One-, Two-, and Three-dimensional Simulations of Oxygen-shell Burning Just before the Core Collapse of Massive Stars

, , , , , and

Published 2019 August 7 © 2019. The American Astronomical Society. All rights reserved.
, , Citation Takashi Yoshida et al 2019 ApJ 881 16 DOI 10.3847/1538-4357/ab2b9d

Download Article PDF
DownloadArticle ePub

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

0004-637X/881/1/16

Abstract

We perform two- (2D) and three-dimensional (3D) hydrodynamics simulations of convective oxygen-shell burning that takes place deep inside a massive progenitor star of a core-collapse supernova. Using a one-dimensional (1D) stellar evolution code, we first calculate the evolution of massive stars with an initial mass of 9–40 M. Four different overshoot parameters are applied, and a CO-core mass trend similar to previous works is obtained in the 1D models. Selecting eleven 1D models that have a coexisting silicon and oxygen layer, we perform 2D hydrodynamics simulations of the evolution for ∼100 s until the onset of core collapse. We find that convection with large-scale eddies and the turbulent Mach number of ∼0.1 is obtained in the models having a Si/O layer with a scale of 108 cm, whereas most models that have an extended O/Si layer up to a few ×109 cm exhibit lower turbulent velocity. Our results indicate that the supernova progenitors that possess a thick Si/O layer could provide the preferred condition for perturbation-aided explosions. We perform the 3D simulation of a 25 M model, which exhibits large-scale convection in the 2D models. The 3D model develops large-scale ( = 2) convection similar to the 2D model; however, the turbulent velocity is lower. By estimating the neutrino emission properties of the 3D model, we point out that a time modulation of the event rates, if observed in KamLAND and Hyper-Kamiokande, could provide important information about structural changes in the presupernova convective layer.

Export citation and abstract BibTeX RIS

1. Introduction

From theory and observations, it is almost certain that the explosions of massive stars as core-collapse supernovae (CCSNe) are generically multidimensional (multi-D) phenomena (see Foglizzo et al. 2015; Janka et al. 2016; Patat 2017 for reviews). To facilitate the neutrino-driven mechanism of CCSNe (Bethe & Wilson 1985), multi-D hydrodynamic instabilities such as neutrino-driven convection and the standing accretion shock instability (Blondin et al. 2003) play a pivotal role in enhancing the neutrino-heating efficiency to trigger the onset of the explosion. In fact, a growing number of self-consistent models in two or three spatial dimensions (2D, 3D) now report the revival of the stalled bounce shock into explosion for a wide mass range of progenitors (see, e.g., Hanke et al. 2013; Takiwaki et al. 2014; Lentz et al. 2015; Melson et al. 2015b; Nakamura et al. 2016; Roberts et al. 2016; Summa et al. 2016; Müller et al. 2017; O'Connor & Couch 2018; Vartanyan et al. 2019 for collective references therein).

These successes, however, provide further motivation for exploring missing ingredients in the neutrino mechanism, partly because the estimated explosion energies obtained in the multi-D models generally do not reach the typically observed value (e.g., ∼1051 erg; Tanaka et al. 2009). Various possible candidates to obtain more robust explosions were recently proposed, including multi-D effects during the final stage of the presupernova evolution (see Couch 2017 for a review), general relativity (GR; e.g., Kuroda et al. 2012, 2016; Müller et al. 2012; Ott et al. 2013), rapid rotation (e.g., Marek & Janka 2009; Suwa et al. 2010; Takiwaki et al. 2016; Summa et al. 2018; Harada et al. 2019) and/or magnetic fields (e.g., Obergaulinger et al. 2006; Mösta et al. 2014; Guilet & Müller 2015; Masada et al. 2015; Obergaulinger & Aloy 2017), and sophistication in the neutrino opacities (Melson et al. 2015a; Bollig et al. 2017; Burrows et al. 2018; Kotake et al. 2018) and in the neutrino transport schemes (e.g., Sumiyoshi & Yamada 2012; Richers et al. 2017; Just et al. 2018; Nagakura et al. 2018). In this work, we focus on the first item listed in the above list.

Couch & Ott (2013) were the first to demonstrate that the inhomogeneities seeded by convective shell burning foster the onset of a neutrino-driven explosion (see also Fernández et al. 2014; Couch & Ott 2015; Müller & Janka 2015; Burrows et al. 2018). This is because the infalling perturbation that could be amplified in the supersonic accretion (Nagakura et al. 2013, 2019; Takahashi & Yamada 2014) enhances the turbulence behind the postshock material, leading to the reduction of the critical neutrino luminosity for shock revival (e.g., Müller & Janka 2015; Abdikamalov et al. 2016). In these studies, the nonspherical structures in the burning shells, although physically motivated, were treated in a parametric manner, due to the paucity of the multi-D stellar evolution models covering the life span of massive stars up to the iron core collapse. Currently, one-dimensional (1D) stellar evolution calculations are the only way to accomplish this (Woosley et al. 2002; Woosley & Heger 2007; Sukhbold et al. 2018), where the errors introduced from the omission of multi-D effects are absorbed into the free parameters of the MLT, namely the mixing length theory, (e.g., Kippenhahn et al. 2012).

Truly multi-D hydrodynamics stellar evolution calculations have been done over several turnover timescales of convection (limited by the affordable computational resources) in selected burning shells (e.g., Meakin & Arnett 2007; Viallet et al. 2013; Campbell et al. 2016; Cristini et al. 2017; Cristini et al. 2019 for different burning shells, and see Arnett & Meakin 2016 for a review). Pushed by the observation of SN1987A, 2D and 3D stellar evolution simulations focusing on the late burning stages have been extensively carried out since the 1990s (Arnett 1994; Bazan & Arnett 1994; Bazán & Arnett 1998; Asida & Arnett 2000; Kuhlen et al. 2003; Meakin & Arnett 2006, 2007; Arnett & Meakin 2011; Chatzopoulos et al. 2014, 2016; Jones et al. 2017).

More recently, ground-breaking attempts to evolve convective shells in 3D prior to the onset of collapse were first reported by Couch et al. (2015) for silicon-shell burning in a 15 M star and by Müller et al. (2016) for oxygen-shell burning in an 18 M star (and also in 11.8, 12, and 12.5 M stars by Müller et al. 2019). Couch et al. (2015) obtained an earlier onset for a neutrino-driven explosion of the 3D progenitor model of the 15 M star compared to that for the corresponding 1D progenitor model. By performing 3D GR simulations with a more advanced neutrino transport scheme, Müller et al. (2017) obtained a neutrino-driven explosion with the seed perturbations. In comparison, this shock was not revived in the corresponding 1D progenitor model. These studies clearly show that convective seed perturbations could potentially have a favorable impact on neutrino-driven explosions. In order to clarify the criteria for precollapse seed perturbation growth, Collins et al. (2018) recently reported a detailed analysis of the convective oxygen- and silicon-burning shells by performing a broad range of 1D presupernova calculations. Using the prescription of the MLT theory in 1D, they pointed out that the extended oxygen-burning shells between ∼16 and 26 M are most likely to exhibit large-scale convective overturn with high convective Mach numbers, leading to the most favorable condition for perturbation-aided explosions. In fact, the 3D progenitor model of the 18 M star (Müller et al. 2016) is in the predicted mass range.

Joining in these efforts, we investigate in this study how the asphericities could grow, in particular driven by the convective oxygen-shell burning in the O- and Si-rich layer. First, we perform a series of 1D stellar evolution calculations with zero-age main-sequence (ZAMS) masses between 9 and 40 M with the HOSHI code developed by Takahashi et al. (2016, 2018). Based on the 1D results, we select about ten 1D progenitors that have extended and enriched O and Si layers, presumably leading to vigorous convection. About 100 s before the onset of collapse, the 1D evolution models are mapped to the multi-D hydrodynamics code (a branch of 3DnSNe; e.g., Nakamura et al. 2016; Takiwaki et al. 2016; Kotake et al. 2018). We perform axisymmetric (2D) simulations for the selected progenitors having an extended O- and Si-rich layer and investigate the features of their convective motion, especially the convective eddy scale and the turbulent Mach number. We then move on to perform a 3D simulation by choosing one of the progenitors that exhibits strong convective activity in 2D. We investigate how the convective features between 3D and 2D differ and discuss its possible implication for the explosion dynamics.

This paper is organized as follows. Section 2 starts with a brief description of the numerical methods employed in our 1D stellar evolution calculation as well as the 2D and 3D hydrodynamics simulations. In Section 3, we present the results of the 1D stellar evolution models in Section 3.1, which is followed in order by the 2D (Section 3.2) and 3D (Section 3.3) results. In Section 4, we summarize with a discussion of the possible implications. The Appendix addresses the comparison of our 1D stellar evolution code with other reference codes and the sensitivity of our results with respect to the different parameters.

