DIMENSION AS A KEY TO THE NEUTRINO MECHANISM OF CORE-COLLAPSE SUPERNOVA EXPLOSIONS

, , , and

Published 2010 August 12 © 2010. The American Astronomical Society. All rights reserved.
, , Citation J. Nordhaus et al 2010 ApJ 720 694 DOI 10.1088/0004-637X/720/1/694

0004-637X/720/1/694

ABSTRACT

We explore the dependence on spatial dimension of the viability of the neutrino heating mechanism of core-collapse supernova explosions. We find that the tendency to explode is a monotonically increasing function of dimension, with three dimensions (3D) requiring ∼40%–50% lower driving neutrino luminosity than one dimension and ∼15%–25% lower driving neutrino luminosity than two dimensions (2D). Moreover, we find that the delay to explosion for a given neutrino luminosity is always shorter in 3D than 2D, sometimes by many hundreds of milliseconds. The magnitude of this dimensional effect is much larger than the purported magnitude of a variety of other effects, such as nuclear burning, inelastic scattering, or general relativity, which are sometimes invoked to bridge the gap between the current ambiguous and uncertain theoretical situation and the fact of robust supernova explosions. Since real supernovae occur in three dimensions, our finding may be an important step toward unraveling one of the most problematic puzzles in stellar astrophysics. In addition, even though in 3D, we do see pre-explosion instabilities and blast asymmetries, unlike the situation in 2D, we do not see an obvious axially symmetric dipolar shock oscillation. Rather, the free energy available to power instabilities seems to be shared by more and more degrees of freedom as the dimension increases. Hence, the strong dipolar axisymmetry seen in 2D and previously identified as a fundamental characteristic of the shock hydrodynamics may not survive in 3D as a prominent feature.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

It was shown some time ago that multi-dimensional instabilities obtain, and are probably central to the core-collapse supernova mechanism (Burrows & Fryxell 1992; Herant et al. 1992, 1994; Burrows et al. 1995; Janka & Müller 1996). However, it has not been clear whether modest effects, particularly in the neutrino sector, that individually might have little consequence, could accumulate to collectively push the core beyond the threshold of instability into explosion (Mezzacappa et al. 1998, 2001, 2007; Bruenn et al. 2006, 2009; Marek & Janka 2009). Moreover, though the explosion energy is only 1051 erg, since more than 1053 erg of neutrinos issue from collapse, the supernova phenomenon might seem to be a "one-percent" effect, necessitating attention to every detail to see the qualitative effect that is the supernova. However, in the crucial early post-bounce phase of a few hundred milliseconds, only a few times 1052 erg in electron-type neutrinos emerge, making the neutrino-driven explosion more like a "tens of percent" effect, challenging the notion that fine detail in the neutrino sector is the key to the viability of the neutrino explosion mechanism.

It may well prove to be the case that the fundamental impediment to progress in supernova theory over the last few decades has not been lack of physical detail, but lack of access to codes and computers with which to properly simulate the collapse phenomenon in three dimensions (3D). This could explain the agonizingly slow march since the 1960s toward demonstrating a robust mechanism of explosion. State-of-the-art two-dimensional (2D) simulations are still ambiguous and problematic (Section 2), though they manifest instability and turbulence and increase the dwell time of matter in the gain region (Bethe & Wilson 1985), and, hence, the efficiency of the neutrino-matter coupling (Murphy & Burrows 2008). The gain region is the low-optical (-neutrino) depth region behind the shock where neutrino heating exceeds neutrino cooling. Moreover, relative to the spherical case, 2D effects have been shown to enlarge the gain region itself and place more matter in shallower climes of the gravitational potential well (from which it is easier to launch). Two-dimensional models also enable accretion in one quadrant to power explosion in another (Burrows et al. 2007c), something impossible in one dimension (1D) and one reason neutrino-driven explosions in 1D are generally thwarted. In 1D, an outward explosion diminishes the mass accretion rate onto the residual core that powers a good fraction of the driving neutrino luminosity during the critical incipient explosive phase.

However, the difference in the character of 3D turbulence, with its extra degree of freedom and inverse cascade to smaller turbule scales than are found in 2D, may further increase the time matter spends in the gain region, so that the critical condition for explosion (Burrows & Goshy 1993; Murphy & Burrows 2008) is more easily achieved. Our thesis in embarking upon this project was that if going from 2D to 3D improves prospects for the neutrino-driven explosion by the same degree as already demonstrated by going from 1D to 2D (see Section 2), then the neutrino-driven mechanism would perhaps be shown to be not only viable, but also robust. In that case, the details of neutrino interactions, general relativity, and nuclear physics would be of secondary importance for demonstrating the mechanism of explosion, though would still be centrally important for determining the actual explosion energies, the magnitude and systematics of pulsar kicks, the residual neutron star masses, and the nucleosynthesis, to name just a few important aspects of the core-collapse supernova phenomenon.

Hence, in this paper, we have conducted a parameter study, similar to that performed by Murphy & Burrows (2008; see Section 3), but directly comparing 1D (spherical), 2D (axisymmetric), and 3D (Cartesian) simulations, while varying the driving neutrino luminosity. To isolate the effect of dimension, we ensured that all other aspects of the runs (equation of state, progenitor model, and neutrino heating algorithm, etc.) were the same. Note that for a given massive-star progenitor, the accretion rate in front of the stalled shock is independent of dimension, since the initial model is spherically symmetric and the post-shock matter is out of sonic contact with the supersonic infalling matter in front of the shock. Therefore, the character of infall, the evolution of the accretion rate and the mass interior to the shock, and by extension the evolution of the neutrino luminosities, are functions more of progenitor than of dimension (until explosion).

