A High-resolution Study of Presupernova Core Structure

, , and

Published 2018 June 15 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Tuguldur Sukhbold et al 2018 ApJ 860 93 DOI 10.3847/1538-4357/aac2da

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/860/2/93

Abstract

The density structure surrounding the iron core of a massive star when it dies is known to have a major effect on whether or not the star explodes. Here we repeat previous surveys of presupernova evolution with some important corrections to code physics and four to 10 times better mass resolution in each star. The number of presupernova masses considered is also much larger. Over 4000 models are calculated in the range from 12 to 60 M with varying mass loss rates. The core structure is not greatly affected by the increased spatial resolution. The qualitative patterns of compactness measures and their extrema are the same, but with the increased number of models, the scatter seen in previous studies is replaced by several localized branches. More physics-based analyses by Ertl et al. and Müller et al. show these branches with less scatter than the single-parameter characterization of O'Connor & Ott. These branches are particularly apparent for stars in the mass ranges 14–19 and 22–24 M. The multivalued solutions are a consequence of interference between several carbon- and oxygen-burning shells during the late stages of evolution. For a relevant range of masses, whether a star explodes or not may reflect the small, almost random differences in its late evolution more than its initial mass. The large number of models allows statistically meaningful statements about the radius, luminosity, and effective temperatures of presupernova stars, their core structures, and their remnant mass distributions.

Export citation and abstract BibTeX RIS

1. Introduction

As a massive star below the pair-instability threshold (∼80 M; Woosley 2017) evolves through its final stages of nuclear burning, its central regions cool by neutrino emission, become degenerate, and tend to decouple from the overlying layers and evolve as separate stars. Although never becoming completely detached except for the lowest-mass stars, the presupernova core takes on a structure similar to that of a white dwarf, mostly composed of iron, surrounded by a dense mantle of oxygen and intermediate-mass elements. As numerous studies have shown, the structure of this configuration, and especially the rate at which the density declines outside the iron core, is strongly correlated with the difficulty of blowing the star up (e.g., Burrows & Lattimer 1987; Fryer 1999). Recent studies have sought to capture this complex structure in just one or two parameters that might predict, albeit approximately, whether a star with a given mass blows up or collapses to a black hole simply from looking at one-dimensional models for stellar evolution (O'Connor & Ott 2011, 2013; Ugliano et al. 2012; Pejcha & Thompson 2015; Ertl et al. 2016; Müller et al. 2016; Sukhbold et al. 2016).

Sukhbold & Woosley (2014) discussed how the advanced burning stages, especially convective carbon and oxygen burning, sculpt this structure and provided a library of presupernova stars consisting of 503 models in the mass range 12–65 M to demonstrate the systematics and its dependence on input physics. Müller et al. (2016) prepared a larger grid of over 2000 models between 10 and 32.5 M and corrected some errors in Sukhbold & Woosley (2014). Their results showed finer structure but similar global systematics (see especially their Figure 6).

More recently, Farmer et al. (2016), using the MESA code, found that the choice of mass resolution, i.e., zoning, "dominates the variations in the structure of the intermediate convection zone and secondary convection zone during core and shell hydrogen burning, respectively" and greatly affects the structure of presupernova stars. They found that a minimum mass resolution of ∼0.01 M was necessary to achieve convergence in the final helium core mass at the ∼5% level. They also found ∼30% variations in the central electron fraction and mass locations of the main nuclear-burning shells, and that a minimum of ∼127 isotopes was needed to attain convergence of these values at the ∼10% level.

Renzo et al. (2017) also used the MESA code to explore the sensitivity of massive star evolution due to variations in the mass loss rate for stars with initial masses between 15 and 35 M. They found variations in the presupernova core-compactness parameter (Section 4.3) of ∼30%, depending upon the choice of the algorithm. In a limited study of resolution in one model, they found roughly 9% variation in final core compactness, smaller than that resulting from uncertainties in the mass loss prescription.

In this paper, we present a new survey similar to Sukhbold & Woosley (2014) and Müller et al. (2016) but using much finer zoning, a greater number of models, and several different mass loss rates. The key differences of the new survey are listed in Section 2, including the correction to an erroneous pair-neutrino loss rate used by Sukhbold & Woosley (2014). Section 2 also offers a detailed discussion of the effects due to zoning, network size, and boundary pressure. In general, we find that the pattern of final core compactness seen in these previous studies is unaltered by finer zoning or a larger reaction network, though the mass limits for different behaviors are shifted by about 10% when the pair-neutrino loss rate is corrected. In Sections 3 and 4, we discuss the general characteristics of the new survey, including its observable properties. With a much greater number of models, clear evidence emerges for multivalued solutions to the presupernova structure of stars in the 14–19 and 22–24 M ranges (see also Müller et al. 2016). In Section 5, we offer an interpretation for this feature through the physics of the advanced-stage evolution in the cores of massive stars. Large variations in core structure are expected for stars of these masses no matter what resolution, mass loss rate, or reaction network is employed in the calculation (for the effects of such variations in individual models, see Farmer et al. 2016; Renzo et al. 2017). Future surveys of supernova, as well as presupernova, models should thus take care to examine, when feasible, a large number of masses in order to sample an outcome that is essentially statistical in nature. More specific results and effects on explodability and remnant masses are discussed in Section 6. Finally, in Section 7, we offer our conclusions.

2. Code Physics and Assumptions

With several important exceptions, the code, physics, and input parameters used here are the same as in Sukhbold & Woosley (2014) and Müller et al. (2016). The (solar) initial composition is from Asplund et al. (2009), as was used in Müller et al. (2016), but with an appropriate correction for the ratio 15N/14N (Meibom et al. 2007). Convective and semiconvective settings, nuclear reaction rates, opacities, and mass loss rates are the same as in both of the earlier works. The mass range studied and lack of rotation are the same as well.

Nuclear burning was handled as usual, using a 19 isotope approximation network until the central oxygen mass fraction declines below 0.03 and a silicon quasi-equilibrium network with 121 isotopes (Weaver et al. 1978) thereafter. Numerous studies that compare this treatment with the energy generation and bulk nucleosynthesis obtained by "coprocessing" with a network of several hundred to over 1000 isotopes have shown excellent agreement (e.g., Woosley et al. 2002; Woosley & Heger 2007), at least up to central oxygen depletion, where the switch to the quasi-equilibrium network is made. By coprocessing, we mean carrying a large network in each zone using the same time step, temperature, and density but not coupling the energy generation from that large network directly to the iterative loop within a time step. After oxygen depletion, the quasi-equilibrium network is more stable, contains the same weak interaction physics, and is roughly an order of magnitude faster. An exception is the treatment of oxygen and silicon burning in stars lighter than about 11 M (Woosley & Heger 2015). For such light stars, it is important to follow neutronization in multiple off-center shells where the quasi-equilibrium approximation has questionable validity and can be unstable. Such light stars are not part of the present survey, which starts at 12 M. Section 2.2 discusses the issue of network sensitivity in greater detail and confirms the validity of the approximation network plus quasi-equilibrium for representative cases.

Key differences in the new study are as follows.

  • 1.  
    The models in this paper employ much finer mass resolution (Figures 1, 2, and 4). For the lightest stars considered (12 M), the increase is roughly a factor of four. Since an effort was made to keep fine resolution even in the larger stars, the factor for the highest masses considered, around 60 M, was closer to 15. In the most massive models, the total number of mass shells was approximately 16,000. Additional studies of individual cases of zoning sensitivity in 15 and 25 M models (Section 2.1) showed little systematic difference. In all cases, zones contained less than about 0.01 M everywhere but were about 10 times smaller than that in the heavy element core (Figure 4). It is this core that is most critical in determining the final presupernova core properties, such as compactness. Surface zoning was a few times 10−5 M, and all temperature and density gradients were well resolved (Figure 2).
  • 2.  
    A major, long-standing coding error in the KEPLER code was repaired. The error affected the axial vector component of the pair-neutrino losses. In particular, the error resulted in the accidental zeroing of the second term involving ${C}_{{\rm{A}}}^{{\prime} 2}$ and ${Q}_{\mathrm{pair}}^{-}$ in the expression (Itoh et al. 1996, Equation (2.1))
    Equation (1)
    Equation (2)
    See Itoh et al. (1996) for the definitions of the quantities. The consequence of this error was that the pair-neutrino loss rate was underestimated by close to a factor of two during carbon burning and somewhat less during later, higher-temperature stages. The error, introduced through inadequate checking of a routine provided by email, was included in 2001 and affected all KEPLER calculations published through 2014. In particular, it affected the often-cited calculations of Woosley et al. (2002), Woosley & Heger (2007), and Sukhbold & Woosley (2014). Because some of the models from Sukhbold & Woosley (2014) were used by Sukhbold et al. (2016), it also affected the outcome of that work for masses larger than 14 M. The bug was repaired, however, in the works of Woosley & Heger (2015) on 9–11 M stars and Woosley (2017) on pulsational pair instability supernovae. Most relevant to this present work, the bug was also repaired for the work of Müller et al. (2016). Because of the strong sensitivity of neutrino and nuclear reaction rates to temperature, the effect of the bug was a slight shift upward in the burning temperature for the late stages of stellar evolution. For a given model, the change in presupernova structure was not great and, as we shall see, within the "noise" of other uncertainties. For the lightest models (<14 M), the difference is hardly noticeable, but it did systematically shift the outcome for higher-mass stars significantly downward. For example, the peak in compactness that occurred for Sukhbold & Woosley (2014) at about 24 M is shifted downward in the work of Müller et al. (2016) and in the present work (Section 4) by about 3 M.
  • 3.  
    The surface boundary pressure is much less than in Müller et al. (2016), which significantly affects the final red supergiant (RSG) properties but does not significantly alter the core structure. Whereas the mass loss rates are varied in the present study, the new calculations do not include Wolf–Rayet models where the envelope is completely lost. All stars studied here retained a substantial hydrogen envelope when they died.

Figure 1.

Figure 1. Masses of individual zones as a function of interior mass for a 15 M zero-age main-sequence star (top) and the corresponding presupernova star (bottom). The gray curves are the zoning for the earlier model of Sukhbold & Woosley (2014), which was very similar to that of Woosley & Heger (2007), Sukhbold et al. (2016), and Müller et al. (2016); the blue curves are for the new survey reported here. The previous zero-age main-sequence model had 1,068 zones; the new one has 4,257. Stars are continually rezoned as they run to accommodate changing gradients in key quantities, but the zoning shown here did not vary greatly, with the exception of fine zones added at the base of the hydrogen envelope around 4.4 M and a few regions of finer zoning at what were once the boundaries of convective shells inside 2 M. The surface zoning was ∼1028 g in all studies (Section 2).

Standard image High-resolution image
Figure 2.

Figure 2. Density and temperature resolution in the same 15 M presupernova model shown in the lower panel of Figure 1. The top panel shows the scale heights (${M}_{x}\equiv {(d\mathrm{ln}x/{dm})}^{-1})$ expressed in solar masses for the density (Mρ; gray curve) and temperature (MT; red curve). The blue dashed curve shows the zoning. The bottom panel shows the actual variation in density between zones and mean atomic weight, $\bar{A}$, plotted also as a function of interior mass. Discontinuities in $\bar{A}$ occur at the boundaries of active and fossil convective shells causing abrupt changes in density that the code attempts to resolve. The large spike in MT at 0.15 M reflects a temperature inversion at the star's center (where dT/dm is negative) and a small region of nearly constant temperature bounding it. Both panels show that the temperature and density gradients are very well resolved throughout the star, except perhaps in the very steep density gradient at the edge of the helium core (4.33 M), where changes of up to 30% occur for the density in a few zones.

Standard image High-resolution image

2.1. The Effects of Zoning

One criterion for adequate zoning is that key variables like temperature and density be well resolved, that is, that they do not vary greatly in going from zone "i" to "i+1." Stellar evolution codes solve linear approximations to nonlinear differential equations. Figure 2 shows the resolution for a standard MZAMS = 15 M presupernova star with 4225 zones. The figure gives the scale heights in Lagrangian units (solar masses here) for the temperature and density. Pressure and other derived quantities, though not plotted, show similar variation. The scale height, e.g., for density is defined as ${M}_{\rho }\equiv {(d\mathrm{ln}\rho /{dm})}^{-1}$. In places, this quantity can abruptly become artificially small because of discontinuous changes in mean atomic weight at the edges of convective shells. Some of the prominent downward spikes in the iron core are where partial photodisintegration has changed $\bar{A}$ appreciably. These abrupt changes in $\bar{A}$ are responsible for most of the discontinuous spikes in the figure. Large upward spikes indicate regions of near-constant temperature or density. An especially large spike near the center of the presupernova star reflects a temperature inversion (note that the scale height can be either positive or negative; the absolute value is plotted). Where the derivative changes sign and passes through zero, a spike results.

