0805.3345/ms.tex
1: %\documentclass[12pt,preprint]{aastex}
2: \documentclass{emulateapj}
3: 
4: %Used for picture environment.
5: \usepackage{color}
6: 
7: \shorttitle{}
8: \shortauthors{}
9: 
10: \begin{document}
11: 
12: \title{Criteria for Core-Collapse Supernova Explosions by the Neutrino Mechanism}
13: 
14: \author{Jeremiah W. Murphy\altaffilmark{1,2,3}}
15: \author{Adam Burrows\altaffilmark{4,1}}
16: 
17: \altaffiltext{1}{Steward Observatory, The University of Arizona,
18:   Tucson, AZ 85721; jmurphy@as.arizona.edu}
19: \altaffiltext{2}{Astronomy Department, University of Washington, Box
20:   351580, Seattle, WA 98195-1580; jmurphy@astro.washington.edu}
21: \altaffiltext{3}{NSF Astronomy and Astrophysics Postdoctoral Fellow}
22: \altaffiltext{4}{Department of Astrophysical Sciences, Princeton University,
23:   Princeton, NJ 08544; burrows@astro.princeton.edu}
24: 
25: \begin{abstract}
26: We investigate the criteria for successful
27: core-collapse supernova explosions by the neutrino mechanism.
28: We find that a critical-luminosity/mass-accretion-rate
29: condition distinguishes non-exploding from exploding models in
30: hydrodynamic one-dimensional (1D) and two-dimensional (2D)
31: simulations.  We present 95 such simulations that parametrically explore
32: the dependence on neutrino luminosity, mass accretion rate, resolution, and dimensionality.
33: While radial oscillations mediate the transition between 1D accretion
34: (non-exploding) and exploding simulations, the non-radial standing
35: accretion shock instability characterizes 2D simulations. We find 
36: that it is useful to compare the average dwell time of matter in the
37: gain region with the corresponding heating timescale, but that tracking the 
38: residence time distribution function of tracer particles better describes
39: the complex flows in multi-dimensional simulations.  Integral
40: quantities such as the net heating rate, heating efficiency, and mass
41: in the gain region decrease with time in non-exploding models, but for
42: 2D exploding models, increase before, during, and after explosion.
43: At the onset of explosion in 2D, the heating efficiency is $\sim$2\% to
44: $\sim$5\% and the mass in the gain region is $\sim$0.005
45: M$_{\sun}$ to $\sim$0.01 M$_{\sun}$.  Importantly, we find that
46: the critical luminosity for explosions in 2D is $\sim$70\%
47: of the critical luminosity required in 1D.
48: This result is not sensitive to resolution or whether the 2D
49: computational domain is a quadrant or the full 180$^{\circ}$.
50: We suggest that the relaxation of the explosion condition in going
51: from 1D to 2D (and to, perhaps, 3D) is of a general character and is 
52: not limited by the parametric nature of this study.
53: \end{abstract}
54: 
55: \keywords{hydrodynamics --- instabilities --- methods: numerical ---
56:   shock waves --- supernovae: general }
57: 
58: \section{Introduction}
59: 
60: Four decades of core-collapse simulations have increased our
61: understanding of the core-collapse mechanism, yet a complete
62: theory of the mechanism has not emerged.  Detection of neutrinos during SN 1987A
63: \citep{bionta87,hirata87} confirmed only the fundamentals of
64: core-collapse supernovae.  Theory suggests that the Fe core collapses to form a protoneutron
65: star (PNS), which launches a shock wave.  Before the bounce
66: shock can explode the star, it is
67: sapped of energy by nuclear dissociation and neutrino losses and
68: stalls into an accretion shock \citep{mazurek82,bruenn85,bruenn89}.
69: Understanding the mechanism that revives the stalled shock into
70: explosion has been the goal of the community for many decades.
