<<

Draft version August 15, 2017 Preprint typeset using LATEX style AASTeX6 v. 1.0 DRAFT FROM August 15, 2017

PRE-NEBULAR LIGHT CURVES OF TYPE I SUPERNOVAE

W. David Arnett1, Christopher Fryer2, and Thomas Matheson3

1Steward Observatory, University of Arizona, 933 N. Cherry Avenue, Tucson AZ 85721 2Los Alamos National Laboratory, Los Alamos NM 3National Optical Astronomy Observatory, Tucson AZ

ABSTRACT We compare analytic predictions of light curves with recent high qual- ity data from SN2011fe (Ia), from KSN2011b (Ia), and the Palomar Transient Factory (PTF) and the La Silla-QUEST variability survey (LSQ) (Ia). Because of the steady, fast cadence of observations, KSN2011b provides unique new in- formation on SNe Ia: the smoothness of the light curve, which is consistent with significant large-scale mixing during the explosion, possibly due to 3D effects (e.g., Rayleigh-Taylor instabilities), and provides support for a slowly-varying leakage (mean opacity). For a more complex light curve (SN2008D, SNIb), we separate the luminosity due to multiple causes and indicate the possibility of a radioactive plume. The early rise in luminosity is shown to be affected by the opacity (leakage rate) for thermal and non-thermal radiation. A general deriva- tion of Arnett’s rule again shows that it depends upon all processes heating the plasma, not just radioactive ones, so that SNe Ia will differ from SNe Ibc if the latter have multiple heating processes.

1. INTRODUCTION issues of frequency-dependent atmospheric Supernovae whose luminosity is primarily physics. This approach, which assumes powered by the decay chains 56Ni(e−, ν)56Co, three–dimensional (3D) incomplete mixing 56Co(e−, ν)56Fe and 56Co(, e+ν)56Fe have during the explosion (Arnett & Meakin generic features. These are prominent in SNe 2016), is thus different from and a natural Ia, and appear in SNe Ibc, that is, in both complement to a 1D, time–dependent stellar thermonuclear supernovae and those core- atmosphere approach such as Blondin, et al. collapse supernovae which have lost their hy- (2013); Dessart, et al.(2015, 2016). drogen envelopes (e.g., Arnett(1982, 1996)). In §2 we introduce the combined problem We compare theoretical light curves (Ar- of parameters and uniqueness, which may be nett 1982; Pinto & Eastman 2000a,b) to more clearly discused in our analytic frame- the best-observed typical SN Ia to date, work. In §3 we discuss SN2011fe, includ- SN2011fe in M101, the ing the early light curve. In §4 we discuss (Pereira, et al. 2013), as well as to KSN2011b, leakage of thermal and non-thermal radiation the best of three SNe Ia (KSN2011b,c; from supernovae in terms of an effective opac- KSN2012a) detected at early times by the ity. In §5 we discuss KSN2011b, the prob- Kepler satellite (Olling, et al. 2015). For lem of converting it to a bolometric scale, these supernovae we have bolometric (or near and show that it has an exceptionally smooth bolometric) light curves, so we may minimize light curve, constraining theoretical models arXiv:1611.08746v2 [astro-ph.HE] 13 Aug 2017 2 of mixing. In §6 we present a more gen- the initial radius R(0), the effective opac- eral derivation of Arnett’s rule (Arnett 1979, ity κt (a measure of leakage rate for radi- 1982). In §7 we discuss the core-collapse ant energy, see §4), and the explosion energy 3 SN2008D (type Ib), and separate the break- Esn. The initial radius of the out, shock, and radioactive heating parts of is R(0)  1013 cm so that its precise value the light curve. §8 contains our summary. is unimportant for the light curve. If the re- maining four parameters are constrained to 2. UNIQUENESS be consistent with a thermonuclear explosion Light curves of Type I supernovae have a of a white dwarf, realistic light curves for SNe characteristic shape which is due to radioac- Ia result. tive heating by gamma-rays and positrons, The situation is more complex for SNe Ibc: cooling by expansion, and cooling by ra- (1) there is a collapsed remnant which may diative loss (Arnett 1982; Pinto & Eastman be active, and (2) there may be fluid dynamic 2000a). The shape of these light curves is de- heating due to interaction of the ejecta with a termined by a parameter mantle or any surrounding gas, for example4.

1 1 This additional physics may imply new pa- y = (2τdτh) 2 /2τNi ∼ (κtMej/vsc) 2 , (1) rameters, further adding to the problem of which is a combination of diffusion time and uniqueness. If we infer y from the shape of expansion time, relative to the a reference the light curve, this only fixes the combina- time (the decay time for 56Ni), or equiva- tion κtMej/vsc (not the individual values, for lently the effective opacity times mass ejected which additional constraints are needed). It divided by the velocity scale. The amplitude might be a good project to examine spec- (peak light) is determined by tral estimates of velocity to constrain velocity structure, for example. So far this seems to L = NiMNiM Λ(x, y) have been done using v(t), which maps into 10 = 2.055 × 10 L MNiΛ(x, y) (2) radius v(r) easily enough, but not mass co- ordinate v(m) with any precision. That is where the dimensionless function Λ is deter- 2 required because EKE = 0.3vsc/Mej. mined by an integration in time, and MNi is the mass of 56Ni in solar units. An esti- 3. LIGHT CURVES FOR SN2011FE mate of radioactive heating by a mass MNi at peak light requires an estimate of the time Because the theoretical models preceded between explosion (synthesis) and peak light. the acquisition of high quality data by The theory is not a one-zone model as in decades, and the physical parameters are Arnett(1979), but involves integration over nearly the same now as then (cf. Table1), both space and time by separation of these the models are a prediction in this sense. It variables (see Arnett(1982), Eq. 11 and 48). The light curves are determined by five as- essentially complete reaction network studies of nu- trophysical parameters, acting in combina- cleosynthesis in supernova shocks, showing that the tion (Eq.1 and Eq.2) 1. They are the mass radioactive nucleus 56Ni was the dominant product, 56 2 with smaller amounts of 55Co and 57Ni. The latter ejected Mej, the mass of Ni ejected MNi, have little effect on the early light curve discussed here (0 ≤ t ≤ 40 days after explosion), however they are important diagnostics (Seitenzahl, Taubenberger & Sim 2009; R¨opke, et al. 2012). 1 The Phillips relation (Phillips 1993) was discov- ered as a feature of B-band photometry, but may be 3 For a typical expansion velocity of ≈ 109 cm realized in the bolometric light curves if y increases s−1, little luminosity is produced before times ≈ 106 with MNi, for example. From Fig.1 we find a “bolo- s for such small radii. metric ∆m ” of ∼ 1.0 mag for SN2011fe. 15 4 SNe Ia may also interact with the circumstellar 2 Truran, Arnett & Cameron(1967) did the first medium (CSM); Hamuy, et al.(2003). 3 is easy to capture the observed behavior with Table 1. Base model for SN2011fe these theoretical models. Fig.1 illustrates this observationally for SN2011fe (Pereira, et Symbol Fig.1 A82 a al. 2013) using (Arnett 1982). b Mej (M ) 1.40 1.45 c MNi(0) (M ) 0.546 ± 0.016 0.5—0.7 R(0) 4.0 × 108 cm 0—1013 cm 2 −1 κt ( cm g ) 0.09 0.08 — 0.1 51 Esn (10 erg) 1.2 ∼ 1 y2 0.573 0.8 — 1.2

aValues from Arnett(1982) for comparison. b Total stellar mass for SNe Ia. c Increase to ∼ 0.58M for variable leakage opacity (see text).

