (6)

with

Equation (7)

where j is the respective species in the ISM (we consider only ionized H; He and free electrons will be neglected), mj is the particle mass of that respective species, k is the Boltzmann constant, T is the temperature of the ISM, and v is the velocity of the particle relative to the medium. For small v (i.e., vgrain ≲ 100 km s−1) the first term in G0 will dominate, leaving $G_{0}(s) \approx 8s/(3 \sqrt{\pi })$. Making these simplifications, we find

Equation (8)

where RSAGB is the distance the dust grain has traveled from the SAGB, and $v_{\rm grain} \equiv \dot{R}_{\rm SAGB}$. Integrating twice and setting RSAGB equal to the traverse distance from the progenitor to Earth, D, gives the transit interval:

Equation (9)

where ζ0 is a dimensionless constant and of the order of unity for γ = 5/3, cs is the sound speed in the ISM, and vgrain, 0 is the initial velocity of the dust grain when it leaves the SAGB.

We approximate the density profile for an SAGB signal by a uniform shell or "top-hat" shape (we will use the "top-hat" profile to avoid confusion with the SN profile discussion). This corresponds to a uniform, steady wind. The top-hat pattern reaches its maximum density value at the leading edge of the signal, retains this value for the duration of the signal, and afterward the density returns to zero. The SAGB phase is characterized by thermal-pulsing of the star's envelope. Since the duration of each pulse (Δtpulse ∼ 1 yr) and interval between pulses (Δtinter ∼ 100 yr) are much shorter than the SAGB phase (ΔtSAGB ∼ 100 kyr ≫ Δtinter > Δtpulse) we assume that the amount of ejected material is approximately constant for the duration of the SAGB phase (see Siess 2010). Furthermore, we assume all parts of the signal experience the same forces from the SAGB to the Earth, so that the duration of the signal remains the same (i.e., Δtsignal, SAGB = ΔtSAGB = 100 kyr).

Because SAGB winds would not be as devastating to the Earth as an SN shock, we forego establishing an inner kill distance, but establish two outer distances: the drag stopping distance, Rdrag, and the magnetic deflection distance, Rmag. The distance, Rdrag, is the range of the SAGB dust grains' e-folding velocity, and the Rmag is the range at which deflection of the dust grain's trajectory by the ISM's magnetic field becomes significant. Using the derivations from Murray et al. (2004), we find

Equation (10)

Equation (11)

The implications of these limits will be discussed in greater detail in Section 6.

3.3. The Radioactivity-based Distance to the Explosion

To estimate the distance, D, to the explosion, we wish to invert Equation (1). When the transit time (the time for the shock to travel from the source to Earth) is negligible (ttravel ≪ τi), the procedure is straightforward, since the only distance dependence is the inverse square dilution of the ejecta, and so $D \propto 1/\sqrt{{\cal F}_{{\rm obs}, i}}$.8 This has been assumed in work to date. However, if D is sufficiently large, then via Equation (2), the distance-dependent transit time can become important and must be included in solving Equation (1); we have done this in all of our results.

Another effect occurs when the radioisotope signal is sampled sufficiently finely to resolve the time history of the deposition signal. This occurs when the signal width (the time from the arrival of the signal's leading edge to the departure of the signal's trailing edge) is larger than the sampling time resolution (Δtsignal > Δtres). In this case, the total radioisotope signal, summed over all time bins, should be used in solving the distance via Equation (1); and as we show in Section 3.5 below, the width of Δtsignal for an SN probes independently the explosion distance. However, the available Knie et al. (2004) data has a time sampling of Δtres ∼ 880 kyr, and shows no evidence for a signal that is extended in time. Thus we infer that the signal width Δtsignal ≲ Δtres, and indeed we find Δtsignal ≲ 880 kyr for most of our possible progenitors. In addition, when solving for distance using the Knie results, we assumed the signal arrived halfway through the sample. Therefore, half of the sampling width is used as a median value rather than assuming the signal arrives right as the sampling window begins or just before it ends.

3.4. Expected Behavior of Time-resolved Signals

Although the time resolution of the deep-ocean 60Fe crust measurements in Knie et al. (2004) data preclude resolution of the time structure of the radioisotope deposition, measurements in sediments can achieve much better time resolution, and so it is of interest to explore the time dependence of the explosion signal. Such work was pioneered by Ammosov et al. (1991) in the context of the 10Be anomalies ∼35 and 60 kyr ago in Antarctic ice cores (Raisbeck et al. 1987).

In developing a model for deposition, we examine a Sedov-Taylor profile for an SN shock (a similar examination could be done for the SAGB case if we had more detailed description of its signal width dependence), which implies an energy-conserving (adiabatic) evolution. This also means the shock will remain self-similar as it progresses and that the majority of the material is concentrated near the leading edge. Although the remnant density profile changes once the remnant transitions to the radiative/momentum-conserving phase, we have chosen to maintain the Sedov profile. We did this, firstly, because the profiles are similar in shape (the radiative profile is a thicker shell profile, see, e.g., Shu 1992, Figure 17.4). Secondly, because the dust will decouple from the gas at some point, either when the shock meets back pressure from encountering the solar wind or at the transition from the adiabatic to the radiative phase when the shock loses its internal radiation pressure (Draine 2011). The exact nature of this decoupling and resulting profile would require detailed calculations that are beyond the scope of this paper.

We assume that the explosion ejecta are well-mixed within the swept-up matter, so that the ejecta density profile follows that of the blast itself. As described in Appendix B, we approximate the Sedov density profile as a "saw-tooth" that drops linearly from a maximum behind the shock radius RSN to zero at an inner radius rin = (1 − epsilon)RSN, with epsilon ≈ 1/6. We note that the leading edge of the blast from an event at distance D arrives at a time ttravel since the explosion which corresponds to geological time tarr; this is given by D = RSN(tarr), whereas the trailing edge of the shell arrives at time tdep given by D = (1 − epsilon)RSN(tdep). Thus we have tdep = tarr/(1 − epsilon)5/2.

With these assumptions we can model the global-averaged flux time profile, $\mathbb {F}$. For radioisotope i, we have:

Equation (12)

as shown in Appendix B; note that here all times are geological and thus increase toward the past.9 This describes a cusp-shaped decline from an initial flux:

Equation (13)

Equation (14)

where the numerical values are mi = 60mu and the yield is appropriate for 60Fe.

To test the profile in Equation (12), one can fit observed time-resolved data to this form, letting tarr and tdep, and $\mathbb {F}_i(D,t_{\rm arr})$ be free parameters. For the Sedov profile, the time endpoints should obey tdep = tarr/(1 − epsilon)5/2, which provides a consistency check for the Sedov (adiabatic) approximation. Moreover, as we show in Section 3.5, the interval tdeptarr provides an independent measure of the explosion distance.

If the radioisotope abundance is sampled over a time interval [t1, t2], then the fluence (without radioactive decays) will be the integral of the surface flux over this interval: ${\cal F}_i(t_1,t_2) = \int _{t_1}^{t_2} \mathbb {F}_i(t) \ dt$, where we have suppressed the dependence on time-independent parameters such as distance. If the observed time resolution Δtres is small compared to tarr and tdep, then the fluence profile ${\cal F}_i(t-\Delta t_{\rm res}/2,t+\Delta t_{\rm res}/2) \approx \mathbb {F}_i(t) \ \Delta t_{\rm res}$ will recover the flux history.