In what follows, we repeat the 1D and 2D study of Murphy & Burrows (2008), but, extending it to 3D, verify that the viability of explosion by delayed neutrino heating is indeed a monotonically increasing function of dimension. In going from 2D to 3D, the enhancements in the efficiency of the neutrino-matter coupling in the gain region and the decreases in the critical luminosity are comparable to those seen in going from 1D to 2D. In a very real sense, this implies that it is ∼50% easier to explode in 3D than in 1D, a huge difference. Since real supernovae occur in 3D, this conclusion may be a major step in unraveling one of the most recalcitrant puzzles in astrophysics.

In Section 2, we review the current status of numerical supernova theory. Then, in Section 3, we summarize the Compressible ASTRO (CASTRO) code and the simple algorithm by which we incorporate neutrino heating and cooling for the purposes of this parameter study. We follow in Section 4 a discussion of our general results, highlighting the role of dimension alone in facilitating explosion. We conclude in Section 5 a summary of what we have learned.

2. SUMMARY OF CURRENT SIMULATION EFFORTS AND RESULTS

Colgate et al. (1961) pioneered the direct (hydrodynamic) mechanism of core-collapse supernovae, in which the bounce shock, launched when collapse is reversed, propagated outward without stalling. Following this, Colgate & White (1966) proposed the neutrino mechanism, wherein the agency of explosion was a burst of neutrino heating behind the shock after bounce. Neutrinos must be included at the high temperatures and densities achieved in collapse. Arnett (1966) and Wilson (1971) refined this model, with the latter challenging its efficacy. These simulations were done in spherical symmetry and employed less sophisticated numerical techniques and cruder microphysics than can be marshalled today. Importantly, current thinking is that the direct hydrodynamic mechanism never works. The neutrino burst that occurs at shock breakout of the neutrinosphere and, to a lesser degree, photodissociation of nuclei into nucleons at the nascent shock debilitate it into accretion. Moreover, and disappointingly, modern simulations with the latest nuclear and neutrino physics, up-to-date progenitor models, and sophisticated numerics demonstrate that if stellar core collapse is constrained to be spherically symmetric most cores cannot supernova (Burrows et al. 1995; Mezzacappa et al. 2001; Rampp & Janka 2000, 2002; Liebendörfer et al. 2001, 2005; Buras et al. 2003; Thompson et al. 2003). However, Kitaura et al. (2006) do obtain a spherical neutrino-wind-driven explosion for the 8.8 M progenitor of Nomoto & Hashimoto (1988), after a slight post-bounce delay (see also Burrows et al. 2007a). Such neutrino-driven winds are generic after explosion (Burrows et al. 1995), but are too weak to generate a 1051 erg explosion. This progenitor can explode in 1D because its envelope is extremely rarified, and the wind emerges almost immediately, but the associated energy is uncharacteristically meager (∼2.5 × 1049 erg, more than an order of magnitude smaller than the canonical value). It explodes spherically because there is no inhibiting accretion tamp and cannot be generic.

In the 1980s, Wilson (1985) suggested the "delayed" mechanism of explosion, wherein the stalled shock was revived by neutrino heating, but after hundreds of milliseconds. This mechanism was championed by Bethe & Wilson (1985), who introduced the concept of the "gain region" behind the shock, where net neutrino energy deposition was positive. Though his calculations were done in 1D, Wilson required "neutron-finger" convection below the neutrinospheres to boost the driving neutrino luminosities. Such instabilities, and the corresponding luminosity boosts, have been called into question by Bruenn & Dineva (1996) and Dessart et al. (2006a).

In the 1990s, the available computer power enabled detailed numerical explorations beyond spherical symmetry into two dimensions (Burrows & Fryxell 1992; Herant et al. 1992, 1994; Burrows et al. 1995; Janka and Müller 1996), with some of these 2D calculations leading to delayed neutrino-driven explosions. These were the first simulations of collapse and explosion that demonstrated the existence and potential centrality of hydrodynamic instabilities and neutrino-driven convection in the supernova phenomenon. Fryer & Warren (2002, 2004) and Fryer & Young (2007) later performed 3D simulations, but used a smoothed particle hydrodynamics code that may not have adequately handled the resultant instabilities and turbulence. Nevertheless, those were some of the first 3D simulations in supernova theory that attempted to include the relevant physics.

Mezzacappa et al. (1998) challenged the notion that multi-dimensional effects were central to the supernova phenomenon, suggesting that precision spherical simulations for which detailed neutrino transport was incorporated would be necessary to demonstrate robust explosions. However, this is not the current view, with multi-dimensional effects first explored in the 1990s now assuming center stage. More recently, using purely hydrodynamic calculations, Blondin et al. (2003) showed that the shock itself, even without neutrino heating and the consequent convection, could be unstable, introducing the "standing accretion shock instability" (SASI). This hydrodynamic phenomenon has been studied analytically and in detail by Foglizzo et al. (2006, 2007). One cannot easily disentangle neutrino-driven convection and the SASI, and it may be that neutrino-driven convection is predominant (Fernandez & Thompson 2009). The major contribution of the Blondin et al. paper was to highlight the need to perform 2D calculations over the full 180° so as not to suppress the interesting dipolar component. Most previous calculations had been performed on a 90° wedge. Otherwise, the "SASI" had been manifest and observed naturally in previous multi-dimensional simulations.