Figure 1. Comparison of bolometric luminosity 56 from analytic models (Arnett 1982) and from With a mass of Ni of 0.546M , after SN2011fe (Pereira, et al. 2013). The analytic 14.14 days (but see §3.1) a maximum lu- 9 models plotted here allow for γ–ray escape, as minosity of 3.00 × 10 L is reached; which did the original ones, by using the deposition es- timate from Monte Carlo simulations of Colgate, fits the data of Pereira, et al.(2013), who 9 Petschek & Kriese(1980). The three curves are quote 3.04 × 10 L . They use MNi = distinguished by a slightly different value for the (0.44±0.08×(1.2/α)M , with α = 1.2 to in- 56 mass of Ni, MNi = 0.546 ± 0.016M , which fer the mass of 56Ni; this expression is from agrees well with that inferred by Pereira, et al. (2013); see text and Table1. The peak luminos- Gonzalez-Gait´an,et al.(2012) and has as- ity increases with MNi. sumptions regarding three-dimensional (3D) effects (see Arnett & Meakin(2016) for a dis- cussion of 3D and resolution issues). We pre- We consider the original case of massive fer to take α = 1 so that the observation- CO white dwarfs igniting 12C under condi- ally inferred value is MNi = 0.528 ± 0.08M , tions of high electron degeneracy as being which agrees better5 with the actual 3D sim- appropriate to SN2011fe and KSN2011B; see ulations of R¨opke, et al.(2012), who found Nugent, et al.(2011) for observational sup- MNi ∼ 0.61M . 12 16 port. Burning C and O to nuclei lying in At peak only 83.8 percent of the instan- 28 40 the range Si– Ca releases almost as much taneous radioactive energy release is being explosive energy as burning the same nuclei deposited in our model, while the rest es- 56 to Ni (qSiCa ∼ qNi), but the ashes (Si–Ca) capes as x-rays and gamma-rays rather than are not radioactive, and do not provide late as thermalized radiation. The mass fractions heating. We adjust the Ni production to fit of 56Ni and 56Co are 0.197 and 0.741 respec- the peak luminosity. We assume that com- tively. Decays of 55Co and 57Ni have a negli- parable amounts of Si–Ca and Ni are pro- gible effect of this part of the light curve. duced, which determines the explosion en- ergy. We adjust the effective opacity to fit the shape of the light curve. The values chosen 5 Accounting for a variable leakage rate (see below) for SN2011fe light are documented in Table1. gives a longer rise to peak light, and MNi ∼ 0.58M . 4

Table 2. Some Initiala values which this was assumed without justification. 15 At much later times, t ∼ 10 cm/vsc ∼ Variable Symbol Value 106 s, maximum light occurs, with thermal

8 −1 photons, x-rays and γ-rays leaking rapidly. Velocity scale vsc 9.28 × 10 cm s Hydro time τh 0.43 s 3.1. Early time light curve 12 Diffusion time τd 1.52 × 10 s b 7 −3 Figure1 shows an excellent agreement of Central Density ρ(0, 0) 1.04 × 10 g cm the original theory with observations. How- 8 Central Temperature T (0, 0) 8.77 × 10 K ever the new data offer the promise of new 14 Optical depth τt(0) 3.73 × 10 understanding: the first three data points are higher than the theoretical ones. In re- aAfter shock emergence. ality the opacity is not constant as assumed; b Uniform, for consistency with γ escape. at early times the temperature and density are rapidly changing, on a time scale t ∼ R(t)/vsc, so variations in opacity are plau- sible. Table2 gives some initial values con- Given Eq.2, and expanding Λ( x, y) at early structed from these parameters. The initial times (t/τNi = xy  1), we have Λ(x, y) → radius divided by the velocity scale gives an x2 (Eq. 45 in Arnett(1982)). The normalized 8 9 expansion time scale τh ∼ 4×10 /10 ∼ 0.4 s. luminosity is then The diffusion time for photons is larger, τd ∼ 2 2 L/NiMNi = t /(2yτNi) , (3) 1.5 × 1012 s. This implies that an enormous expansion must occur before photons leak out which is the same form used by Firth, et al. 2 readily, and that the temperature is deter- (2015) but with n = 2 and a = 1/(2yτNi) . If 2 2 mined by a balance between adiabatic cool- y = (τdτh/2τNi) ∝ κtMej/vsc, is not strictly ing and radioactive heating. After the ex- constant, but decreases slowly with time6, plosion, the Hugoniot-Rankine relations for n > 2 results. Such a decrease in opacity is a shock wave imply that the enthalpy and also needed to make the light curve in Fig.2 kinetic energy are comparable. Spherical ex- merge into that in Fig.1 at peak luminosity. pansion will reduce the internal energy by a This differs from the more limited “blast factor τh/t, so that within a minute the inter- wave” approximation usually used to support nal energy is reduced to 1.2×10−4 of its post a t−2 behavior: here the effective tempera- shock value. This energy is converted into ture Teff and the velocity at the photosphere kinetic energy by work done by pressure in vphotos are not required to be constant. They 3 2 are only constrained to radiate the luminos- expanding the matter, 10 Mejvsc → Esn, and we have an expanding Hubble flow. The en- ity leaked from inside. ergy released by radioactive decay equals the Nugent, et al.(2011) found the initial rise reduced thermal energy after about 40 s; after of SN2011fe to be well-described by a radia- that the internal energy results from radioac- tive diffusion wave with an opacity due to tivity. “Early” observations (t > 1 hour) of Thomson scattering, κ = κth; Figure2 con- SNe Ia tell us primarily about thermal, not firms this, using the parameter values of Ta- 2 −1 hydrodynamic effects. ble1, but increasing κ → κth ≈ 0.2 cm g . This implies that (after 1 hour) the spatial Now the earliest three points lie on the theo- distribution of internal energy closely tracks retical curve, but the luminosity in the peak that of radioactivity, because that internal energy is caused by radioactive heating. This resolves an issue in Arnett(1982), Eq. 13, in 6 From y2 = 1.273 in Fig.2 to y2 = 0.573 in Fig.1. 5 is too low because the opacity is too high; n > 2 if a power-law fit is made for constant κ ∼ 0.09 cm2 g−1 at peak is needed to model a (i.e., L ≈ a tn). the escape of thermal energy. At later times We note that this behavior of n may be a (t > 30 days after explosion) the shape of general property of the decoupling of super- the light curve is controlled by increasing thermal photons with continuing expansion gamma-ray escape, and the numerical value rather than the distribution of 56Ni, as sug- of the effective opacity for thermal radiation gested by Piro & Nakar(2013), or a combi- is not so important. nation of the two. It remains to disentangle these effects; they both deal with the effects of radioactive decay heating.