2. Setup and Numerical Methods

In this section, we briefly summarize the numerical setups of our stellar evolution calculations in 1D (Section 3.1), 2D (Section 3.2), and 3D (Section 3.3).

2.1. 1D Stellar Evolution

We calculate the 1D evolution of massive solar-metallicity stars with ZAMS masses between 9 and 40 M up to the onset of collapse of the iron core. The calculations are performed using an up-to-date version of the 1D stellar evolution code, HOngo Stellar Hydrodynamics Investigator7 (HOSHI) code (e.g., Takahashi et al. 2016, 2018). In the code, a 300 species nuclear reaction network8 is included, the rates of which are taken from JINA REACLIB v1 (Cyburt et al. 2010) except for 12C(α, γ)16O (see Takahashi et al. 2016 for details).

The mass of the helium (He), carbon–oxygen (CO), and iron (Fe) cores as well as the advanced stage evolution depend on the treatment of convection. We use the Ledoux criterion for convective instability. Inside the convective region, we treat the chemical mixing by means of the MLT using the diffusion coefficients as described in Takahashi et al. (2018).

In order to take into account the chemical mixing by convective overshoot, an exponentially decaying coefficient,

Equation (1)

is included, where Dcv,0, Δr, fov, and HP0 are the diffusion coefficients at the convective boundary, the distance from the convective boundary, the overshoot parameter, and the pressure scale height at the convective boundary, respectively (e.g., Takahashi et al. 2016). We consider the following four models to see the impacts of different overshoot parameters (fov) on the 1D stellar evolution. First, we consider two cases during the H- and He-core burning phases (fov = 0.03 or 0.01; see Table 1). The former and the latter values are determined based on the calibrations to early B-type stars in the Large Magellanic Cloud (LMC; Brott et al. 2011) and the main-sequence width observed for AB stars in open clusters of the Milky Way Galaxy (Maeder & Meynet 1989; Ekström et al. 2012), respectively. We name the former as model "L" after the LMC and the latter as model "M" after the Milky Way Galaxy. Similarly, in order to investigate the impact in more advanced stages, we test two different overshooting parameters of fov,A = 0 or 0.002 (see Table 1) for the advanced stages including the core carbon-burning phase. The convective overshoot during advanced stages is considered for both core and shell convective regions. The subscript "A" is added to the models with fov,A = 0.002. When a star model has a ZAMS mass of x M and belongs to model L(A) or M(A), we set the model name to be xL(A) or xM(A). We also name the set of models contained within models xL(A) or xM(A) as Set L(A) or Set M(A).

Table 1.  Stellar Evolution Model Sets and the Corresponding Overshoot Parameters

Model fov fov,A
L 0.03 0
LA 0.03 0.002
M 0.01 0
MA 0.01 0.002

Note. fov is the overshoot parameter during the H- and He-burning phases. fov,A is the parameter during the more advanced stages, namely after the He-burning phase. For the meaning of the notations "L" and "M," see the text. Convective overshoot during the advanced stages is considered for model sets with the subscript "A."

Download table as:  ASCIITypeset image

We note that the stellar evolution and the final structure also depend on the metallicity and rotation. However, the main purpose of this study is to investigate precollapse inhomogeneities for canonical CCSNe with solar metallicity. We leave the investigation of the metallicity and rotation dependence to a future study.

The mass-loss rate at different evolution stages is important for determining the final mass and the He- and CO-core masses for high-mass stars. We adopt the Vink et al. (2001) mass-loss rate of a main-sequence star when the effective temperature is higher than log Teff = 4.05, and the surface hydrogen mass fraction XH is higher than or equal to 0.3. The mass-loss rate of Nugis & Lamers (2000) is adopted for Wolf–Rayet (WR) stars where the effective temperature is higher than log Teff = 4.05 and the surface hydrogen mass fraction XH is lower than 0.3. When the effective temperature is lower than log Teff = 3.90, we adopt de Jager et al. (1988) mass-loss rate.

2.2. Multi-D Stellar Hydrodynamics Simulations

We compute 2D and 3D models with our hydrodynamics code, the 3 Dimensional nuclear hydrodynamic simulation code for Stellar EVolution (3DnSEV), which is a branch of the 3DnSNe code (see Nakamura et al. 2016; Takiwaki et al. 2016; Sasaki et al. 2017; Kotake et al. 2018 for recent code development). Similar to the base code, 3DnSEV solves Newtonian hydrodynamics equations using spherical polar coordinates as follows:

Equation (2)

Equation (3)

Equation (4)

Equation (5)

where $\rho ,{\boldsymbol{v}},p,{e}_{\mathrm{tot}},{\rm{\Phi }}$ are the density, velocity, pressure, total energy density (sum of the internal energy and kinetic energy), and gravitational potential, respectively. Xi denotes the mass fraction of the ith isotopes and epsilonburn is the energy generated by the change in composition, γburn, due to nuclear burning. C is the energy loss by neutrino emission. The subgrid-scale physics is handled by implicit numerical diffusion instead of solving filtered hydrodynamic equations and by creating a subgrid model for the dissipation of kinetic energy as a large-eddy simulation. A piecewise linear method with the geometric correction of the spherical coordinates is used to reconstruct variables at the cell edge, where a modified van Leer limiter is employed to satisfy the condition of total variation diminishing (TVD; Mignone 2014). The numerical flux is basically calculated by an HLLC solver (Toro et al. 1994). For the numerical flux of isotopes, the consistent multifluid advection method of Plewa & Müller (1999) is used. The models are computed on a spherical polar coordinate grid with a resolution of nr × nθ × nϕ = 512 × 64 × 128 (3D) and nr × nθ = 512 × 128 (2D) zones. The radial grid is logarithmically spaced and covers the center up to the outer boundary of 1010 cm. For the polar and azimuthal angles, the grid covers all 4π sr. To focus on the convective activity mainly in the oxygen shell, the inner 108 cm is solved in spherical symmetry. We include self-gravity assuming a spherically symmetric (monopole) gravitational potential. Such a treatment is indispensable for reducing the computational time; the nonlinear coupling between the core and the surrounding shells (e.g., Fuller et al. 2015) is beyond the scope of this study.

We use the "Helmholtz" equation of state (EOS; Timmes & Swesty 2000). Neutrino cooling is taken into account (Itoh et al. 1996) as a sink term in the energy equation. A nuclear reaction network of 21 isotopes9 (aprox21; Paxton et al. 2011) is implemented, where the inclusion of 54Fe, 56Fe, and 56Cr is crucial for treating low-electron fractions Ye ≳ 0.43 in the presupernova stage. The network is as large as that of Couch et al. (2015) and a little larger than the 19 isotopes of Müller et al. (2016). When the temperature is higher than 5 × 109 K, the chemical composition is assumed to be in nuclear statistical equilibrium (NSE). To avoid the temperature variations caused by numerical instability, we set an artificial upper bound in our multi-D runs, in such a way that the (absolute) sum of the local energy generation rates by thermonuclear reactions and weak interactions does not exceed 100 times the local neutrino cooling rate. To correctly treat the neutronization of heavy elements from Si to the iron group and the gradual shift of the nuclear abundances, one needs to use a sufficient number of isotopes (∼100; Arnett & Meakin 2011), which is currently computationally and technically very challenging. Because the NSE region appears mainly in the Fe core may not significantly affect the convection in the O layer on which we focus.

About 100 s before the onset of collapse, the 1D evolution models are mapped to our multi-D hydrodynamics code, and we follow the 2D and 3D evolution for ∼100 s until the onset of collapse. When we start the multi-D runs, seed perturbations to trigger nonspherical motions are imposed on the 1D data by introducing random perturbations of 1% in density on the whole computational grid. We terminate the 2D/3D runs when the central temperature exceeds 9 × 109 K, because the core is dynamically collapsing at this time.

3. Results

3.1. 1D Stellar Evolution Models

In total, 100 stellar evolution models are calculated in 1D. Four sets of models are constructed, to which different overshoot parameters are applied (Table 1). Each set consists of 25 models to cover the initial mass range of 9–40 M. These 1D models are evolved from the ZAMS stage up to the onset of collapse, which is determined using the threshold central temperature of TC ∼ 109.9 K. Details of our 1D evolution models (e.g., comparison with reference stellar evolution codes) are given in the Appendix.