The figure shows that there are roughly 100–1000 zones per scale height everywhere in the star, except in the steep gradient at the edge of the helium core, where the actual density varies by six orders of magnitude over an interval of ∼0.2 M. Here the zoning is worst. In one location, the density changes by a maximum of 30% from one zone to the next. Zone masses here were ∼5 × 10−4 M. Because this location moves in mass as the star evolves, still finer zoning would have significantly lengthened the calculation. We conclude that the relevant physical quantities are well resolved in the new study. In fact, except at the edge of the helium core, they are overresolved. This was done in order to explore the sensitivity of the outcome, e.g., the extent of the convective shells, and to see if fine zoning alone would lead to a "converged" answer.

Convergence assumes the existence of a well-defined solution to the stellar structure equations. As the numerical resolution is increased, if the subgrid physics (e.g., convection) is coded in a zoning-independent way, the answer from a given calculation should approach this solution and give a constant answer. Certainly, it is possible for a stellar model to be inadequately zoned. Envision Figure 2 with zoning of 0.1 M. No one today would think of trying to use just a few hundred zones in a presupernova model for a 25 M star, though substantial progress was once made that way (Weaver et al. 1978). But given the power of modern machines, how many zones are enough?

Typical KEPLER models by Sukhbold & Woosley (2014) and Müller et al. (2016) used about 1200 zones, roughly independent of the star's total mass, though the zoning was by no means uniform. Zones were concentrated at the center, where temperature-dependent burning required fine resolution; in the steep density gradient at the base of the red giant convective envelope; and at the surface. Continuous rezoning kept the gradients in density, temperature, and zone radius well resolved. For a large range of masses, characteristics of the presupernova star—such as its helium, carbon, and iron core masses; luminosity and radius; and "compactness"—varied smoothly with mass and were consistent with numerous past studies. In some mass ranges, however, especially in the range 14–19 M, the compactness parameter often varied wildly. These variations were attributed by Sukhbold & Woosley (2014) to the interaction of convective carbon- and oxygen-burning shells. The characteristics of those burning shells were irregular in location, extent, duration, and intensity and, for some masses, so was the final presupernova core structure. But might some of that variation also be attributed to inadequate resolution? Would a set of finer-zoned models show less variation or reveal structure in the noise?

Most of this paper is about the results of repeating the survey of Sukhbold & Woosley (2014) with finer zoning (Figure 1). In this section, however, we briefly examine, for just four masses, the question "What is enough?" In part, the answer must have a pragmatic aspect. If running with 3000 and 30,000 zones gives answers that differ in some important quantity by 5%, but changing the stellar mass by 0.01 M or some bit of uncertain stellar physics by a small fraction of its error bar alters the answer by 50%, perhaps 3000 zones is "good enough." Resources would be better spent studying the variations that depend on these other variables.

There is a deeper issue, though, when the answer is deterministic but unpredictable. It is an inherent aspect of chaos that tiny changes upstream produce large differences in outcome. As the noted chaos theorist Edward Lorentz once said, "Chaos exists when the present determines the future, but the approximate present does not approximately determine the future."6 As noted by Sukhbold & Woosley (2014) and Müller et al. (2016) and, as will be explored in greater depth here, for some ranges of mass, presupernova evolution is like that. Seemingly minuscule changes in initial conditions may determine whether a star in an important mass range explodes as a supernova or collapses to a black hole. How different could two outcomes be?

Table 1 gives the final presupernova properties of four sets of models with varying resolution for initial masses of 15.00, 15.01, 25.00, and 25.01 M. The spatial resolution ranges from about 1200 to 16,000 mass shells. The most finely resolved models used roughly twice as many time steps to reach the presupernova stage as the least-resolved models. The "compactness" parameter, ξ2.5 (O'Connor & Ott 2011), is inversely proportional to the radius enclosing the innermost 2.5 M of the presupernova core, while M4 and μ4 (Ertl et al. 2016, further discussed in Section 4.3) are, respectively, the Lagrangian location of the mass shell with entropy per baryon s = 4 kB and the derivative of the mass at that location evaluated over a mass interval of 0.3 M. The motivation for using these parameters is discussed in Section 4.3.

Table 1.  Presupernova Properties with Varying Resolution and Mass

MZAMS Nzones Nsteps MpreSN LpreSN RpreSN Teff,preSN MHe MFe ξ2.5 M4 μ4
(M)     (M) (1038 erg s−1) (1013 cm) (K) (M) (M)   (M)  
15.00B 1258 59992 12.935 3.57 6.04 3425 4.309 1.552 0.152 1.665 0.079
15.00C 2153 60162 12.866 3.58 5.97 3450 4.314 1.580 0.178 1.680 0.089
15.00D 4225 63435 12.804 3.59 6.06 3420 4.326 1.581 0.176 1.718 0.085
15.00E 8138 77578 12.485 3.69 6.19 3410 4.409 1.494 0.094 1.496 0.044
15.00F 15912 87612 11.717 4.02 6.64 3363 4.683 1.413 0.111 1.591 0.047
25.00A 874 38558 16.047 7.88 8.56 3505 8.342 1.609 0.303 1.915 0.103
25.00B 1885 38202 16.026 9.59 9.78 3445 8.351 1.620 0.301 1.911 0.102
25.00C 3358 40799 15.690 9.72 9.79 3455 8.427 1.606 0.301 1.916 0.102
25.00D 6515 45929 15.902 9.59 9.77 3445 8.382 1.630 0.301 1.910 0.102
25.00E 12734 58975 15.816 9.70 9.85 3440 8.432 1.643 0.303 1.903 0.104
15.01B 1260 59794 12.922 3.59 6.05 3425 4.319 1.561 0.164 1.675 0.083
15.01C 2149 60641 12.701 3.64 6.05 3435 4.361 1.513 0.189 1.742 0.088
15.01D 4190 63675 12.574 3.67 6.16 3410 4.386 1.357 0.126 1.457 0.067
15.01E 8103 76827 12.157 3.83 6.37 3390 4.523 1.506 0.100 1.517 0.046
15.01F 15952 77854 11.845 3.95 6.55 3373 4.638 1.438 0.159 1.599 0.070
25.01A 876 38416 16.138 7.88 8.56 3505 8.269 1.606 0.301 1.911 0.102
25.01B 1897 38165 16.319 9.57 9.80 3440 8.348 1.610 0.298 1.909 0.101
25.01C 3335 41211 15.141 9.85 9.82 3460 8.498 1.613 0.300 1.910 0.101
25.01D 6466 46494 15.235 9.81 9.89 3445 8.485 1.606 0.302 1.907 0.103
25.01E 12268 67699 13.264 10.34 9.86 3495 8.720 1.600 0.315 1.901 0.110

Note. All quantities are measured at the presupernova stage. Here Nzones is slightly larger at the beginning of calculation.

Download table as:  ASCIITypeset image

The zoning for the typical 15 M case in Figure 1 corresponds to the moderately zoned case (15.00D) in Table 1. Initial zoning in other 15.00 and 15.01 M cases was scaled at all locations in the star by roughly factors of two. For the 25.00 and 25.01 M models, the number of zones was proportionately larger. That is, Model 25.00C had roughly 25/15 as many zones as Model 15.00C, etc. There was no Model 25.00F; instead, there was a Model 25.00A, which had about half the zoning of 25.00B. The continuous rezoning parameters were set so as to preserve, within about 10%, the initial zoning. The number of zones given in Table 1 is for the presupernova star.

Two closely adjacent masses were studied at 15 and 25 M in order to compare the effects of fine zoning with slight variations in any other parameter of the problem. One might have chosen instead to vary the time step, overshoot, rate for 12C(α, γ)16O, semiconvection, mass loss, or rotation. Although no substitute for a full survey of such dependencies, varying the mass slightly in a situation where the solution is chaotic might be expected to send the calculation down one path or another. Stated another way, if varying the mass by less than 0.1% results in an answer significantly different than the one obtained by increasing the zoning by a factor of two, perhaps there is no reason to use much finer zoning for that mass. On the other hand, if all models give the same answer, one might have confidence in the uniqueness of that model, at least for an assumed set of stellar physics.

The two choices, 15 and 25 M, illustrate these two possibilities nicely. In all cases, the observable properties of the presupernova star, its luminosity and radius, and hence its effective temperature, are well determined. Changing opacities, composition, etc. would certainly cause variation, but the bulk observables of our models are well determined. There is a weak but noticeable trend to produce smaller presupernova stellar masses and larger helium core masses when the zoning is finer. As will be discussed later, this is a consequence of variation in the time spent as a red or blue supergiant, which, in turn, is sensitive to an uncertain treatment of semiconvection. A small number of episodic mixing events at the edge of the convective hydrogen-burning core can temporarily "rejuvenate" the core and, by affecting the gravitational potential of the hydrogen-burning shell, eventually cause the transition from blue to red to occur at different times. This trend does not emerge because the models are converging with higher resolution but rather as a result of enhanced mixing due to the uncertain treatment of convection physics.

The measures of presupernova core structure, on the other hand, vary by factors of two for the 15 M models but scarcely at all for the 25 M models. The reasons for the difference will be discussed in Section 5, but we note that any reasonable zoning suffices to get the structure of the 25 M model correct, but no reasonable zoning shows convergence for the 15 M model. Not knowing beforehand what the outcome will be for an arbitrary mass, here we used the maximum zoning that time and the desire to survey thousands of models allowed. This is most like the "D" models in Table 1. There is little motivation, however, for carrying out still finer resolution studies when a trivial change in mass, 15.00–15.01 M, causes more variation than an order of magnitude increase in zoning.

2.2. The Effects of Networks

Uncertainty in nuclear energy generation and neutronization during the various episodes of burning can be another source of variability in presupernova models. In this section, the effect of using either the standard 19 isotope approximation network plus silicon quasi-equilibrium (Section 2) or a much larger nuclear reaction network is explored. It should be noted that while the standard network carries the abundances of only 19 species, it has the power of a network roughly twice that size, since reactions through trace species like 23Na, 27Al, etc. are carried in a steady-state approximation (Weaver et al. 1978). Reactions coupling 54Fe to 56Ni are also included, as are approximations to energy generation by the CNO and pp cycles and to 14N(α, γ)18F(${e}^{+}\nu {)}^{18}{\rm{O}}$(α, γ)22Ne.7

The large network used for comparison here is drawn from an extensive nuclear database maintained and used for nucleosynthesis studies for decades (e.g., Woosley et al. 2002). Key reaction rates, such as 14N(p, γ)15O, 3α, 12C(α, γ)16O, the (α, γ) and (α, p) reactions on α-nuclei, and the weak interaction rates are the same in both networks. The network is "adaptive" (Rauscher et al. 2002) in the sense that isotopes are added or subtracted at each step to accommodate the reaction flow (e.g., there is no need to include a detailed network for the iron group during hydrogen burning). Usually, the settings on the network size are quite liberal, resulting in from 700 to 1200 isotopes being carried in the early and late stages of evolution, respectively (Woosley & Heger 2007). Here, because we are only interested in stellar structure and not, e.g., the s-process, a smaller number was carried, typically 300 isotopes extending up through the element selenium. All of the iron-group isotopes in the quasi-equilibrium network were included, and many more, so that core neutronization during and after silicon burning was equivalently calculated.