3.2. Opacity Consider a simple leaky-bucket model for radiation loss7; frequency regions of low opac- ity are “holes” and determine the leakage rate. The highest energy photons (gamma and x-rays) have their longest mean-free paths for scattering, and these interactions fill the holes because of their relatively weak frequency dependence (see §4.3). At early times the gamma and x-rays are downscat- tered into the thermal range, so the the Figure 2. Comparison bolometric luminos- relevant scattering cross section for filling ity from analytic models (Arnett 1982) and the holes is that for Thomson scattering, SN2011fe (Pereira, et al. 2013). Here the ef- 2 −1 2 −1 and κth = 0.20 cm g . At late times fective opacity is κt = 0.2 cm g , appropriate for early times when Thompson scattering dom- a floor on the opacity is derived from the inates. The early points (t < 6 days) agree with Klein-Nishina transport cross section to be the analysis of Nugent, et al.(2011), but bang 2 −1 κKN = 0.067 cm g = κth/3. Fitting time is about 2 days earlier than in Fig.1. After about 8 days the analytic model has an effec- the shape of the peak in Figure1 gave us 2 −1 tive opacity which is too large around peak light κt ≈ 0.09 cm g which lies between these 2 −1 (κt = 0.2 > 0.09 cm g ), but at 30 days af- limits. Apparently the effective opacities for ter explosion, with gamma-escape important, the energy leakage in SNe Ia are (1) small, (2) light curve again approaches the observed values again. vary little, (0.067 < κt < 0.2), and (3) are insensitive to composition, a surprising result given the complexity of the spectra and the The “bang time” (time of explosion) in Fig- problem. This is consistent with the discus- ure2 is now earlier by 2 days. Firth, et al. sion in Dessart, et al.(2014), §3.2. It seems (2015) found a rise time of 18.98 ± 0.54 days that leakage of energy rather than line for- for 18 SNe Ia, ranging from 15.98 to 24.7 mation is the key issue for bolometric light days. For SN2011fe our value of 14.14 + 2 ∼ curves; spectra are the opposite. 16.1 days agrees well with the Pereira, et al. It appears that a modest and plausible vari- (2013) value of 16.58 ± 0.14 days. ation in the effective opacity (escape proba- Firth, et al.(2015) find the exponent of bility) will allow an excellent description of the power law fit to the early rise to be n = 2.44 ± 0.13. Although Figure2 begins as 2 L ∝ t , merging upward to Figure1 requires 7 This discussion was heavily influenced by Pinto an steeper increase in the transition and thus & Eastman(2000a,b). 6 the SN2011fe light curve, e.g., something like maximum light is approached non-thermal Figure1 with Figure2 for the first few days heating becomes less localized and mean-free- would result from a slowly varying (almost paths become longer. If we regard this as a constant) opacity. second channel for energy flow, the effective opacity would be (roughly) the inverse aver- 4. OPACITY AND ENERGY LEAKAGE age over both thermal and non-thermal chan- In Section3 we found that the effective nels, with the lower opacity channel carrying opacity to be the remaining parameter, from more energy. the original five, which needed to be used to adjust the light curves to fit the data 4.2. Conditions at maximum light from SN2011fe. The analytic solutions for Table3 shows the values of selected vari- the bolometric light curve require a rate at ables at maximum light for the models pre- which photon energy, both thermal and non- sented in Figure1. The temperatures lie in thermal, leaks from the ejected mass; this the range 1 × 104 K to 2 × 104 K, which are may be quantified by a “leakage time”. commonly found in normal . Due to the large radiative conductivity, the SN has a rel- 4.1. Leakage time atively shallow temperature gradient. The thermal leakage is estimated by a ra- Unlike the temperatures, the densities diative diffusion model, which involves the in- (of order 3 × 10−13 g cm−3) are consider- tegrated effect of the opacity over the struc- ably lower than encountered in stellar at- ture and time (Arnett 1982; Pinto & Eastman mospheres. These low densities suggest the 2000a,b,c). Although the local opacity will question: how valid is the assumption of col- vary sensitively with the local density, tem- lisional equilibrium? As discussed in Pinto perature and photon frequency, the leakage & Eastman(2000b), thermal velocity v  c time integrates these effects over a dynamic implies that collisional equilibrium is difficult structure (Pinto & Eastman 2000b), giving a to attain in supernovae, and thermal equilib- smoother variation in global properties. The rium comes from radiative interactions. term “opacity” in the analytic solutions is a a placeholder for a leakage time. This “effec- Table 3. Variables at Peak Light tive opacity” was taken to be a constant8 (for analytic simplicity) which approximates the variable value same leakage of radiative energy as a phys- b tmax − tbang 14.14 days ically correct simulation would (in practice, L 3.00 × 109 L this may only mean that the choice of effec- max tive opacity fits the observations). m(bol) -18.97 −13 −3 As seen above, this should include effects ρc 4.56 × 10 g cm 4 due to x-ray and γ-ray transport, not just Tc 1.92 × 10 K 4 deposition as in Colgate, Petschek & Kriese Teff 1.06 × 10 K

(1980). As maximum light is approached, en- τhν (thermal) 46.6 ergy transport by superthermal photons (x- τγ (superthermal) 15.5 rays and γ-rays) becomes comparable to that D(γ-deposit) 0.838 by “thermal photons” (see Pinto & Eastman 56Ni 0.199 (2000b) for discussion of thermalization). As 56Co 0.740 57Ni 3.12 × 10−5 8 Pinto & Eastman(2000b) estimated κ ≈ 0.1 cm2 g−1, which is almost identical to our value inferred from the SN2011fe light curve. Table 3 continued 7 Table 3 (continued)

variable value 57Co 0.0237

aSee Figure1. b Add ∼ 2 days to account for extension seen in Figure2.