71: 
72: Since the pioneering work of
73: \citet{wilson85} and \citet{bethe85}, the favored shock revival
74: mechanism has been the delayed-neutrino mechanism (or simply neutrino mechanism), in which neutrinos
75: heat the post-shock region and restart the shock's outward progress
76: after hundreds of milliseconds of delay.  Detailed one-dimensional (1D) simulations using state-of-the-art equations of state
77: (EOSs), neutrino-matter cross sections, and neutrino transport have
78: shown that the neutrino mechanism fails to produce explosions in 1D
79: \citep{liebendorfer01a,liebendorfer01b,rampp02,buras03,thompson03,liebendorfer05b},
80: except the least massive of the massive stars \citep{kitaura06,burrows07d}. 
81: Recent 2-dimensional (2D) simulations, and the accompanying
82: aspherical instabilities, have suggested that the neutrino mechanism may
83: yet succeed, though it fails in 1D
84: \citep{herant94,janka95,burrows95,janka96,burrows07a,kitaura06,buras06a,buras06b,marek07,ott08}.
85: Thus, the fundamental question of core-collapse theory remains; how is
86: accretion reversed into explosion?
87: 
88: Exposing the core-collapse mechanism will
89: require detailed three-dimensional (3D) radiation-hydrodynamic
90: simulations.  However, the core-collapse problem is messy,
91: with many subtle nonlinear couplings and feedbacks 
92: on local and global scales, and extracting the essence of the mechanism from even 1D radiation-hydrodynamic simulations has
93: proven very difficult.  Hence, revealing the core ingredients and conditions of the mechanism will
94: likely require a
95: two-front attack.  On the one hand, multi-dimensional
96: radiation-hydrodynamic simulations will give fully consistent
97: explosions, producing observationally testable energies,
98: neutron star masses, nucleosynthetic yields, and more.
99: On the other hand, simplified models that nevertheless retain the important physics,
100: but allow adjustment of important parameters will help reveal what is important.  In this paper, we pursue the latter
101: philosophy.  We present a systematic parameterization of the conditions
102: and criteria for explosion by the neutrino mechanism, emphasizing the
103: effect that going to 2D from 1D has on the critical neutrino luminosity
104: required for explosions.
105: 
106: \citet{burrows93} suggested a simple framework for determining the
107: conditions for successful explosions by the neutrino mechanism.
108: They approximated the stalled shock and accretion phase as a steady-state problem, transforming the governing partial differential
109: equations into ordinary differential equations.  By
110: parameterizing the electron-neutrino luminosity, $L_{\nu_e}$, and the
111: mass accretion rate, $\dot{M}$, they identified a critical $L_{\nu_e}$-$\dot{M}$ curve that
112: distinguishes steady
113: state accretion solutions (lower luminosities) from explosions (high luminosities).  This implied
114: that global conditions, not local conditions, mediate the transition
115: from accretion to explosion, which in turn suggests that core-collapse
116: explosion is a global instability.
117: More than a decade later, \citet{yamasaki05,yamasaki06}
118: reproduced these results and performed a
119: linear stability analysis of the steady-state solutions that showed
120: unstable solutions near the critical luminosity.
121: In this paper, 
122: we extend this previous work by performing a suite of hydrodynamic
123: simulations in both 1D and 2D.
124: 
125: Assuming that the concept of a critical luminosity applies to 2D simulations, many have noted that critical luminosities for 2D
126: simulations are likely to be lower than in 1D simulations.  In their 1D steady-state
127: solutions, \citet{yamasaki05,yamasaki06} investigated the
128: effects of rotation and convection on the critical luminosity, but
129: these analyses lacked a self-consistent treatment of what is an inherently 2D/3D phenomenon.
130: Early 2D simulations using flux-limited diffusion noted a trend toward
131: explosions aided by neutrino-driven convection
132: \citep{burrows95,janka96}, while 1D simulations failed to
133: explode for the same neutrino luminosity.  More recently,
134: \citet{blondin03} identified a new instability that may prove crucial
135: for the viability of the neutrino mechanism.  It is the standing accretion
136: shock instability (SASI), which may be an advective-acoustic \citep{foglizzo00,foglizzo02,foglizzo07} or purely
137: acoustic cycle \citep{blondin06}.  Whichever explains the SASI, recent simulations have suggested
138: that it may facilitate the neutrino
139: mechanism \citep{marek07,buras06b}.  However, these simulations have yet
140: to demonstrate a reliable explosion mechanism for a wide range of progenitor
141: masses that yields explosion energies consistent with Nature.