So far we have calculated the observed fluence without the effect of decay. Since all the atoms were created at the same time, the observed fluence is reduced by a factor of $e^{-(t_{\rm arr}+t_{\rm travel})/\tau _i}$. The effects of uptake (Section 4.1) and dust depletion (Section 4.2) introduce further factors of Ui and fi. We thus arrive at the observed fluence:

Equation (15)

One can show that the total integrated fluence ${\cal F}_{{\rm obs}, i}(t_{\rm arr},t_{\rm dep})$ takes precisely the value in Equation (1). This reflects the number conservation (aside from decays) of the atoms in the SN ejecta. This also implies that, in a time-resolved measurement, the total fluence is conserved, which means that the area under a fluence versus time curve will be constant for fixed explosion parameters. This implies that fluence measurements of 60Fe/Fe will show lower values when measured over fewer bins, and finer sampling will show higher fluence over more bins (see Figure 1).

Figure 1.

Figure 1. Sample time-resolved calculations of the observed fluence, ${\cal F}_{\rm obs, {}^{60}{\rm Fe}}$, for three SNe and an SAGB. Each progenitor is at a different distance: 25-M SN at 130 pc, 21-M SN at 59 pc, ECSN at 67 pc, and 6.5-M SAGB at 79 pc. Of note, each of these progenitors would produce the same measured 60Fe signal by Knie et al. (2004); Fitoussi et al. (2008), but with a finer time-resolution (10 kyr in this case), the shape of the signal is readily discernible. Also note that since the plots produce the same observed fluence, the areas under the curves are the same.

Standard image High-resolution image

3.5. A Resolved Signal Timescale Probes the Distance to the Explosion

A time-resolved signal not only encodes information about the shape of the blast density profile, but also about the distance. The relation D = (1 − epsilon)RSN(tdep) also allows us to write tdep = tarr/(1 − epsilon)5/2 and thus that

Equation (16)

where we define a dimensionless parameter, α, that relates the signal width in terms of the arrival time. For our profile, α = (1 − epsilon)−5/2 − 1 ≈ 0.577. We see that the radioisotope width grows in proportion to the blast transit time to Earth, ttravel, which itself depends on distance. Thus a measurement of Δtsignal from a time resolved radioisotope signal gives an measure of the explosion distance. Within this Sedov model, D = RSN(tarr), and so we can solve for the distance based on "blast timing:"

Equation (17)

Equation (18)

This distance measure is independent of the "radioactivity distance" and its associated uncertainties, notably due to uptake, dust fraction, and radioisotope yields. Moreover, as characteristic for Sedov blast waves, the blast-timing distance in Equation (17) scales as small powers of the timescale, as well as the energy and density. This will weaken the uncertainties in distance estimate.

Having two independent distance estimates allows a consistency check for the model. Alternatively, if we adopt one of the distance estimates as the correct value, we can deduce the parameters in the other. For example, adopting the blast-timing distance we can use the observed fluence and solve for the product of radioisotope yield, uptake, and dust fraction: $M_{{\rm ej}, i} U_i f_i \propto D^2 \ {\cal F}_{{\rm obs}, i}$. Given a geophysical estimate of uptake, this allows for a measure of the yield and thus a direct probe of the nucleosynthesis output and thus the nature of the explosion (see Figure 1 for examples).

4. DEPOSIT FACTORS

The delivery of astrophysical debris to the solar system and its incorporation into geological and lunar samples is clearly complex. In this section we consider several factors we are aware of that can have a substantial influence on the observed signals.

4.1. Uptake

Uptake in the Fe-Mn crust involves a complex chemical process that incorporates material into the crust with a low accumulation rate ∼2 mm/Myr. Usual deep-ocean sediments, on the other hand, do not make such a geochemical selection, and have greater accumulation rates ∼3–4 mm/kyr (Feige et al. 2012). The Fe uptake factor was calculated by Knie et al. (2004) using the relative concentrations of Fe and Mn in water and the Fe-Mn crust and the uptake of Mn (4%), leading to an estimate for the Fe uptake, UFe = 0.6%. However, recent studies have suggested that UFe = 0.5–1 (Bishop & Egli 2011; Feige et al. 2012). Using the smaller estimate of UFe, Knie et al. (2004) calculated an SN distance of D ≈ 40 pc; a reasonable distance considering the Local Bubble is ∼200 pc in diameter in the Galactic plane extending 600 pc perpendicular to the plane (Fuchs et al. 2006) and superbubbles in the Large Magellanic Cloud are typically ∼100–200 pc in diameter (for a single round of star formation, Chu 2008). Changing the uptake factor has the immediate effect of changing the implied distance to the explosion. This can be roughly understood if we ignore the effect of the debris decays in transit, in which case the signal follows the inverse square law and we have $D \propto \sqrt{U_{i}/{\cal F}_{{\rm obs}, i}}$. The effect of decays en route softens this dependence somewhat.

If UFe is an order of magnitude larger, the implied distance increases by a factor ≳ 2. With a Fe-Mn crust uptake factor of UFe = 0.5–1, the implied distances are around D ∼ 200 pc. This seems an unlikely distance, given that the solar system is roughly in the center of the Local Bubble (see Berghöfer & Breitschwerdt 2002, Figure 2), and an SN would have had to occur outside the Local Bubble in order to produce the signal (assuming the progenitor is in the Galactic plane).

However, implicit in the Knie et al. (2004) calculation is the assumption that the dust fraction, fFe = 1. As we will show in Section 4.2, this is most likely not the case, and the combination of a higher uptake value (we chose UFe = UAl = 0.5) with a smaller dust fraction (fi ≪ 1) can still yield reasonable progenitor distances.

4.2. Dust Condensation

It was shown in Fields et al. (2008), that for an SN further than D ∼ 10 pc, the solar wind would keep the SN blast plasma outside of 1 AU, and thus the Earth will not find itself inside gas-phase SN debris. However, refractory SN ejecta will be condensed into dust grains. As discussed in Athanassiadou & Fields (2011), we expect these grains to be entrained in the SN blast as it reaches the heliosphere, but then decouple at the SN-solar wind shock and move essentially ballistically through the inner solar system. Once the dust decouples from the gas in the shock, it can travel great distances. At this point, both SNe and SAGBs behave the same, as they are subject to the same drag stopping distance, Rdrag discussed in Section 3.2. This is more than sufficient to reach the Earth in spite of the solar wind, and indeed should carry dust grains beyond the SN remnant when it finally comes to rest. Thus, for the D > 10 pc events of interest, the amount of any radioisotope i that comes to Earth will be proportional to the fraction, fi, of the isotope that reaches Earth via dust, as seen in Equations (1) and (15).

Determining fi requires examining a number of factors (the results are summarized in Table 2).

  • 1.  
    How much of the isotope condenses into dust at departure from source?
  • 2.  
    How much of that dust survives the interstellar journey from the source to the solar system?
  • 3.  
    How much of the remaining dust can filter through the heliopause and enter the solar system?
  • 4.  
    How much of the filtered dust can overcome the solar wind/radiation pressure and reach the Earth's orbit?