4.3. LTE Fe opacities and holes Chris Fontes has calculated new LTE opac- ities for these low densities, which are shown in Figure3. These opacities are produced by the OPLIB database team (Magee et al. Figure 3. Logarithm of Fe opacity as a function 1995) with new calculations for the low den- of photon energy for various combinations of den- sities needed for light-curve calculations, and sity and temperature using an extension of the shown in Fig.3. Following Pinto & Eastman OPLIB database (Magee et al. 1995) (courtesy Chris Fontes). Note that as the ejecta expands (2000b) we accept that LTE is a flawed ap- (and the density and temperature decrease), the proximation in detail, but use it to suggest opacity in a given band can both decrease and in- qualitative behavior. In a simplified leakage crease. At low density (ρ = 10−12 g cm−3) gaps scheme for transport of energy, we assume open in the atomic opacity, which decreases to- ward the IR frequencies. a single opacity for the transport, but the opacities vary dramatically with photon en- ergy. For leakage schemes, the transport is thermalization is weakening, and affect the dominated by the dips in the opacity, not the leakage and transport of energy. Such effects peaks. The energy where this dip occurs de- have been subsumed in the ”effective opac- pends both on the temperature and density of ity”. the supernova ejecta, and the leakage opacity This appears to support the inference from may vary both up and down as the ejecta ex- §3 that the effective opacity may be con- pand, ranging within the visible bands from trolled by scattering processes, Thomson and below 0.001 cm2 g−1 to above 10 cm2 g−1. Compton, which close leaks. As the densities approach supernova val- ues (ρ < 10−12 g cm−3), the opacity in Fig.3 4.4. Leakage and transport shows gaps with values log10 κ ∼ −2, around To better understand the effect of such and below the LTE temperature of 1eV. opacities, we have implemented an opacity Guided by Pinto & Eastman(2000b) we as- switch in the simplified transport code of sume that this property is set by atomic Bayless, et al.(2013) that allows the opac- structure and will be qualitatively true for ity to move up or down by an order of mag- non-LTE distributions of states; see also nitude as the temperature drops below 1 eV Dessart, et al.(2014, 2015, 2016), who find (1.16×104 K). Figure4 shows the light curves that their spectra are surprisingly insensitive from a standard model assuming κ = 0.1 to the opacity controling the radiation flow. as well as the opacity switch that increases Non-thermal radiation may affect the de- and decreases the opacity. These light-curves gree of ionization around peak light, when demonstrate the sensitivity to opacity. With 8 a full transport solution, we might expect inevitably resulting from instabilities in the opacity variation to produce observable fea- explosion. Theoretical models of supernovae tures in the light curve (which we do see), which are spherically symmetric have no an- but the basic trends with the slopes will not gular resolution to deal with this broken sym- change dramatically. Notice that for the in- metry. Mixing is treated in a ad-hoc man- creased opacity case, the lightcurve begins to ner. Observed young supernova remnants rise earlier, as in Figure2. have a pronounced filamentary structure, a heterogeneity which may have developed dur- ing this epoch of instability. How might this happen? We summarize the discussion in Arnett & Meakin(2016) and Arnett, et al.(2015). Explosions are rife with instabilities. Most notable are Rayleigh- Taylor and Richtmeyer-Meshkov instabilities, which also appear in simulations of convec- tion by bottom heating. Heat gives rise to plumes which develop a mushroom cloud shape. In a convective region the upward mo- tion is bounded and turned back. Because of the astronomically large Reynolds number, the flow is turbulent and gives rapid mixing, which in high-resolution simulations is essen- tially complete in two turn-over times (four transit times). Suppose the heating is vio- lent, and the motion is not contained (an Figure 4. Comparison of bolometric luminos- explosion). Then only one transit time is ity from SN2011fe (Pereira, et al. 2013) (solid available for mixing, which is global yet in- dots) using our diffusion code with three opac- ity prescriptions: (a) constant κ = 0.1 cm2 g−1 complete. Turbulence is frozen by expansion. (solid), (b) increase at low temperatures: κ in- A fundamental feature of turbulence is high creases to 1 cm2 g−1 below 1 eV (dotted), and vorticity: vortex filaments are the “sinews” (c) decrease at low temperatures: κ decreases of turbulence (Landau & Lifshitz 1959; Pope 2 −1 to 0.01 cm g below 1 eV (dashed). Increas- 2000). Further expansion of the supernova ing the opacity broadens the light-curve, decreas- ing narrows it. These explosions assume an ex- may lead to a regime of thermal instability, plosion mass of 1.4 M , an explosion energy of in which denser regions radiate, cool, and are 51 1.3 × 10 erg and a nickel mass of 0.6 M further compressed, leaving the sort of fila- mentary structure seen in young supernova Until these issues are convincingly resolved, remnants. Is it possible to test this idea with the leakage time should be considered an observations of supernova light curves? The adjustable parameter, strongly constrained Kepler observations of supernovae may bear by observation, constrained less strongly by on this question. microscopic theory, but not yet one solidly founded in experiment and simulation. 5. SUPERNOVA KSN2011B While the Kepler satellite does not have 4.5. Filaments and mixing wavelength coverage comparable to that In addition to issues associated with line which became available for SN2011fe, the Ke- formation and leakage, there are the difficul- pler coverage is broadband, has high cadence ties associated with non–spherical variation in time, has small errors from statistical fluc- 9

knowledge of the calibration of the filter. We 1.2 can construct a Kepler magnitude and a flux in the Kepler filter for each spectrum that are 1.0 all consistent relative to each other, but there 0.8 is still an uncertain absolute scaling. Actual

max bolometric corrections are hard to do without 0.6

L/L calibrating the Kepler system. 0.4 Kepler luminosity SN 2011fe It appears that SN2011fe and KSN2011b Kepler 2011B are similar enough (Fig.5) that we may use 0.2 this similarity in the data to establish a cor- 0.0 respondence between the bolometric scale of 0 10 20 30 Days Relative to Explosion SN2011fe and the Kepler supernovae.

5.1. A Calibration Figure 5. Comparison of Kepler-normalized luminosities (i.e., in the Kepler filter bandpass) From Fig.5, we infer that the bolometric for SN2011fe and KSN2011b (Olling, et al. 2015). light curve of KSN2011b would be similar to When normalized at peak light the two SNe Ia look very similar. The time axis was translated SN2011fe. There are many caveats that need to match the point of explosion. We took the to be considered for such an inference, such time of explosion for SN 2011fe from Pereira, et as the fact that SN2011fe and KSN2011b al.(2013) (2.6 days before the first spectrum) and from Olling, et al.(2015) for Kepler 2011b may have different bolometric characteristics. (18.1 days before maximum). Given the striking similarity of the two light curves in the Kepler band, though, we be- tuations, and the light curves track the data lieve that this is a plausible inference. Using from SN2011fe well; see Fig. 1 in Olling, et the bolometric luminosties from Pereira, et al.(2015). We suggest that the Kepler data al.(2013), we establish a ‘bolometric correc- provide a constraint on mixing in SNe Ia that tion’ for our derived Kepler-magnitude light is unattainable even with the high quality curve of SN 2011fe. We then apply the same UVOIR data available for SN2011fe. corrections to KSN2011b to derive a bolomet- The spectra of SN 2011fe in Pereira, et ric light curve. We take the flux and treat it al.(2013) were obtained using the Super- as a relative luminosity (L/Lmax); this ratio Integral Field Spectrograph (Lantz et also removes concerns about absolute scal- al. 2004) on the 88” telescope on Mauna Kea ing. Fig.6 shows the time evolution of the and are presented as fully calibrated. We bolometric L/Lmax for both SN2011fe and multiplied the spectra by the filter function KSN2011b. for Kepler (a wide filter, ∼ 4400 − 8800 A)9. Now that we have a calibration, how does The Kepler system is not fully calibrated, so the best Kepler supernova compare with the there is no measured zero point to place Ke- analytic curves? KSN2011b, from Olling, et pler magnitudes onto a known scale. We used al.(2015) is shown in Fig.6. Random pho- a zero point similar to that for standard fil- tometric errors of KSN2011b are small. The ters, a procedure which gives a reasonable an- uncertainties in the SN 2011fe bolometric lu- swer, but still contains some arbitrary scal- minosities and the bolometric corrections ap- ing. Even evaluating the flux directly from plied to the KSN2011b Kepler magnitudes the spectra through the filter requires some dominate the error bars. There is evidence for a well-defined and smooth light curve, but systematic biases might affect details of 9 http://keplerscience.arc.nasa.gov/the-kepler- shape fitting. space-telescope.html Again we see a deviation between observa- 10

IIn events. The smoothness of 1D theoret- 1.2 ical light curves result from ad-hoc “box- Kepler 2011B car” mixing; simulations without such mix- 1.0 SN 2011fe Model ing are not smooth (Pinto & Eastman 2000a; 0.8 Dessart, et al. 2015, 2016). Explosions are

max unstable to 3D mixing which can leave an im- 0.6

L/L print on the young as well 0.4 as the light curve (Arnett & Meakin 2016). SN2011fe does exhibit larger fluctuations in 0.2 its light curve, consistent with its observa- 0.0 tional error bars. 10 0 10 20 Days Relative to Maximum 6. ARNETT’S RULE