142: Furthermore \citet{burrows06}, \citet{burrows07a},
143: \citet{burrows07b}, \citet{burrows07d}, and \citet{burrows07c}, using
144: VULCAN/2D \citep{livne93,livne04} and multi-group
145: flux-limited diffusion (MGFLD), obtain
146: successful explosions only for the 8.8-M$_{\sun}$ model by the
147: neutrino mechanism alone.
148: Recently, \citet{ott08} compared MGFLD and multi-angle transport in VULCAN/2D.  While they note some interesting
149: differences between the transport schemes, explosions remain elusive.
150: Moreover, it
151: has not been demonstrated that the concept of a critical
152: $L_{\nu_e}$-$\dot{M}$ condition pertains to 2D simulations, and if it
153: does, how do the critical luminosities of 1D and 2D simulations compare?
154: 
155: To address these questions, we conduct 1D and 2D simulations for
156: various values of $L_{\nu_e}$.  In the past, there have been a few
157: investigations on the systematic effects of neutrino luminosity on the
158: explosion mechanism \citep{janka96}, neutron star kicks
159: \citep{scheck06}, and the SASI \citep{ohnishi06}, but none has
160: thoroughly investigated both $L_{\nu_e}$ and $\dot{M}$ in 1D and 2D
161: simulations to address the viability of a critical luminosity condition
162: for explosion.
163: In addition to $L_{\nu_e}$ and $\dot{M}$, we compare simulations with
164: different spatial
165: resolutions.  Using the code, BETHE-hydro \citep{murphy08},
166: we simulate the core-collapse, bounce, and post-bounce phases in
167: time-dependent 1D and 2D simulations.  These simulations have no
168: inner boundary and include the PNS core.  A
169: finite-temperature EOS
170: that accounts for nucleons, nuclei,
171: photons, electrons, positrons, and all the appropriate phase
172: transitions is used \citep{shen98}.  Employing 11.2- and 15-M$_{\sun}$ progenitors \citep{woosley02,woosley95} as initial conditions, a wide range of $\dot{M}$ is
173: sampled (from $\sim$0.08 M$_{\sun}$/s to $\sim$0.3 M$_{\sun}$/s).
174: Finally, we use standard
175: approximations for neutrino heating and cooling that enable a
176: straightforward parameterization of $L_{\nu_e}$ \citep{bethe85,janka01}.
177: 
178: The basic equations and the numerical techniques are presented in \S
179: \ref{section:numerics}.
180: In \S \ref{section:progenitor}, we discuss the progenitor models
181: and describe the suite of simulations performed.  In \S
182: \ref{section:shock}, we analyze the evolution of the shock for 1D and 2D
183: simulations and discuss the role shock oscillations play in the explosion.  In \S \ref{section:grid}, we
184: investigate the effect that the grid, specifically spatial resolution and
185: the angular extent of the domain, has on explosions.  In \S
186: \ref{section:timescales}, we revisit the
187: condition for explosion as expressed by the heating and advection
188: timescales, and in \S \ref{section:conditions} describe other
189: conditions at explosion.  In \S \ref{section:critical}, we quantify
190: the differences in the critical luminosity 
191: condition between 1D and 2D simulations.
192: Finally, in \S \ref{section:conclusions}, we summarize our conclusions.
193: 
194: \section{Equations and Numerical Techniques}
195: \label{section:numerics}
196: 
197: The basic equations of hydrodynamics are
198: the conservation of mass,
199: momentum, and energy: 
200: \begin{equation}
201: \label{eq:mass_lag}
202: \frac{d \rho}{d t} = - \rho \vec{\nabla} \cdot \vec{v} \, ,
203: \end{equation}
204: \begin{equation}
205: \label{eq:mom_lag}
206: \rho \frac{d \vec{v}}{d t} = - \rho \vec{\nabla} \Phi - \vec{\nabla} P
207: \, ,
208: \end{equation}
209: and
210: \begin{equation}
211: \label{eq:ene_lag}
212: \rho \frac{d \varepsilon}{d t} = - P \vec{\nabla} \cdot \vec{v} 
213: + \mathcal{H} - \mathcal{C} \, .