Buras et al. (2006a, 2006b), using their code MuDBaTH, obtained an explosion in 2D with the 11.2 M progenitor of Woosley & Heger (2006). However, this model's explosion was underpowered by an order of magnitude. More recently, Marek & Janka (2009) obtained an explosion of a rotating 15 M progenitor (their model M15LS-rot), a more representative star. However, they experienced a long delay to explosion of more than 600 ms, by which time the mass in the gain region was too low to absorb enough neutrino energy to achieve more than ∼2.5 × 1049 erg by the end of their simulation. Moreover, the shock wave was not followed beyond ∼600 km and they used a soft nuclear equation of state with an incompressibility at nuclear densities, K, of 180 MeV. Their model with K = 263 MeV did not explode. An incompressibility nearer 240 ± 20 MeV is currently preferred by measurement (Shlomo et al. 2006). In addition, Marek and Janka's high-resolution run of the same model (M15LS-rot-hr) did not explode, though it was continued to nearly the same final post-bounce epoch as their exploding model M15LS-rot.

The Oak Ridge supernova group, led by A. Mezzacappa and S. Bruenn, is paralleling in many ways with the work of Marek & Janka (2009), but is using its CHIMERA code (Bruenn et al. 2006, 2009; Mezzacappa et al. 2007) and Bruenn's 1D multi-group, flux-limited diffusion algorithm (Bruenn 1985) in the multiple 1D "ray-by-ray" transport approximation.3 As do Marek & Janka (2009), the ORNL team employs the lower-K, softer nuclear equation of state (EOS). They have found robust explosions for a variety of progenitors (Bruenn et al. 2006, 2009; Mezzacappa et al. 2007) and have suggested that the mechanism is a combination of neutrino heating, nuclear burning (of infalling oxygen at the shock), the SASI, and inelastic neutrino–electron and neutrino–nucleon scattering. However, the latter should be a subdominant 5%–10% effect. The potential role of nuclear burning on infall had been studied and eliminated earlier (Burrows & Lattimer 1985; H.-T. Janka 2009, private communication) and though included in many other previous simulations has not been seen to have played a positive role. Infalling oxygen should burn long before reaching the shock wave that has stalled near ∼150–200 km and, moreover, the amount of nuclear energy available is rather small. Why the results of these two groups differ so substantially (when their approaches are ostensibly so similar), in both explosion energy when explosions are claimed, and in whether they see explosions at all, is a puzzle.

Those using the fully 2D radiation/hydrodynamics code VULCAN (Livne 1993; Livne et al. 2004, 2007; Burrows et al. 2006, 2007c, 2007d), apart from obtaining neutrino-driven explosions in the contexts of a 8.8 M progenitor (Burrows et al. 2007a) and accretion-induced collapse (Dessart et al. 2006b), do not witness neutrino-driven explosions for most progenitors. VULCAN/2D incorporates an implicit, time-dependent, multi-energy-group transport scheme that has both multi-angle (Sn) and flux-limited (MGFLD) variants in two spatial dimensions. VULCAN is the first and (to date) only multi-group supernova code to operate in 2D in both the hydro and radiation sectors. Interestingly, using a multi-angle variant of VULCAN, Ott et al. (2008) have shown that for non-rotating models the results for the multi-angle and MGFLD simulations are similar. VULCAN also incorporates "2.5D" MHD and, hence, is the only code currently employed in supernova studies with both multi-group transport and MHD capabilities. With it, Burrows et al. (2007b) have explored MHD-driven jet explosions (LeBlanc & Wilson 1970; Bisnovatyi-Kogan et al. 1976; Akiyama & Wheeler 2005; Wheeler & Akiyama 2007), but concluded that such explosions, which require very rapid rotation at bounce that is very unlikely to be generic, might obtain only in the rare "hypernova" case and could not explain the typical supernova.

However, using VULCAN and after simulating for around a second after bounce, Burrows et al. (2006, 2007d) observed vigorous dipolar g-mode oscillations that damp by the asymmetric emission of sound waves that steepen into shock waves. This acoustic power has been adequate to explode all progenitors, but would be aborted if the neutrino mechanism, or some other mechanism, obtained earlier. This acoustic mechanism is controversial, takes a long time (many seconds) to achieve explosion energies of ∼1051 erg, and has not been seen by other groups. However, no other group has calculated for the physical time necessary to witness vigorous g-mode excitation. Unlike all VULCAN simulations, most other simulations have been done with the inner core in 1D.4 As a result, we suggest that other groups have not simulated long enough to see this phenomenon. Moreover, we suspect that installing a spherical inner core will partially suppress the excitation of a core g-mode that is fundamentally dipolar, not spherical, and whose amplitude should be largest at the center. A solid criticism of the acoustic mechanism comes from Weinberg & Quataert (2008), who using approximate analytic techniques, suggest that the nonlinear excitation of dissipative daughter modes via a parametric resonance might render the amplitudes of these dipolar g-mode oscillations too small to power a supernova. Unfortunately, the wavelengths of these daughter modes are smaller than reasonable grid spacings and cannot easily be captured by extant supernova codes. Hence, whether a very delayed acoustic power mechanism for explosion is at all viable in any circumstance remains to be seen.