Table 2. Summary of Relevant Dust Filtering Processes

Resulting Dust Fraction Repository Material Density Elemental Fraction Fraction Fraction Fraction
Fraction in Condensed Surviving Passing into Reaching
Material into Dust Interstellar Heliosphere Earth's Orbit
Passage
fFe, SN ≈ 0.01 Fe 7.9 0.1a 1c 1a 0.1a,d 1g
  FeS 4.8 0.9a 1c 0.5a 0a, d 1g
fFe, SAGB ≈ 0.2 Fe 7.9 0.48b 1c 1 0.1a,d 1g
  [Fe, Mg]2SiO4 3.5 0.43b 1c 1 0.25d,e 1g
  [Fe, Mg]SiO3 3.5 0.09b 1c 1 0.25d, e 1g
fAl, SN ≈ 0 Al2O3 4 1a 1c 0.01a 0a, d 1g
fAl, SAGB ≈ 1 Al2O3 4 1a 1c 1 1d, f 1g

Notes. Density is given in g cm−3. Mass fractions from each filter process are determined using the relative size number distributions, N(a), given in Silvia et al. (2010) and Mathis et al. (1977), multiplying by volume, V(a), and density, ρ, and integrating over the relevant radii [amax, amin]. This method is described in Mathis et al. (1977): $m = \int ^{a_{{\rm max}}}_{a_{{\rm min}}} \rho \ N(a) \ V(a) \ da$. aSilvia et al. (2010). bPollack et al. (1994), Table 1 B. Given the temperature of the Local Bubble (T > 1000 K), it is assumed all FeS has been converted to Fe. cMatsuura et al. (2011) and Spitzer & Jenkins (1975). dSlavin et al. (2010). eMathis et al. (1977). fHoppe & Zinner (2000). gBurns et al. (1979) and Gustafson (1994).

Download table as:  ASCIITypeset image

In order to determine the isotope fraction that condenses into dust, we use recent observations of SN 1987A. Herschel observations in the far infrared and sub-millimeter wavelengths were modeled in Matsuura et al. (2011) with different elemental abundances and dust compositions. In both models studied, the synthesized Fe mass was nearly identical to the Fe condensed into dust. This suggests that after less than 30 yr (much less time than required to travel to a nearby solar system), practically all of the iron from the SN is in the form of dust. Furthermore, when comparing condensation temperatures (Spitzer & Jenkins 1975), one finds that TC, Al = 1800 K > TC, Fe = 1500 K, suggesting Al would condense into dust at the same time as Fe, if not sooner. Based on this reasoning, we assume 100% of Al and Fe condenses into dust for both SNe and SAGBs.

While refractory elements seem to condense rapidly after ejection, only the dust that survives transport to the solar system will reach the Earth. Dust leaving from SAGBs will be subject to shocks from neighboring star systems as well as sputtering from radiation and collisions with other dust grains. However, for the purposes of this paper, we will assume these affects are negligible compared to other filtering effects examined. Therefore, we will assume all of the dust from SAGBs is able to pass from the SAGB to the solar system (a more in-depth discussion of interstellar effects on dust grains is discussed in Murray et al. 2004).

Conversely, SN remnants are likely to be much harsher environments for dust, leading to predictions of very small survival probabilities for some dust species and thus for some radioisotopes. Dust formed from ejecta in a newborn SN remnant will encounter a reverse shock as the remnant transitions from the free expansion to the Sedov/adiabatic phase. The reverse shock propagates from the outer edge of the remnant back to the source and is generally stronger than the interface between the outer edge of the remnant and general ISM. The reverse shock causes large-scale sputtering/destruction of grains resulting in gas phase emission from previously refractory elements (e.g., see emission from Cassiopeia A, Rho et al. 2008). Silvia et al. (2010, 2012) studied this interaction and found grains ≲ 0.1 μm to be most affected; ∼1% of Al2O3 (corundum), ∼50% of FeS (troilite), and ∼100% of Fe (metallic iron) previously condensed into dust survives.

Once the dust reaches the solar system, it must pass through the heliosphere to reach the Earth. Linde & Gombosi (2000) suggested a cut-off grain size of 0.1–0.2 μm for filtering by the heliosphere for grains with speeds of 26 km s−1 corresponding to the Sun's motion through the local ISM. This filtering is less severe for faster dust grains (Athanassiadou & Fields 2011), but for our larger SN distances we find slower speeds in Table 3. Since magnetohydrodynamic simulations by Slavin et al. (2010) showed penetration but strong deflection of 0.1 μm grains, we chose a minimum grain size of 0.2 μm for entering the solar system. For SN dust, this cut-off means that negligible amounts of Al2O3 and FeS grains will enter the solar system, while ∼10% of Fe grains will be large enough to pass through. For SAGB dust, we assume the Fe is in elemental Fe and silicate grains (distributed according to Pollack et al. 1994 with FeS assumed to be Fe); Al will be in Al2O3, but with a larger grain size (Hoppe & Zinner 2000). The Fe size distribution is assumed to be the same as for SN (Sterken et al. 2013), and silicate grains are assumed to follow the Mathis et al. (1977) distribution (dN/daa−3.5) ranging from 0.5–350 nm (Weingartner & Draine 2001). This means ∼10% of Fe, ∼25% of silicates, and ∼100% of Al from SAGBs will enter the solar system.

Table 3. Predicted Parameters for Possible 60Fe Signal Sources

Progenitor Distance to Source, D, Time en route, ttravel, Signal Width, Δtsignal, Arrival Speed, varr,
pc Myr kyr km s−1
6.5-M SAGB 79$_{-8}^{+13}$ 2.8 100 25
7.0-M SAGB 66$_{-7}^{+11}$ 2.3 100 26
7.5-M SAGB 73$_{-8}^{+12}$ 2.6 100 25
8.0-M SAGB 97$_{-9}^{+14}$ 3.5 100 24
8.5-M SAGB 110$_{-10}^{+15}$ 4.2 100 23
9.0-M SAGB 110$_{-10}^{+15}$ 4.1 100 23
15-M CCSN 94$_{-12}^{+19}$ 0.44 250 84
19-M CCSN 120$_{-13}^{+18}$ 0.74 430 61
20-M CCSN 71$_{-9}^{+15}$ 0.22 130 130
21-M CCSN 59$_{-8}^{+13}$ 0.14 80 170
25-M CCSN 130$_{-13}^{+17}$ 0.98 570 52
8-10-M ECSN 67$_{-8}^{+12}$ 0.61 351 43

Notes. Errors are only for variances in the Knie et al. (2004) decay-corrected fluence value and do not include variations in nuclear reaction rates (SNe) or delayed super-wind phase (SAGBs). These parameters are calculated with the Fe uptake factor, UFe = 0.5.

Download table as:  ASCIITypeset image

Lastly, the dust grain must overcome the Sun's radiation pressure once it is in the solar system. For this, we consider the parameter β (Burns et al. 1979) that characterizes the ratio of the Sun's radiation force, Fr, to the gravitational force, Fg:

Equation (19)

where Cr is 7.6 × 10−5 g cm−2 and Qpr is the efficiency of the radiation on the grain (we will assume Qpr ∼ 1 for the size of grains we are interested in, Gustafson 1994). From Sterken et al. (2013), only dust grains with β ≲ 1.3 will reach Earth's orbit; based on the densities of the minerals considered, if the grain can enter the solar system, it will be able to reach Earth's orbit.

Combining each of these factors, we find for SNe: fFe, SN ≈ 0.01 and fAl, SN is negligible. For SAGBs: fFe, SAGB ≈ 0.2 and fAl, SAGB ≈ 1. In spite of the number of considerations in determining these quantities, there are still others that could be included, namely a velocity dependence on the filtering by the heliopause. We would expect dust grains with a sufficiently high velocity (i.e., vgrain > vesc, ☉) to ignore size filtering limitations, but including these effects will be left for a future work.