214: \end{equation}
215: $\rho$ is the mass density, $\vec{v}$ is the fluid velocity, $\Phi$ is
216: the gravitational potential, $P$ is the isotropic pressure,
217: $\varepsilon$ is the specific internal energy, and $d/dt = \partial
218: / \partial t + \vec{v} \cdot \vec{\nabla}$ is the Lagrangian time
219: derivative.
220: In this work, the neutrino heating, $\mathcal{H}$, and cooling, $\mathcal{C}$, terms
221: in eq. (\ref{eq:ene_lag}) are assumed to be
222: \begin{equation}
223: \label{eq:heating}
224: \mathcal{H} = 1.544 \times 10^{20} L_{\nu_e} \left ( \frac{100 {\rm
225:       km}}{r} \right )^2 \left ( \frac{T_{\nu_e}}{4 {\rm Mev}} \right
226: )^2 \left [ \frac{\rm erg}{{\rm g} \, {\rm s}} \right ] \, ,
227: \end{equation}
228: and
229: \begin{equation}
230: \label{eq:cooling}
231: \mathcal{C} = 1.399 \times 10^{20} \left ( \frac{T}{2 {\rm MeV}}
232: \right )^6
233: \left [ \frac{\rm erg}{{\rm g} \, {\rm s}} \right ] \, .
234: \end{equation}
235: Note that these approximations for heating and cooling by
236: neutrinos \citep{bethe85,janka01} depend upon local quantities and
237: predefined parameters.  They are $\rho$, temperature ($T$),
238: the distance from the center ($r$), the electron-neutrino temperature ($T_{\nu_e}$), and the electron-neutrino luminosity ($L_{\nu_e}$), which
239: is in units of $10^{52}$ erg s$^{-1}$.  By using
240: eqs. (\ref{eq:heating}) and (\ref{eq:cooling}), we gain considerable
241: time savings by approximating the effects of detailed neutrino transport.  For
242: all simulations, we set $T_{\nu_e} = 4$ MeV.  In eq. (\ref{eq:heating}), it has
243: been assumed that $L_{\nu_e} = L_{\bar{\nu}_e}$ and that the mass
244: fractions of protons and neutrons sum to one.  Therefore, the sum of
245: the electron- and anti-electron-neutrino luminosities is $L_{\nu_e
246:   \bar{\nu}_e} = 2 L_{\nu_e}$.
247: Closure for eqs. (\ref{eq:mass_lag}-\ref{eq:ene_lag}) is obtained
248: with an EOS appropriate for matter in or near nuclear statistical
249: equilibrium (NSE) \citep{shen98}, and the effects of photons, electrons,
250: and positrons are included. As such, the
251: EOS has the following dependencies:
252: \begin{equation}
253: \label{eq:eos}
254: P = P(\rho,\varepsilon,Y_e) \, ,
255: \end{equation}
256: where $Y_e$ is the electron fraction.
257: Therefore, we also solve the equation:
258: \begin{equation}
259: \label{eq:advec}
260: \frac{d Y_e}{dt} = \Gamma_e \, ,
261: \end{equation}
262: where $\Gamma_e$ is the net rate of $Y_e$ change.
263: 
264: Using BETHE-hydro
265: \citep{murphy08}, we solve eqs. (\ref{eq:mass_lag}-\ref{eq:ene_lag}) in one and two
266: dimensions by the Arbitrary Lagrangian-Eulerian (ALE) method. To
267: advance the discrete equations of hydrodynamics by one timestep, ALE
268: methods generally use two operations, a Lagrangian hydrodynamic step
269: followed by a remap.  The structure of BETHE-hydro's
270: hydrodynamic solver is designed for arbitrary-unstructured grids, and the
271: remapping component offers control of the time evolution of the grid.