However, the delayed neutrino mechanism seems compelling, if only it can be shown to work robustly and to give canonical explosion energies of ∼1051 erg generically. Burrows & Goshy (1993) suggested that the neutrino-driven explosion is a critical phenomenon. There is a critical luminosity at a given mass accretion rate into the shock above which there is no accretion solution and the mantle explodes. Recently, Murphy & Burrows (2008) have shown that this critical luminosity decreases and the neutrino-matter coupling efficiency increases by ∼30% in going from 1D to 2D. Murphy & Burrows (2008) have diagnosed the increase in heating efficiency in 2D (vis-à-vis 1D) as the increase in the average dwell time (Thompson et al. 2005) of matter in the gain region due to neutrino-driven convection and the SASI, before it settles into the cooling region and onto the inner core. What has not been shown, and is our goal with this paper, is the corresponding effect in going from 2D to 3D.

The absence of inelastic scattering terms and corrections for general relativistic effects in VULCAN code might compromise its conclusions and be the reason those using VULCAN are not seeing neutrino-driven explosions in the generic case, or when other groups obtain such explosions (however tepid). However, the inelastic terms are small, and relativity's effects nearly cancel. A more plausible explanation for the ambiguity rife in the field and the marginality (and rarity) of neutrino-driven explosions among published simulations is that the neutrino mechanism may not work well in 2D. It may be that the neutrino mechanism is truly viable only in 3D and that 3D effects have been the missing ingredient needed to explain supernova explosions with delayed neutrino driving. This is the thesis of our paper. We posit that the viability of the neutrino-driven mechanism for core-collapse supernova explosions and the energy of explosion (when they explode) are monotonic functions of dimension. We note that it is only recently, with the advent of computers of sufficient size and speed, that theorists have been able to test such an hypothesis. Furthermore, we suggest that the ambiguity of the extant 2D results is a symptom of its marginality in 2D. Moreover, as did Murphy & Burrows (2008), we conjecture that the increased dwell time in the gain region in going to 3D, due to both the extra degree of freedom for non-radial motion and the character of the cascade in 3D turbulence to small turbule sizes,5 renders the heating efficiency and all the other ancillary quantities that support explosion adequate to transform "marginality and ambiguity" into "robustness and viability" (Section 4).

3. CASTRO CODE AND GENERAL METHODOLOGY

To address these issues in a full 3D context, we have employed CASTRO (Almgren et al. 2010), a new multi-dimensional radiation/hydrodynamic code. CASTRO is an Eulerian, structured grid, compressible, radiation/hydrodynamics code that incorporates adaptive mesh refinement (AMR). It uses a second-order unsplit piecewise-parabolic Godunov methodology for general convex equations of state. The most general treatment of self-gravity in CASTRO uses multigrid to solve the Poisson equation for the gravitational potential. For the calculations presented here, the monopole approximation is used.

AMR in CASTRO uses a nested hierarchy of rectangular grids that are simultaneously refined in both space and time. CASTRO uses a recursive integration procedure in which coarse grids are advanced in time, fine grids are advanced multiple steps to reach the same time as the coarse grids, and the data at different levels are then synchronized. For regridding, an error estimation procedure evaluates where additional refinement is needed and grid generation procedures dynamically create or destroy fine grid patches to achieve the desired local resolution. CASTRO is implemented within the BoxLib framework that handles data distribution, communication, memory management, and I/O for parallel architectures.

To most closely parallel the approach of Murphy & Burrows (2008), while generalizing to three dimensions, our calculations were done with CASTRO using the simple neutrino heating and cooling algorithm described in Murphy & Burrows (2008). We solve the fully compressible equations of hydrodynamics:

Equation (1)

where ρ, T, p, g, and u are the fluid density, temperature, pressure, gravitational acceleration, and velocity. The total energy is given by $E = e + \frac{1}{2}u^2$, where e represents the internal energy. The equation of state provides closure and is a function of ρ, T, and electron fraction, Ye. While a sophisticated nuclear equation of state is used (Shen et al. 1998a, 1998b), the neutrino heating and cooling rates behind the stalled shock wave via the super-allowed charged-current reactions involving free nucleons assume a given electron neutrino and anti-electron neutrino luminosity (taken to be the same). This luminosity is varied from simulation to simulation, but is held constant during a simulation. The neutrino heating, $\mathcal {H}$, and cooling, $\mathcal {C}$, rates are those derived in Janka (2001), used in Murphy & Burrows (2008), and given by

Equation (2)

and

Equation (3)

where $L_{\nu _e}$ is the electron neutrino driving luminosity, $T_{\nu _e}$ is the electron neutrino temperature, r is the distance from the center of the star, Yn and Yp are the neutron and proton fractions, and $\tau _{\nu _e}$ is the electron neutrino optical depth. Note that these approximations assume that the $L_{\nu _e}=L_{\bar{\nu }_e}$.

This approach, using Equations (2) and (3) in place of full transport, enabled the extensive (and computationally expensive) parameter study in 1D, 2D, and 3D we report here. The single progenitor we employ for all our runs is the non-rotating, solar-metallicity 15 M red-supergiant model of Woosley & Weaver (1995), but we need only the inner ∼10,000 km for our simulations of core collapse. The calculations commence at the onset of infall and are carried out to and after bounce. The explosion time is formally determined when the average shock radius reaches 400 km and does not recede during subsequent evolution (see Table 1). Many of the 1D and 2D simulations are carried to beyond a second after bounce (up to ∼1.4 s). The 3D simulations were all carried out to beyond 400 ms after bounce. The Liebendörfer et al. (2005) scheme for determining the electron fraction on infall was used, which involves tying Ye to the mass density achieved for densities below the trapping density.