Network sensitivity was examined for four different masses of stars, 15.00, 15.01, 25.00, and 25.01 M. Two studies were carried out to test network sensitivity. Table 2 edits the evolution of the 15.00 star, similar to Model 15.00D in Table 1, at different times in its evolution using three approximations to the energy generation and neutronization. The Approx case corresponds to using the 19 isotope approximation network and quasi-equilibrium network, as is standard in the rest of this paper. The Coproc case carries the large network of about 300 isotopes along with the approximation network in "coprocessing mode." That is, the approximation and quasi-equilibrium networks are used for energy generation in the stellar structure calculation, but the network is also carried along in passive mode. The same time step is used at the same temperature and density to evolve, zone by zone, a much larger number of isotopes in parallel with the approximation network. Subcycling is done within a given time step to follow the abundances of trace isotopes accurately. Coprocessing is only performed in the code up to the point where the transition to quasi-equilibrium occurs in a zone, typically at oxygen depletion X(16O = 0.03). The output from the large coprocessing network is used to continually update the electron mole number, Ye, which feeds back into the structure calculation. Because Ye evolves slowly, no iteration is required. The energy generation from the large network can also be used to check the validity of the approximation network, though no correction is applied. In the third case, Full, the big network is coupled directly to the structure calculation throughout the entire life of the star, including silicon burning and core collapse. This might be deemed the most accurate approach, but it is slow, requiring an order of magnitude more computer resources, and prone to instability during silicon burning because of strongly coupled flows.

Table 2.  Central Properties for Different Network Treatments for the 15.00 M Model

Network $\mathrm{log}{T}_{{\rm{c}}}$ $\mathrm{log}{\rho }_{{\rm{c}}}$ ${Y}_{{\rm{e}},{\rm{c}}}$ $\mathrm{log}{S}_{{\rm{n}}}$ $\mathrm{log}| {S}_{\nu }| $
  (K) (g cm−3)   (erg g−1 s−1)
H burning
Approx 7.564 0.822 0.70042 5.104 3.910
Coproc 7.564 0.822 0.70196 5.104 3.910
      0.70196 5.107  
Full 7.565 0.820 0.69991 5.109 3.914
H depletion
Approx 7.676 1.138 0.50581 5.376 4.114
Coproc 7.676 1.138 0.50616 5.376 4.114
      0.50616 5.469  
Full 7.677 1.136 0.50489 5.387 4.129
He burning
Approx 8.249 3.127 0.49998 5.600 1.375
Coproc 8.250 3.123 0.49932 5.621 1.383
      0.49932 5.631  
Full 8.249 3.122 0.49932 5.621 1.371
He depletion
Approx 8.406 3.426 0.49998 5.624 2.560
Coproc 8.406 3.426 0.49932 5.627 2.562
      0.49932 5.667  
Full 8.407 3.422 0.49932 5.667 2.577
C ignition
Approx 8.703 4.544 0.49998 0.405 4.750
Coproc 8.704 4.546 0.49933 0.434 4.756
      0.49933 0.368  
Full 8.700 4.513 0.49933 0.159 4.729
C depletion
Approx 9.081 6.602 0.49998 6.193 8.198
Coproc 9.080 6.598 0.49894 6.181 8.202
      0.49894 6.108  
Full 9.079 6.580 0.49897 6.063 8.212
O ignition
Approx 9.176 6.895 0.49998 10.63 9.123
Coproc 9.176 6.899 0.49896 10.60 9.116
      0.49896 10.58  
Full 9.176 6.944 0.49899 10.57 9.188
O depletion
Approx 9.343 7.056 0.49998 11.53 11.05
Coproc 9.346 7.037 0.49242 11.60 11.12
      0.49242 11.64  
Full 9.352 6.990 0.49309 11.81 11.27
Si ignition
Approx 9.477 8.068 0.48679 12.14 11.21
Coproc 9.477 8.087 0.48609 12.18 11.19
Full 9.477 8.084 0.47945 11.51 11.29
Si depletion
Approx 9.583 7.699 0.47060 12.71 13.01
Coproc 9.571 7.706 0.46987 12.67 12.97
Full 9.477 7.739 0.46809 12.40 12.92
Presupernova
Approx 9.885 9.747 0.43848 (16.11) 16.24
Coproc 9.880 7.762 0.43747 (16.09) 16.20
Full 9.869 10.10 0.43266 (16.83) 16.94

Note. Here Sn is the energy generation rate excluding neutrino losses and is negative during the presupernova stage due to photodisintegration. Values in parentheses denote $\mathrm{log}| {S}_{{\rm{n}}}| $. The two values of Ye,c and Sn given for coprocessing calculations until core oxygen depletion (until quasi-equilibrium) are from the approximation network (top) and coprocessed big network (bottom).

Download table as:  ASCIITypeset image

In Table 2, up until oxygen depletion, two different numbers are given for the coprocessing run (Coproc). The first is the energy generation from the approximation network, which is used to calculate the stellar model, and the second is the output from the passively carried big network. Hydrogen burning and depletion correspond to central hydrogen mass fractions of 0.4 and 0.01. Helium burning and depletion are when the central helium mass fraction is 0.5 and 0.01. Carbon "ignition" is actually evaluated during the Kelvin–Helmholtz contraction between helium depletion and real carbon burning when the central temperature is 5 × 108 K, and carbon depletion is when the central carbon mass fraction is 1%. Similarly, "oxygen ignition" is when the central temperature during the Kelvin–Helmholtz contraction is 1.5 × 109 K and oxygen has yet to burn, and oxygen depletion is when the central oxygen mass fraction is 0.03. Silicon ignition is at 3.0 × 109 K, and silicon depletion is when the silicon mass fraction is 1%. Presupernova is when the core-collapse speed exceeds a maximum of 900 km s−1.

The table shows near-exact agreement in energy generation and Ye calculated using all three approaches up until at least oxygen ignition. The slight differences in energy generation at carbon ignition do not matter because neutrino losses (far right column) dominate at the time examined. By oxygen depletion, though, things are starting to mildly diverge. This divergence has three causes. One is the electron capture that goes on in the late stages of oxygen burning and starts to appreciably affect Ye. This change is not followed by the Approx network. Carrying Ye in the coprocessing run addresses most of this divergence; the values of Ye in the coprocessing and full network runs are nearly the same. Another effect, more difficult to disentangle, is the fact that, at late times, the stellar structure in the different calculations starts to diverge. It diverges, in part, because of the different nuclear physics, but even more because of the chaotic nature of the evolution. As was seen in the zoning study (Section 2.1), in certain initial mass ranges, two models with even slightly different physics will differ appreciably in final outcome. The third effect is the instantaneous difference due solely to the differing treatments of nuclear physics. This is what we are actually trying to study, but it is difficult to separate from cumulative effects due to structural changes. Even so, at silicon depletion, the three values of Ye agree very well, and the energy generation differs by a factor of two, which is mostly due to the higher temperature in the Approx case and the very sensitive temperature dependence of the silicon-burning reactions.

These small differences probably cause very little change in the presupernova model because the star has a certain amount of fuel to burn and the total energy release is set by known nuclear binding energies. The different burning rates only modestly affect the temperature at which the silicon burns in steady state (Woosley et al. 2002). We conclude that the approximation–plus–quasi-equilibrium approach is adequate for our survey and introduces errors that are small compared with other variations that occur when other uncertain quantities, or even the zoning, change.

Table 3 examines the effect of using either the large network or the approximation network plus quasi-equilibrium to study stars of 15.00, 15.01, 25.00, and 25.01 M, the same masses considered in the zoning sensitivity study (Section 2.1). Only one zoning was considered here, but the resolution was about three times greater in all four cases than in Sukhbold & Woosley (2014). Fewer time steps were taken during hydrogen burning than in the zoning sensitivity study. An overall limit of 5 × 1010 s was placed on the time step, as opposed to 1 × 1010 s in the survey and zoning sensitivity study. Here the emphasis is on comparing runs with different networks, not on the highest spatial and temporal resolution. Consequently, the 15.00 model in Table 3 is not directly comparable with the one in Table 1.

Table 3.  Presupernova Properties with Varying Network Size

MZAMS Network Nzones Nsteps MHe MCO 12Cign. MFe ξ2.5 M4 μ4 Ye,c
(M)       (M) (M)   (M)   (M)    
15.00 Approx 4233 32013 4.319 2.949 0.215 1.527 0.187 1.760 0.087 0.438
15.00 Coproc 4227 33294 4.319 2.946 0.215 1.537 0.179 1.742 0.084 0.437
15.00 Full 4185 39361 4.383 3.014 0.202 1.407 0.139 1.469 0.068 0.433
15.01 Approx 4235 33091 4.308 2.937 0.211 1.373 0.143 1.449 0.088 0.434
15.01 Coproc 4248 35736 4.298 2.925 0.216 1.387 0.135 1.421 0.097 0.432
15.01 Full 4194 36083 4.390 2.998 0.214 1.511 0.195 1.761 0.089 0.437
25.00 Approx 4717 30080 8.663 6.711 0.179 1.604 0.311 1.913 0.107 0.444
25.00 Coproc 4721 30143 8.670 6.722 0.180 1.580 0.306 1.891 0.106 0.443
25.00 Full 4589 42642 8.179 6.322 0.179 1.806 0.240 1.807 0.084 0.438
25.01 Approx 4696 30317 8.633 6.720 0.176 1.617 0.312 1.912 0.108 0.444
25.01 Coproc 4681 30120 8.640 6.670 0.176 1.564 0.289 1.854 0.110 0.443
25.01 Full 4531 43698 8.411 6.513 0.180 1.835 0.252 1.837 0.088 0.438

Note. All quantities are measured at the presupernova stage, except 12Cign (Section 2.2). Here Nzones is slightly larger at the beginning of calculation, and Ye,c is evaluated at the center.

Download table as:  ASCIITypeset image

Table 3 gives, in addition to the presupernova zoning and time steps, the masses of the iron, carbon–oxygen, and helium cores; the central mass fraction of carbon evaluated just prior to carbon ignition; the central value of the electron mole number, Ye, in the presupernova star; and the descriptors of presupernova core structure, ξ2.5, μ4, and M4. Though not given, the presupernova masses, radii, luminosities, and effective temperatures are in the same range as the same-mass stars in Table 1.

Based solely on the changes in the core structure and iron core mass for the 15.00 and 15.01 M models, one might conclude that using a large network made a substantial difference, and all future runs should take much greater care with the nuclear physics. Though the models do indeed differ, this would not be a valid inference. As shown in Section 2.1 and discussed elsewhere in the paper, changing anything for models in some mass ranges can give qualitatively different answers for core structure. The differences between runs with large and small networks is within the range of differences resulting from increased or decreased zoning (Table 1) or a small change in the star's mass. More telling is the fact that the central carbon abundances, the helium and carbon–oxygen core masses, and, most critically, the central value of Ye all agree very well. The approximation network is doing a fine job representing the nuclear physics. It is just that other, uncontrollable factors introduce large variations in the core structure of a 15 M, and to a lesser extent, 25 M star. We conclude that the 19 isotope approximation network and quasi-equilibrium hypothesis are quite adequate for surveys like this. Instead, the focus should be placed on studying the impact of stellar physics on the statistical properties over the entire mass range.

2.3. Boundary Pressure

Especially during its post-main-sequence evolution, the radius of a model star is sensitive to surface boundary conditions. Different codes treat the surface in different ways. Many use a boundary condition on pressure or density that sometimes includes a reduction of gravity by the Eddington factor and the inertia term from wind acceleration (e.g., Appenzeller 1970; Heger 1998; Paxton et al. 2015). Others solve the wind equation for conditions at the sonic radius (Grassitelli, private communication, 2018). Still others fit the surface to a stellar atmosphere of varying complexity (e.g., plane-parallel gray or tables; e.g., Chieffi & Straniero 1989; Paxton et al. 2011). KEPLER uses a constant boundary pressure, Pbound, that does not vary during the evolution. The advantage of such an approach is its simplicity, but care must be taken that Pbound is not so large as to greatly alter the solution.

Ideally, Pbound only influences the structure of a tiny mass near the surface. No reasonable value of Pbound affects main-sequence evolution, for example, because the pressure gradient there is very steep. The gradient becomes shallower during the RSG phase, though, and even slightly different radii there can significantly alter the mass loss rate. The timing and development of the surface convection zone is also affected. Both can have a significant effect on the star's location in the HR diagram and, to a lesser extent, the helium core mass and presupernova structure. Traditionally, studies with KEPLER have used Pbound = 50–100 dyne cm−2 (e.g., Woosley et al. 2002; Woosley & Heger 2007; Sukhbold & Woosley 2014). An exception is the work of Müller et al. (2016), where a much larger value, ∼3500 dyne cm−2, was used.

In the present study, a value of 50 dyne cm−2 is employed. Is this low enough? To explore the sensitivity of the results to boundary pressure, a standard 15.00 M, Model S15.00D in Table 1, was calculated using a variety of boundary pressures from 10 to 6400 dyne cm−2 maintained throughout the evolution. The resulting HR diagrams and radial histories are shown in Figure 3. For Pbound ≲ 500 dyne cm−2, the trajectory in the HR diagram is essentially identical for six different choices of ${P}_{\mathrm{bound}}$. The presupernova radius and luminosity do not vary. For larger values of Pbound, though, especially Pbound ≳ 1000 dyne cm−2, significant variation is found.