Figure 1 shows the total mass (red) and masses of the He (blue), CO (green), and Fe (magenta and cyan) cores at the onset of collapse as a function of the ZAMS mass (MZAMS). The top panel shows results for Sets L and LA, while results for Sets M and MA are shown in the bottom panel. We note that Sets L and LA result in very similar total, He-core, and CO-core masses, and differ only in the Fe-core mass. This is because the He- and CO-core masses are mostly determined by the size of the core convection during the H- and He-burning phases, respectively, and are largely independent of the overshoot during more advanced stages. The same is true for Sets M and MA. The He-core mass is defined as the largest enclosed mass with hydrogen mass fraction less than 10−3. Similarly, the CO-core mass is defined as the largest enclosed mass with He mass fraction less than 0.1, and the Fe-core mass is defined as the largest enclosed mass with the sum of the mass fractions of Z ≥ 21 elements larger than 0.5.

Figure 1.

Figure 1. The final stellar mass M for Set L (top panel) and Set M (bottom panel) as a function of the ZAMS mass. The red, blue, green, magenta, and dashed cyan lines correspond to the total mass, He-core mass, CO-core mass, Fe-core mass of Sets LA and MA, and the Fe-core mass of models L and M, respectively.

Standard image High-resolution image

The total mass at the collapse is determined by the mass-loss history. Because the mass loss is relatively weak, the total mass monotonically increases with the ZAMS mass for models below MZAMS ≲ 20–25 M in both Sets L and M. The mass-loss rate increases with increasing luminosity, and thus with increasing ZAMS mass. The increasing mass-loss rate explains the flat and even decreasing trends seen in the 20–30 M models in Set L. At the same time, models in the same mass range show a stochastic trend for Set M. This is caused by the bistability jump of the mass-loss rate, which results from the discontinuous rate increase along the decreasing effective temperature (e.g., Vink et al. 2000). For more massive models above MZAMS ≳ 30 M, the mass-loss rate becomes so efficient to remove most of the H envelope during the He-burning phase. Therefore, the total masses of these models coincide with their He-core masses. This is why the total mass again shows a monotonic increase in this massive end of the ZAMS mass range. The most massive models (the 32, 35, and 38 M models of Set L and the 35, 38, and 40 M models in Set M) finally retain only a small amount of hydrogen of 0.26–0.29 M in their envelopes, which will correspond to these being observed as late-type WN stars (Crowther 2007).

As an exception, model 40L (MZAMS = 40 M model in Set L) has lost not only the entire H envelope but also most of the He layer. This is due to the even stronger WR wind mass loss during the helium- and carbon-burning phases. The He mass remaining on the surface is 0.24 M. We apply the mass-loss rate of Nugis & Lamers (2000) for the H-deficient stars. However, there is a large uncertainty in the estimation of the WR wind mass-loss rate. Among H-deficient stars especially, the mass-loss rate of He-deficient WC stars can be larger than the rate in Nugis & Lamers (2000) by a factor of ∼10 (Yoon 2017). The remaining He mass can be even less if we consider more efficient WR wind mass loss. Therefore, we expect that the star will most probably be observed as a He-deficient WC star, and moreover, it will be observed as a Type Ic supernova when this star explodes.

The features of the distributions of the final stellar mass and the He- and CO-core masses as a function of ZAMS mass are also seen in the results obtained in previous works using KEPLER code (Woosley & Heger 2007; e.g., Figure 4 of Ebinger et al. 2019 for a concise summary).

The He- and CO-core masses monotonically increase with ZAMS mass except for model 40LA, which is affected by the strong WR wind during the He-burning phase. The mass of the helium layer, which is shown as the difference between the He-core and the CO-core masses, also increases with the ZAMS mass. As mentioned earlier, the He- and CO-core masses are insensitive to the overshoot parameter after the He burning, fov,A. Thus, the difference in the CO-core mass between Sets LA and L (and similarly between Sets MA and M) is less than 0.7%. Furthermore, the difference in the He-core mass is less than 0.1%. Note that model 18MA exceptionally forms a CO-core mass about 3% larger than model 18M. This results from the emergence of a narrow convection in the outer layer of the CO core, in which a small amount of He is contained. Because this narrow convection is activated only after the oxygen-core depletion, the structure outside the He layer of model 18MA is mostly the same as that of model 18M, like other models.

We will compare Set L with Set M. Below MZAMS ≲ 25 M, the final mass is not that sensitive to the overshoot parameter for the H- and He-core convection. On the other hand, models above 25 M show a scatter within a factor of ∼0.3. Set L tends to show larger He- and CO-core masses than Set M. This is simply due to the larger overshoot parameter applied during the H- and He-burning phases. For most of the models, the ratio of the He-core masses is about a factor of 1.2–1.3, and the CO-core mass ratio is somewhat larger than the He-core mass ratio. As an exception, the He-core mass ratio reaches 2.16 for the 9 M models. This is due to the merging of He layer to the H envelope by the second dredge-up. A larger overshoot for Set L brings about more effective convective mixing to make more massive He and CO cores. Small core-mass ratios for 40 M models are due to the strong mass loss occurring in model 40L.

To select models that will show strong convective activities in the SiO-rich layer, we utilize two measures, the SiO-coexistence parameters of fM,SiO and fV,SiO. Both of them are defined based on mass fraction distributions of 16O and/or 28Si. The fM,SiO is a product of the mass fractions of 16O and 28Si weighted by the enclosed mass between 108 cm and 109 cm:

Equation (6)

where cM is a scaling coefficient, X(AZ) is the mass fraction of isotope AZ, and Θ(x) is the step function, which satisfies Θ(x) = 1 for x ≥ 0 and 0 for x < 0, ρ is the density, and r8 is the radius in units of 108 cm. Therefore, the value becomes large in a model that has a layer mainly composed of both oxygen and silicon. Such a layer would be the most preferable site to host strong turbulence powered by oxygen-shell burning. This definition of fM,SiO has an uncertainty on how the local density strength of the turbulence depends. Hence, we also test another indicator, fV,SiO, in which the product of the mass fraction is weighted not by the enclosed mass but by the enclosed volume instead:

Equation (7)

where cV is a scaling coefficient. The scaling coefficients are arbitrarily chosen. We calculate these two measures at every time step from 120 to 80 s before the last step of the calculations to see the characteristics at times close to the onset of the multidimensional simulations.

The result is shown in Figure 2, wherein cM = 3.2 × 10−10 and cV = 0.025 are applied. We do not see clear dependencies among different treatments of the overshoot. Some models in the ZAMS mass range ≳22 M show large (≳0.6) fM,SiO values. In the volume-weighted case, the ZAMS mass range showing the models having large (≳0.9) fV,SiO values is 13–28 M. From this result, we selected 11 models, in which either fM,SiO or fV,SiO, or possibly both of them, shows a large value. Models showing the seven highest fM,SiO values are models 28M, 23LA, 25M, 28LA, 27LA, 27M, and 22L. Models showing the six highest fV,SiO values are models 13LA, 28M, 21MA, 25M, 16MA, and 18MA. Among them, models 25M and 28M show large values for both fM,SiO and fV,SiO. The actual values of the parameters are shown in the second and third columns of Table 2.

Figure 2.

Figure 2. SiO-coexistence parameters fM,SiO (top panel) and fV,SiO (bottom panel). The red and orange circles represent models in Sets LA and L, respectively. The blue and cyan squares are for models in Sets MA and M, respectively.

Standard image High-resolution image

Table 2.  2D Model Properties and SiO-coexistence Parameters

Model fM,SiO fV,SiO $\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$ r($\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$) Layer max dc/HP
        (108 cm)      
Low Ma
13LA 0.27–0.28 0.95–0.96 0.018 11.6 O/Si 12 6.22
16MA 0.24–0.24 0.90–0.91 0.015 3.9 O/Si 4 3.20
18MA 0.57–0.58 0.91–0.91 0.131 3.1 Si/O 14 1.06
21MA 0.47–0.47 0.91–0.95 0.134 3.0 Si/O 8 4.42
23LA 0.75–0.80 0.78–0.80 0.069 11.5 O/Si 4 5.20
High Ma
22L 0.57–0.61 0.77–0.82 0.108 9.4 Si/O 2 2.50
25M 0.75–0.79 0.91–0.94 0.160 5.8 Si/O 3 3.65
27LA 0.59–0.66 0.76–0.76 0.179 45.0 O/Si 2 4.56
27M 0.58–0.65 0.37–0.40 0.134 4.7 Si/O 10 2.44
28LA 0.60–0.68 0.37–0.42 0.117 5.3 Si/O 8 1.81
28M 0.83–0.90 0.90–0.95 0.369 14.6 O/Si 2 4.08

Note. SiO-coexistence parameters fM,SiO and fV,SiO are obtained from the result of the 1D evolution simulations. $\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$ represents the maximum turbulent Mach number obtained at a radius of r($\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$) at the end of the 2D simulations. "Layer" represents the composition of the convective region. max represents the value where ${c}_{{\ell }}^{2}$ has a peak (see Equation (9)). dc/HP represents the width of the convective region normalized by the local scale height. These quantities are all estimated at the last step of the simulations. See the text for a more detailed definition.