Table 1. Explosion Times and Accretion Rates

$L_{\nu _e,\bar{\nu }_e}$ Model texp $\dot{M}_{\rm exp}$ Model texp $\dot{M}_{\rm exp}$ Model texp $\dot{M}_{\rm exp}$  
(1052 erg s−1) 1D (ms) (M s−1) 2D (ms) (M s−1) 3D (ms) (M s−1)  
1.7 1d:L_1.7  ⋅⋅⋅   ⋅⋅⋅  2d:L_1.7  ⋅⋅⋅   ⋅⋅⋅  3d:L_1.7 415 0.28  
1.8       2d:L_1.8  ⋅⋅⋅   ⋅⋅⋅         
1.9 1d:L_1.9  ⋅⋅⋅   ⋅⋅⋅  2d:L_1.9 1045 0.14 3d:L_1.9 254 0.41  
1.95       2d:L_1.95 444 0.24        
2.0 1d:L_2.0  ⋅⋅⋅   ⋅⋅⋅  2d:L_2.0 240 0.29        
2.1 1d:L_2.1  ⋅⋅⋅   ⋅⋅⋅  2d:L_2.1 229 0.33 3d:L_2.1 154 0.54  
2.3       2d:L_2.3 217 0.39        
2.5 1d:L_2.5 748 0.15 2d:L_2.5 174 0.43        
2.6 1d:L_2.6 679 0.18 2d:L_2.6 165 0.48        
2.7 1d:L_2.7 568 0.21 2d:L_2.7 163 0.49        
2.8 1d:L_2.8 365 0.27 2d:L_2.8 155 0.53        
2.9 1d:L_2.9 178 0.41              
3.0 1d:L_3.0 168 0.47              
3.1 1d:L_3.1 159 0.51              

Notes. The explosion time, texp, is calculated when the average shock radius, 〈Rshock〉, reaches 400 km and does not recede during subsequent evolution. The accretion rate at the time of explosion is calculated just exterior to the shock where the infalling envelope is spherical and is given by $\dot{M}_{\rm exp} = 4\pi \rho v r^2$. A "..." indicates that the model did not explode during the allotted maximum simulation time (1.4 s). The bold numbers are those for models which both exploded within 1.4 s and were calculated in either both 1D and 2D or both 2D and 3D. As this table indicates that there were models that were calculated in 1D, 2D, and 3D, but the corresponding 1D models did not explode within 1.4 s.

Download table as:  ASCIITypeset image

Interior to a radius of 200 km, all of our simulations have zones smaller than a kilometer. Exterior to 200 km, the resolution dynamically follows the shock via adaptive grids. In 3D, our Cartesian domain is a cube of length 10,000 km. The domain is discretized at the coarsest level with a 3043 uniform grid, and several levels of adaptive refinement are used so that cells at the finest level are a factor of 64 finer in each direction. This results in a resolution of ∼0.5 km (which we refer to as the "effective" resolution) in the interior 200 km of the star and in other regions at the highest refinement level. For our axisymmetric simulations, we employ a 2D domain that is 10,000 km by 5000 km, discretized with a uniform coarse grid of 256 by 128 cells. All of our 2D simulations cover the full 180° angular range from north to south pole. Employing the same refinement criteria used in 3D yields an effective resolution of ∼0.6 km for our 2D simulations. For our spherically symmetric (1D) simulations, we can obtain, and have tested, higher resolution runs. Additional resolution leads to minimal differences (≲10 ms) in the reported explosion times. The 1D results presented in this paper are for a radial domain of 256 coarse cells (with similar refinement to that employed in the multi-dimensional simulations), distributed over 5000 km for an effective resolution of ∼0.3 km.

4. RESULTS

Depicted in Figure 1 are critical curves in driving luminosity versus accretion rate ($\dot{M}$) through the shock for calculations performed in 1D, 2D, and 3D, all else being equal. The luminosity is in units of 1052 erg s−1 and the mass accretion rate is in solar masses per second (M s−1). $\dot{M}$ is 4πr2ρ|vR| in the infalling material just exterior to the shock wave. Above each curve, the core explodes and below each curve it does not. As shown in Murphy & Burrows (2008), the 1D curves approximately recapitulate the 1D analytic theory found in Burrows & Goshy (1993). The important result in this paper is the position of the 3D curve. It is ∼15%–25% below the corresponding 2D curve, which is in turn ∼30% below the 1D curve. Moreover, the advantage of going to 3D is larger for higher driving luminosities. The upshot is that the magnitude of the driving neutrino luminosity necessary to "supernova" a given core for a given mass accretion rate through the shock is reduced in going from 1D to 3D by ∼40%–50%, a rather large, perhaps enabling, effect.

Figure 1.

Figure 1. Critical curves in electron–neutrino driving luminosity ($L_{\nu _e}$) vs. accretion rate ($\dot{M}$) for calculations performed in 1D (black), 2D (blue), and 3D (red). The luminosity is in units of 1052 erg s−1 and the mass accretion rate ($\dot{M}$) is in solar masses per second. Importantly, we include results in 3D. Note that the driving electron neutrino luminosity is always accompanied by an associated anti-electron neutrino luminosity (L$_{\bar{\nu }_e}$). See Section 4 for a discussion of the meaning of this plot and its salient features.