5. LIVE RADIOISOTOPE DATA

5.1. Terrestrial Measurements of 60Fe

The primary data value for our analysis is the decay-corrected 60Fe fluence measured by Knie et al. (2004) in a deep-ocean Fe-Mn crust. They found an isotopic ratio of ${{}^{60}{\rm Fe}}/{\rm Fe} = 1.9 \times 10^{-15}$ within the crust; this corresponds to a decay-corrected fluence of ${\cal F}_{{\rm arr},60} = (2.9 \pm 1.0) \times 10^{6} \ \rm atoms\,cm^{-2}$. This may be used to determine the distance from the progenitor. At the time of the original measurements, the half-life of 60Fe was estimated to be 1.49 Myr and that of 10Be was estimated to be 1.51 Myr, resulting in an arrival time, tarr = 2.8 Myr ago. Since then, the half-lives have been refined: current best estimates being $\tau _{1/2, {}^{60}{\rm Fe}}$ = 2.62 Myr and $\tau _{1/2, {}^{10}{\rm Be}} = 1.387$ Myr, and this places the signal arrival at 2.2 Myr ago. To update the Knie 60Fe fluence, we use ratios to convert the previous results to the updated values that are similar in method to those employed by Bishop & Egli (2011) and Feige et al. (2012): ${\cal F}_{\rm 60, update}/{\cal F}_{\rm 60, previous} = e^{-t_{\rm 60, update}/\tau _{\rm 60, update}}/e^{-t_{\rm 60, previous}/\tau _{\rm 60, previous}}$, which gives the following updated, decay-corrected fluence in the crust.

Equation (20)

The 60Fe measurement in the crust has been verified by Fitoussi et al. (2008), but within this same work, no comparable 60Fe signal was detected in a sea sediment sample. Fitoussi et al. (2008) suggested several reasons for the non-detection in the sediment, including differences in uptake and divergences in the sediment from the global background. In addition, we note that the fluence calculation assumes an even distribution of dust over the Earth's surface. However, the Earth's wind patterns are not uniform, nor is Earth's precessional axis necessarily orthogonal to the progenitor's position. Consider, for example, a spherical dust grain of radius 0.2 μm falling at terminal velocity (∼0.1 m s−1) through a 1500 m-thick jet stream flowing horizontally at 100 km hr−1 with a density of 4 × 10−4 g cm−3 (the assumption of falling at terminal velocity should be valid as the Earth's atmosphere will have dissipated most of the dust grain's remaining interstellar kinetic energy). As the dust grain falls, the pressure from the jet stream will quickly accelerate the grain horizontally to the same velocity and push the grain ∼300 km before it falls out of the jet stream. In view of the non-uniformity of the jet stream's flow as well as other terrestrial winds, anisotropies in the observed fluences are expected. Furthermore, one should also consider the source's orientation to the Earth's precessional axis; the Fe-Mn crust used by Knie et al. (2004) is from 9°18' N, 146°03' W (∼1, 000 mi/1,600 km SE of Hawaii), and the sediment used by Fitoussi et al. (2008) is from 66°56farcm5 N, 6°27farcm0 W (∼250 mi/400 km NW of Iceland). The crust sample's location relative to the equator would make it more likely to receive a signal over a range of arrival angles while the northern hemisphere could be partially shielded from a more southerly progenitor. Rather than the Fe-Mn crust signal being due to a misinterpretation of the global background, the absence of a sediment signal could be due to the geometry of the source's position. If this is the case, a sediment sample from the southern hemisphere (e.g., ELT49-53, 38°58farcm7 S, 103°43farcm1 E and ELT45-21, 37°51farcm6 S, 100°01farcm7 E used by Feige et al. 2013) should have an 60Fe signal.

5.2. Lunar Measurements of 60Fe

In addition to sea sediment and Fe-Mn crusts, lunar surface (regolith) samples can also be used to search for a nearby progenitor signal. The lunar surface is not affected by wind or water erosion, but, as pointed out by Feige et al. (2013), the sedimentation rate is low (precluding the possibility of time-resolved measurements) and regular impacts by a range of impactors (Langevin & Arnold 1977), continually churn up the regolith, mixing different levels. Lunar samples would be better suited to providing a "first hint" of a signature (Feige et al. 2013). Apollo core samples were analyzed by Cook et al. (2009), Fimiani et al. (2012), and Fimiani et al. (2014); in particular, these authors found both the Apollo 12 sample 12025 and Apollo 15 sample 15008 to have an 60Fe signal above the background. Nishiizumi et al. (1979) and Nishiizumi et al. (1990) found that the Apollo 12 and 15 cores, respectively, showed little to no large-scale mixing just prior to, during, and/or since the potential arrival of the signal, meaning no large impactor could have ejected part of the regolith thus diluting the signal. However, as we show in the next section, another issue arises with the dust's arrival at the Moon's surface that can dilute the signature.

5.3. Lunar Regolith and Dust Grains: Vaporization

As noted above, there are putative detections of a non-meteoric 60Fe lunar anomaly, which seem to verify the presence of the deep-ocean signal; an amazing confirmation. However, having argued that the 60Fe will arrive in the form of high-velocity dust grains, we now consider the implications for the deposition onto the lunar surface.

Cintala (1992) made a detailed study of impacts on the lunar regolith and found semi-empirical fits to the volume of vapor produced as a function of impactor size and velocity and of the target composition. It was found that impactor velocity is the dominant factor and that, for vgrain > 100 km s−1, the volume of target material that is vaporized is ∼10–100 times the volume of the impactor itself. More recently, Cremonese et al. (2013) showed that micrometeor impacts can be a non-negligible source of the lunar vapor atmosphere. They use the Cintala (1992) model and find that the contribution may be 8% of the photo-stimulated desorption at the subsolar point, becoming similar in the dawn and dusk regions and dominant on the night side. Moreover, Collette et al. (2014) did laboratory experiments to simulate micrometeorite impacts, studying the neutral gas created as a result. They find that the number of neutrals produced per unit impactor mass scales as ${\sim } v_{\rm grain}^{2.4}$, and they conclude that complete vaporization is expected for speeds exceeding 20 km s−1.

With these results in mind, we now consider the conditions surrounding the arrival of the dust signal in question at the lunar surface. After entering the solar system, the dust grains continue to the Earth/Moon at essentially the same speed. However, whereas the Earth's atmosphere slows the arriving dust grains prior to reaching the surface, the Moon's tenuous atmosphere has practically no influence, and the dust grains' velocities are unchanged before arrival at the lunar surface. The dust grains are estimated to be ∼0.2 μm in size, and moving at ∼20–100 km s−1 depending on the progenitor's distance. Thus the grains behave as "micrometeorites" moving at very high speeds similar to those examined above.

Consider a silicate dust grain (ρgrain ∼ 3.5 g cm−3) impacting the lunar surface (ρregolith ∼ 1.6 g cm−3) at vgrain ∼ 20 km s−1. The grain arrives with kinetic energy $E_{\rm grain} = m_{\rm grain}v_{\rm grain}^2/2 \sim 0.23 \ \rm erg$, where m = 4πρgraina3/3 is the mass of a spherical grain of radius a. The dust grain will penetrate the regolith to a depth comparable to its diameter and will vaporize some of the surrounding material; we will assume Vvapor = 10 Vgrainmvapor ≈ 4.6 mgrain. Given the grain's high speed and shallow penetration, the vaporization will happen quickly (i.e., very little expansion occurs before the entire mass is vaporized). Moreover, since the initial density of the vaporized regolith is much greater than the density of the Moon's atmosphere, the gas will behave as if it is expanding isentropically into a vacuum. The grain's kinetic energy will go into vaporizing the grain and regolith and into the thermal and kinetic energy of the resulting gas. To determine the vaporization energy, the standard enthalpy of formation for the lunar regolith is ∼1.5 × 1011 erg g−1 = (3.9 km s−1)2. This is an approximate value for both of the lunar regolith's main constituents, silica, SiO2, and aluminum oxide, Al2O3, (Nava & Philpotts 1973), and includes both the vaporization and dissociation energies for the molecules. Therefore, the total energy consumed in vaporization is 0.1 erg, leaving 0.13 erg for the thermal and kinetic energy, Evapor, of the gas.