Figure 6. Comparison of the bolometric lumi- It has been shown, for one-zone models nosity of SN 2011fe (Pereira, et al. 2013) and (Arnett 1979), with constant opacity and ra- adjusted data (see text) based on KSN2011b dioactive heating by 56Ni and 56Co decay, (Olling, et al. 2015). The light curves presented that by Pereira, et al.(2013) are also obtained by con- volving filter functions with their spectra. They L /M ≈  (t ). (4) estimate the errors in their magnitudes using the peak ej NiCo peak flux calibration procedure for their spectra and where L is the luminosity at maxi- we adopt the same 1σ errors per epoch. The peak errors for KSN2011b incorporate those errors as mum light, Mej is the ejected mass, and the bolometric correction is applied. The choice NiCo(tpeak) is the instantaneous rate of en- of explosion time (the precise zero of the x-axis) ergy production from radioactivity at that depends upon the algorithm used, and may be uncertain by a day or so (see §3.1). time. From a solution by separation of vari- ables (radius and time), a more accurate ex- pression was derived, tions and analytic light curves at very early Lpeak/Mej = NiCo(tpeak)D, (5) times, which we attribute to use of a constant effective opacity. After this first deviation, where D is the deposition function of Col- the curves are strikingly similar, as the higher gate, Petschek & Kriese(1980), which is a cadence of the Kepler data makes clear. Al- function of “optical” depth for γ-rays (see though the observed supernovae might not be § IV.a in Arnett(1982)). For SN2011fe, D = identical, so that fitting with different the- 0.838 at maximum light (Fig.1), which cor- oretical parameters would be legitimate in responds to a correction to Eq.4 of ≈ 20 per- principle, we have used exactly the same the- cent (increase) in estimated radioactive mass. oretical parameters as in Fig.1 in order to illustrate just how similar these light curves 6.1. A Derivation seem to be. Here we present a new derivation which is Because of the steady, fast cadence of ob- more concise, general, and hopefully clearer. servations, KSN2011b provides unique new We emphasize that, as in Arnett(1982), it information on SNe Ia: the smoothness is the instantaneous total heating rate of the of the light curve. While KSN2011b plasma which is constrained, not the radioac- tracks SN2011fe within its error bars, tive decay rate, which may explain at least the KSN2011b light curve is noticeably part of the discrepancy with Dessart, et al. smoother. There is no indication of any (2015, 2016). In addition we add a brief ap- bumps due to collision of ejecta with circum- pendix summarizing a modern approach to supernova matter, as with type Ibn and solving for Arnett(1982) light curves, so that 11 errors in implementation may be easily cor- 2017). Using Eq.8 and 10 we have rected. L/M ∼ (t)D − E(d ln L/dt + d ln κ/dt(11)). The luminosity due to radiative diffusion may be written as At any extremum in L, dL/dt = 0, so at that 4 3 time we have L = aT R /τdif , (6) L/M = D − E d ln κ/dt, (12) where

τdif ∼ κM/βcR, (7) for any plasma heating rate D. This will be true at multiple peaks and at dips as well. and β ≈ 13.7 is a dimensionless form fac- This is a more general form of Eq.5. tor from integration of the diffusion equation As emphasized in Arnett(1982) SNe Ia are over space. It includes the geometric fac- special in that the kinetic energy source and tor for spherical geometry, and is very slowly synthesis of radioactivity are more directly varying with mass-density structure; see Ar- linked than in SNe Ibc. Heating from other nett(1980), Table 2 and §VI. causes would have the same qualitative ef- At any extremum in L, either a peak or a fects as radioactive heating. The D in Eq. 12 dip, dL/dt = 0. The time derivative of Eq.6 may be generalized to be the sum of all pro- is cesses heating the ejecta. It may be affected d ln L/dt = 4d ln T/dt + 4d ln R/dt by relativistic jets (the fraction of their en- − d ln κ/dt. (8) ergy which is thermalized), , pul- sars, fluid accretion (fall-back) onto a neu- For a slowly-varying escape time, the term tron , and so on, as well as radioactiv- d ln κ/dt will be small; see discussion in §4. ity. SNe Ia models are simple in that they This is also consistent with the good fits to have only one heating source and no collapsed the light-curve data shown in Fig.1. remnant. A second equation involving the luminosity In Fig.1 there are two extrema in luminos- is first law of thermodynamics, which may be ity: at the time of explosion (bang time) and written after spatial integration as at maximum light. The first peak at very early time is due to shock breakout, and oc- L/M ∼ (t)D − dE/dt − P dV/dt. (9) curs at such small radii that it has negligible For homologous expansion (Hubble flow) effect on the luminosity for SNe Ia. Eq.3 gives the “initial” luminosity after radiative d ln V/dt = 3d ln R/dt equilibrium and Hubble flow have been es- and for a radiation-dominated plasma, tablished (see §3). For SNe Ibc, the initial radius is not so 4 E = 3PV = aT V, small. The first peak in SN2008D (Fig.7) may have a contribution from compressional so heating as the supernova shock break-out dE/dt+P dV/dt = E(4d ln T/dt+4d ln R/dt). occurs, as well as from radioactivity in (10) the outer, fast ejecta. SNe Ibc may be This requires a 3D simulation of high resolu- showing evidence in light curves for tion to compute accurately (Arnett & Meakin fluid dynamic heating/cooling in non- 2016), but deviations from Hubble flow may homologous flow that is comparable to be thought of as compressional heating and the radioactive heating rate at peak. expansional cooling from an inelastic collision Multiple-peaked light curves like SN2016gkg within the supernova ejecta (Arnett(1980), (Tartaglia, et al., 2016) might also be inter- § VIII), or with surrounding matter (Smith preted in this way. 12