272: Taken together, these features enable the use of
273: time-dependent arbitrary grids to avoid some unwanted features of
274: traditional grids.
275: 
276: For the calculations presented in this paper, we use this flexibility
277: to avoid the singularity of spherical grids in two dimensions.  While spherical
278: grids are generally useful for core-collapse simulations, the
279: convergence of grid lines near the center place extreme constraints on
280: the timestep via the Courant-Friedrichs-Levy condition.  A common remedy is to
281: simulate the inner $\sim$10 km in 1D or to use an inner boundary
282: condition.   Another approach, which has been used in VULCAN/2D simulations \citep{livne93,livne04}, is to avoid the singularity with a grid
283: that is pseudo-Cartesian near the center and smoothly transitions to a
284: spherical grid at larger radii.  We use a similar grid, the butterfly
285: mesh \citep{murphy08}, for the
286: simulations in this paper.
287: 
288: For 1D simulations, 700 radial zones are distributed from the center
289: to 4000 km.  The innermost 100
290: zones have a resolution of 0.34km, and the remaining 600 zones are
291: spaced logarithmically from 34 km to
292: 4000 km.  After each Lagrangian hydrodynamic solve, the flow is
293: remapped back to the original grid, which in effect produces an
294: Eulerian calculation.  For 2D simulations, we use a butterfly mesh interior to 34 km and a
295: spherical grid exterior to this radius.  For all 180$^{\circ}$
296: simulations, the innermost pseudo-rectangular region is 50 by 100
297: zones, and the region that transitions from Cartesian to spherical
298: geometry has 50 radial zones
299: and 200 angular zones.  The outermost spherical region has 200 angular
300: zones and 300 or 150 radial zones which are logarithmically spaced
301: between 34 km and either 4000 km or 1000 km.  The butterfly portion has an
302: effective radial resolution of 0.34 km with the shortest cell edge
303: being 0.28 km.  The grids for the 90$^{\circ}$ simulations are similar,
304: but with half the number of zones.  In 2D simulations, we do not remap every
305: timestep.  Since the timestep is limited by the large sound speeds in
306: the PNS core, we can afford to perform several Lagrangian hydrodynamic
307: solves before remapping back to the original grid.  By remapping
308: seldomly, we gain considerable savings in computational time.
309: 
310: We investigate the systematics of the neutrino mechanism by parameterizing
311: $L_{\nu_e}$, $\dot{M}$, resolution, and the dimensions of the simulation.
312: Since this requires a large set of simulations, we make
313:   several approximations to expedite the calculations. 
314: For one, gravity is calculated via $\vec{g} = -G M_{\rm int}/r^2$, where $M_{\rm int}$ is
315:     the mass interior to the radius $r$. 
316: During collapse, rather than solving the rate equations for $Y_e$ due to
317:     neutrino transport, we use a $Y_e$ vs. $\rho$ parameterization.  \citet{liebendorfer05a}
318: observed that 1D simulations including neutrino transport produce
319: $Y_e$ values during collapse which are essentially a function of
320: density alone.  This allows for a parameterization of
321: $Y_e$ as a function of $\rho$.  To change $Y_e$, we use results of 1D SESAME
322: \citep{burrows00,thompson03} simulations to define the function
323: $Y_e(\rho)$, and we employ the
324: prescription of \citet{liebendorfer05a} to calculate local values of $\Gamma_e$.  As a result, the most
325: important effects of electron capture during collapse
326: are included without the need for expensive, detailed neutrino transport.
327: 
328: \section{Progenitor and Simulation Models}
329: \label{section:progenitor}
330: 
331: To sample a range of $\dot{M}$, we initiate the simulations with
332: two core-collapse progenitor models that have different density profiles: a
333: 15 M$_{\sun}$ progenitor with a shallow density profile exterior to the Fe
334: core \citep{woosley95}, and a 11.2 M$_{\sun}$ progenitor with a steeper
335: exterior density profile \citep{woosley02}.