As the gas expands, the thermal portion vanishes asymptotically, with all the energy becoming kinetic. Thus, after this cooling, the asymptotic expansion speed of the vaporized material is:

Equation (21)

[For further discussion, see Zel'dovich & Raizer (1967, p. 101–104, 844–846)].

The vapor speed is much larger than the lunar escape velocity, vesc, Moon = 2.4 km s−1. This suggests that much of the vaporized material, including the dust impactor with its 60Fe material, would escape from the Moon. This would imply that the Moon has an uptake factor UMoon ≪ 1. Thus, we should not be surprised that the lunar results for 60Fe are lower than expected naively from the terrestrial Fe-Mn crust results. While lunar samples confirm the signal found in the Fe-Mn crust, they are less suitable for determining the fluence given the difficulties in determining UMoon.

5.4. 244Pu Measurements

Several searches for live 244Pu have been performed, beginning with Wallner et al. (2000) looking at Fe-Mn nodules. Studies of top layer sea sediment by Paul et al. (2001); Paul et al. (2003), and Paul et al. (2007) have shown there is a very low background in 244Pu, making 244Pu an excellent candidate to confirm an 60Fe signal from an extra-solar source (presumably from a CCSN). Wallner et al. (2000, 2004) reported the detection of a single 244Pu atom in the same Fe-Mn crust sample used by Knie et al. (1999) and Knie et al. (2004) covering the entire time interval of 1–14 Myr. Separately, Raisbeck et al. (2007) looked in sea sediment for a 244Pu signal, but did not find any evidence for a signal. It should be noted, however, that the Raisbeck study was using the previous arrival time (2.8 Myr ago) for his search. The samples were dated using magnetic polarization analysis and did not cover a large enough time interval to include the appropriate dating interval using the new value for the 10Be lifetime (Meynadier et al. 1994).

5.5. 26Al Measurements

Feige et al. (2013) reported on searches for 10Be and 26Al using ∼3 kyr time intervals in samples from sea sediments ELT 49-53 and ELT 45-21. In the case of 26Al, the measurements showed only variations consistent with fluctuations around the background level, and found no evidence for an extra-solar signal. The paper also reported that 53Mn measurements are planned.

6. RESULTS

In Figure 2 we compare our model predictions with the 60Fe data of Knie et al. (2004) and Fitoussi et al. (2008), showing that the model matches the results within the uncertainties for an SN or SAGB occurring 2.2 Myr ago. We note that the sampling was continuous through the entire data range, and straddled the signal arrival. In addition, the value for the 880-kyr time resolution was less than the 440-kyr sample, as expected due to the additional stable Fe in the wider sample.

Figure 2.

Figure 2. Evidence from Knie et al. (2004) and Fitoussi et al. (2008) for an anomalous peak in the 60Fe isotope fraction ∼2.2 Myr ago, compared with simulations of a possible signal from an SN explosion. We plot the results using ECSN yields; other progenitors yield similar results.

Standard image High-resolution image

Using the decay-corrected Knie et al. (2004) fluence of 60Fe (Section 5.1), and 60Fe yields from various source candidates (Section 2), we have solved Equation (1) for the distance to the source. Distances and other parameters for some of the possible sources appear in Table 3 and Figure 3. We see that, for sources at distances ∼100 pc that are typical of our subsequent estimated distances, the en route time and the signal width are ${\cal O}$(Myr), so it is possible that the signal could be time-resolved in future measurements, and thus it is of interest to model the signal shape.

Figure 3.

Figure 3. Estimated distances for possible progenitors, for UFe = 0.5. SN candidates are circles and SAGB candidates are squares. The solid error bars represent uncertainty in the fluence measurement (Knie et al. 2004). The dashed error bars represent additional uncertainty in 60Fe yields due to nuclear reaction rates in SNe (Tur et al. 2010) and a delayed super-wind phase in SAGBs (Doherty et al. 2013). Of particular note are the TNSN/Type Ia SN and the KN/NS–NS merger models, which are too close to have produced the detected 60Fe signal.

Standard image High-resolution image

6.1. Core-Collapse and Electron-Capture Supernovae

Figure 3 shows the calculated distances for our examined CCSNe and ECSN; they range from ∼60–130 pc. All CCSNe from our set lie outside of the kill distance and within the fadeaway distance for both their average fluence values and errors. Similarly, the ECSN lies outside the kill distance and within the fadeaway distance (the ECSN kill and fadeaway distances are shorter due to its lower explosive energy). The ECSN upper error is outside the fadeaway distance, but because SN dust can still travel great distances after decoupling, this is not an absolute limitation. Based on these distances, either a CCSN or an ECSN could have produced the measured 60Fe signal.

6.2. Thermonuclear Supernovae

TNSN produce so little 60Fe that it would require a TNSN to have been at a distance of ∼0.6 pc in order to produce the signal measured by Knie et al. (2004). This is an implausibly short distance, and any uncertainty in the fluence measurement would not change this determination. At that range, the TNSN would have killed nearly all life on Earth, so we can exclude a TNSN as the source of the 60Fe signal (in this case, the descreening kill distance for a TNSN is ∼10 pc and the ionizing radiation kill distance from 1048 erg of γ-rays is ∼20 pc, Smith et al. 2004). Adopting the largest yield ($M_{\rm ej,{}^{60}{\rm Fe}}\sim 10^{-7} \, M_\odot$) from Seitenzahl et al. (2013) extends the distance to ∼6 pc, which is still inside the kill radius and does not change this conclusion.

6.3. Kilonovae

Our calculations give a possible KN distance of ∼5 pc. Of the little that is known observationally or even theoretically about KNe, we are unaware of any estimates of their ionizing radiation output. In addition, the strength and shape of the shock from ejected material is highly dependent on the orientation of the merger. Thus, we are unable to estimate the corresponding kill distance either by direct exposure or descreening. The ejecta from KNe are certainly energetic (explosive velocities ∼0.3 c, Goriely et al. 2011), and one might imagine decompressing neutron star matter initially emitting in the UV or at shorter wavelengths. However, the observed radiation for the KN candidate associated with GRB 130603B is very red at times ≳ 8 hr (Berger et al. 2013). Moreover, while the KN shock and radiation is expected to be much more isotropic than the GRB, more study of the geometry of the resulting blast is needed to determine a definitive kill distance like that used for TNSN. Consequently, a biohazard argument cannot rule out a KN explosion as the source of the 60Fe anomaly.

However, a much better discriminator for a KN source would be the 244Pu/60Fe ratio. The single 244Pu atom detected by Wallner et al. (2000, 2004) yields a surface fluence of 3  ×  104 atoms cm−2 for the period 1–14 Myr ago. Looking at the yields from Goriely et al. (2011) again, we can infer the yield for A = 244 should be at least on the order of the yield for A = 60 (i.e., $({{}^{244}{\rm Pu}}/{{}^{60}{\rm Fe}})_{\rm KN} \,{\ge}\, 1$).10 Based on this assumption and the surface fluence for 60Fe during the signal passing (1.41  ×  106 atoms cm−2/0.5  =  3  ×  106 atoms cm−2), then

even though 244Pu was measured over 10 times the time period as 60Fe (Note: this assumes the dust fraction for Pu is the same as Fe, fPufFe).