6.2. A Thought Experiment opacity is not strictly constant, just slowly Consider a thought experiment which has varying, and gives an increase in ∆rise = observational implications: suppose a plume tmax − tbang of 2 days, upon integration of of 56Ni rapidly expands so that all γ-rays the SNe 2011fe (Ia) light curve. This brings escape. This energy is lost, and does not ∆rise from 14.14 days to 16.1 which agrees heat the plasma or contribute to the light well with Pereira, et al.(2013) who observe curve, except later when kinetic energy from 16.58 ± 0.14 for SN 2011fe. Near maximum the positron channel of 56Co decay becomes light, Eq. 12 may be written as significant. There would be an inconsistency L/MD ≈ 1 − (E/)d ln κ/dt, (13) in the amount of radioactive mass estimated from the pre-nebular light curve and the ra- which implies that opacity decreasing in time dioactive tail. In a less extreme case, we increases ∆rise by shifting the peak to later note that at early times γ-rays are always time. trapped, and only escape freely as expan- 6.4. SNe Ibc sion makes this possible. There will be an enhancement of the light curve before the γ- In contrast, Dessart, et al.(2015, 2016) rays easily escape, and possibly an early peak found errors of 50 percent using Eq.4 for in the light curve before the dominant one. models of SNe Ibc; these errors they at- Such a plume might be expected to have high tributed to the use of a constant mean opac- entropy, so that the 56Ni might be formed ity. However their conclusions may be weak- by the “α-rich freezeout”, possibly produc- ened, not because of doubt about the CFM- ing detectable amounts of radioactive 44Ti. GEN results, but because their approxima- This is an example of how 3D mixing can de- tion to Arnett(1982) deviates from the origi- viate qualitatively from 1D, and why Eq.5 is nal; they use 100 percent trapping of gamma- to be preferred over Eq.4. See §7 below. rays (D = 1), and the integral approach of Katz, Kushnir & Dong(2013). Dessart, et al. 6.3. SNe Ia (2015, 2016) also state that they also must Blondin, et al.(2013) found agreement add a new 56Co decay from Valenti, et al. to within 10 percent with “Arnett’s rule” (2008) (which they correct further), rather (apparently Eq.4) for their set of delayed- than the original in § VI.a of Arnett(1982). detonation models of SNe Ia, using the CFM- As this rule of Katz, Kushnir & Dong(2013) GEN code. In §5 we showned that the light and Eq.5 are both based on the first law of curve for the , KSN2011b, thermodynamics, it may be suspected that was a smooth curve. The bolometric light deviations between the two are most likely curve directly measures the rate of leakage due, at least in part, to the equations be- of thermal energy from the supernova (see ing incorrectly applied, or to using different §4), and because the curve is smooth, this physics (e.g., D = 1). See AppendixB for a leakage may be approximated by a slowly simpler mathematical formulation of Arnett varying average opacity. This contrasts to (1982). the complex variation of the local opacity seen in §4.3 and CFMGEN. Evidently inte- 7. SN2008D gration over frequency, ionization, and mean- The similarity in light curves, of thermonu- free-path tames the unruly local behavior of clear and core-collapse supernovae suggests the opacity to give a better behaved leakage that we attempt to use the procedures for rate (as averaging over the Kolmogorov cas- SNe Ibc as well. The approach of Arnett cade does for turbulence). (1982) gives an opportunity to separate ra- We saw in §3.1, however, that the leakage dioactive from non-radioactive heating in the 13 Table 4 (continued) light curves of SNe Ibc. To illustrate this point we examine the light curve of the well observed SNe Ib 2008D. Variable Symbol Value Supernova 2008D was discovered at the KE EK 2–4 bethe time of explosion, which coincided with an ejected mass Mej 3—5 M X-ray event due to the break-out of the su- initial 56Ni M 0.05—0.1 M pernova shock wave (Soderberg et al. 2008). Ni This was the first successful observation of the previously predicted break-out of a super- nova shock (Colgate & McKee 1969; Arnett 1971; Chevalier 1976), and to be associated Table4 summarizes parameters derived with a bare core, or stripped-envelope super- from Soderberg et al.(2008). Modjaz, et nova (Arnett 1977). Modjaz, et al.(2008) al.(2008) have further discussed the derived presented a detailed discussion of the obser- parameters, and most notably found signifi- vational data (optical, infra-red, and X-ray) cantly different results for initial radii with of this event, and Bianco, et al.(2014) sum- shock breakout theories of Waxman et al. marized the results for 64 similar events, in- (2007) or Chevalier & Fransson(2008). cluding this one. We base our discussion on the data sets in these papers, which contain a more detailed history and references. SN2008D occurred in NGC 2770 (D = 31± 2 Mpc) and had a peak of MV = −17.0±0.3 mag and MB = −16.3±0.4 mag (Modjaz, et al. 2008). The rise time in the V-band was 16.8 ± 0.4 days and in the B-band 18.3 ± 0.5 days, near the long end of the range for SNIbc. The observation of the X-ray outburst determines a more precise value for the explosion time, which is of value in fitting the data to theoretical models (see Section 3.1). Integrating the observed bands, or fitting a blackbody give similar bolomet- Figure 7. Bolometric light curves for SN2008D (circles) compared to analytic models fitted to ric corrections, so that at 19.2 days after ex- the luminosity peak at 20 days (red line). V- plosion, the bolometric luminosity peaks at magnitudes from this Ib supernova (Bianco, et 42.2±0.1 −1 8 al. 2014) are bolometrically corrected using Fig. 9 L = 10 erg s or L = 4.12 × 10 L . of Modjaz, et al.(2008), and are shown as cir- cles. The X-ray spike found by Soderberg et al. Table 4. Some values of variables defined by (2008) is shown (purple) at the origin in time Soderberg et al.(2008) for SN2008D. and allows an accurate determination of explo- sion time. The difference between this purely radioactive part (red line) and the observations Variable Symbol Value (circles) allow us to estimate the excess due to additional processes (blue line and pluses). 10 initial radius R∗ 7 × 10 cm −1 velocity (phot) vphot 11, 500 cm s In order to compare bolometric light curve −1 velocity (scale) vsc 4, 743 cm s shapes to observational data, we correct the V-band magnitudes (Bianco, et al. 2014) Table 4 continued using graphical extraction of the bolomet- ric corrections from Fig. 9 in Modjaz, et al. 14

(2008). From 1 to 30 days after outburst, What might cause this intermediate peak? the variation is not large. The possibilities are intriguing but not In Fig.7 an estimate of the bolometric unique. For example: light curve for SN 2008D is presented. There (1) Shock heating of a clump of matter are three different pieces: (a) an X-ray burst which was ejected to a radius r > 1014 cm (purple) at explosion time, (b) newly-found, prior to core collapse, perhaps due to pul- underlying intermediate peak at about 2 days sations and eruptive mass loss (Arnett & (blue), and (c) a large peak at about 16-20 Meakin 2016), could be responsible. days (red). We proceed to analyze each piece. (2) A 56Ni plume of modest mass due to Rayleigh-Taylor instabilities (for example, a 7.0.1. Main diffusive stage mass of 0.028M in the plume, with 0.012M The major diffusive part (c) is represented of 56Ni and 0.3 bethe of kinetic energy can re- by an analytic light curve (red) with the pa- produce the excess luminosity in Fig.7, with rameters given in Table4, and with κt = little effect on the main peak). 0.09 as used previously for SN2011fe and (3) Many other possibilities which may KSN2008b. For simplicity no attempt was be imagined due to activity in the newly- made to correct for effects of varying leakage collapsed core (see, e.g., §6.1). (opacity). In this approximation the evolu- (4) The breakout might be more complex tion begins after the shock reaches the sur- than now imagined, and relax in a broader face; there is no shock break-out included. peak, not a δ-function. The first two are likely to give variations 7.0.2. Break-out stage from event to event, as might the last two. The break-out (a) gives a spike in X-rays New instruments such as LSST should clar- (Soderberg et al. 2008), followed by cool- ify the extent of such fluctuations in the light ing of this near-surface region by radiative curve of SN Ibc at early times (t < 8 days). diffusion and by expansion, quickly merg- In this paper we have attempted precise com- ing to the lowest-order diffusion solution Ar- parisons with a few selected data sets of high nett(1980, 1982); Pinto & Eastman(2000a). quality; this has the disadvantage that it may The break-out spike in luminosity (purple) is not span the space of natural variation, which much brighter than the V-band luminosity, may be large for SNe Ibc (Wheeler, Johnson because of large bolometric error (the spike & Clocchiatti 2015). radiation begins as X-ray and cools to UV, Table4 gives a moderately different set of and so is not UVOIR). parameters from Table5 because: (1) the bang time was measured, (2) the shape of the 7.0.3. Intermediate peak light curve was fit, including the excess (in- If we subtract the analytic model from termediate peak), and (3) the leakage opac- the bolometric luminosity, there is an ex- ity was assumed the same as in SN2011fe cess which coresponds to a peak located at (§3). For consistency a small correction for 1 < t ≤ 8 days. This is piece (b) referred to the shift in bang time due to variable leak- above and seen in Fig.7 as blue. After this age (opacity) should be made; for simplicity time the total luminosity is well represented it was not. by the analytic model (red), which is due only The net result of fitting the inferred dif- to radioactive decay. This intermediate peak fusion peak (blue) separately seems to be a (in blue) at 1 < t < 8 days, following the ini- smaller estimate of ejected mass and explo- tial X-ray spike, does not appear in SNe Ia. sion energy for this SN Ib (compare Table5 This use of the analytic model allows us to to Table4), but these differences are not probe the underlying physics in more detail. drastic. If the total mass at core collapse is 15