Download table as:  ASCIITypeset image

For later convenience, we separate the SiO-rich layer into the "Si/O" layer and the "O/Si" layer. The "Si/O" layer has a larger Si mass fraction than O mass fraction in the layer, i.e., X(28Si) ≥ X(16O) and X(16O) ≥ 0.1, whereas the "O/Si" layer has the relation 0.1 ≤ X(28Si) < X(16O). Then, we can classify these 11 models into two groups having different structures of the SiO-rich layer. One group has an extended O/Si layer instead of the O/Ne layer above the Si/Fe layer. The other group has a Si/O layer between the inner Si/Fe layer and the outer O/Ne layer. The former consists of models 13LA, 16MA, 18MA, 21MA, 23LA, 27LA, and 28M. The top panel of Figure 3 shows the mass fraction distribution of model 13LA as a function of radius. The radius of the outer boundary of the O/Si layer is ∼3 × 109 cm. This layer was originally formed as an O/Ne layer. Neon burning has started after the core silicon-burning phase, transforming neon into oxygen and silicon. In the case of models 18MA, 21MA and 23LA, a thin Si/O layer exists between the Si/Fe layer and the O/Si layer with a width of less than 3 × 108 cm.

Figure 3.

Figure 3. Mass fraction distributions of models 13LA (top panel) and 25M (bottom panel) as a function of radius at the last step. The red, black, cyan, blue, magenta, green, and orange lines correspond to the mass fractions of p, He, C, O, Ne, Si, and iron-peak elements with Z ≥ 21 denoted as "Fe," respectively.

Standard image High-resolution image

Models in the latter group have a layered structure, in which the innermost Fe core is surrounded by the Si/Fe, Si/O, and O/Ne layers. Models 22L, 25M, 27M, and 28LA comprise this group. The bottom panel of Figure 3 shows the mass fraction distribution of model 25M, an example of the latter group. The model has a Si/O layer between 3 × 108 cm and 1.1 × 109 cm. For other models in this group, the width of the Si/O layer is typically several times 108 cm. We show the mass fraction distributions as a function of radius for models other than 13LA and 25M in Figure 17.

3.2. 2D Stellar Hydrodynamics Simulations

In order to investigate the convective activities in a multidimensional space, we perform 2D hydrodynamics simulations of oxygen-shell burning. In the previous subsection, we picked up 11 models that show large SiO-coexistence parameters, fM,SiO and/or fV,SiO. Profiles of these models at ∼100 s before the end of the 1D calculations are taken as the initial conditions. The 2D calculations are continued until the central temperature reaches 9 × 109 K, by which point the stars have started runaway collapse due to the gravitational instability.

Following Müller et al. (2016), we evaluate the angle-averaged turbulent Mach number as an indicator of the turbulence strength,

Equation (8)

where ρ is the density; vr, vθ, and vϕ are the radial, tangential, and azimuthal velocities; $\langle {v}_{r}\rangle $ is the angle-averaged radial velocity; cs is the sound velocity; and Ω is the solid angle. The maxima of $\langle {{Ma}}^{2}{\rangle }^{1/2}(r)$ evaluated at the end of the simulations $\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$ are shown in the fourth column of Table 2, and $r(\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2})$ represents the radii where $\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$ are obtained.

Based on the Mach number, we divide our 2D models into two groups, either showing "low Ma" or "high Ma." The criterion of high Ma is set as $\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}\geqslant 0.1$, because the turbulence with such a high Mach number potentially fosters the perturbation-aided explosion (Müller & Janka 2015; Müller et al. 2016). It is noted that models 18MA and 21MA are exceptionally classified into low Ma despite their "large" Mach numbers. We discuss this later in this section. The column "Layer" in Table 2 represents the dominant chemical composition in the convective layer, i.e., the Si/O layer or the O/Si layer. We also show the mass fraction distribution of models 13LA and 25M in Figure 3. For other models, the mass fraction distributions are shown in Figure 17 in the Appendix. See the last part of the previous subsection for the definition of the layer.

Time evolution of convective motion—In Figure 4, we show the time evolution of the turbulent Mach number and the Si mass fraction for representative models from low Ma (13LA, top) and high Ma (25M, middle, and 27LA, bottom). The color visualizes the angle-averaged turbulent Mach number $\langle {{Ma}}^{2}{\rangle }^{1/2}$ (left) and the 28Si mass fraction X(Si) (right). Note that the outer radial frame of the panels for model 27LA is set to 8 × 109 cm in order to show how the outer edge of the convective region keeps moving outward and reaches this radius at the end.

Figure 4.

Figure 4. The time and radial distributions of the angle-averaged turbulent Mach number $\langle {{Ma}}^{2}{\rangle }^{1/2}$ (left panels) and the 28Si mass fraction (right panels). Top, middle, and bottom panels are for models 13LA, 25M, and 27LA, respectively.

Standard image High-resolution image

The model 13LA has no Si/O layer. This star has an Fe core at the central region of R ≲ 2 × 108 cm, which is surrounded by the convective Si/Fe layer (R ∼ 2–4 × 108 cm) and the convective O/Si layer (R ≳ 4 × 108 cm). The Si mass fraction at the Si/Fe layer is ∼0.5 (see the top panel of Figure 3). As shown in the top-right panel of Figure 4, the Si mass fraction is small compared to that of the 25M model that is shown in the middle right panel of Figure 4. Reflecting this structure, the turbulent Mach number is lower than 0.1 in the inner Si/Fe layer and in the outer O/Si layer throughout the simulation (see the top-left panel of Figure 4). However, oxygen burning slightly enhances the 28Si mass fraction in the base region of the O/Si layer of ∼4–8 × 108 cm. Note that $\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$ in Table 2 is estimated at the end of the simulations, and it does not refer to the peak seen at ∼4 × 108 cm at ∼30 s.

Readers may be confused by the models 18MA and 21MA because they have the Si/O layer in Table 2 but they are classified into the low-Ma group. Actually, models 18MA and 21MA have a thin Si/O layer and $\langle {{Ma}}^{2}{\rangle }^{1/2}$ ∼ 0.13 in the Si/O layer, but only after the last ∼10 s of the simulations. This is because the turbulence is triggered by gravitational contraction, which amplifies the temperature at the bottom of the Si/O layer, enhancing the oxygen-burning rate. The turbulence powered by gravitational contraction has too short a time to form an extended convective region, which is in contrast to the shell convection powered by hydrostatic burning. This is why we selected these models as members of low Ma.

Models 22L, 25M, 27LA, 27M, 28LA, and 28M are categorized into models with high Ma. Convective motion with such a strong turbulence develops by oxygen burning in the Si/O layer in these models. We pick out two models in which it is easy to explain the typical dynamics of the convection.

Model 25M consists of the central Fe core (R ≲ 2 × 108 cm), the Si/Fe layer (R ∼ 2–3 × 108 cm), the Si/O layer (R ∼ 3–10 × 108 cm), and the O/Ne layer (R ≳ 10 × 108 cm). It is noteworthy that, despite the Si/Fe layer seeming to have a homogeneous chemical composition (see the lower panel of Figure 3), the outer part of R ∼ 2.5–3 × 108 cm is actually composed of a small amount of oxygen with X(16O) < 0.01. The oxygen-free region of R ∼ 2–2.5 × 108 cm becomes convective within a short timescale of ∼10 s from the start of the simulation, though the shell convection is not extended further. At the bottom of the outer Si/O layer, hydrostatic oxygen-shell burning takes place. In this case, nuclear burning drives high turbulent velocity with $\langle {{Ma}}^{2}{\rangle }^{1/2}\gt 0.1$, which is sustained for 20–110 s (see the middle-left panel of Figure 4). Accordingly, turbulent mixing homogenizes the 28Si mass fraction in the region of R = 3–10 × 108 cm. Furthermore, oxygen burning also takes place in the oxygen-containing outer region of the Si/Fe layer. Despite the large mean molecular weight, the heating due to oxygen burning is strong enough to lift the silicon-rich material up into the surrounding Si/O layer. As a result, the silicon mass fraction in the Si/O layer is significantly enhanced. This silicon enhancement repetitively takes place at ∼50 and 80 s, which accompanies the enhancement of the convective Mach number as well. It seems that repetitive mixing follows the oscillation of the outer edge of the Si/Fe convection. This will be because, with the small oxygen mass fraction, the temperature fluctuation originally caused by the oscillation is enhanced by the O burning, resulting in a large density fluctuation that triggers convection in the Si/O layer. Indeed, the temperature rise at the bottom of the Si/O layer where the O mass fraction is ∼0.08–0.1 reaches 8% at maximum, which is much higher than the temperature change solely due to the oscillation, which is less than ∼1%, measured in the outer region of the Si/F and Si/O layers.