Additionally, KN occur infrequently (∼10 Myr−1 galaxy−1, Goriely et al. 2011) compared to CCSN & ECSN (∼30 kyr−1 galaxy−1).11 If we approximate the Milky Way as a thin cylinder of radius 10 kpc and thickness 200 pc, the rates (Γ) of a KN occurring within ∼5 pc or a CCSN/ECSN occurring within ∼75 pc of the Earth are:

After inverting these quantities, we can expect a nearby SN every ∼1 Myr compared to a nearby KN every ∼107 Myr ≫ 1/H0 (the Hubble time). This makes a KN an unlikely source for the 60Fe signal. However, this result should be revisited as specific yields for 60Fe become available and especially if a signal from strong r-process isotopes is detected (e.g., 146Sm, 182Hf, and 244Pu).

6.4. Super-AGB Stars

Figure 3 plots three of our six examined SAGBs (all are listed in Table 3). Their distances range ∼60–110 pc; similar to those of CCSNe. With errors, all SAGBs lie well within the distance for dust stopping due to drag (∼90 kpc) but well outside the magnetic deflection distance (∼1 pc). While it is tempting to rule out SAGBs as a source under the assumption that any dust would be quickly deflected, we have decided not to rule out SAGBs (see, e.g., Frisch 1995; Cox & Helenius 2003; Florinski et al. 2004; Frisch et al. 2012) since there is uncertainty in the strength, direction, and uniformity of the Local Bubble's magnetic field. Depending on the nature of the Local Bubble's magnetic field, charged dust particles could travel with very little deflection. Instead, we will be examining an alternate search in a future work.

6.5. 26Al Results

In Figure 4, we plot 26Al predictions for various progenitors, and the expected background in the accelerator mass spectrometry (AMS) data from Feige et al. (2013) using the same 3-kyr sampling intervals used in this experiment. Because fAl, SN ≈ 0, we do not expect any signal to be present if the source was an SN. The calculated signal from a 15-M CCSN with fAl, SN = 1 is plotted simply as an example if the dust fraction was significantly higher. Additionally, while we would expect some Al from SAGBs to reach Earth, it would not be visible above the variations in the 26Al background.

Figure 4.

Figure 4. Model predictions compared with the 26Al AMS data from Feige et al. (2013). Note that the 15-M SN is an example only and included to demonstrate the consequence if fSN, Al = 1 instead of the fSN, Al ∼ 0 we expect.

Standard image High-resolution image

6.6. 53Mn Results

In anticipation of AMS 53Mn measurements mentioned by Feige et al. (2013), in Figure 5 we plot predictions for 53Mn based on the distance determined by the 60Fe fluence. Since the survival and grain size for Mn from an SN has not been described to our knowledge, we plotted a range of SN progenitors (since SAGBs are not expected to produce 53Mn) and dust fractions, using the largest possible signal source (21-M CCSN), a mid-range source (15-M CCSN), and the lowest source (ECSN). We varied the dust fraction from an order of magnitude above to an order of magnitude below fSN, Fe. As can be seen for the 15- and 21-M CCSN with fSN, MnfSN, Fe, a signal should be readily detectable the given the AMS detection threshold of ∼10−15 53Mn/55Mn (Poutivtsev et al. 2010). However, for the ECSN and most CCSNe with fSN, Mn < fSN, Fe (even a 21-M case could be difficult to detect depending on the fluctuations in the 53Mn/55Mn background), it should be improbable for an SN progenitor to be detected with 53Mn.

Figure 5.

Figure 5. Model predictions for upcoming 53Mn AMS measurements. The vertical axis has been broken into three parts in order to show the peak values of each configuration. Note that the average background 53Mn/55Mn level (Feige et al. 2013) is shown ahead of the SN's arrival. With an AMS sensitivity of ∼10−15 53Mn/55Mn (Poutivtsev et al. 2010), progenitors such as a 21-M SN should be detectable across a range of fSN, Mn values, whereas an ECSN progenitor should not be detectable.

Standard image High-resolution image

7. DISCUSSION AND CONCLUSIONS

Since the discovery and confirmation of the terrestrial 60Fe signal by Knie et al. (1999) and Knie et al. (2004), several experiments have tried to find a corroborating signal either in lunar samples or with other isotopes. To date, none of these experiments provided a definitive signal on the order of that originally reported. This paper attempts to provide a context for these observations and seek other possible progenitors besides a CCSN whose properties could be consistent with the observations. We also anticipate future observations with a hope for time-resolved signals.

From our list of candidates, we can rule out a TNSN as it would be too close (∼0.6 pc) to both create the 60Fe signal and not to kill most life on Earth. We also rule out a KN as a potential source. The KN would have been ∼5 pc away from the Earth, and while more study of the geometry of a KN is required to determine a definitive kill distance, the low amount of 244Pu (a strong r-process element) detected to date contrasted with the high number of r-process elements per merger makes a KN a low probability. Additionally, KNe/Neutron Star Mergers are very rare, making it unlikely for the solar system to have passed within 5 pc of one.

Although SAGB stars are outside the magnetic field deflection distance, we have decided not to rule out SAGB stars based solely on this stipulation. Depending on the strength, direction, and uniformity of the Local Bubble's magnetic field and the charge on the dust grains, it may be possible for an SAGB to have produced the measured 60Fe signal. Since an SAGB would likely have evolved to the white dwarf stage by now, we plan to investigate this possibility in a future work.

All variations of CCSNe and ECSN remain possible sources. Of these, ECSN would be the most likely, firstly, because they arise from the lowest-mass and thus most common core-collapse progenitors. Additionally, Fuchs et al. (2006) listed members of the Sco-Cen association and their masses (using their listed magnitudes and mass-to-magnitude relation) which included the range: MSco-Cen = 2.5–8.2 M, compared to MECSN = 8–10 M. Since more massive stars evolve faster than lower-mass stars, it is reasonable to expect the signal progenitor to be near the upper end of the mass range. Lastly, the continued lack of a definitive 244Pu signal, in spite of multiple attempts, is also consistent with the possibility of an ECSN as the progenitor due to its lack of strong r-process products.

Several caveats are important to bear in mind. Probably most importantly, our ability to test different explosion candidates is only as good as the radioactive yield predictions. These challenging calculations are continually improving, but are subject to significant uncertainties, including stellar evolution, hydrodynamics, and nuclear physics. Indeed, two key nuclear cross sections alone can lead to 60Fe and 26Al yield variations by factors up to ∼10 (Tur et al. 2010). This alone suggests that all of the CCSN candidates should be revisited as yields improve. Seitenzahl et al. (2013) showed that TNSN yields are sensitive to the number of initiation sites and the transition from deflagration to detonation. Clearly, improved radioisotope yield calculations for any of our explosive sources could dramatically change the landscape of possible scenarios. Thus, we implore future nucleosynthesis studies to include (at least) 60Fe and the other radioisotopes we have discussed here.

Other important uncertainties similarly invite future work. As we have seen, Fe uptake in Fe-Mn crusts represent another topic that invites future study. Uptake has a dramatic impact on our results: the inferred distance to the explosion scales as $D \propto U_{i}^{1/2}$. Additionally, the local interstellar density and magnetic field plays a key role in the propagation of the signal (whether from an SN or SAGB) and in the duration of the time profile of the radioisotope flux. Finally, as we have seen, the observability of different radioisotopes is highly sensitive on the formation and survival of SN dust of different compositions and sizes. We have relied on theoretical calculations (Silvia et al. 2010, 2012) which imply, among other things, that 26Al is unlikely to be observable terrestrially despite its SN abundance comparable to 60Fe. Further such theoretical studies relevant to other radioisotopes, and observational corroboration, are critically needed.