336: In Fig. \ref{mdotvstime}, the mass accretion rate, $\dot{M}$, vs. time at 250 km (just above the
337:   stalled shock) is shown for 1D non-exploding models. The solid and dashed lines show the
338:   time-dependent accretion rate for the 11.2 and 15.0 M$_{\sun}$ models.
339:   While the outer portions of the Fe core accrete through the shock, $\dot{M}$ is as high as 10
340:   M$_{\sun}$/s and decreases to 2 M$_{\sun}$/s within 50 ms of bounce for both
341:   masses.  After the Fe core is fully
342:   accreted, the accretion rates for the two models
343:   deviate.
344: It takes 50 ms to fully accrete the
345:   Si-burning shell using the 11.2-M$_{\sun}$ model and 100 ms using
346:   the 15-M$_{\sun}$ model.
347: Afterward,
348: $\dot{M}$ declines slowly to 0.2 M$_{\sun}$/s (0.08 M$_{\sun}$/s for the 15
349: M$_{\sun}$ model).
350: Together, these two models slowly sweep through a
351:   range of accretion rates from 0.3 M$_{\sun}$/s to 0.08 M$_{\sun}$/s,
352:   enabling a study of the $\dot{M}$ dependence.
353: 
354: The results presented in the following sections are derived from 95
355: simulations that as a group, represent a
356: parameterization of $L_{\nu_e}$, $\dot{M}$, resolution, and dimensionality.
357: Table \ref{table:sequences} lists these simulations into 14 sequences.  Each sequence of simulations is
358: distinguished by the progenitor model (column 2), dimensionality
359: (column 3), number of radial zones (column 4), radius of the outer
360: boundary (column 5), and a range of electron-neutrino luminosities
361: (column 6).  The sequence names encode the information contained in columns 2
362: through 5.  The first
363: four characters indicate the progenitor model.  Next are the
364: dimensionality labels:  ``1D'' for 1D simulations, ``2D'' for 2D
365: simulations that incorporate the full of $\theta$ (180$^{\circ}$),
366: and ``Q'' for 2D simulations that simulate the quadrant that lies
367: between the pole and equator (90$^{\circ}$).
368: For the 2D simulations, the final character represents the
369: resolution:  ``1'' for the lowest resolution of 250 effective radial
370: zones within 4000 km, ``2'' for a finer resolution of 400 effective
371: radial zones in 4000 km, and ``3'' for the highest resolution of 400
372: effective radial zones within 1000 km.  On the whole, 
373: Table \ref{table:sequences} represents a parameter study for the
374: conditions near explosion in $L_{\nu_e}$, $\dot{M}$, dimensionality, and
375: resolution.
376: 
377: \section{Radial Shock Oscillations in 1D and the SASI in 2D}
378: \label{section:shock}
379: 
380: When \citet{burrows93} reported a critical luminosity vs.
381: accretion rate condition for successful explosions, they did so based
382: upon steady-state solutions.
383: It was left to subsequent calculations to show that
384: the concept of a critical luminosity condition was
385: appropriate for time-dependent simulations.
386: With plots of the evolution of the shock radius, we show that a critical
387: luminosity condition distinguishes non-exploding and exploding models
388: in time-dependent 1D and 2D simulations, and that the
389:   critical luminosity for 2D simulations is lower than
390:   the critical luminosity for 1D simulations.
391: In addition, we report global oscillations near the
392:   critical luminosity in both 1D and 2D simulations, but we show that
393:   the oscillations are of a quite different character.
394: 
395: The time evolution for the shock radius, $R_{\rm shock}$, of 1D
396: simulations is presented in Fig. \ref{rvst_1d}.
397: In the top panel, we display the radii for many models of the 15.0-1D
398: sequence, and each is labeled by $L_{\nu_e}$.  The range of
399: luminosities shown, $L_{\nu_e} = 2.2$-2.9 (in units of $10^{52}$ erg s$^{-1}$), highlight values near the critical luminosity for
400: this model.  The fact that several luminosities lead to explosion at
401: different times is an early indication that the critical luminosity depends upon
402: the mass accretion rate. In \S \ref{section:critical}, we
403: elaborate on the accretion rate dependence.  Similarly, the bottom
404: panel shows $R_{\rm shock}$ for the range of luminosities surrounding
405: the critical luminosity, $L_{\nu_e}
406: = 1.1$-1.7, of the 11.2-1D sequence.