In model 27LA, the turbulent activity in the O/Si layer starts to increase at ∼45 s (see the bottom-left panel of Figure 4). During the simulation time of ≳200 s, the high Mach number region extends outward, finally reaching ∼6 × 108 cm, which roughly corresponds to the composition jump between the CO-rich and the He-rich layers. At the same time, the turbulent Mach number grows with time in almost the entire region in the convective layer. As a consequence, the turbulent Mach number exceeds ∼0.15 in the wide outer region of R ∼ 30–50 × 108 cm at the end of the simulation. Initially, the 28Si mass fraction decreases with radius in the O/Si layer (see the mass fraction distribution of model 27LA in Figure 17). The convection powered by the oxygen burning mixes material in the slightly silicon-enriched region, which is initially located below ∼8 × 108 cm, into the outer, slightly silicon-poor region after ∼70 s. However, the convective mixing in the 2D simulation is still not efficient enough to achieve the homogeneous chemical distribution. This is due to the limitation of the calculation time, because the total time of this simulation covers only about one convection-turnover time. Model 28M shows similar convective properties to model 27LA.

High/low Mach number and the chemical distribution—Models 13LA, 16MA, 18MA, 21MA, and 23LA belong to low Ma, i.e., they do not show strong turbulence in their convective regions during the simulations. These low-Ma models have characteristic chemical composition profiles. A main characteristic is no or thin Si/O layer. Models 13LA and 16MA do not have a Si/O layer and have an extended O/Si layer on the Si layer (see the top panel of Figure 3 and top-left panel of Figure 17, respectively). Model 23LA does not have a Si/O layer at the beginning of the 2D simulation. The turbulent Mach number of these models in the O/Si layer is low. Models 18MA and 21MA have a Si/O layer at the beginning of the 2D calculations but the width is less than ∼1 × 108 cm. Although the turbulent Mach number exceeds 0.1 at the bottom of the Si/O layer for a few seconds before the termination of the simulations, the turbulence in this layer does not develop before this time.

Models 22L, 25M, 27LA, 27M, 28LA, and 28M belong to high Ma. The main characteristic of the chemical profiles is an extended Si/O layer. Models 22L, 25M, and 28M have a Si/O layer with the width of ∼8 × 108 cm. Note that the Si/O layer of model 28M has merged to the O/Ne layer before the end of the simulation. Models 27M and 28LA also have a Si/O layer, although their width is thinner than that of the three models.

Model 27LA is an exception. This model does not have a Si/O layer but the chemical composition profile is similar to the last step of model 28M.

We briefly discuss the relation to the SiO-coexistence parameters. All models in high Ma are selected using a large fM,SiO value. Although models 25M and 28M are also selected using a large fV,SiO value, they also have a large fM,SiO. The models having a large fM,SiO rather than a large fV,SiO are associated with high Ma.

2D distribution—Figure 5 shows the 2D distributions of the radial turbulent Mach number, Mar = ${v}_{r}/{c}_{s}-\langle {v}_{r}\rangle /\langle {c}_{s}\rangle $, and the 28Si mass fraction taken at the last step of the simulations for models 13LA, 25M, and 27LA. The turbulent Mach number of model 13LA (top panels in Figure 5) develops only within the level of Mar ∼ 0.01. The spherical boundary is clearly observed at r ∼ 3 × 108 cm, where the turbulent Mach number becomes almost zero. Inside the boundary, convection is developed in the Si/Fe layer. A more extended but even weaker turbulent motion is also developed in the O/Si layer above the boundary. The outer boundary of the O/Si convection may be defined at ∼6 × 108 cm, but there is only a diffuse Mar ∼ 0 region in this case. A thin and nearly spherical band with X(28Si) ∼ 0.2 surrounding the inner boundary between the Si/Fe and the O/Si layers exists. As a result of the low-velocity turbulence in the O/Si layer, this silicon-rich material is slowly mixed into the inner region of the O/Si layer at R ≲ 6 × 108 cm.

Figure 5.

Figure 5. 2D distributions of the turbulent Mach number of the radial velocity Mar (left panels) and 28Si mass fraction (right panels). Top, middle, and bottom panels correspond to models 13LA, 25M, and 27LA, respectively.

Standard image High-resolution image

Model 25M (middle panels in Figure 5) develops convective motion with a high turbulent Mach number in the Si/O layer ranging from R ∼ 3 × 108 cm to R ∼ 10 × 108 cm. At the end of the simulation, outflows stream in three directions: the northern pole direction, ∼45° from the polar axis, and ∼135° from the polar axis, and inflows are sandwiched by the outflows. These convective flows have turbulent Mach numbers larger than ∼0.1. The 28Si mass fraction is roughly homogenized inside the convective region, having the value of 0.3–0.4, though some fluctuations are observed especially near the outer boundary.

Model 27LA (bottom panels in Figure 5) has an extended O/Si layer distributed from R ∼ 5 × 108 to 50 × 108 cm. A large-scale convective motion is developed in this layer; a broad conical outflow with the opening angle of ∼45° is formed in both polar regions, and between them, a thick inflow is formed around the equatorial plane. The convective Mach number reaches ∼0.12. The large-scale outflow mixes the silicon-rich material into the O/Si layer. The silicon mass fraction in most parts of this layer is initially ∼0.1, while the outflow has a higher fraction of ∼0.16.

Width of the convective region—We briefly discuss the width of the convective region divided by the local scale height. We define a convective region as a region having $\langle {{Ma}}^{2}{\rangle }^{1/2}\gt 1/3\,\times \langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$ including $r(\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2})$. We determined the factor of 1/3 to avoid including the neighboring convective region because we obtain a small turbulent Mach number even at the convection boundary. The width of the convective region is compared with the pressure scale height HP at $r(\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2})$. This results in 1.8–4.6 for the high-Ma models, which are listed in Table 2. We should note that the above definition does not specify the width of the convective region correctly for low-Ma models. In models 16MA and 18MA, the specified region contains the Si layer inside the O/Si and Si/O layer, respectively. In model 21MA, the calculated region contains part of the Fe core, the Si and Si/O layer, and part of the O/Si layer. The turbulent Mach number at the boundary determined by the abundance distribution is not small enough compared with $\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2}$ to specify the boundary of the convective region for these models.

Typical scale of the convection—In addition to the Mach number, the dominant angular wave number in spherical harmonics also characterizes the convection. This is related to the typical size of the convective flow. This quantity is important because a large-scale convective flow can amplify the explodability of core-collapse supernovae (Müller et al. 2016). The power spectrum of the radial turbulent velocity at $r(\langle {{Ma}}^{2}{\rangle }_{\max }^{1/2})$ is calculated as

Equation (9)

where Yℓm(θ) is the spherical harmonics function of degree and order m. ℓmax in the table represents the value at which ${c}_{{\ell }}^{2}$ has a peak.

Figure 6 shows the power spectrum ${c}_{{\ell }}^{2}$ at three different times for models 13LA (top), 25M (middle), and 27LA (bottom). For model 13LA, ${c}_{{\ell }}^{2}$ has a maximum at  = 12, but the spectrum is rather flat and less energetic. The radius of the highest Mach number in the O/Si layer is 1.16 × 109 cm. Although convective mixing occurs in the inner region of R ≲ 6 × 108 cm, the turbulent velocity is lower than that in the outer region at the last step. Large-scale convection in the O/Si layer is not developed probably because of small turbulent motion. For model 16MA, max is equal to 4 and the trend of the power spectrum is similar to model 13LA.

Figure 6.

Figure 6. Power spectra of the radial velocity dispersion ${c}_{{\ell }}^{2}$ at $r(\langle {{Ma}}^{2}{\rangle }^{1/2})$ for models 13LA (top panel), 25M (middle panel), and 27LA (bottom panel) in the 2D simulations.

Standard image High-resolution image

For the other low-Ma models, models 18MA and 21MA show large max (see Table 2), and they have a thin Si/O layer. This trend is roughly consistent with the analysis of the convective eddy scale relating to the scale of the convective layer and the typical radius of the layer (e.g., Müller et al. 2016). In these models, the SiO-rich layer is thin compared to the radius of the layer.