∼ 3.5M , as we infer from Fig.7, SN2008D which is easily removed by the relax- could be the stripped core of a star of initially ation of the assumption of strictly con- 12—15 M , for example. stant leakage time (opacity) at these times. This also slightly affects the Table 5. Parameters for model (red curve) of SN2008D inferred time from explosion to peak (Fig.7) light. 4. Kepler data for KSN2011b may already indicate the presence of turbulent mix- Variable Symbol Value ing during the explosion. It provides a Mass ejected Mej 2.0 M observational support for a slowly vary- 56 Initial Ni MNi(0) 0.082 M ing leakage time (mean opacity). Initial radius R(0) 1.0 × 1010 cm 9 −1 Velocity scale vsc 1.38 × 10 cm s 5. Arnett’s model may be generalized to 2 −1 Opacity κt 0.09 cm g SNIbc; for SN 2008D, inclusion of ra- 51 Explosion energy Esn 2.3 × 10 ergs dioactive plumes or pre-collapse erup- tions are among the possible causes of a excess radiation. Total stellar mass at core collapse is ∼ 3.5M , that is, Mej plus the mass of the collapsed core. 6. Arnett’s rule deals with the radioactive heating of the plasma, and so includes 8. CONCLUSION effects of gamma-ray escape. We have applied the Arnett(1982) model to a wide variety of high-quality data on type 7. The SNe Ib, 2008D, is consistent with Iabc supernovae. In summary: the explosion of a stripped-envelope star of initially 12–15M , possibly with 1. There is excellent agreement between a small radioactive plume. the best SNIa bolometric data and an- alytic models (which are bolometric).

2. There is an important problem of uniqueness; multiple combinations of We wish to thank an exceptionally cogent parameters may all explain a given light and critical referee for encouragement, and curve. Light curves should be used, the Kepler team for providing a machine- but with care, as probes of the su- readable copy of their data for KSN2011b,c pernova engine for core collapse super- and KSN2012a, Federica Bianco for machine- novae (SNIbc). A light curve fit might readable data for SN2008D, and Gautham have little to do with details of core col- Narayan and Charles Kilpatrick for interest- lapse. ing discussions of their work. We thank the Theoretical Astrophysics Program (TAP) at 3. The best very early SNIa data show the University of Arizona, and Steward Ob- a deviation from the analytic models, servatory for support. 16

APPENDIX

A. NUCLEAR DATA The data for the nuclear decay of 56Ni and 56Co in the analytic light curves is documented in Table A1.

Table A1. Nuclear physics dataa

Variable Symbol Nucleus Value

56 Half-life τ 1 Ni 6.075 days 2 56Co 77.236 days 56 Total gamma energy Eγ Ni 1.750 MeV 56Co 3.610 MeV 56 Positron KE Eβ+ Co 0.120 MeV 56 10 Energy release rates Ni Ni 3.9805 × 10 erg/g/s 56 9 Co(γ) Co 6.4552 × 10 erg/g/s + b 56 8 Co(β ) Co 2.1458 × 10 erg/g/s

aSlightly updated and extended version of Nadyozhin(1994), using the NNDC Chart of Nuclides www.nndc.bnl.gov/chart. b Mean kinetic energy of positron emission; positrons are assumed to slow before annihilation, and these gamma rays are assumed to escape.

B. LIGHT CURVE EQUATIONS where D is the deposition function of A more versatile method for solution of the Colgate, Petschek & Kriese(1980) for Arnett(1982) model is to use four coupled or- γ-rays, and τdif = κtMej/(13.7cR(0)), 4 dinary differential equations, which we docu- and φ(t) = a[(T (t)R(t))/T (0)R(0)] . ment below. This is mathematically equivalent to Eq. 1 of Katz, Kushnir & Dong(2013). B.1. Equations for Arnett (1982) model 2. Constant homologous expansion (Hub- 1. The first law of thermodynamics, ble flow):

dE/dt + P dV/dt = deposit − ∂L/∂m dR/dt = vsc,

1 becomes, upon integrating over the co-   2 where v = 10E /3M , moving spatial variables (Arnett 1982) sc KE ej and assuming dominant radiation pres- 56 sure, 3. N decay:

dφ(t)/dt = R(t)/R(0) dXNi/dt = −XNi/τNi,  h MNi (NiXNi + CoXCo)D 4. 56Co decay:

+ i  +CoXCo − φ(t)/τdif , dXCo/dt = −dXNi/dt − XCo/τCo, 17

See also Table A1. B.2. Changes and extensions These equations are integrated by Runge- To include another radioactivity, add a new Kutta methods to a global accuracy of better decay equation (like 3 annd 4), and add the 8 than one part in 10 . For the same physi- new plasma heating to 1. cal assumptions concerning escape of γ-rays, To add a new source of energy (, i.e., the function D, these equations and the accretion disk, etc.), define a plasma heating integral approach of Katz, Kushnir & Dong rate as a function of time and add to 1. (2013) should give consistent results (fitting To do energy tests or a Katz, Kushnir & the light curve with this approach may give Dong(2013) analysis, add new equations and more information, as seen in §7). variables (integrands) to integrate the energy types over time. Typos do happen, but try to validate and verify before publishing (e.g., reproduce Fig- ure1 for the same parameters).