Standard image High-resolution image

Note that in our study each run is performed for a given neutrino and anti-neutrino luminosity, while the accretion rate onto the inner core evolves with time, running from high to low values of $\dot{M}$ in a way determined by the initial inner density profile of the chosen progenitor star (in this case the 15 M progenitor of Woosley & Weaver 1995). Hence, the $\dot{M}$ is not fixed during a run and if another progenitor model had been chosen the temporal evolution of $\dot{M}$ would have been different. When the $\dot{M}$ reaches the value for a given fixed L$_{\nu _e}$ at which the core explodes, we identify this L$_{\nu _e} - \dot{M}$ pair as a point on the critical curve (Figure 1). Performing this exercise for a range of luminosities maps out the corresponding critical curve in whatever number of dimensions is being studied.

Since before explosion the trajectory in L$_{\nu _e}(t)-\dot{M}(t)$ space followed by a given progenitor is a weak function of dimension, depending mostly on the initial progenitor density profile, which in turn determines $\dot{M}$ and thereby the inner core density, temperature, and Ye profiles (given an EOS, a neutrino transport algorithm, and quasi-hydrostatic equilibrium), changing the dimension of a simulation is not met with significant compensating shifts in the actual $L_{\nu _e}(t)$ versus $\dot{M}(t)$ trajectory with which to compare our critical curves. The result is an undiminished and monotonic advantage at higher dimensions of significant magnitude. The effect of dimension is seen to be not merely a few percent, but nearly a factor of 2 in a central aspect of the collapse context, the driving neutrino luminosity. This new result leads us to suggest that lack of access to 3D computational capabilities has been a major retarding factor in progress during the last few decades toward the solution to the core-collapse supernova problem.

Plotted in Figure 2 are various curves depicting the temporal evolution of the average radius of the shock (in kilometers) for various driving electron neutrino luminosities and for simulations in 1D, 2D, and 3D. This figure shows that for luminosities for which the 1D and 2D model shocks remain stagnant for long periods, the 3D models explode much earlier. For instance, at $L_{\nu _e} = 1.9\times 10^{52}$ erg s−1, in 1D the core does not explode even after 1.4 s, the 2D core explodes after around ∼0.8–1.0 s, but the 3D core explodes only after ∼250 ms. In all cases, the time to explosion (if there is an explosion) is shorter at higher dimension than at lower dimension. Bruenn et al. (2006, 2009) witness explosions in all their recent 2D simulations, but these results have not been reproduced by others (cf. Marek & Janka 2009). Moreover, all their explosions pile up at similar early times, as does the 3D simulation, they are currently tending. This suggests that whatever is causing their explosions does not much distinguish between 2D and 3D in the way we so clearly see in our suite of simulations. The reason for this is unclear, but we note that when we obtain early explosions (in this paper, at the highest driving neutrino luminosities), the difference in the time to explosion in 2D and 3D is similarly significantly reduced.

Figure 2.

Figure 2. Depicted here are curves showing the temporal evolution of the average radius of the shock (in kilometers) for various driving electron neutrino luminosities and for simulations in 1D, 2D, and 3D. The 1D models are rendered as black dashed lines, the 2D models as solid blue lines, and the 3D models as thick solid red lines. Each line is indexed by the corresponding electron–neutrino luminosity of the simulation, in units of 1052 erg s−1. Implied is an equal anti-electron–neutrino luminosity. Time is in seconds after bounce. See the text in Section 4 for a discussion of this figure.

Standard image High-resolution image

Table 1 clearly demonstrates that the time to explosion is shorter at higher dimension than at lower dimension and provides a more extended compilation of 1D, 2D, and 3D exploding models and the approximate times at which they explode. The table is arranged so that overlapping horizontal rows, though done for a different number of dimensions, have the same driving luminosities. This format clearly reveals that, all else being equal, the time to explosion is significantly shorter at higher dimension. For example, the 1D model at $L_{\nu _e} = 2.5\times 10^{52}$ erg s−1 explodes around ∼0.75 s while the corresponding 2D model explodes near 0.2 s. Interestingly, this is the time at which the 3D model at the much lower luminosity of 1.9 × 1052 erg s−1 explodes. The time it takes the 1.9 × 1052 erg s−1 model in 3D to lift the average shock radius from ∼200–300 km to ∼1200 km during the early explosion phase is ∼200 ms, at which point the shock is moving at a speed of ∼30,000 km s−1. As indicated in the figure, early in the incipient explosion phase the average shock radius gradually, but steadily, accelerates.

In the panels of Figure 3, we compare representative entropy color maps of simulations for the same neutrino luminosities and times after bounce, but for different numbers of dimensions. The top two panels contrast 1D (left) and 2D (right) runs, both for a luminosity of 2.5 × 1052 erg s−1 and at ∼468 ms after bounce. Note that while the 2D run is exploding, the 1D run is not, though the physical model and inputs are otherwise identical. The positions of the shocks in these simulations, as indicated by the abrupt color transition, are very different. The bottom panels compare models in 2D (left) and 3D (right), both for a luminosity of 1.9 × 1052 erg s−1 and at ∼422 ms after bounce. While the shock in the 2D simulation is still stalled, the shock in the 3D simulation has launched dramatically. These snapshots together and in conjunction demonstrate pictorially the qualitative dependence on dimension of the outcome of collapse. Since supernovae occur in three spatial dimensions, we conclude that this dimensional dependence is of direct relevance to the viability of the neutrino heating mechanism for core-collapse supernova explosions. Moreover, the magnitude of the "advantage" of going to 3D (see Figure 1) should dwarf that of any refinements in the microphysics, such as inelastic neutrino-matter scattering, nuclear burning upon infall, or general relativity. The latter are at most "∼10%" effects and are small compared with the ∼40%–50% effects we identify here that are associated with dimension, in particular in going to 3D.