In models 25M and 27LA, which are high-Ma models, the power spectrum peaks at max = 3 and 2, respectively, and the spectrum decreases with increasing above that. Models 22L and 28M show a similar power spectrum to model 27LA. For these models, large-scale convective eddies have developed. On the other hand, models 27M and 28LA indicate larger values probably owing to the thin SiO-rich layer. Indeed, the previous three models show a larger width of the convective region normalized by the scale height compared to the latter two models (see Table 2). Note that these models develop shell convection in the Si/Fe layer as well. However, the convective region is always confined inside the layer. This will be because the timescale of the silicon burning is shorter than the convective turnover time, so that the mean molecular weight of the convective blob soon increases, suppressing the convective motion.

From the results shown above, it is discerned that fM,SiO will be a more suitable measure than fV,SiO to discriminate a model that develops convection with high turbulent velocity and a small max. First, we have shown that high-Ma models are selected based on the high fM,SiO values. Moreover, models showing small max ≤ 3 are all selected based on fM,SiO (models 22L, 25M, 27LA, and 28M), and only one of the two models showing max = 4 (model 16MA) is selected based on fV,SiO.

3.3. 3D Stellar Hydrodynamics Simulation

A 3D hydrodynamics simulation is conducted using the 1D model 25M as the initial condition. The size of the convective region of model 25M would be suitable for investigating the multidimensional effects of the structure of a presupernova star on the supernova explosion.10 The hydrodynamical evolution is followed for ∼100 s until the central temperature reaches 9 × 109 K, at which point the Fe core is unstable enough to collapse.

Figure 7 shows the time evolution of the 28Si mass fraction distribution of model 25M. The initial distribution of the 28Si mass fraction is spherically symmetric (top-left panel). After the start of the simulation, the convection in the Si/O layer develops from the inner region. We see that Si-enriched plumes go up into the Si/O layer (top-right panel). The convective motion reaches a steady flow by ∼20 s. The inhomogeneous 28Si mass fraction distribution introduced by the convection at 30 s is shown in the middle-left panel. After a while, the turbulent velocity becomes small. We will discuss the mechanism of this weakening later.

Figure 7. The time variation of the 28Si mass fraction distribution of model 25M at t = 0 s (top left), 10 s (top right), 30 s (middle left), 75 s (middle right), 90 s (bottom left), and 105 s (bottom right). An animated version of this figure is available, showing the time variation from t = 0 to 105 s. The animation duration is 13 s.

(An animation of this figure is available.)

Video Standard image High-resolution image

At ∼70 s, the Si/O layer gradually contracts, triggering the strong oxygen-shell burning at the bottom of the Si/O layer. This strong burning drives high-velocity turbulence and expands the Si/O layer. We see some 28Si-enriched plumes flow from the inner region of the Si/O layer at 75 s (see red region in the middle right panel). As a result, the high-velocity turbulent flow mixes with the surroundings and increases the Si mass fraction in the whole Si/O layer. The 28Si mass fraction in the Si/O layer slightly increases from about 0.36 at 30 s to 0.38 at 90 s (bottom-left panel). The convective motion in the Si/O layer continues until the last step of the simulation. We see inhomogeneous 28Si mass fraction distribution at the last step (bottom right panel).

In Figure 8, the time evolution of $\langle {{Ma}}^{2}{\rangle }^{1/2}$ and the 28Si mass fraction obtained from the 3D simulation is shown. As shown by the left yellow region in the top panel, the convective motion with $\langle {{Ma}}^{2}{\rangle }^{1/2}\sim 0.1$ is obtained in the Si/O layer at ∼20 s. The bottom panel shows that the 28Si mass fraction is enhanced in this layer by that time. After a while, the Si/O layer slightly expands and the convective motion weakens. The averaged Si mass fraction in the Si/O layer does not change significantly from 20 to 70 s.

Figure 8.

Figure 8. The angle-averaged radial and time distribution of the Mach number $\langle {{Ma}}^{2}{\rangle }^{1/2}$ (top) and the 28Si mass fraction (bottom) of model 25M.

Standard image High-resolution image

By ∼70 s, the Fe core contracts and silicon-shell burning as well as oxygen-shell burning is enhanced. Because of this, the turbulent Mach number becomes large not only at R ∼ 3–12 × 108 cm but also at R ∼ 2 × 108 cm. In addition, the Si-rich material at the bottom of the Si/O layer is carried into the Si/O layer through this convective motion, and the 28Si mass fraction in this region increases during ∼70–80 s.

From ∼90 s, the convective motion becomes weak again. The whole Si/O layer gradually contracts toward core-collapse. Meanwhile, the base temperature of the Si/O layer increases, boosting the oxygen-shell burning. This results in the enhancement of the convective motion at the bottom of the Si/O layer. The Si/O layer at the end of the simulation is distributed from 3.0 × 108 cm to 10.5 × 108 cm. The maximum radial convective Mach number is $\langle {{Ma}}^{2}{\rangle }^{1/2}=0.087$, which is obtained at R = 3.6 × 108 cm.

Figure 9 shows the power spectrum of the radial convective velocity ${c}_{{\ell }}^{2}$ at R = 5.8 × 108 cm taken at three different times. The radius, which is located in the middle of the convective Si/O layer, is determined from the radius where the maximum $\langle {{Ma}}^{2}{\rangle }^{1/2}$ is obtained in the 2D simulation. Similar to the 2D case, the power spectrum in the 3D case is calculated as

Equation (10)

The power spectrum ${c}_{{\ell }}^{2}$ in the 3D simulation peaks at  = 2. This is consistent with the finding in Müller et al. (2016) for the 18 M star. The small maximum mode means that the convective motion is dominated by a large-scale flow. The main difference between 2D and 3D power spectra is the weaker turbulence in the 3D simulation. Note that the stronger turbulence in 2D compared to 3D is not surprising because the turbulent energy cascade could occur artificially, as previously identified, from small to large scales along the coordinate symmetry axis. This is mainly due to the reduced degree of freedom in which the energy can dissipate.

Figure 9.

Figure 9. The power spectrum of the convective radial velocity ${c}_{{\ell }}^{2}$ at R = 5.8 × 108 cm for model 25M taken at three different times of 90.0, 100.0, and 105.6 s in the 3D simulation.

Standard image High-resolution image

Finally, we briefly present the result of the neutrino emission at the precollapse stage of the 3D simulation. Using the same method in Yoshida et al. (2016), we calculate the time evolution of the luminosity and the spectrum of neutrinos emitted via the pair-neutrino process for model 25M. In Figure 10, the emission rate and the average energy of neutrinos obtained from the 1D and the 3D simulations are compared.

Figure 10.

Figure 10. Time variation of the emission rate (top panel) and average energy (bottom panel) of antineutrinos produced through the pair-neutrino process for model 25M. The thick solid and thin dashed curves correspond to the results of 3D and 1D simulations, respectively. The red and blue curves correspond to ${\bar{\nu }}_{e}$ and ${\bar{\nu }}_{\mu ,\tau }$.

Standard image High-resolution image

The overall features of the neutrino spectra for the 1D and 3D simulations are in common. The neutrino emission rate and the average temperature increase with time toward the core-collapse after ∼30 s in both simulations. The emission rates of νe and ${\bar{\nu }}_{e}$ are larger than those of νμ, τ, and ${\bar{\nu }}_{\mu ,\tau }$. The average energy of ${\bar{\nu }}_{e}$ is slightly smaller than that of ${\bar{\nu }}_{\mu ,\tau }$. In the 3D simulation, however, the decrease in the emission rate and the average energy of neutrinos is observed in 70–90 s.

Furthermore, we evaluate the time evolution of the event rate of ${\bar{\nu }}_{e}$ using KamLAND (e.g., Gando et al. 2013), assuming that the initial mass of 25 M at the distance of 200 pc is at the ongoing core-collapse phase, emitting neutrinos via the pair-neutrino process.

KamLAND is a one-kiloton size liquid-scintillation-type neutrino detector (see, e.g., Gando et al. 2013). We take the neutrino oscillation into account in a simple manner: the survival probability of ${\bar{\nu }}_{e}$ is set to be 0.675 and 0.024 in the normal and inverted mass ordering, respectively (Yoshida et al. 2016). The live-time-to-runtime ratio and the total detection efficiency are set to be 0.903 and 0.64, respectively (Yoshida et al. 2016). Figure 11 shows the evolution of the neutrino event rate using KamLAND. As an overall feature, the rate increases with time independent of mass ordering. In the 3D simulation, however, the decrease in the event rate due to the oxygen-shell burning at the bottom of the Si/O layer is seen at 70–90 s. The time variation in the precollapse neutrino detection thus can be used as the indicator of the multidimensional convective activity deep inside the star, although it is practically impossible to detect such a time variation with current one-kiloton-size neutrino detectors.