To confirm the origin of the 60Fe signal and pin down its source will require measurements of 60Fe at other sites, other sources (e.g., magnetosomes in addition to crust and sediment), and other radioisotopes. Lunar regolith measurements provide unique confirmation of the terrestrial 60Fe signal. However, we find absolute measurements will be difficult because high-velocity dust vaporizes on the lunar surface and much of the incident material will then escape the Moon. This said, we eagerly await detailed presentation of lunar measurements hinted at by Fimiani et al. (2014). We are also looking forward to 53Mn measurements as mentioned by Feige et al. (2013) and 244Pu measurements by Piran et al. (2014, see Wallner et al. 2014, citation therein) that will be helpful in discriminating between the remaining possible progenitors.

Looking ahead to further measurements, the behavior of dust condensation, survival, and filtering will be a key factor in narrowing the remaining pool of possible progenitors. The dust fraction includes several filtering processes that can all affect the resulting distance calculation ($D \propto f_{i}^{1/2}$). Studies of dust formation have focused on silicates, iron, and corundum, but formation processes with other elements (especially Mn, Ca, Ti, Zr, Tc, and Pu) could be used to differentiate the remaining possibilities given the varying yield ratios between these elements for each progenitor. The search for other isotopes is not simply a matter of choosing those with high lifetimes ($\tau _{i} \sim {\cal O}({\rm Myr})$), but also those with low backgrounds, high condensation temperatures (in order to form dust grains), and large grain sizes (a ≳ 0.2 μm). Of particular interest would be 41Ca and 53Mn. While perhaps not ideal candidates with regards to background levels, they have long lifetimes and can be condensed at high temperatures (≳ 1100 K) into Perovskite (CaTiO3), Melilite (Ca2Al2SiO7, Ca2MgSi2O7), and Alabandite (MnS) (see Field 1975). Other possible isotopes with long lifetimes such as 93Zr, 97Tc, 99Tc, and 107Pd, as well as strong r-process elements such as 146Sm, 182Hf, and 244Pu could be used to constrain CCSNe if more details of their dust condensation are determined, but, regardless, any other candidate isotope would need to form grains large enough to survive escape from its progenitor and enter the solar system. It would be a remarkable coincidence if the only isotope that is capable of carrying an extra-solar signal (i.e., 60Fe) is the first one examined.

With observations of additional isotopes, it is possible not only to identify a specific event or progenitor, but also to: (1) provide a better measure of the distance to the source, (2) directly probe individual radioisotope nucleosynthesis, (3) constrain the nearby SN rate, (4) guide astrophysical searches for the SN remains (i.e., pulsars), and (5) model the explosion light curve and to assess the possible damage to the terrestrial biosphere. Finally, we have seen that the measurement of time-resolved radioisotope profiles provides direct information of the blast passage through the solar system and an independent measurement of the distance to the progenitor. The authors are optimistic that new data will make such questions tractable in the near future.

Since our first submission, we have been made aware of the superb thesis by Jenny Feige that covers many similar topics and was completed independently of this work (Feige 2010). We are pleased to acknowledge the Vera Laboratory including Jenny Feige, as well as the Technische Universität München (TUM) Group including Shawn Bishop. We also thank our reviewer whose thoughtful and thorough comments on the manuscript greatly improved this work. We are grateful to Thomas Johnson and Craig Lundstrom for illuminating discussions of isotope geology. The work of John Ellis was supported partly by the London Centre for Terauniverse Studies (LCTS), using funding from the European Research Council via the Advanced Investigator Grant 267352. Brian Fields thanks Stuart Shapiro for his enlightening discussions about the Sedov solution and Friedrich-Karl Thielemann and Ivo Seitenzahl for their informative discussions of supernova radionuclide synthesis.

APPENDIX A: LIST OF VARIABLES

Variable - Description [common value or unit of measure]

a - radius of dust grain [μm]

A - atomic number [dimensionless]

α - signal width parameter [≈0.577]

β - ratio of Sun's radiation force to gravitational force on a particle [dimensionless]

c - speed of light [∼3 × 105 km s−1]

cs - speed of sound in ISM [km s−1]

Cr - constant from combination of solar flux, gravitational constant, among others [7.6 × 10−5 g cm−2]

D - distance from progenitor to Earth [pc]

δ - shell thickness of uniform shell for an SN remnant [dimensionless]

ECCSN - energy deposited into ejecta by a CCSN [∼1051 erg]

EECSN - energy deposited into ejecta by an ECSN [∼1050 erg]

EKN - energy deposited into ejecta by a KN [∼1049 erg]

ETNSN - energy deposited into ejecta by a TNSN [∼1051 erg]

Egrain - kinetic energy of dust grain [erg]

Evapor - thermal and kinetic energy in vapor [erg]

epsilon - shell thickness of saw-tooth shell [dimensionless]

f - dust fraction, fraction of isotope that passes from progenitor to Earth [dimensionless]

Fdrag - drag force on dust grain [dyne]

Fg - force of gravity on dust grain [dyne]

Fmag - force of magnetic field on dust grain [dyne]

Fr - force from solar radiation pressure on dust grain [dyne]

${\cal F}_{{\rm arr}, i}$ - decay-corrected (or arrival) fluence of an isotope [atoms cm−2]

${\cal F}_{{\rm interstellar}, i}$ - total fluence of an isotope across spherical signal front [atoms cm−2]

${\cal F}_{{\rm obs}, i}$ - observed fluence of an isotope [atoms cm−2]

${\cal F}_{{\rm surface}, i}$ - total fluence of an isotope regardless of uptake [atoms cm−2]

$\mathbb {F}$ - material flux, fluence per time [atoms cm−2 kyr−1]

G(x) - density profile function [dimensionless]

G0(s) - collisional drag function [dimensionless]

γ - ratio of specific heats [dimensionless]

Γ - progenitor rate [Myr−1]

H0 - Hubble Constant [∼70 km s−1 Mpc−1]

i - (as subscript) 'for a given isotope' (e.g., 60Fe, 26Al, etc.)

k - Boltzmann constant [1.38 × 10−16 erg K−1]

m - mass [g]

M - mass of progenitor [M]

Mej, i - total mass of an isotope in the ejecta [M]

mu - atomic mass unit [∼1.66 × 10−24 g]

n - number density (e.g., of ISM, dust grain, etc.) [cm−3]

N - number of dust grains [dimensionless]

${\cal N}_{i}$ - number of atoms of an isotope [dimensionless]

PSW - pressure of solar wind [dyne cm−2]

Qpr - efficiency of solar radiation on dust grain [dimensionless]

Rdrag - distance at which drag effects are significant on a dust grain [pc]

Rfade - distance at which an SN shock transitions into a sound wave [pc]

Rkill - distance from the Sun an SN progenitor can produce a shock that penetrates to Earth's orbit [pc]

Rmag - distance at which magnetic deflection effects are significant on a dust grain [pc]

RSAGB - radius of leading edge of SAGB dust shell [pc]

RSN - radius of leading edge of SN remnant [pc]

ρ - mass density (e.g., of ISM, dust grain, etc.) [g cm−3]

ρ0 - density in front of shock [g cm−3]