Figure 3.

Figure 3. In these panels, we compare representative entropy maps of two simulations for the same driving neutrino luminosities and times after bounce, but for different numbers of dimensions. The top two panels contrast 1D (left) and 2D (right) runs, both for an electron-type neutrino luminosity of 2.1 × 1052 erg s−1 and at ∼0.468 ms after bounce. The bottom panels compare models in 2D (left) and 3D (right), both for an electron-type neutrino luminosity of 1.9 × 1052 erg s−1 and at 0.422 ms after bounce. The same color map is used for all four panels. Note that while the two top panels are for the same luminosity and epoch after bounce, only the 2D simulation has exploded. Similarly, while the two bottom panels are at the same luminosity and time after bounce, the development of the 3D simulation is qualitatively different from that of the corresponding 2D run. Also note the different general morphologies of the 3D (bottom) and 2D (top) models that explode. See the text in Sections 4 and 5 for a discussion.

Standard image High-resolution image

Figure 4 depicts entropy scatter plots near and after the time of explosion of the corresponding higher-dimensional run for (1) 1D and 2D models (top) and (2) 2D and 3D models (bottom). In each case, the lower-dimensional model of a given comparison (either top or bottom) has not exploded during the time spanned by the plots. The top panels are for L$_{\nu _e} = 2.1\times 10^{52}$ erg s−1 and the bottom panels are for L$_{\nu _e} = 1.9\times 10^{52}$ erg s−1. The times to explosion are given in Table 1. Figure 4 clearly indicates that, all else being equal, higher entropies are achievable at higher dimension. Figure 5 depicts the corresponding mass-weighted average entropy in the gain region versus time for the same 1D (black), 2D (blue), and 3D (red) models and makes the same point. Importantly, the average entropy for the 3D run in the runup to explosion is ∼1.5 units higher than in 2D. This expands the physical extent of the gain region, makes it less bound, and creates what seems to be a more "explosive" situation (cf. Figures 13). The higher average entropies can be traced to higher average dwell times of a Lagrangian parcel of matter in the gain region (Murphy & Burrows 2008). We plan to present a more detailed analysis of this and other effects in a subsequent paper.

Figure 4.

Figure 4. Scatter-plot entropy profiles vs. radius for two different driving luminosities, comparing 1D and 2D (top) and 2D and 3D (bottom). The 1D (red) and 2D (black) comparison at the top is for L$_{\nu _e}$ = 2.1 × 1052 erg s−1 and the 2D (red) and 3D (black) comparison at the bottom is for L$_{\nu _e}$ = 1.9 × 1052 erg s−1. For both top and bottom, two different times after bounce are given—the times for both left panels are near the onset of the explosion of the associated higher-dimensional model. Note that the peak entropies and average entropies are always higher for the higher-dimensional runs and that for a given dimension the higher the driving luminosity the higher the peak entropy achieved. Also note that different horizontal radius scales are employed for the left and right panels and that the associated model names are given on the plots. See the discussion in Section 4.

Standard image High-resolution image
Figure 5.

Figure 5. Mass-weighted average entropy in the gain region vs. time after bounce (in seconds) for the 1D (black), 2D (blue), and 3D (red) models with $L_{\nu _e}$ = 1.9 × 1052 erg s−1. This figure, representative of corresponding figures at other driving luminosities, demonstrates the higher entropies achieved in the gain region behind the shock in going to 3D. After ∼0.2 s after bounce, the difference between the 2D and 3D model-average entropies is ∼1.5 units. See Section 4 for a discussion of the meaning and relevance of this figure.

Standard image High-resolution image

In addition, as Figures 3 and 6 indicate, even though we do see pre-explosion and blast asymmetries in 3D, we do not see a strong ℓ = 1 axially symmetric "SASI" oscillation, such as is obtained in 2D in this paper and by others (Blondin et al. 2003; Bruenn et al. 2006, 2009; Burrows et al. 2006, 2007d; Buras et al. 2006a, 2006b; Marek & Janka 2009). Rather, the free energy available to power instabilities seems to be shared by more and more degrees of freedom as the dimension increases (Iwakami et al. 2008). In 1D, as demonstrated by Murphy & Burrows (2008), one sees a strong radial oscillation when the luminosity is near critical. However, 2D models do not manifest this radial mode, but instead execute an ℓ = 1, m = 0 dipolar oscillation. In 3D, this dipolar oscillation exists (Blondin & Mezzacappa 2007; Fernandez 2010), but in the linear limit competes with the m = {−1, 1} modes. The ℓ = 1, m = 0 mode is even less in evidence in the nonlinear limit, which is achieved early in our simulations. Hence, the strong dipolar symmetry seen in 2D and identified as the fundamental characteristic of the "SASI" may not survive in 3D. At the very least, the ℓ = 1, m = 0 mode does not dominate in our non-rotating 3D models. Whether rotation can change this conclusion remains to be seen (Guilet et al. 2010).