Figure 11.

Figure 11. The evolution of the neutrino event rate of model 25M with KamLAND. The thick solid and thin dashed curves correspond to the results of the 3D and 1D simulations, respectively. The red and blue curves correspond to the normal and inverted orderings.

Standard image High-resolution image

Next, we examine the detectability of precollapse neutrinos by Hyper-Kamiokande. Hyper-Kamiokande is a planned water-Čerenkov-type neutrino detector, which has a huge fiducial mass of 190 kton (Abe et al. 2018). If a supernova explodes in the vicinity of Earth, high-quality data of supernova neutrinos with well-resolved time and energy bins are expected to be obtained (e.g., Takahashi et al. 2001; Mirizzi et al. 2016). However, because the threshold energy of Hyper-Kamiokande is expected to be at the same level as Super-Kamiokande (∼4.79 MeV corresponding to 3.5 MeV for recoil electrons; Sekiya 2013) and is higher than that of a liquid-scintillation-type detector, Hyper-Kamiokande is less preferable for observations of neutrinos produced through the pair-neutrino process during the precollapse stage. Yoshida et al. (2016) discussed the possibility of neutrino observations using delayed γ-rays from Gd in Gd-loaded Hyper-Kamiokande (Beacom & Vagins 2004), because the energy threshold will be reduced to 1.8 MeV, an energy similar to the case of KamLAND, in this case. Considering a moderate detection efficiency of 50%, the detection rate is about 178 times as large as the rate of KamLAND. When the enhancement factor of 178 for the detection rate is applied, the event rate is expected to be 3–14 s−1 in the normal mass ordering. So, the time variation due to the convective motion with a timescale of seconds could be observed by Hyper-Kamiokande.

We should note that we consider only the pair-neutrino process for presupernova neutrinos in this study. For a few minutes before core-collapse, the main source of ${\bar{\nu }}_{e}$ will be β decays rather than the pair-neutrino process (Kato et al. 2017). βdecays mainly take place in the innermost Fe core, and the ${\bar{\nu }}_{e}$ energy is similar to pair neutrinos. We expect that the neutrino event rate by β decays also decreases for 70–90 s because the central temperature and density decrease during this period. So, even when we take into account the neutrino event rate by β decays, the time variation of presupernova neutrino events would give us information about the dynamics in the SiO-rich layer or a collapsing Fe core.

4. Summary and Discussions

We performed 1D, 2D, and 3D simulations of the oxygen-shell burning just before the core-collapse of massive stars. First, we calculated the 1D evolution of massive solar-metallicity stars with ZAMS mass of 9–40 M. We considered four cases of convective overshoot parameters in hydrogen and helium burning and the following stages.

From these, we searched for massive stars having O- and Si-rich layers located in the range of 108–109 cm, because the oxygen-burning shell is expected to give rise to large asphericity in the mass accretion rates onto the protoneutron star, favoring the onset of neutrino-driven explosions. For the enclosed mass weighted SiO-coexistence parameter fM,SiO, the stars with MZAMS ∼ 20–30 M have SiO-enriched environment in 108–109 cm. For the volume-weighted SiO-coexistence parameter fV,SiO, the mass range indicating a SiO-rich layer increases to 13–30 M.

We selected 11 models showing a large SiO-rich layer from the result of the massive star evolution and performed 2D hydrodynamics simulations of the evolution for ∼100 s until the central temperature reaches 9 × 109 K. We investigated the time evolution of the Mach number of convective velocity and analyzed the power spectrum of the radial convective velocity. Based upon the analysis, we classified the 11 models into five models with low turbulent Mach number (low Ma) and six models with high turbulent Mach number (high Ma).

High-velocity turbulence is obtained in four models having a Si/O layer with range up to ∼1 × 109 cm and two models having an extended O/Si layer. These models commonly have large fM,SiO values. All models with large fM,SiO except for 23LA exhibit high-velocity turbulence in the 2D simulations. In addition, most of the small max models were selected based on fM,SiO. Our results suggest that high density in the SiO-rich layer could be conducive to producing vigorous convection just before the core collapses (see the difference in Equations (2) and (3) and that fM,SiO could be a more suitable measure for convection with high-velocity turbulence and large-scale eddies compared to fV,SiO.

We performed a 3D hydrodynamics simulation for model 25M for ∼100 s until the central temperature reaches 9 × 109 K. The time evolution of the convection properties is qualitatively similar between the 2D and 3D simulations. The convection is dominated by large-scale flows with either max = 2 in the 3D case or max = 3 in the 2D case. The main difference is the smaller turbulent velocity in the 3D simulation.

Using the 3D result of the hydrodynamics, we have evaluated the time evolution of the neutrino emission through electron–positron pair annihilation. The emission rate and the average energy of neutrinos evolve similarly in 1D and 3D simulations. However, in the 3D model, the neutrino emission rate shows significant variation due to the strong oxygen-shell burning at ∼70 s. The multi-D effect of the convective burning would be observed in presupernova neutrino events by present and future neutrino detectors such as KamLAND and Hyper-Kamiokande.

In this study, we were only able to compute a single 3D model of the 25 M star due to the high numerical cost. However, more systematic and long-term 3D simulations employing a variety of the progenitors are needed (e.g., Müller et al. 2019). In order to accurately deal with the neutronization of heavy elements to the iron-group nuclei, we need to not only implement a bigger nuclear network but also a complete set of neutrino opacities (Kato et al. 2017) in our multi-D models, again albeit computationally expensive. Employing the 3D progenitor models in core-collapse supernova simulations is also mandatory to see how much the precollapse inhomogeneities could foster the onset of neutrino-driven explosions (Couch et al. 2015; Müller et al. 2019). The impact of rotation (not to mention magnetic fields!) on the burning shells during the last stage of massive stars is yet to be studied in 3D (see Chatzopoulos et al. 2016 for 2D evolution models with rotation). All in all, this study is only the very first step toward the more sophisticated and systematic multi-D presupernova modeling.

We acknowledge the anonymous referee for reading our manuscript carefully and giving us many useful comments to improve this manuscript. We thank C. Nagele for proofreading. This study was supported in part by the Grants-in-Aid for the Scientific Research of Japan Society for the Promotion of Science (JSPS, Nos. JP26707013, JP26870823, JP16K17668, JP17H01130, JP17K14306, JP18H01212), the Ministry of Education, Science and Culture of Japan (MEXT, Nos. 26104007, JP15H00789, JP15H01039, JP15KK0173, JP17H05205, JP17H05206 JP17H06357, JP17H06364, JP17H06365, JP24103001 JP24103006 JP26104001, JP26104007), by the Central Research Institute of Explosive Stellar Phenomena (REISEP) of Fukuoka University and the associated research projects (Nos.171042, 177103), and by JICFuS as a priority issue to be tackled by using the Post "K" Computer. T.T. was supported by the NINS program for cross-disciplinary study (grant numbers 01321802 and 01311904) on Turbulence, Transport, and Heating Dynamics in Laboratory and Solar/Astrophysical Plasmas: SoLaBo-X. K. T. was supported by the Japan Society for the Promotion of Science (JSPS) Overseas Research Fellowships. The numerical simulations were done using XC50 at Center for Computational Astrophysics at National Astronomical Observatory of Japan and XC40 at Yukawa Institute for Theoretical Physics in Kyoto University.

Software: HOSHI (Takahashi et al. 2016, 2018), 3DnSNe (Nakamura et al. 2016; Takiwaki et al. 2016; Kotake et al. 2018).

: Appendix

We show some evolutionary properties of the massive star models. We show the Hertzsprung–Russell (HR) diagram of 10 different ZAMS mass models for Sets LA and MA in Figure 12. Stars with MZAMS < 30 M for Set LA and with MZAMS < 35 M for Set MA end as red supergiants. Stars with heavier MZAMS evolve to a WR star.

Figure 12.

Figure 12. HR diagram of Sets LA (panel (a)) and MA (panel (b)). Green dotted lines in panels (a) and (b) are the 20 M models of Stern (Brott et al. 2011) and GENEC (Ekström et al. 2012), respectively.