407: 
408: Prior to the accretion of
409: the Si/O interface, the shock stalls for both sequences at radii that are
410: slightly dependent upon $L_{\nu_e}$.
411: The accretion of this interface and the abrupt drop in $\dot{M}$
412: trigger an immediate explosion for the highest luminosity, $L_{\nu_e}
413: = 1.7$ (2.9) for the 11.2-1D (15.0-1D) sequence and
414: oscillations in $R_{\rm shock}$ for the other luminosities considered.
415: These radial oscillations either decay or maintain large amplitudes
416: until the model explodes.  As an example, the oscillation periods of
417: the 15.0-1D sequence range from $\sim$90 ms
418: ($L_{\nu_e}= 2.2$) to $\sim$170 ms ($L_{\nu_e} = 2.8$), and the
419: timescales for decay range from 450 ms ($L_{\nu_e} = 2.2$) to 1.0 s
420: ($L_{\nu_e} = 2.5$).  These radial oscillations and their large
421: amplitudes near explosion support, but do not prove, the hypothesis
422: that a global instability is responsible for the transition between
423: accretion and explosion.
424: 
425: Near the critical luminosity, others have reported large radial oscillations in 1D simulations.
426: \citet{ohnishi06} executed time-dependent 1D simulations of the
427:     steady state solutions in \citet{yamasaki05} and observed similar
428:     pulsations, with amplitudes of order 10\%.
429: For some luminosities, we report similar amplitudes (Fig. \ref{rvst_1d}), but just as
430: many models have large amplitudes of order unity.
431: Furthermore, we find, as do they, that the amplitudes, periods, and $L_{\nu_e}$ are correlated.
432: \citet{buras06a} simulated the collapse, bounce, and
433: postbounce phases for the 15 M$_{\sun}$ model using a Boltzmann
434: transport algorithm in 1D and a ``ray-by-ray'' derivative of the same
435: transport algorithm for 2D simulations.  Generically, they did not
436: obtain explosions.  However, for a few runs, they
437: omitted the velocity-dependent terms in their transport
438: formulation, and this resulted in explosions.  Explosions occurred
439: soon after the shock stalled in their 2D simulations, aborting any
440: obvious oscillations in shock
441: radius.  On the other hand, like our simulations, accretion of the Si/O interface initiates
442: pulsations with amplitudes $\sim$25\% in their 1D simulations
443: oscillations. 
444: .
445: 
446: In reporting the temporal morphology of shocks in 2D simulations, we decompose the shock position,
447: $R_{\rm shock}(\theta,t)$, into spherical harmonics:
448: \begin{equation}
449: R_{\rm shock}(\theta,t) = \sqrt{4 \pi} \sum_{\ell} a_{\ell}(t) Y_{\ell
450:   0} \, ,
451: \end{equation}
452: where
453: \begin{equation}
454: \label{eq:al}
455: a_{\ell}(t) = \frac{1}{\sqrt{4 \pi}}\int R_{\rm shock}(\theta,t) Y_{\ell 0} d \Omega \, ,
456: \end{equation}
457: are the time-dependent coefficients, and the normalization for
458: $a_{\ell}$  in eq. (\ref{eq:al}) ensures that $a_0$ is the average
459: shock radius.
460: As is usual,
461: $Y_{\ell 0}$ are the spherical harmonics, azimuthal symmetry
462: dictates $m = 0$, and the spherical harmonics are normalized to satisfy
463: \begin{equation}
464: \int Y^*_{\ell 0} Y_{\ell 0} d \Omega = 1 \, .