Figure 6.

Figure 6. Blast morphology. These panels depict an evolutionary sequence (top left, top right, bottom left, and bottom right) of the $L_{\nu _e} = 1.9\times 10^{52}$ erg s−1 model in 3D. The top left panel depicts a phase well before the onset of explosion. The two surfaces in each panel are ρ = 1012 g cm−3 (blue interior) and Ye = 0.47 (outer) for four different times after bounce (0.156, 0.201, 0.289, and 0.422 s). The scale is more than 2000 km on a side. Note that the crude axis of the explosion is not along any of the three Cartesian directions and that there is no obvious ℓ = 1 asymmetry. See Section 5 for a discussion of the possible implications of this figure and of Figure 3 for the relevance of the ℓ = 1, m = 1 "SASI" mode in the context of core-collapse supernovae.

Standard image High-resolution image

5. CONCLUSIONS

In this paper, we have performed 1D, 2D, and 3D numerical simulations of the collapse, bounce, and explosion of a massive-star core to isolate the effect of spatial dimension in the context of the neutrino heating mechanism of core-collapse supernova explosions. We have found that both the viability of explosion (measured by the driving neutrino luminosity needed to overcome a given mass accretion rate at the shock and its associated accretion pressure tamp) and the position of the "critical curve" are monotonically increasing functions of dimension, with 3D more viable than 2D by 15%–25% and 3D more viable than 1D by almost a factor of 2. Some had thought that 3D would prove similar to 1D, but our calculations do not support this expectation. Moreover, we have discovered that the time to explosion is significantly shorter in 3D than 2D, all else being equal. These results suggest that a key missing ingredient in the recipe for this astrophysically central phenomenon has been access to 3D simulation tools and that the neutrino-driven explosion mechanism is fundamentally 3D. Two-dimensional simulations have to date yielded marginal explosions and/or ambiguous answers (for a discussion, see Janka et al. 2007).

Furthermore, we find that the prominent ℓ = 1, m = 0 dipolar mode of shock oscillation, often identified in 2D as a central aspect of the "SASI," is little in evidence in 3D. The free energy available to excite hydrodynamic instabilities seems, instead, to be shared by more degrees of freedom, diminishing the amplitude of this axisymmetric dipolar component. Whether this conclusion is a function of progenitor and/or rotation has yet to be determined, but the preliminary indications are that this sloshing mode, so visible in 2D simulations, does not survive as a central feature of core collapse when three spatial dimensions are allowed for.

We emphasize that in this study, in order to isolate the crucial effect of dimension, we employed a very simple neutrino "transport" approach. This has enabled us to discover what we suggest is a key aspect of the core-collapse supernova phenomenon. However, detailed neutrino transport must be incorporated to determine the true systematics with stellar progenitor of the explosion energies, nucleosynthesis, residual neutron star (and black hole) masses, pulsar kicks, and blast morphologies. If we had included such transport at this stage in the development of computer hardware and the computational arts, a single 3D simulation would have required many, many years of continuous execution on the largest existing and available massively parallel platforms. As a result, the insights we have achieved via our more modest study might have been delayed or obscured for years. This highlights the continuing benefits of a multi-pronged, varied, and flexible strategy to address this most important, though refractory, astrophysical problem.

The authors acknowledge fruitful past collaborations with, conversations with, or input from Jeremiah Murphy, Christian Ott, Louis Howell, Rodrigo Fernandez, Manou Rantsiou, Tim Brandt, Dave Spiegel, Eli Livne, Luc Dessart, Todd Thompson, Rolf Walder, Stan Woosley, and Thomas Janka. They also thank Hank Childs and the VACET/VisIt Visualization team(s) for help with graphics and with developing multi-dimensional analysis tools. J.N. and A.B. are supported by the Scientific Discovery through Advanced Computing (SciDAC) program of the DOE, under grant DE-FG02-08ER41544, the NSF under the subaward ND201387 to the Joint Institute for Nuclear Astrophysics (JINA, NSF PHY-0822648), and the NSF PetaApps program, under award OCI-0905046 via a subaward 44592 from Louisiana State University to Princeton University. Work at LBNL was supported in part by the SciDAC program under contract DE-FC02-06ER41438. The authors thank the members of the Center for Computational Sciences and Engineering (CCSE) at LBNL for their invaluable support for CASTRO. J.N. and A.B. employed computational resources provided by the TIGRESS high performance computer center at Princeton University, which is jointly supported by the Princeton Institute for Computational Science and Engineering (PICSciE) and the Princeton University Office of Information Technology; by the National Energy Research Scientific Computing Center (NERSC), which is supported by the Office of Science of the US Department of Energy under contract DE-AC03-76SF00098; on the Kraken and Ranger supercomputers, hosted at NICS and TACC and provided by the National Science Foundation through the TeraGrid Advanced Support Program under grant TG-AST100001; and on the Blue Gene/P at the Argonne Leadership Computing Facility of the DOE.

Footnotes

  • Also used by Marek & Janka (2009), and first introduced into supernova theory in more primitive form by Burrows et al. (1995).

  • The VULCAN modelers used an enabling Cartesian grid in the inner 30–50 km.

  • In 2D, the turbulent cascade is artificially inverted and flows to large scales.

Please wait… references are loading.
10.1088/0004-637X/720/1/694