Figure 3.

Figure 3. Top panel: HR diagram for variations of Model 15.00D (Table 1) in which the surface boundary pressure is varied from 10 to 6400 dyne cm−2 in roughly factor of two steps. The inset shows that the evolution is insensitive to this boundary condition until the star becomes an RSG. The main figure shows nearly identical results for pressures ≲500 dyne cm−2 but significant variations for larger values. Bottom panel: radius as a function of time during the final evolution of the same models. Only the last 1.4 Myr, when the stars become RSGs, are shown. Models with boundary pressures less than 500 dyne cm−2 have nearly identical radii as supergiants until very close to the end, while those with bigger pressures vary. The time spent as RSGs is similar.

Standard image High-resolution image

Even the models with small surface pressures show some variation in final properties. For example, for Pbound between 10 and 400 dyne cm−2, the final star mass varied between 12.59 and 12.61 M, while the helium core mass varied between 3.61 and 3.67 M. Some of this variation is a consequence of the irreducible noise in running any model twice with even small variations in the physics or mass, but part could be a residual sensitivity to ${P}_{\mathrm{bound}}$.

Closely related to the treatment of surface boundary conditions is surface zoning. In our recent studies (Sukhbold & Woosley 2014; Müller et al. 2016), rezoning was allowed to continue all the way to the surface of the star throughout its evolution. This often resulted in coarse zoning near the photosphere during the RSG stage. Fine zoning on the main sequence was lost to dezoning when the temperature and density gradients were shallow. See Figure 1 for a sample comparison with the older model. In the new models, rezoning was not allowed in the outer 250 zones, typically 0.3 M. Thus, the fine surface zoning shown in Figure 1 was maintained throughout the evolution for all models. Typical zoning near the surface is less than 5 × 10−4 M.

The use of finer surface zoning and a reduced boundary pressure generally gave larger radii and increased mass loss during the RSG stage, especially as compared with Müller et al. (2016). Because the helium core mass was not appreciably affected, though, the final core compactness, nucleosynthesis, and remnant masses are not altered by this change.

3. The New Survey

Three grids of models were calculated for initial main-sequence masses in the range 12–60 M using the zoning described in Section 2 and three choices of mass loss rate. Since all stars retained at least a small envelope mass as presupernova stars, the relevant mass loss rate was that of Nieuwenhuijzen & de Jager (1990). The first set of models uses that rate unmodified (${\dot{M}}_{{\rm{N}}}$). This results in so much of the envelope being lost by stars above about 27 M that the residual envelope, typically a few M, becomes unstable and develops density inversions of more than an order of magnitude. We suspect that massive stars with extremely large radii and low-mass envelopes become unstable in nature at this point and lose their remaining hydrogen rapidly (e.g., Sanyal et al. 2015, 2017). Mass loss is not the focus here, though. In order to have a greater range of helium core masses, a second set of models used half of the standard mass loss (${\dot{M}}_{{\rm{N}}}/2$) appropriate for solar metallicity and encountered no instability up to 40 M. Finally, a third set was calculated with one-tenth the standard mass loss (${\dot{M}}_{{\rm{N}}}/10$) to provide a sparse mass grid between 12 and 60 M. Considering that mass loss prescriptions are probably uncertain (Renzo et al. 2017; Beasor & Davies 2018), the factors of 2 and 10 could reflect a sensitivity study, or they might be appropriate for stars with reduced metallicity.

Altogether, 1,499 models were calculated in the mass range 12–27 M using the standard mass loss rate and 2799 models from 12 to 40 M using half that value. The grid spacing was 0.01 M for both sets. The third set, with one-tenth the standard mass loss rate, consisted of only 49 models between 12 and 60 M with a mesh size of 1 M. Lower-metallicity stars were not considered, but, except for zero, metallicity would have affected the answer chiefly by changing the mass loss and hence hydrogen envelope mass of the presupernova star. As we shall see, the helium core mass is not entirely independent of the remaining envelope mass, but it is insensitive.

The code was compelled to spend a much greater number of time steps during hydrogen and helium burning by imposing a limit on the maximum time step of 1010 s. A minimum of roughly 10,000 steps was thus spent burning hydrogen on the main sequence. This helped to weaken the impact of random semiconvective mixing events that affected the final envelope mass and, to a smaller extent, the helium core mass. It also improved the accuracy of the treatment of convection and burning as "split" operations (convection and burning are not implicitly coupled in the code). Very little difference was noted in a few test cases when this maximum step was increased to 5 × 1010 s.

Minor changes to improve the stability of the code when transitioning to silicon quasi-equilibrium and the convergence criteria were also incorporated. These had insignificant effects on the outcome but improved code performance. Models for the main survey were all calculated on identical processors using the same version of the code and compiler so as to reduce any possible noise introduced by machine architecture (Ohio Supercomputer Center 1987).

One major goal of this work was to explore the effect of resolution on calculations of presupernova structure. The number of zones used in the main survey was typically 4000 × (MZAMS/12 M). Stars more massive than 30 M thus had over 10,000 initial mass shells. In all cases, the maximum mass zone anywhere in the star was, at all times, ≲0.01 M, but zoning was by no means uniform. The initial zoning (Figure 4), which continued to characterize most of the star until the presupernova stage, had much finer zoning in the helium core, down to about 0.001 M at the star's center.

Figure 4.

Figure 4. Total number of zones in the presupernova star as a function of main-sequence mass for the model sets with a standard mass loss rate ${\dot{M}}_{{\rm{N}}}$ (blue), reduced rate of ${\dot{M}}_{{\rm{N}}}/2$ (green), and reduced rate of ${\dot{M}}_{{\rm{N}}}/10$ (black). In order to maintain fine resolution in stars of increasing mass, the number of shells becomes greater. The lower curves show the zoning inside the helium, carbon–oxygen, and iron cores. These new models retain more zones only in the iron core, as the earlier studies did in the whole star.

Standard image High-resolution image

Except as noted above for hydrogen burning, the time-step criteria were not varied from the previous studies. Nevertheless, the finer resolution of the new runs did result in taking more time steps as smaller zones experienced mass loss or were cycled through the large density contrast at the base of the hydrogen convective shell. Abundances within a given zone convectively coupled to many other zones also required more time steps to change. The typical number of steps for a new calculation is 45,000, about twice the earlier studies' ∼25,000.

The presupernova structure and composition tables for all models from the main sets with ${\dot{M}}_{{\rm{N}}}$, ${\dot{M}}_{{\rm{N}}}/2$, and ${\dot{M}}_{{\rm{N}}}/10$ are available through the Harvard Dataverse archive.8

4. Results

4.1. Observable Properties

The observable properties of the new models are summarized in Figures 5 and 6 and Table 4. Final (presupernova) masses are sensitive to both the uncertain mass loss rate prescription (Nieuwenhuijzen & de Jager 1990) and the uncertain history of the stellar radius. The results vary considerably. For lower-mass stars, the spread in final mass is smaller, since the star only loses a small amount of mass. For bigger stars, though, a large fraction of the hydrogen envelope may be lost. Most of the mass loss occurs during helium burning, and most of that occurs during the late stages when the central helium mass fraction is less than 0.5.

Figure 5.

Figure 5. Stellar masses during helium burning and for presupernova stars. The upper panel shows the results for standard mass loss (${\dot{M}}_{{\rm{N}}}$) at helium ignition (pink) when 1% of the helium has burned to carbon, helium burning (orange) when the central helium mass fraction is 50%, helium depletion (red) when the central helium mass fraction is 1%, and the presupernova star (blue). The presupernova and helium depletion masses are almost indistinguishable. Most of the mass loss, and most of the dispersion in the presupernova mass, arises due to radius expansion after substantial helium has already burned. The lower panel shows the larger presupernova masses expected when the mass loss rate is reduced by a factor of two. Models with reduced mass loss generally have lower dispersion, but both show the dispersion increase with increasing initial mass (thus increasing the amount of mass lost). The presupernova helium core masses show much less variation (see Figure 7).

Standard image High-resolution image
Figure 6.

Figure 6. Top panel: location in the HR diagram for the presupernova stars that used the standard mass loss rate (${\dot{M}}_{{\rm{N}}}$). Bottom panel: Teff at helium depletion (red) and for the presupernova stars (blue). The presupernova radii vary almost linearly from 5 × 1013 to 10 × 1013 cm as the mass increases from 13 to 26 M. The stars all die with a nearly constant effective temperature between 3400 and 3600 K. Observations would select RSGs prior to helium depletion, so the helium depletion curve (red) in the lower panel is a lower bound to what is likely to be observed for Teff.

Standard image High-resolution image

Table 4.  Properties of Three Main Model Sets

MZAMS MpreSN MHe MCO logLpreSN Teff,preSN Teff,He dep RpreSN
(M) (M) (M) (M) (erg s−1) (K) (K) (1013 cm)
${\dot{M}}_{{\rm{N}}}$ (binned by 1 M)
12 11.19 3.35 2.20 38.39 3440 3670 4.97
13 11.72 3.79 2.53 38.47 3420 3610 5.49
14 12.22 4.20 2.85 38.54 3410 3570 5.97
15 12.73 4.61 3.19 38.60 3405 3545 6.41
16 13.23 5.01 3.53 38.65 3405 3535 6.80
17 13.74 5.42 3.87 38.69 3410 3535 7.15
18 14.12 5.84 4.24 38.73 3420 3530 7.45
19 14.60 6.24 4.60 38.77 3425 3530 7.72
20 14.87 6.66 4.98 38.81 3430 3535 8.07
21 14.97 7.09 5.35 38.85 3430 3535 8.51
22 15.09 7.49 5.71 38.90 3430 3540 8.95
23 15.11 7.90 6.05 38.94 3435 3550 9.36
24 14.72 8.34 6.44 38.98 3450 3565 9.73
25 14.49 8.75 6.80 39.01 3470 3590 10.01
26 14.27 9.15 7.15 39.04 3500 3620 10.20
${\dot{M}}_{{\rm{N}}}/2$ (binned by 1 M)
12 11.82 3.39 2.22 38.40 3460 3685 4.95
13 12.59 3.82 2.55 38.47 3450 3630 5.44
14 13.35 4.24 2.88 38.54 3440 3595 5.90
15 14.12 4.65 3.22 38.60 3440 3580 6.31
16 14.87 5.07 3.58 38.65 3440 3570 6.71
17 15.62 5.49 3.93 38.70 3450 3565 7.04
18 16.37 5.91 4.30 38.74 3460 3565 7.32
19 17.01 6.36 4.69 38.78 3465 3560 7.65
20 17.66 6.79 5.08 38.82 3465 3560 8.04
21 18.27 7.22 5.45 38.87 3460 3560 8.49
22 18.69 7.67 5.85 38.92 3455 3560 9.00
23 19.25 8.07 6.20 38.96 3455 3565 9.42
24 19.58 8.52 6.59 39.00 3455 3565 9.86
25 19.95 8.94 6.94 39.03 3455 3570 10.23
26 20.14 9.38 7.31 39.06 3460 3575 10.60
27 20.36 9.80 7.67 39.09 3465 3585 10.92
28 20.66 10.19 7.99 39.11 3470 3590 11.19
29 20.61 10.65 8.38 39.14 3480 3600 11.50
30 20.72 11.08 8.73 39.17 3495 3610 11.75
31 20.83 11.49 9.10 39.19 3510 3625 11.97
32 20.85 11.95 9.48 39.21 3525 3640 12.18
33 20.98 12.37 9.86 39.23 3540 3655 12.37
34 21.03 12.85 10.27 39.26 3560 3675 12.55
35 21.08 13.33 10.70 39.28 3580 3690 12.74
36 21.36 13.69 11.01 39.29 3590 3705 12.86
37 21.40 14.18 11.46 39.31 3610 3725 12.99
38 21.63 14.56 11.80 39.33 3625 3735 13.15
39 21.80 14.98 12.18 39.34 3635 3750 13.29
${\dot{M}}_{{\rm{N}}}/10$ (not binned)
12 11.88 3.21 2.09 38.36 3480 3725 4.69
15 14.76 4.46 3.07 38.58 3465 3615 6.05
20 19.48 6.66 4.96 38.81 3500 3590 7.76
25 24.00 8.86 6.89 39.02 3495 3605 9.88
30 27.75 11.39 8.95 39.17 3510 3615 11.73
35 31.55 13.89 11.05 39.29 3535 3625 13.17
40 35.25 16.23 13.19 39.37 3575 3660 14.23
45 36.65 21.41 18.22 39.53 3610 3650 16.73
50 41.40 21.21 17.87 39.52 3635 3710 16.28
55 44.27 23.73 20.20 39.58 3655 3715 17.26
60 47.31 25.64 22.00 39.62 3680 3740 17.79

Note. All quantities were measured at the presupernova stage except Teff,He dep, which was measured when the helium mass fraction in the core was depleted to 1%. All Teff values were rounded to multiples of 5.