ρ1 - density behind shock [g cm−3]

s - velocity parameter [dimensionless]

t - elapsed time [Myr]

tarr - time from today in the past that the leading edge of the signal arrived [Myr]

tdep - time from today in the past that the trailing edge of the signal departed [Myr]

ttravel - time for isotope to transit from progenitor to Earth [Myr]

T - temperature of ISM [K]

TC - condensation temperature [K]

Δtinter - delay between envelope pulses of SAGB [∼100 yr]

Δtpulse - duration of envelope pulse of SAGB [∼1 yr]

Δtres - time resolution of samples [kyr]

ΔtSAGB - duration of SAGB phase [∼100 kyr]

Δtsignal - signal width, time for signal to pass Earth and duration of ejecta deposition on Earth [kyr]

τi - mean lifetime of an isotope, τi = τ1/2, i/ln 2 [Myr]

τ1/2, i - half-life of an isotope [Myr]

Ui - uptake, fraction of deposited isotope that is incorporated into sampled material [dimensionless]

v - speed [km s−1]

varr - velocity of a dust grain upon arrival at Earth [km s−1]

vesc - escape velocity [km s−1]

vgrain - speed of dust grain/impactor [km s−1]

vgrain, 0 - initial speed of dust grain [km s−1]

v - speed at infinity [km s−1]

vSN - speed of leading edge of the SN remnant [km s−1]

V - volume [cm3]

${\cal V}$ - voltage of dust grain [V]

X - mass fraction [dimensionless]

ξ0 - SN proportionality constant [dimensionless] (Zel'dovich & Raizer 1967)

ζ0 - SAGB proportionality constant [dimensionless]

APPENDIX B: BLAST EXPANSION AND RADIOISOTOPE FLUX PROFILE

We model astrophysical explosions as spherically symmetric, and we are interested in distances sufficiently large that the swept up interstellar mass is much larger than the ejecta mass. We treat a blast wave as adiabatic (energy-conserving) and thus adopt the Sedov–Taylor solution. The Sedov blast wave evolves in a self-similar manner. This means in particular that gas properties as a function of radius r maintain the same shape when plotted in terms of the similarity variable

where the shock radius at t is given by Equation (2). In particular, the density profile is

where the density immediately behind the shock is:

and the dimensionless density profile function is thus normalized to G(1) = 1. Note that mass conservation implies that the total mass $M_{{\rm swept}}=4\pi \rho _0 R_{\rm SN}^3/3$ swept up in the blast is equal to the total mass in the blast profile

and so setting Mtotal = Mswept implies that

Equation (B1)

We consider two approximations to the full Sedov profile. For a uniform shell approximation, we have G(x) = 1 for x ∈ [1 − δ, 1] and zero otherwise, which gives the location of the inner shell radius via

and thus

whereas going to first order in δ we would find δ = 1/12. For a "saw-tooth" approximation, the blast material is in a thin shell with a profile that linearly decreases from a maximum behind the shock to zero at coordinate x0 ≡ 1 − epsilon. Thus we have G(x) = Ax + B, with the constraints that G(1) = 1 and G(x0) = 0 at the inner radius, which gives

Equation (B2)

Choosing γ = 5/3, to first order we find $\int _0^1 x^2 G(x)\ dx = \int _{1-\epsilon }^1 x^2 G(x)\ dx \approx \epsilon /2$. From Equation (B1) we find the dimensionless shell thickness epsilon ≈ 1/6, twice the value in the uniform shell. As seen in Figure 6, the saw-tooth density profile more closely matches the exact Sedov density profile compared to the uniform shell profile, making the saw-tooth profile more appropriate for modeling the signal spreading for our SN distances.

Figure 6.

Figure 6. Comparison of the uniform shell and saw-tooth profiles to exact solution profile, for γ = 5/3⇒δ ≈ 0.083 (uniform shell), epsilon ≈ 0.17 (saw-tooth). In addition, the paper used a ratio to approximate the observed SN crossing velocity. The right chart compares this approximation to the exact solution profile.

Standard image High-resolution image

The (radial) velocity profile is

with $\dot{R}_{\rm s}$ the shock speed, and the dimensionless velocity profile function normalized to U(1) = 1. To a good approximation, the velocity is linear, and we will adopt the approximation U(x) ≈ x. This leads to a "Hubble law" relation

Figure 6 compares this linear velocity profile with the exact Sedov solution. Our approximation is necessary to maintain the self-similarity of the saw-tooth profile, and, while different than the exact solution, should be sufficient for the region we are most interested in (0.8 ⩽ r/R ⩽ 1). For a more detailed description of the analytical Sedov solution, see Book (1994).

Turning to the explosive ejecta, we note that if a number ${\cal N}_i$ of atoms of species i were distributed with uniform density at time t, then the mean number density in i would be:

We will assume that, at times of interest, the ejecta is well-mixed into the blast wave, with a constant mass fraction at all radii. That is, we assume that the ejecta density profile follows that of the blast itself. This means that the highest ejecta density is just behind the shock, with a value:

and the ejecta density profile is:

Combining the ejecta density profile with the "Hubble law" velocity approximation gives the global-averaged ejecta flux onto the surface of the Earth (i.e., 1/4 the interstellar flux, not including radioactive decay), evaluated at distance r = D:

Equation (B3)

with the time-independent prefactor:

Equation (B4)

We see here explicitly that a time-resolved flux directly encodes the blast density profile and thus probes the propagation of the radioisotope ejecta from explosion to Earth.

Using our saw-tooth approximation for the blast density profile G (Equation (B2)), and using the Sedov result $\dot{R}_{\rm s}/R_{\rm SN}= 2/5t$, we find a flux profile in time of:

Equation (B5)

We note that the leading edge of the blast from an event at distance D arrives at a time ti given by D = RSN(tarr). Thus we can recast D/RSN(t) = (t/tarr)−2/5 in terms of the initial arrival time. The trailing edge of the shell arrives at time tdep given by D = (1 − epsilon)RSN(tdep). Thus we have tdep = tarr/(1 − epsilon)5/2. This means that we can write epsilon = 1 − (tarr/tdep)2/5, and we can express the global-averaged flux time profile as:

Equation (B6)

This is the sum of two power laws in t, leading to a steep cusp at early times ttarr that flattens at late times ttdep.

Footnotes

  • We adopt a fiducial TNSN 60Fe yield of ∼10−9M, which is consistent with the classic W7 result (Nomoto et al. 1984) and is larger than almost all Seitenzahl et al. (2013) models.

  • It is possible to have Ui > 1 if, for example, a marine sample can chemically scavenge the element of interest so efficiently that it collects more than the amount deposited in the water column directly over the sample.

  • Note, our kill distance does not include the case of a gamma-ray burst given their narrowly beamed emission.

  • The Coulomb force can be large, however, it is not for our selected grain parameters, so we neglect Coulomb forces (see constraints for (ϕ2ln Λ) in Draine 2011, Section 26.1.1).

  • This "radioactivity distance" is analogous to the usual luminosity distance: the yield plays the role of luminosity, and radioisotope fluence the role of flux (Looney et al. 2006).

  • In Equation (2) time increases toward the future, but note that elsewhere, unless explicitly stated, times are used in the geological sense, and thus increase toward the past.

  • 10 

    More likely, A = 244 yields are 10–100 times larger than A = 60 yields given the A  ∼  240 yields and the fact that the fission recycling sources are centered around A  ≃  280–290 region, Goriely et al. (2011).

  • 11 

    We would like to thank the reviewer for suggesting this addition to the KNe discussion.

Please wait… references are loading.
10.1088/0004-637X/800/1/71