Standard image High-resolution image

As shown in Section 2, we considered four sets of models with different overshoot parameters. The overshoot parameter during the hydrogen and helium burning affects the evolution on the HR diagram. We compare the HR diagram of models 20LA and 20MA with that of the 20 M star models of Stern (Brott et al. 2011) and GENEC (Ekström et al. 2012), respectively (see the green lines in Figure 12). The main difference in the HR diagram between model 20LA and model 20MA is the main-sequence (MS) bandwidth, i.e., the difference in the effective temperature between ZAMS and the hydrogen-burning termination. The MS bandwidth of model 20LA is almost identical to the Stern model. On the other hand, model 20MA is almost identical to GENEC. As explained in Section 2, the overshoot parameters of Sets LA and MA are determined based on the calibrations to early B-type stars in the LMC similar to the Stern model and the MS width observed for AB stars in open clusters of the MW Galaxy similar to the GENEC model, respectively. Except for the MS width, we do not see any large differences in the HR diagram among these models. The evolution of the 20 M star models of Stern, GENEC, MESA (Paxton et al. 2011; Martins & Palacios 2013), and Starevol (Decressin et al. 2009) are compared in Martins & Palacios (2013). The HR diagram of model 20MA is also similar to that of MESA and Starevol.

We show the evolution of the relation between the central density and the central temperature in Figure 13. All stars except for the ZAMS mass of 9 M in Sets MA and M end with the collapse of an Fe core. The 9 M star in Set MA does not bring about Ne ignition (panel (b): red line). We also confirmed no Ne ignition in the 9 M star in Set M until the central density becomes log ρC = 8.5. These stars are expected to end as an ONe white dwarf or an electron-capture supernova. The off-center Ne ignition occurs for the 9 M star in Set LA (panel (a): red line) and the 11 M star in Set MA (panel (b): green line). The similar off-center neon burning is also seen in a 10 M star in Sets MA and M. The low-mass MA and M models ignite silicon at an off-center region.

Figure 13.

Figure 13. The evolution of the central density log ρC and the central temperature log TC of Sets LA (panel (a)) and MA (panel (b)).

Standard image High-resolution image

Whether the carbon-core burning occurs convectively or radiatively depends on the stellar mass. The convective carbon-core burning occurs in the stars with MZAMS ≤ 18 M in Sets LA and L and with MZAMS ≤ 21 M in Sets MA and M. We suggest that the maximum CO-core mass for convective carbon burning is between 4.6 and 4.9 M for our models.

We calculate the massive star evolution taking account of a weak overshoot after the helium burning for models in Sets LA and MA. Despite the small value compared with the hydrogen and helium burning, the overshoot in the advanced stage affects the advanced evolution in a complicated way. We show the Kippenhahn diagram of models 25MA and 25M in Figure 14 for comparison. In these models, the carbon-core burning occurs radiatively and the first carbon-shell burning (C(i)) ignites at Mr ∼ 1.3 M. In model 25MA, the convective region of the C(i) burning extends both inward and outward with the help of overshoot. The convective C layer extends in the range 1.0–2.3 M. Then, the second carbon-shell burning (C(ii)) ignites at ∼2.1 M, and the convective region again extends inward and outward. The inner convection boundary of the second carbon-shell burning reaches 1.74 M. This inner boundary restricts the region of the following burning. The Si core is formed through O-core burning (O(c)) and the first (O(i)) and second (O(ii)) oxygen-shell burning. The third oxygen-shell burning (O(iii)) extends the Si layer up to 1.67 M, but the boundary is still below the inner boundary of the second carbon-shell burning. The Si core (Si(c)) and the following four silicon-shell burning (S(i)–(iv)) form an Fe core of ∼1.56 M. The left panel of Figure 15 shows the mass fraction distribution of model 25MA. There is a SiO-rich layer between the Si layer and the O/Ne layer, but it is very thin (the width is about 108 cm).

Figure 14.

Figure 14. Kippenhahn diagram of models 25MA (left panel) and 25M (right panel) from the ignition of the central carbon burning until the calculation termination.

Standard image High-resolution image
Figure 15.

Figure 15. Mass fraction distributions of the last time step of models 25MA (left panel) and 25M (right panel). The red, black, cyan, blue, magenta,green, and orange lines correspond to the mass fractions of p, He, C, O, Ne, intermediate elements with Z = 14–20 denoted as "Si," and iron-peak elements with Z ≥ 21 denoted as "Fe," respectively.

Standard image High-resolution image

Model 25M indicates an evolution different from model 25MA from the first carbon-shell burning (C(i)). The first carbon-shell burning ignites at ∼1.3 M, and the convective region extends outward to 2.9 M. The convection does not extend inward in this model. Then, the second carbon-shell burning (C(ii)) starts at 2.9 M. Because the second carbon-shell burning starts at a large radius, the O-core burning (O(c)) and the first oxygen-shell burning (O(i)) extend more effectively outward and form a larger Si core. The Si core (Si(c)) and the following three silicon-shell burning (S(i)–(iii)) form an Fe core of 1.68 M. The second oxygen-shell burning (O(ii)) makes a large Si/O layer between the top of first oxygen-shell burning (1.9 M) and the bottom of the second carbon-shell burning (2.9 M). The mass fraction distribution of model 25M is shown in the right panel of Figure 15. We see a large Si/O layer in this figure.

The compactness parameter, ξ2.5, is considered as a quantity that correlates to the dynamics during the gravitational collapse (O'Connor & Ott 2011). It is defined as the inverse of the radius, inside which the mass of 2.5 M is enclosed,

Equation (11)

Several works have reported that a small compactness parameter is favorable for supernova explosions (Horiuchi et al. 2014; Nakamura et al. 2015; Pejcha & Thompson 2015; Sukhbold et al. 2016), although the criterion has not yet converged (Ugliano et al. 2012; Ertl et al. 2016; Suwa et al. 2016; Burrows et al. 2018). This parameter is also important to predict the neutrino emission (Horiuchi et al. 2017). The tight correlation of ξ2.5 to the CO-core mass has been shown (Limongi & Chieffi 2018), and a detailed study on the compactness was performed by Sukhbold & Woosley (2014) and Sukhbold et al. (2018). Effects of convective boundary mixing in the advanced evolutionary stages on ξ2.5 was studied by Davis et al. (2019).

Figure 16 shows the compactness parameter ξ2.5 of our models. Although the results show a large scatter, ξ2.5 roughly increases with ZAMS mass. The scatter is reduced when the x-axis is changed from the ZAMS mass to the He-core mass (panel (b)) as shown in Sukhbold et al. (2018). As for the correlation between ξ2.5 and the He-core mass, the scatter is small for the models with MHe ≲ 5 M. Furthermore, we do not see a clear dependence of ξ2.5 on fA in this mass range. For more massive models, the scatter is larger, and the different overshoot parameters influence ξ2.5 in a complicated manner.

Figure 16.

Figure 16. Compactness parameter ξ2.5 as a function of MZAMS (panel (a)) and of MHe (panel (b)).

Standard image High-resolution image

Because of the limited number of our models, it is difficult to make a detailed comparison between our results on the compactness parameter (Figure 16) with Sukhbold et al. (2018) (see their Figure 8). We do see a diversity of ξ2.5 in MHe ≳ 10 M; however, the range of ξ2.5 for MHe ≲ 6 M is almost within the same range as in Sukhbold et al. (2018). Although a more systematic study needs to be done as in Sukhbold et al. (2016, 2018), the overall feature (i.e., the increasing trend of the compactness parameter with the ZAMS masses) is roughly in accordance with Sukhbold et al. (2018).

Finally, Figure 17 shows the mass fraction distributions as a function of the radius at the last step of nine models, the 2D hydrodynamics simulations for which were performed but are not shown in Figure 3.

Figure 17.
Standard image High-resolution image
Figure 17.

Figure 17. First six panels: same as Figure 3 but for models 16MA (top left), 18MA (top right), 21MA (middle left), 22L (middle right), 23LA (bottom left), and 27LA (bottom right). Last three panels: same as Figure 3 but for 27M (top left), 28LA (top right), and 28M (bottom).

Standard image High-resolution image

Footnotes

  • "HOSHI" is a noun in Japanese meaning star.

  • The employed nuclear species are as follows: 1n, 1,2H, 3,4He, 6,7Li, 7,9Be, 8,10,11B, 11–16C, 13–18N, 14–20O, 17–22F, 18–24Ne, 21–26Na, 22–28Mg, 25–30Al, 26–32Si, 27–34P, 30–37S, 32–38Cl, 34–43Ar, 36–45K, 38–48Ca, 40–49Sc, 42–51Ti, 44–53V, 46–55Cr, 48–57Mn, 50–61Fe, 51–62Co, 54–66Ni, 56–68Cu, 59–71Zn, 61–73Ga, 63–75Ge, 65–76As, 67–78Se, and 69–79Br. The isomeric state of 26Al is also included.

  • 10 

    Note that strong turbulent activity still continues for model 27LA at the last step of the simulation. We shall perform a 3D hydrodynamics simulation for other models including model 27LA in future work.

Please wait… references are loading.
10.3847/1538-4357/ab2b9d