Download table as:  ASCIITypeset image

Since the luminosity does not vary greatly during helium burning, the amount of mass lost and its uncertainty depends mostly upon the radial history of the star: R0.8 in the Nieuwenhuijzen & de Jager (1990) formalism, i.e., the amount of time the helium-burning star spends as a blue supergiant (BSG) or RSG. It is well known that semiconvective and overshoot mixing play a key role in determining the ratio of time spent as each (Lauterborn et al. 1971). In the present calculations, small changes in individual semiconvective mixing episodes affect the timing of the development of a deep surface convection zone, and hence the movement of the star to the red. With more semiconvection, the star spends a greater fraction of its helium-burning lifetime as a BSG and hence loses less mass, ending its life with a larger hydrogen envelope still intact. However, RSGs have the converse behavior. Less semiconvection, e.g., strictly Ledoux convection, favors a longer time as an RSG and thus reduces the threshold for a massive star to lose its envelope. Though the spread in Figure 5 looks large, the uncertainty is usually a small fraction of the total mass lost. For example, a 25 M star with the standard mass loss rate might end up as a 12 or 15 M star. That is, it might lose 10–13 M, a range of about 25%. Lower-mass models have a smaller variation.

The final masses for stars calculated using ${\dot{M}}_{{\rm{N}}}$/2, of course, are larger. Bigger stars on the main sequence then retain more envelope by the time they die, and we were thus able to determine the core structure of more massive stars. For the most massive models considered, the final hydrogen envelope mass was roughly 4 M for the ${\dot{M}}_{{\rm{N}}}$ models and 5 M for the ${\dot{M}}_{{\rm{N}}}/2$ models. For smaller envelope masses, the radius expanded beyond 1.5 × 1014 cm and became unstable in the KEPLER code due to recombination. In these cases, the envelope could only be retained on the star by the addition of a large, unrealistic boundary pressure that resulted in gross density inversions in the outer envelope. We suspect that this is a real instability in nature, that once the envelope mass decreases below a critical value in a very massive star and the radius extends beyond 10 au, the remaining envelope is lost through nonsteady processes (Sanyal et al. 2015, 2017). The luminosity in the hydrogen envelope is a substantial fraction of the Eddington value. This interesting possibility is deferred for a later study.

Since a larger pressure can give a smaller radius and reduced mass loss, different choices of boundary pressure can lead to significant differences in the final mass. The boundary pressure used in Müller et al. (2016), for example, several thousand dyne cm−2, was larger than what we now believe reasonable and much larger than that used in the present study, 50 dyne cm−2. This probably accounts, at least partly, for the larger final masses found in the Müller et al. (2016) study and their ability to study higher-mass stars using a single mass loss prescription.

This boundary pressure does not appreciably affect the compactness of the cores for stars of given main-sequence masses, however. This is because the large spread in final star masses (Figure 5) is not reflected in the helium core mass (Section 4.2) or the luminosity of the star (Figure 6). Since the core structure, which is the main focus of this paper, is chiefly sensitive to the helium core mass and not the envelope mass, the spread in final masses seen in Figure 5 does not affect our major conclusions.

Figure 6 and Table 4 give the final luminosities and effective temperatures of our presupernova models. The luminosity of the presupernova star depends chiefly on the helium core mass and is approximately given by LpreSN ≈ 5.77 ×1038 (MHe/6 M)1.5 erg s−1. The effective temperature, essentially bounded by the Hayashi limit, remains pegged at close to ∼3500 K for all presupernova stars. Thus, the radius of the star increases as ${M}_{\mathrm{He}}^{0.75}$. Except for a few presupernova stars, most RSGs will be observed during their helium-burning stage, where their effective temperatures will be hotter. Our calculations show a systematic ∼200 K offset between the effective temperatures at helium depletion and presupernova stars. Earlier in helium burning, the temperature will be even hotter; thus, Teff ≈ 3550 K should be a lower bound to what is seen. Larger values are also expected for subsolar metallicity and lower values for presupernova stars. These numbers are in good accord with measurements of the hottest RSGs (Levesque et al. 2005; Davies et al. 2013).

4.2. Core Masses

Figure 7 shows the helium, carbon–oxygen (CO), and iron core masses for all of our presupernova models with ${\dot{M}}_{{\rm{N}}}$ and ${\dot{M}}_{{\rm{N}}}/2$. The dispersion in helium and CO core masses for a given main-sequence mass is small, much less than the hydrogen envelope masses inferred in Figure 5. This implies that the helium and CO core masses are only slightly affected by the assumed mass loss rate and probably not very sensitive to metallicity. The helium core mass as a function of main-sequence mass is approximately ${M}_{\mathrm{He}}\approx 6.46\ {({M}_{\mathrm{ZAMS}}/20)}^{1.27}$, which can be combined with the previous relation in Section 4.1 to give LpreSN ≈ 6.5 × 1038 (MZAMS/20 M)1.92 erg s−1. Since presupernova core compactness is chiefly a function of helium core mass (Sukhbold & Woosley 2014), these results suggest a near-universal dependence of presupernova core structure on initial mass, provided mass loss does not remove the entire hydrogen envelope.

Figure 7.

Figure 7. Helium, CO, and Fe core masses for all presupernova stars from sets with ${\dot{M}}_{{\rm{N}}}$ and ${\dot{M}}_{{\rm{N}}}/2$. Despite significant variations in mass loss (Figure 5), the final helium and CO cores are well determined by the star's initial mass and a standard choice of stellar physics. Note multiple branches for the iron core mass below 19 M.

Standard image High-resolution image

Iron core masses increase, on average, with increasing stellar mass, reaching a maximum of about 2.0 M for the most massive stars studied (>40 M). Still larger iron cores, up to about 2.5 M, characterize more massive stars in the pulsational pair-instability range (70–140 M; Woosley 2017). For stars below 23 M, the iron core mass is markedly multivalued for stars with nearly the same initial mass. This reflects the operation of multiple shells of carbon and oxygen burning, as will be discussed further in Section 5. The two major branches of iron core masses below 20 M, which are most of the stars that leave neutron star remnants, might result in bimodality in the neutron star mass function.

4.3. Core Structure

4.3.1. Measures of "Explodability"

Early theoretical studies of supernovae noted a strong correlation of the rapidly declining density outside the iron core with the degree of difficulty encountered in trying to explode the star using the neutrino energy transport (e.g., Burrows & Lattimer 1987; Fryer 1999). O'Connor & Ott (2011) introduced a simple, single-parameter measure of this density decline called the "compactness parameter,"

Equation (3)

where R(Mbary = M) is the radius of the Lagrangian mass shell enclosing mass M in the presupernova star. The fiducial mass is often chosen as the innermost 2.5 M, so that for a wide range of initial masses, it not only encloses the iron core but samples enough of the overlying shell material around it. Though the parameter is defined to be evaluated at the time of bounce, it is more often measured at the time of presupernova (when the collapse begins), since the systematics are insensitive to slightly different fiducial choices (Sukhbold & Woosley 2014). Subsequent studies by Ugliano et al. (2012), O'Connor & Ott (2013), and Sukhbold et al. (2016) showed strong correlation between the ease with which model stars exploded and the ξ parameter, in the sense that stars with small ξ, i.e., steep density gradients outside the iron core, were easier to explode using a standard, albeit approximate, set of supernova physics. Although this is a useful starting point, a single parameter conveys limited information about the structure of the core, and more physics-based representations followed.

In particular, Ertl et al. (2016) suggested an alternative two-parameter characterization based upon M4, the mass coordinate in solar masses, where the entropy per baryon reaches 4.0 kB, and the radial gradient, μ4, of the density at that point. In practice, μ4 was obtained by evaluating the change in mass over the change in radius between several mass shells separated by 0.3 M in the vicinity of M4, i.e.,

Equation (4)

where dm = 0.3 M. Smaller values of μ4 thus imply steeper density gradients (less change in enclosed mass when the radius changes). The quantity M4 has long been used to locate the steep density decline often associated with a strong oxygen-burning shell in the presupernova star (e.g., Woosley & Heger 2007). Where the entropy per baryon abruptly rises at nearly constant temperature, the density declines. When the core collapses, the arrival at the neutrinosphere of lower-density matter reduces the "ram pressure" and facilitates the launch of an outgoing shock. Besides its location, it is helpful to characterize the rate of the density change, which is the role played by μ4.

It is reasonable that higher neutrino luminosities and smaller accretion rates on the proto-neutron star should favor explosion. Ertl et al. (2016) made the case that μ4, multiplied by its radius (i.e., the radius at which M4 is found) and divided by a collapse timescale, is a surrogate for the accretion rate at the time of explosion. The product μ4M4 is similarly a surrogate for the accretion luminosity. This assumes, as the models suggest, that the radius where the neutrino luminosity is generated does not vary greatly with mass. Explosion is thus favored by small μ4 (accretion rate) and large μ4M4 (luminosity). Ertl et al. (2016) determined a simple condition expressed as a straight line,

Equation (5)

such that models that lay below the line (lower μ4) were more likely to explode than those above. The values k1 and k2 were determined from several large-scale simulations and leading models for SN 1987A. For one representative set, k1 = 0.194 (M)−1 and k2 = 0.058 ("N20" model in their Table 2).

More recently, Müller et al. (2016) adopted a different approach using a semi-analytic model for the explosion based upon presupernova properties. By using a fuller description of the presupernova star than afforded by just one or two parameters, they were able to approximately estimate not only the success or failure of the explosion but also the explosion energy, ejected nickel mass, and compact remnant mass. This approach does not include, in its present form, several important pieces of explosion physics, such as proto-neutron star cooling, self-consistent fallback, and explosive nuclear burning, and thus it is not as powerful as the parameterized simulations such as Sukhbold et al. (2016). Its main power lies is its ability to rapidly calculate general trends in explosion properties that depend on presupernova structure. Their approach uses five physically motivated parameters to determine the outcome. Here we analyze our new presupernova models using their standard values: βexpl = 4, ζ = 0.7, αout = 0.5, αturb = 1.18, and τ1.5 = 1.2 (see their Table 1). For the definitions of these parameters and further technical details, see Müller et al. (2016).

4.3.2. Results for the Compactness Parameter

Figure 8 shows the compactness parameter, ξ2.5, i.e., ξ for our three model sets, measured at a fiducial mass of 2.5 M at the time the presupernova collapse speed reaches 900 km s−1. One notable feature seen in Sukhbold & Woosley (2014) is the relatively large "scatter" of the compactness parameter between 18 and 22 M, which is now shifted to roughly 14 and 19 M (panel (a)) with updated neutrino losses (Section 2). As has long been noted (e.g., Timmes et al. 1996), ∼20 M marks the transition region between exoergic convective central carbon burning at lower masses and endoergic radiative burning at higher ones. That is, at high mass, the energy produced by carbon fusion does not exceed the neutrino losses in the star's center, and it proceeds from helium depletion to oxygen ignition on a short Kelvin–Helmholtz timescale. As a result, subsequent burning—especially in shells, where the net energy generation is positive—proceeds differently. The carbon convective shells move inward (with increasing initial mass) until this transition point and are weaker and greater in number (Barkat 1994).

Figure 8.