REFERENCES

Abarzhi, S. I., 2010, Phil. Trans. R. Soc. A, 368, Colgate, S. A., 1969, ApJ, 157, 623 1809 Colgate, S. A., Petschek, A. G., & Kriese, J. T., Abarzhi, S. I., & Rosner, R., 2010, Physica Scripta, 1980, ApJL, 237, L133 T142, 014012 Dessart, L., Hillier, D. J., Blondin, S., Khokhlov, Arnett, W. D., 1971, ApJ, 163, 11 A., 2014, MNRAS, 441, 3249 Arnett, W. D. & Falk, S. W., 1976, ApJ, 210, 733 Dessart, L., Hillier, D. J., Woosley, S. E., Livne, E., Arnett, W. D., 1977, 8th Texas Symposium on Waldman, R., Yoon, S-C., Langer, N., 2015, Relativistic Astrophysics, 302, Ann. N. Y. Acad. MNRAS, 453, 2189 Sci., 90 Dessart, L., Hillier, D. J., Woosley, S. E., Livne, E., Arnett, W. D., 1979, ApJL, 230, L37 Waldman, R., Yoon, S-C., Langer, N., 2016, Arnett, W. D., 1980, ApJ, 237, 541 MNRAS, 458, 1618 Arnett, W. D., 1982, ApJ, 253, 785 Falk, S. W., 1978, ApJ, 226, L133 Arnett, W. D., 1988, ApJ, 331, 377 Falk, S. W., & Arnett, W. D., 1973, ApJ, 180, L65 Arnett, D., 1996, Supernovae and Nucleosynthesis, Falk, Sydney W. & Arnett, W. David, 1977, ApJS, Princeton University Press, Princeton NJ 33, 515 Arnett, W. D. & Meakin, C., 2011, ApJ, 733, 78 Fesen, R. A., & Milisavljevic, D., 2016, ApJ, 818, 17 Arnett, W D., & Meakin, C., 2016, Reports of Prog. Firth, F. E., Sullivan, M., Gal-Yam, A., Howell, D. Phys. A., Maguire, K., Nugent, P., Piro, A. L., Baltay, Arnett, W. D., Meakin, C. A., Viallet, M., C., Feindt, U., Hadjiyksta, E., McKinnon, R., Campbell, S. W., Lattanzio, J. C. & Moc´ak,M., Ofek, E., Rabinoowitz, D., Walker, E. S., 2015, 2015, ApJ, 809, 30 MNRAS, 446, 3895 Bayless, A. J., Even, W., Frey, L. H., Fryer, C. L., Fryxell, B. A., Arnett, W. D., & M¨uller,E., 1991, Roming, P. W. A., Young, P. A. 2015, ApJ, 805, ApJ, 367, 619 98 Gonzalez-Gait´an,S., Conley, A., Bianco, F. B., et Bianco, F. B., Modjaz, M., Hicken, M., Friedman, al., 2012, ApJ, 745, 44 A., Kirshner, R. P., Bloom, J. S., Challis, P., Hamuy, M., Phillips, M., Suntzeff, N., Maza, J., Marion, G. H., Wood-Vasey, W. M., & Rest, A., 2003, IAUC, 8151 2014, ApJS, 213, 19 Hoeflich, P., Khoklov, A., & M¨uller,E., 1991, A&A, Blondin, S., Dessart, L., Hillier, D. J., Khokhlov, A. 248, L7 M., 2013, MNRAS, 429, 2127 Katz, B., Kushnir, D., & Dond, S., 2013, Chevalier, R. A., 1976, ApJ, 208, 826 arXiv:1301..6766v1. Chevalier, R. A. & Kirshner, R. P., 1978, ApJ, 219, Kirshner, R. P. & Chevalier, R. A., 1977, ApJ, 218, 931 142 Chevalier, R. A. & Fransson, C. 2008, ApJ, 683, Landau, L. D. & Lifshitz, E. M., 1959, Fluid L135 Mechanics, Pergamon Press, London. Clayton, D. D., 1968, Principles of Stellar Evolution Landau, L. D. & Lifshitz, E. M., 1969, Statistical and Nucleosynthesis, McGraw-Hill, New York Physics, 2nd Ed., Addison–Wesley, Reading MA. 18

Lantz, B., Aldering, G., Antilogus, P., et al. 2004, Smith, Nathan, arXiv:1612.02006, to be published Proc. SPIE, 5249, 146 in the Supernova Handbook Magee, N. H., Abdallah, J., Jr., Clark, R. E. H., Smitka, M., et al., in preparation Cohen, J. S., Collins, L. A., Csanak, G., Fontes, Soderberg, A. M., Berger, E., Page, K. L, et al. C. J., Gauger, A., Keady, J. J., Kilcrease, Merts, 2008, Nature, 453, 469 A. L. 1995, in ASP Conf. Ser. 78, Astrophysical Swisher, N. C., Kuranz, C. C., Arnett, D., Applications of Powerful New Databases, ed. S. J. Hurricane, O., Remington, B. A., Robey, H. F., & Adelman & W. L. Wiese (ASan Francisco, CA: Abarshi, S. I., 2015, Physics of Plasmas, 22, ASP). 51 102707 Landau, L. D. & Lifshitz, E. M., 1969, Statistical Tartaglia, L., Fraser, M., Sand, D. J., Valenti, S., Physics, 2nd Ed., Addison–Wesley, Reading MA. Smartt, S. J., McCully,, C., et al., 2016, arXiv: Modjaz, M., Li, W., Butler, N., et al., 2009, ApJ, 1611.00419v1 702, 226 Truran, J. W., Arnett, W. D., & Cameron, A. G. M¨uller,E., Fryxell, B. & Arnett, D., 1991, A&A, W., 1967, Can. J. Phys., 45, 231 251, 505 Valenti, S., Benetti, S., Cappellaro, E., Patat, F., Munari, U., et al., 2013, New Astronomy, 20, 30 Mazzali, P., Turatto, M., Hurley, K., Maeda, K., Nadyozhin, D. K., 1994, ApJS, 92, 527 Gal-Yam, A., Foley, R. J., Filippenko, A. V., Nugent, P. E., Sullivan, M., Cenko, S. B., et al., Pastorello, A., Challis, P., Frontera, F., 2011, Nature, 480, 344 Harutyunyan, A., Iye, M., Kawabata, K., Olling, R., Mushotzky, R., Shaya, E., Rest, A., Kirshner, R. P., Li, W., Lipkin, Y. M., Matheson, Garnavich, P., Tucker, B., Kasen, D., Margheim, T., Nomoto, K., Ofek, E. O., Ohyama, Y., Pian, S., & Filippenko, A., 2015, Nature, 521, 332 E., Poznanski, D., Salvo, M., Sauer, D. N., Pereira, R., Thomas, R. C., Aldering, G., et al., Schmidt, B. P., Soderberg, A., & Zampieri, L., 2013, A&A, 554, A27 2008, MNRAS, 383, 1485 Phillips, M., 1993, BAAS, 182, 2907 Waxman, E., M?esza?ros, P., & Campana, S. 2007, Pinto, P. A., & Eastman, R., 2000a, ApJ, 530, 744 ApJ, 667, 351 Pinto, P. A., & Eastman, R., 2000b, ApJ, 530, 757 Wheeler, J. C., Johnson, V., & Clocchiattii, A., Pinto, P. A., & Eastman, R., 2000c, New 2015, MNRAS, 450, 1295 Astronomy, 6, 307 Woosley, S. E. & Weaver, T. A., 1986, ARA&A, 24, Pope, S. B., 2000, Turbulent Flows, Cambridge 205 University Press, Cambridge, GB Zheng, W., et al., ApJL, 778, L15 Piro, A. L., & Nakar, E., 2013, ApJ, 769, 67 Zingale, M., Nonaka, A., Almgren, A. S., Bell, J. B., Riess, A., 1999, AJ, 118, 2675 Malone, C. M., & Woosley, S. E., 2011, ApJ, 740, R¨opke, F. K., Kromer. M., Seitenzahl, I. R., et al., 8 2012, ApJL, 750:L19 Seitenzahl, I. R., Taubenberger, S., & Sim, S. A.. 2009, MNRAS, 400, 531