Figure 8. (a) Compactness parameter for a fiducial mass of 2.5 M evaluated at the presupernova stage shown for models below <19 M as a function of initial mass for the ${\dot{M}}_{{\rm{N}}}$ set. Even where rapid variations occur, the scatter is not wholly random. Points tend to concentrate in one of several solutions. (b) Same points as in panel (a), now shown as a function of helium core mass. The branching solutions are more clearly seen. The range of ξ2.5 seen in this mass range includes stars that might explode or implode, and thus both outcomes are possible for stars with nearly identical initial mass. (c) Compactness for all three values of mass loss. Slight shifts between the sets are due to the growth of the helium core by hydrogen-shell burning, resulting in a slightly larger compactness for the models with reduced mass loss. (d) Same points as in panel (c), now shown as a function of helium core mass. Much of the shift seen in panel (c) is gone, since the helium core mass is a chief determinant of the core structure.

Standard image High-resolution image

Quite noticeable in the new results and the prior work of Müller et al. (2016) are concentrations of points in the compactness parameter plot, which were not clearly seen in our earlier studies due to much coarser increments in mass space. Points are not randomly scattered between some local maximum and minimum value. Note, for example, the existence of two, and possibly three, discrete solutions for helium core masses near 4.8 M (panel (b)). Multiple branches also exist at other masses but with less clarity. For a helium core mass of 5.6 M, the compactness parameter for stars with nearly identical masses varies by a factor of three, from ξ2.5 = 0.1 to 0.3. Such a large variation spans the range of stars that might explode or collapse into black holes and suggests that both outcomes are possible for relatively low-mass stars of almost the same initial mass.

When all three new sets are plotted as a function of initial mass (panel (c)), the models with lower mass loss rates are slightly shifted with respect to models with higher mass loss rates. The shifts are hardly noticeable at a lower initial mass (<20 M), but they grow with increasing mass, and for our most massive models, the ${\dot{M}}_{{\rm{N}}}/10$ set is shifted by a few M with respect to the ${\dot{M}}_{{\rm{N}}}/2$ set. The primary reason for this shift is that for a given initial mass, models with lower mass loss rates attain slightly larger He cores and slightly different compositions. There is also a general underlying scatter, similar to that of the total star mass shown in Figure 5, some of which results from the slight change of He core mass due to the semiconvective effects discussed in Section 4.1.

Panel (d) of Figure 8 shows the compactness as a function of He core mass for all of our models. Since the initial mass scales cleanly with He core mass (Figure 7), the structure of the curve stays nearly identical; however, the abovementioned shifts and scatters are mostly gone. The He core mass and, within it, the CO core mass are the chief determinants of core structure, not the final supernova mass, which includes the hydrogen envelope. Uncertainties in mass loss rates are thus not particularly relevant to the compactness, except for determining the main-sequence mass above which the entire envelope is lost.

As was noted by O'Connor & Ott (2011) and studied many times since, the compactness plotted against mass is nonmonotonic and highly structured, with two distinct peaks in the vicinity of 21 and 35 M (Sukhbold & Woosley 2014). A shallow dip occurs near 50 M, followed by a slow rise at still higher masses (Sukhbold & Woosley 2014), until finally, the pulsational pair domain is reached near 70 M. As discussed in Sukhbold & Woosley (2014) and Sukhbold (2016), this structure is crafted out of an overall gradually rising curve by the operation of an advanced-stage shell-burning episode of one fuel modulating the core-burning episode of the next fuel. In particular, the structure near the first peak is primarily driven by the effect of shell carbon burning on core oxygen burning, while the structure near the second peak is driven by the effect of shell oxygen burning on core silicon burning. At very high mass, oxygen burning ultimately ignites a strong shell outside of 2.5 M, and therefore the compactness stays high.

Figure 9 compares the new results with those of Sukhbold & Woosley (2014) and Müller et al. (2016) and illustrates these points. Most notable is a shift of the new models above ∼14 M downward in mass by about 2–3 M compared with that of Sukhbold & Woosley (2014). This is a consequence of fixing the bug in the neutrino losses as discussed in Section 2 and not due to differing resolution. Models below 14 M are affected very little, but the more massive stars change appreciably. The region of variability that was between 18 and 22 M in Sukhbold & Woosley (2014) is now shifted down to 14–19 M. If anything, that variability is now greater with the finer resolution, perhaps due to the larger number of models surveyed.

Figure 9.

Figure 9. Presupernova compactness compared with previous results for solar-metallicity stars obtained by Sukhbold & Woosley (2014; top panel) and Müller et al. (2016; bottom panel). The new models differ little from those of Sukhbold & Woosley (2014) below 14 M, but the peak in compactness formerly at 23 M is shifted downward to about 20.5 M. The "noise" noted by Sukhbold & Woosley (2014) from 18 to 22 M is also shifted downward to 14–19 M; now, with finely incremented models, it shows several concentrations of points suggestive of a multivalued solution. The bottom panel shows the comparison with the work of Müller et al. (2016). Since the ${\dot{M}}_{{\rm{N}}}$ and ${\dot{M}}_{{\rm{N}}}/2$ sets are nearly identical at lower mass (Figure 8), for clarity, we have plotted the ${\dot{M}}_{{\rm{N}}}$ set for lower mass and the ${\dot{M}}_{{\rm{N}}}/2$ set for higher mass. The agreement between these models and those from Müller et al. (2016) is substantial, even though the current survey uses more models and carries more zones and time steps per model. Reducing the mass loss (see text) shifts the results toward slightly lower main-sequence mass.

Standard image High-resolution image

Indeed, the comparison with Müller et al. (2016), who used the corrected neutrino loss rate and sampled many more models, is excellent, both in the location of structures and the range of ξ2.5 within those structures. This agreement persists despite the use of 4–10 times more mass shells in each of the new models and a radical decrease in the surface boundary pressure. The latter affected the mass lost by the star but did not appreciably affect the helium core masses. There is no reason to believe that still finer zoning, smaller time steps, or a different reaction network will greatly alter these results, unless the code physics itself (reaction rates, semiconvection, rotation, etc.) is changed. The studies of Müller et al. (2016) and the present work are mutually confirming.

4.3.3. New Results and the Ertl Parameterization

An important subsidiary question is whether ξ2.5 is really the best measure for presupernova core structure. Might some of the variability seen in Figure 8 be simply because of the choice of a single arbitrary point in the star to sample its structure? Perhaps other parameterizations might give less variability and more reliable predictions? In particular, ξ2.5 is sensitive to recent shell activity in the vicinity of 2.5 M that might not always describe well what went on deeper inside.

The Ertl parameterization of our results is given in Figure 10. As previously discussed (Section 4.3), the points beneath the dashed line represent models that are more likely to explode in a simple neutrino-transport scheme. Multivalued solutions are clearly seen, especially for stars with μ4M4 less than 0.25. In some ranges, the difference between solutions is enough to significantly affect the probable outcome of the explosion. Stars with μ4M4 less than 0.25 typically have masses less than 20 M. For more massive stars—and, in particular, for μ4M4 greater than 0.25—explosion seems quite unlikely. The reasons for the multivalued solution are discussed in Section 5.

Figure 10.

Figure 10. "Explodability" according to the sample "engine" model from Ertl et al. (2016). The Ertl parameters, μ4 and μ4M4, are shown for the new models with standard mass loss rates. See text for definitions. Points beneath the dashed line are more likely to explode than those below. Multiple solutions are clearly visible for μ4M4 below 0.25 (M)−1. The scatter in this figure is a lot less than that in Figure 9.

Standard image High-resolution image

Also striking is the much tighter clustering of points in the Ertl representation of our new models (Figure 10), in contrast with an equivalent plot of "compactness" (Figure 8). The results are less noisy and show less overall variability. This is due to the obvious correlation between two Ertl parameters on one hand, but it also presumably reflects the better representation of core structure by a two-parameter, physics-based representation than a single parameter anchored to a single point in the core. It also suggests that much of the "noise" in Figure 8, especially for initial masses of 14–19 M, may not be so much a consequence of "nonconvergence" of the models but rather a poor representation of the results.

5. Interpretation

Why is the core structure of presupernova stars nonmonotonic and multivalued for some ranges of mass? The short answer is that the advanced-stage evolution of massive stars involves two to three carbon-burning shells (plus central core burning for models below ∼19 M) and one or two oxygen shells (plus core burning). Combinations of these shells lead to variable outcomes but not a continuum of all possibilities, because the shells have finite sizes and their number is an integer. The transitions are abrupt, and small changes upstream in the strength or extent of one or more shells can send a star down one path or another. This is especially true for stars below 19 M, where there are more carbon-burning shells.

Globally, ξ2.5, μ4, and M4 all increase gradually with mass. Higher-mass stars have greater entropy in their middles and are less degenerate in their final stages. Greater degeneracy leads to an increased central concentration of the mass and "core convergence." For lighter stars, the presupernova structure more resembles a white dwarf embedded in a low-density envelope, where the density declines rapidly at the edge of the central "white dwarf." This effect is clearly at work in the lightest stars surveyed. Below 13 M, compactness defined at a mass of 2.5 M has little meaning as a measure of explodability, since the fiducial point is not even inside of the helium-burning shell. For stars in this mass range, the helium-burning shell always lies outside 2 × 109 cm, thus guaranteeing a small compactness parameter. For a 10 M presupernova star, 2.5 M is even outside the helium core and in the hydrogen envelope. For all parameterizations of core structure considered in this paper, these light stars will always explode and need no further discussion here.

Degeneracy remains an important consideration, though, even for the larger stars. A 12 M presupernova star has a core that is degenerate (degeneracy parameter η ≡ μ/kT > 0) out to 1.46 M. For 30 M, degeneracy extends out to 1.87 M. Even for 100 M, the iron core is degenerate out to 2.1 M. Most of the increase in degeneracy occurs between carbon depletion and oxygen depletion in the core when the core is efficiently cooled by the neutrinos (Sukhbold & Woosley 2014). Degeneracy affects the core structure, making the (thermally adjusted) Chandrasekhar mass relevant for all stars that are likely to explode as common supernovae. On the other hand, the shells outside what will be the iron core in the presupernova star—and, in particular, the carbon- and oxygen-burning shells—are always nondegenerate. Strong burning thus leads to expansion that affects both ξ2.5 and μ4.

The upper panel of Figure 11 shows a generally strong correlation between the compactness parameter, ξ2.5, and M4, except at very high and low masses. A correlation is expected, since a sharp density decline close to the iron core requires a large radius to enclose 2.5 M, thus small ξ2.5. The sharp density decline implies a deeper location where the entropy per baryon exceeds 4.0 kB and thus a smaller M4 and a steeper mass gradient at that point as well. The reversal of the plot for the most massive models is a consequence of both the increasing central entropy and the outward migration of the first oxygen shell. This migration causes a nonmonotonic dependence on mass for both M4 and ξ2.5, as shown in Figure 8, but M4 starts to decline at a slightly lower initial mass due to increasing entropy. Therefore, when the location of M4 recedes from ∼2.5 to 2 M, the ξ stays roughly constant and results in the reversal near ∼35 M.

Figure 11.

Figure 11. Compactness enclosing the innermost 2.5 M at the time of presupernova for the ${\dot{M}}_{{\rm{N}}}/2$ set of models plotted as a function of M4, the Lagrangian coordinate of the first mass shell where the entropy per baryon exceeds 4.0kB. For lower-mass models, three major branches are clearly apparent below M4 ≈ 1.75 M. The lower panel shows a zoomed-in version of the upper panel, characterizing models with MZAMS < 20 M and with color contours denoting their corresponding helium core masses.

Standard image High-resolution image

The lower panel of Figure 11 shows that the behavior of compactness and M4 for stars lighter than about 20 M is clearly multivalued. Three distinct branches are apparent for the lighter stars. There is no reason to believe that a more dense grid of calculations would randomly fill out the spaces between the branches.

Figure 12 delves deeper into this correlation between compactness and M4 and helps to understand why both are multivalued relations of mass for moderate-mass stars (13 M < MZAMS < 19 M). The first panel shows M4 in the presupernova star as a function of helium core mass. A cleaner pattern results from using the helium core mass instead of the main-sequence mass, since it eliminates the variations due to envelope mass loss (Figure 5). As the figure shows, M4 is usually pegged to the "oxygen shell," by which we mean the location in the presupernova star where the energy generation from oxygen fusion (excluding neutrino losses) is at maximum. This is frequently, though not always, interior to the boundary of the "silicon core," inside of which the silicon mass fraction is greater than the oxygen fraction. Though one might naively assume that the oxygen-burning shell is at the edge of the silicon core, this is not generally true. There are often gradients on the silicon-to-oxygen ratio left behind by receding convection or radiative burning. In these cases, the oxygen shell lies inside the silicon core. On the other extreme, in some stars, the oxygen shell can be so active as to merge with the carbon- and neon-burning shells, producing one large convective shell where all three are burning vigorously and the silicon core and oxygen shell are coincident. When this occurs, the compactness is usually small (Figure 16 of Sukhbold & Woosley 2014). Many of the pink, brown, and green points in the top panel of Figure 12 are of this sort.

Figure 12.

Figure 12. Top: M4 vs. helium core mass for presupernova stars with main-sequence masses between 13 and 19 M. Pink points, which usually underlie green stars or brown pentagons, indicate M4. Green stars mark the locations of the most active oxygen-burning zone, and brown pentagons are where the silicon mass fraction first exceeds the oxygen mass fraction, going inward. For clarity, brown and green symbols not closely affiliated with M4 within 0.01 M have not been plotted. Usually, the M4 point is found at or near the maximum oxygen-burning zone. Several clusters of points, sometimes lying in nearly straight lines, are apparent. Middle: compactness parameter, ξ2.5, plotted for the same helium cores, equivalent to panel (b) of Figure 8. Note also a clustering of points here and a high degree of correlation with M4 in the top panel. Bottom: μ4 as a function of helium core mass. Note the almost completely congruent pattern with ξ2.5, since both measure the most active burning shell at the time of presupernova.

Standard image High-resolution image

It is not surprising that the oxygen-burning shell, the strongest burning shell during the late stages of stellar evolution, is usually the location of a jump in entropy. The oxygen fusion rate is very temperature-sensitive, so oxygen burns at a nearly constant temperature. The overlying star expands, decreasing the local density and thus increasing the entropy. The quantity μ4 in the Ertl parameterization is a measure of the strength of this burning.

The middle panel of Figure 12 shows that the behavior of ξ2.5 in this same mass range (equivalent to panel (b) of Figure 8) correlates reasonably well with that of M4, including, approximately, the location and extent of clusters of solutions. Within a cluster, there is usually a quasi-linear relation with a well-defined slope. A larger value of M4 implies a larger value of ξ2.5. The farther out in the star the strong burning shell, the less centrally concentrated its density. This correlation is not perfect, though. Sometimes long lines in the M4 plot, e.g., the long string of points from MHe = 4.1 to 5.2 M, break into several segments with different ξ2.5. This is partly due to the arbitrary pinning of ξ2.5 to a single mass shell, but also because specifying M4 alone does not measure the strength of the burning there.

Instead, as might be expected, the bottom panel of Figure 12 shows that ξ2.5 correlates better with μ4, the mass gradient at M4. This correlation is very strong, since both quantities are sensitive to the strength of the most active burning shell in the last days of the star's life. Why, though, are M4 and μ4 not randomly scattered between their extrema? Why the patterns? The top panel of Figure 12 suggests a reason. Here M4 traces the location of the strongest oxygen-burning shell. Oxygen burns there because it is at the deepest location that has not already depleted oxygen by prior shell burning. That location is, in turn, set by the extent of oxygen-burning shell(s) during the previous evolution, which in turn depends upon the entropy structure set up during carbon shell burning.

Sukhbold & Woosley (2014) pointed out this correlation between the core compactness and the extent of the first oxygen-burning convective shell (their Figure 14). Figure 13 shows a related quantity for the new model set. The figure shows the mass of the "silicon core," the mass interior to which oxygen has burned out, at the time silicon burning ignites in the star's center. Here a specific central temperature, 3.0 × 109 K, was chosen in order to make the plot, but the conditions reflect what the oxygen-burning shell (or shells) have accomplished prior to silicon ignition. Visual inspection of the convective history of these 700 models shows that the silicon core mass at silicon ignition is very nearly equal to the maximum extent of the second convective oxygen-burning shell for stars below ∼14.6 M (helium core mass 4.2 M) and the first oxygen convective shell for more massive stars up to at least 20 M. In this plot, we see the clearest evidence yet for regular but nonmonotonic and occasionally multivalued behavior. Two helium cores of very nearly 4.2 M can give rise to presupernova cores with structures in one of two well-defined states. Slight shifts in mass, composition, or even numerical approach (including zoning) can send the star one way or another.

Figure 13.

Figure 13. Size of the oxygen-depleted core (oxygen less than silicon or iron) as a function of helium core mass at silicon ignition (when the central temperature first reaches 3.0 × 109 K). This is approximately the maximum extent of the first oxygen convective shell. The function has multiple trajectories that correlate with the major features in the plot of M4 vs. helium core mass in the top panel of Figure 12.

Standard image High-resolution image

Further analysis of the convective histories reveals the systematics behind this behavior—at least in the 1D code, if not in nature.

  • 1.  
    The little jump at MHe = 3.7 M (MZAMS = 13.2 M) from MSi = 1.7 to 1.6 M reflects the operation of the third convective carbon-burning shell. Below this mass, the third shell ignites inside the former full extent of the second shell and outside the effective Chandrasekhar mass. Thus, the core oxygen burning starts to burn in a smaller extent while this outermost carbon shell operates. For masses above this value, the third shell ignites at the outer boundary of the previous shell and within the Chandrasekhar mass. Core oxygen burning now has to wait until this third shell is complete; therefore, the base of this third shell sets (approximately) the extent of the first oxygen convective shell and thus the base of the second one (top left panel of Figure 14).
  • 2.  
    The much larger decrease in MSi from 1.7 to 1.4 M at MHe = 4.2 M (MZAMS = 14.6 M) is due to the diminished significance of the second oxygen-burning shell. The silicon core shrinks to the extent of the first oxygen-burning shell, though the second shell continues to be sporadically important for a time (bottom left panel of Figure 14).
  • 3.  
    The jump and wild variations starting at MHe = 5.2 M and extending up to 5.7 M (MZAMS = 17.2–19.0 M) mark the transition from convective carbon burning to radiative burning. After the transition near ∼19 M, carbon no longer burns exoergically at the center of the star, and there are only two convective shells before oxygen burning rather than three or more. During this transition, carbon core and shell burning vary greatly in location and extent. The large spread in MSi and, ultimately, in compactness reflects the irregularity of this transition (right panels of Figure 14).

Figure 14.

Figure 14. Convection histories of four sample models from the range of initial masses from 14 to 19 M, representing key structural changes that are responsible for the significant variations in the final presupernova core properties. Each panel shows the evolution of the innermost 4.5 M material roughly during the last 1000 yr of its life, i.e., from core carbon burning until presupernova. Colors denote energy generation (red) and energy loss (blue) gradients. Hatched black regions are convective episodes. The x-axis is shown as the log of the time until core collapse, tcc. The initial mass of each model is denoted in the top left corners of the panels. Inserted mini plots illustrate the compactness parameter (purple) and the Lagrangian location of entropy per baryon equal to 4.0 kB (green), both as a function of helium core mass corresponding to the abovementioned initial mass range. The crosses inside each mini plot denote the values corresponding to the model. Top left: The lowest-mass stars have four or more convective carbon-burning episodes followed by three oxygen-burning episodes (including a central core-burning episode). The shell helium burning lies within the 2.5 M location, the point on which the compactness parameter, ξ2.5, is measured, and thus these models have small compactness and small M4. Bottom left: Central carbon burning weakens with increasing mass, and as a result, the shell carbon-burning episodes gradually "migrate" inward (with increasing initial mass). Once the third shell carbon burning ignites within the effective Chandrasekhar mass, the core oxygen burning has to wait until the overlying carbon-burning episode is finished, which leads to significantly weakened or, in some cases, absent immediate second oxygen shell burning. In these models, oxygen eventually burns later when silicon is already ignited in the core. This marks the transition of M4 to the "second" branch. Due to its sensitivity to the last major shell-burning episode, the compactness parameter has multiple solutions depending on the final configuration of oxygen, neon, and carbon burning (i.e., together burn vigorously in one shell or separately). Top right: Around ∼18 M, it becomes increasingly harder for carbon to burn convectively near the center. The tiny convective central-burning episode is followed by a long-lasting radiative flame, and as a result, with increasing initial mass, the base of the convective carbon-burning shells starts to migrate outward. Bottom right: Eventually, at high enough initial mass, all central carbon burning is radiative. Models between 17 and 19 M have rapidly changing core structures and thus rapidly varying ξ2.5 and M4. The same format plots for all 800 models between 12.00 and 19.99 M are available online (see footnote 9).

Standard image High-resolution image

Other shell interactions result in the weaker multiple branches seen in higher-mass stars in Figure 8. Despite the specific masses given in the list above, none of these transitions are abrupt, and, given slight nudges, the star may oscillate from one to the other solution when it is close to boundaries. This leads to the "multivalued" behavior. Although complicated and probably sensitive to the one-dimensional treatment of the problem, the conclusion is that the presupernova structure in stars from 13 to 19 M results from an interplay of convective carbon- and oxygen-burning shells after carbon ignition.

Putting these various factors together, a deterministic picture emerges. Multiple carbon-burning stages—core burning below 19 M and two or more episodes of shell burning—act to sculpt an entropy distribution in the core so that oxygen shell burning, when it occurs, ignites at and extends to various mass shells. Whether the number of carbon shells is two or three or four and the number of oxygen shells is one or two strongly affects the core structure of a presupernova star in the mass range where a significant number of events are observed. These changes in the extent of convective shells, which amplify small differences in the earlier evolution, may end up determining whether the star explodes or makes a black hole.

6. Compact Remnants and Explosions

Since the new models differ from those of Sukhbold & Woosley (2014) and Sukhbold et al. (2016) and, to a lesser extent, Müller et al. (2016), it is worth revisiting some of the conclusions of those papers regarding compact remnants using the new models.

Figure 15 shows the models now expected to explode based on the criteria of Ertl et al. (2016) and Müller et al. (2016). This figure can be compared with Figure 4 of Ertl et al. and Figure 6 of Müller et al., which it very closely resembles. The "N20" engine parameterization was used to make the comparison with Ertl et al. (2016), and the "standard" choices of five parameters were used for the Müller et al. (2016) method.

Figure 15.

Figure 15. The compactness parameter shown in Figure 8 is plotted again twice, color-coded as to the success or failure of the explosion based upon the parameterization of Ertl et al. (2016; top) and the semi-analytical method of Müller et al. (2016; bottom). Red symbols denote successful explosions by these criteria, and black symbols denote failures. The very good agreement between these two different approaches suggests that both are good representations of core structure, though the criterion of Müller et al. (2016) favors slightly more explosions for the parameters chosen.

Standard image High-resolution image

The most significant difference with the Ertl criterion is the shift of the pattern for models above about 14 M to slightly (∼10%) lower masses. Changes in average quantities, like the remnant masses, are small, however. Figure 16 shows a comparison of IMF-weighted remnant mass distributions of imploding models assuming a Salpeter (1955) initial mass function with a power of −2.35. The usual two cases are considered: (a) collapse of the helium core but ejection of the hydrogen envelope and (b) collapse of the entire presupernova star. Case (a) is more appropriate for stars where the envelope is lost to a binary companion or, for very massive stars, to winds. The envelope might also be lost to a very weak explosion that did not unbind the helium core (Quataert & Shiode 2012; Lovegrove & Woosley 2013; Fuller 2017). Case (b) is for more robust explosions. In each case, there is an element of uncertainty because the final mass depends on the mass loss rate, e.g., for RSGs in case (b) and Wolf–Rayet stars in case (a). In making Figure 16, but not in computing the averages, only the mass range covered by our ${\dot{M}}_{{\rm{N}}}$ mass loss survey, 12–27 M, was included.

Figure 16.

Figure 16. The IMF-weighted black hole mass distributions for the new ${\dot{M}}_{{\rm{N}}}$ set are shown in comparison with the results from Sukhbold et al. (2016) over the same initial mass range between 12 and 27 M. The differences are small and well within the range of variation seen for different explosion models. The average black hole masses for the older study are 15.9 and 6.95 M for two scenarios, while the new models yield 14.3 and 6.50 M.

Standard image High-resolution image

Astronomers, of course, observe black holes coming from all masses of stars, not just 12–27 M. To compute the averages for solar-metallicity stars (and the answer will be sensitive to metallicity), the new results were supplemented with the prior models of Sukhbold et al. (2016) for stars more massive than 27 M. For just the limited mass range 12–27 M, the average black hole masses from Sukhbold et al. (2016) were previously 6.95 and 15.9 M, respectively, for the helium core and full star assumptions. The corresponding new numbers are 6.50 and 14.3 M. Considering the entire mass range of stars that experience iron core collapse, and using the N20 parameters in Sukhbold et al. (2016), the previous averages were 9.35 and 13.7 M. The equivalent new numbers are 8.25 and 13.7 M. A less than 1% adjustment has been applied to Table 6 of Sukhbold et al. (2016) based upon a slightly different way of interpolating the grid.

Similar corrections can be estimated for the neutron star gravitational masses. Lacking an explosion model, it is assumed, based on prior experience (i.e., Ertl et al. 2016), that the baryonic mass of the resulting neutron star is usually equal to M4. After appropriately correcting for neutrino mass loss (same as in Müller et al. 2016), the new average neutron star gravitational mass in the range 12–27 M is 1.45 M. Explosions below 12 M were calculated by Sukhbold et al. (2016) using presupernova models in which the neutrino bug had been fixed. They can thus be combined with the current set. Assuming neutron star production above 27 M is negligibly small; the new global average neutron star mass is 1.38 M. In Table 6 of Sukhbold et al. (2016), the corresponding global average was 1.41 M.

The agreement of Figure 15 with the earlier work (Figure 6 of Müller et al. 2016) is excellent, and it is thus expected that the explosion outcomes will also be very similar. Indeed, plots of neutron star mass (Figure 17), black hole mass, and explosion energy as a function of initial mass (not shown) are virtually indistinguishable from the panels in their Figure 2. This suggests that the difference in envelope structure and zoning between the series of Müller et al. (2016) and the present paper did not have much influence on the core structure and statistical explosion properties.

Figure 17.

Figure 17. Neutron star mass distribution resulting from analyzing the new model set with normal mass loss (${\dot{M}}_{{\rm{N}}}$) for stars in the mass range 12–27 M using the approach of Müller et al. (2016). Top: baryonic mass of the expected neutron star plotted as a function of M4, the mass where the entropy per baryon reaches 4.0kB in the presupernova star (Section 5). Neutron stars on the gray dashed line have a mass before neutrino losses equal to M4. Many points are found here because M4 is usually the "mass cut" in a successful explosion, unless the final entropy profile is significantly more shallow. Bottom: gravitational mass (adjusted from the baryonic mass in the same way as in Müller et al. 2016) of the expected neutron star as a function of main-sequence mass. More massive neutron stars are typically made by more massive main-sequence stars with a greater ξ2.5 and μ4, but the most massive points on this plot come from models with MZAMS ∼ 14–15 M, the most massive models with a significant second oxygen shell burning. The existence of multiple branches for these quantities (Figures 12 and 13) results in a large range of neutron star masses being accessible by stars of nearly the same main-sequence mass for some initial masses. Note also the cluster of neutron stars with mass near 1.65 M coming from the most massive supernova progenitors.

Standard image High-resolution image

Figure 15 shows that a significantly larger fraction of stars in the interesting mass range 15–20 M explode using the Müller et al. (2016) formalism, even though "N20" is one of the more energetic formulations of the Ertl et al. (2016) model. Conservatively, this could be regarded as an uncertainty in outcome until more realistic simulations of the actual explosion can be done. The various energies in the Ertl et al. (2016) model came, however, from calibrating a central engine with 1D neutrino transport, a shrinking proto-neutron star, and fallback to SN 1987A using various presupernova models and might, for now, be considered the more realistic of the two. This calibration to SN 1987A has its own uncertainties, though, especially since the structure of the presupernova star could be different in more realistic binary merger models (e.g., Menon & Heger 2017 and references therein).

Figure 17 gives the expected neutron star masses that result when our new presupernova models are analyzed using the formalism and standard parameters of Müller et al. (2016). Fallback is neglected in their analysis, so the baryonic mass of the proto-neutron star is equal to the mass coordinate where the neutrino-driven engine shuts off a few hundred ms after shock revival. As mentioned before, this correlates strongly with M4 (top panel). There are branches of neutron star masses above the M4 line, which primarily originate from less massive progenitors (MZAMS < 15 M), reflecting the operation of a prominent second oxygen-burning shell and the resulting shallow entropy profile. The most massive neutron stars have a baryonic mass that is bounded by the base of the convective carbon-burning shell, which is located much further out than M4 in the presupernova star. The bottom panel of Figure 17 shows the expected neutron star (gravitational) mass as a function of main-sequence mass. These results agree quite well with those of Müller et al. (2016; see their Figure 2(c)), including the existence of very massive neutron stars produced for main-sequence stars of 14–15 M and the highly variable nature of the solution between 14 and 19 M. For the majority of models, the resulting neutron star mass is tightly correlated with M4 (Figure 12) and the oxygen-depleted core at the time of silicon ignition (Figure 13), and in certain initial mass ranges, the explosion of models with nearly identical mass (or slightly different input physics) can result in very different neutron stars.

It is not expected that the nucleosynthesis and light curves calculated by Sukhbold et al. (2016) will be significantly altered by using the new models, though further exploration is certainly encouraged. In particular, the deficiency of light s-process elements seen in Sukhbold et al. (2016) will persist in the new model set, since most of the production of these elements is due to the most massive stars (Brown & Woosley 2013), which still fail to explode. The fraction of solar-metallicity stars above 9 M exploding as supernovae was 74% based on the N20 parameterization (Table 6 of Sukhbold et al. 2016) and is now 65%. Above 18 M, the fraction of supernovae reduces to only 8%, which actually is in slightly better accord with the results from Smartt (2009, 2015), as compared to the previous study. Since the envelope masses have not changed appreciably and the explosion energies are also expected to be unchanged, the light curves will be unaltered. The fractions given in their Table 6 for supernovae above 12, 20, and 30 M will also probably not change within the variations already seen for the different central engine characteristics.

7. Conclusions

The full evolution of over 4000 massive stars of solar metallicity in the mass range 12–60 M has been studied using unprecedented resolution, both in zones per star and in number of stars within a given mass range. The mass loss rate was varied, and an important bug in the neutrino loss routine was repaired that caused a significant variation in the presupernova properties from those calculated by Sukhbold & Woosley (2014) and prior works. Our chief conclusions are as follows.

  • 1.  
    The pattern of core compactness seen in previous studies (e.g., Sukhbold & Woosley 2014; Müller et al. 2016) is robust. The "noise" in these studies was not a consequence of inadequate zoning but reflected real variability. The range of variation and the location of peaks (Figure 8) in the new study are virtually identical to those seen by Müller et al. (2016), even though the new models use 4–10 times the zoning and a much smaller surface boundary pressure. The results are also qualitatively similar to those of Sukhbold & Woosley (2014), but important peaks in the compactness plot are shifted downward by about 10% in initial mass (Figure 9) due to the corrected neutrino loss rate. For the stellar physics used and spherically symmetric nature of the calculation, the variations and peaks seen in Figure 8 are now well determined.
  • 2.  
    The large variation in core structure seen for stars between 14 and 19 M by Sukhbold & Woosley (2014) is not completely random. For a larger set of models with finer spacing in initial mass, several branches of solutions emerge. This behavior was also seen by Müller et al. (2016). The branches apparently result from variations in the locations of the oxygen-burning shells in the presupernova star, which in turn "remember" the location of several carbon-burning shells of variable extent. Some noise is introduced by the fact that the carbon abundance and size of the carbon–oxygen core are not precisely monotonic with mass (Figure 7). This, in turn, reflects the operation of semiconvection at the boundary of the hydrogen and helium convective core. In the mass ranges 14–19 and 22–24 M, the presupernova core structure is more sensitive to events in the last year of a star's life (and sometimes the last hours) than to the star's initial mass. Presupernova structure in these mass ranges does not necessarily follow the Vogt–Russell theorem (Russell 1917; Vogt 1926). It may be more comparable to weather on Earth.
  • 3.  
    Even accounting for these systematics, the branches of the solutions are still noisy. The level of noise is reduced if one characterizes the presupernova star by its helium core mass, not its starting mass or final total mass. The scatter is also reduced in the Ertl two-parameter characterization of core structure rather than the O'Connor-Ott compactness parameter.
  • 4.  
    Given the large number of models, it is possible to give statistically meaningful results for the radius, luminosity, and effective temperature of supernova progenitors (Section 4.1).
  • 5.  
    The mass distributions of neutron stars and black holes resulting from supernovae in the mass range studied are not greatly altered from our earlier surveys (Section 6). The average gravitational mass of neutron stars, including all masses of supernovae, is now 1.38 M. The average black hole mass, if only the helium core implodes, is 8.61 M. If the entire presupernova star collapses, the average black hole mass is 13.5 M. The fraction of stars above 9 M that explode rather than collapsing is estimated to be 65%. The fraction above 18 M is 8%. Nucleosynthesis and light curves are basically unchanged.

Binary mass exchange and rotation undoubtedly play key roles but have been neglected in this study. To first order, stars that end up with the same final helium core mass will have similar presupernova compactness and fates. Similar systematic variations and multivalued solutions are expected to persist. The statistical averages of compact remnant masses may vary, however, and certainly the light curves will be different. Our models are publicly available to those wanting to estimate outcomes based upon their own distribution of helium core masses, carbon–oxygen core masses, etc. In the future, we will consider rotating models, but this was principally a study of how resolution affects the solution to a well-defined, frequently studied problem.

Our results suggest a slightly different strategy to the study of presupernova evolution and supernova modeling than sometimes used in the past. Given the variation in outcome for stars of nearly the same mass, or the same mass with different codes, a statistically meaningful sample of models must be calculated before drawing strong conclusions about supernova mass ranges, remnant mass distributions, nucleosynthesis, etc. Historically, researchers have sometimes focused on the calculation of just a few masses, e.g., 15, 20, and 25 M, and sought to test the sensitivity to changes in physics in just those cases. Here we see that calculating a statistically meaningful sample may be just as important as getting the physics precisely right. The size of such a sample depends upon the need to resolve regions of rapid variability found with a given code and physics, but a minimum initial mass grid of 0.1 M is reasonable within such regions.

Some of the models calculated here had merged carbon and oxygen convective shells at the end. Many did not. Others were only separated by a single thin zone from being coupled. Similar to the numerical artifact seen when helium burning is calculated with too little semiconvection, these mathematical bifurcations may not be real and might be overcome, or at least smoothed out (Alexakis et al. 2004), by increasing the convective overshoot or doing a multidimensional calculation (e.g., Meakin & Arnett 2007; Viallet et al. 2013; Arnett & Meakin 2016; Cristini et al. 2017; Jones et al. 2017). In general, linked convective shells give a structure with a smaller compactness parameter, ξ, and favor explosions (see, e.g., Collins et al. 2018). Further study with other representations of convective overshoot mixing and multidimensional codes are thus encouraged.

Semiconvection and overshoot mixing remain major uncertainties in studies of this sort and play important roles in determining the helium and carbon–oxygen core masses. How KEPLER treats semiconvection is described in Weaver & Woosley (1993). It is expected that other studies using different treatments will find results that differ in important details from those presented here. The overall pattern, multipeaked structure, and range of variability of the compactness parameter and other measures of core structure should persist, however. Clusters of solutions due to variable numbers and extents of the carbon and oxygen shells should also be a common feature. Further exploration is again encouraged.

All of the new presupernova models presented here are available online. Also online are plots of the convective histories of 800 models between 12 and 20 M of the $\dot{M}$ set (see footnote 9). Other auxiliary presupernova models (e.g., those presented in Sections 2.1 and 2.2) are available upon request to the authors.

We thank Raphael Hirschi for a careful review of this work and for providing many useful suggestions. We also thank Bernhard Müller for helping us to process our new progenitor results through his semi-analytical explosion modeling and feedback on this manuscript. Thomas Janka and Thomas Ertl provided valuable insight into the supernova explosion models and their dependence on progenitor properties. We thank Todd Thompson and John Beacom for useful comments on the manuscript. We also appreciate the efforts of Gang Gao in discovering the neutrino bug in the KEPLER code. All numerical KEPLER calculations presented in this work were performed on the RUBY cluster at the Ohio Supercomputer Center (Ohio Supercomputer Center 1987). TS is partly supported by NSF PHY-1404311 to John Beacom. SW is supported by NASA NNX14AH34G. AH is supported by an ARC Future Fellowship, FT120100363, and, in part, by the National Science Foundation under grant No. PHY-1430152 (JINA Center for the Evolution of the Elements).

Footnotes

Please wait… references are loading.
10.3847/1538-4357/aac2da