465: \end{equation}
466: 
467: These temporal coefficients for the
468: 15.0-2D3 sequence are plotted in 
469: Fig. \ref{sphharm}. The average shock
470: radius, $a_0$ or $\langle R_{\rm shock} \rangle$, is shown in the top
471: panel, and a dot indicates the time of explosion, $t_{\rm exp}$ for
472: the exploding models.  Accompanying are error bars
473: indicating an estimated uncertainty of $t_{\rm exp}$ (see \S
474: \ref{section:conditions} for the definition of $t_{\rm exp}$).  The $\ell = 1$ and
475: $\ell = 2$ components normalized by $a_0$ are shown in the middle and
476: bottom panels.  First of all, note that the range of luminosities bracketing the critical
477: luminosity for the 2D
478: simulations in Fig. \ref{sphharm} are lower than the range for 1D
479: simulations in Fig. \ref{rvst_1d}.  Further comparison indicates that, unlike 1D simulations, 2D simulations show very little oscillation
480: in $\langle R_{\rm shock} \rangle$ prior to explosion. In fact, oscillations
481: in $\langle R_{\rm shock} \rangle$ are miniscule for non-exploding
482: models and only reach $\sim$10\% around
483: $t_{\rm exp}$ for exploding simulations.
484: 
485: Instead,
486: SASI originated oscillations in $\ell=1$ and $\ell=2$ components dominate the non-exploding and exploding models.
487: The amplitudes of the dipole component, $a_1/a_0$, for $L_{\nu_e}=1.5$
488:   remain low at $\sim$2.5\%, from bounce until 500 ms.  Around 500 ms, the amplitude 
489:   grows to $\sim$10\%.
490: On the other hand, models with $L_{\nu_e} = 1.6$-1.9 show a steady rise to
491:   $\sim$10\% from bounce.
492: Exploding models, for which $L_{\nu_e} =1.8$, 1.9, and 2.0, show an abrupt
493: increase in amplitude to $\sim$20\%.
494: For non-exploding models, $a_1/a_0$ saturates at $\sim$10\%, and
495:   for $L_{\nu_e} = 1.5$ and 1.6 it decreases to $\sim$5\%.
496: At early times, simulations with $L_{\nu_e} = 1.5$ and $L_{\nu_e}
497: = 2.0$ show the lowest and highest SASI amplitudes, respectively, and are
498: easily distinguished from the other simulations.  While this would suggest a
499: correlation in SASI amplitude with $L_{\nu_e}$, the other models do
500: not show a similar monotonic trend.  Rather their amplitudes occupy
501: the region between $L_{\nu_e} = 1.5$ and $L_{\nu_e} = 2.0$ but show no
502: discernible trend with luminosity.
503: The quadrupole term, $a_2/a_0$, is generally below $\sim$5\% during non-exploding
504: phases.  For times just before explosion and afterward, $a_2/a_0$
505: can reach $\sim$10\%.
506: 
507: Two important results are apparent in Figs. \ref{rvst_1d} and
508: \ref{sphharm}.  A critical luminosity separates exploding and
509: non-exploding models in time-dependent 1D and 2D simulations, and
510: the critical luminosity for 2D simulations is $\sim$70\% of the critical luminosity
511: for 1D models. More evidence will be presented in \S
512: \ref{section:critical} that the critical luminosity depends upon $\dot{M}$.
513: For 1D simulations, we observe pulsations near the critical
514: luminosity, but in 2D simulations, radial oscillations are all but
515: absent.  Instead, non-radial SASI oscillations dominate.
516: 
517: \section{Effects of the Grid}
518: \label{section:grid}
519: 
520: The conclusions of the previous section might depend upon the
521: grid structure used in 2D simulations.
522: To address this concern, we investigate the dependence of these
523: results on resolution and the range of polar angle, $\theta$, included
524: in the simulations.
525: Encoded in the sequence names (Table \ref{table:sequences}) are the types of grids used for
526:   2D simulations.  For calculations that encompass the full
527:   range of $\theta$ from pole to pole (180$^{\circ}$) $\theta$ is divided into 200 zones,
528:   and for simulations that include only the region between a pole and
529:   the equator (90$^{\circ}$) $\theta$ is divided into 100 zones.
530: Sequences that use the 180$^{\circ}$ grid have `2D' in their name,
531:   and for sequences that use the 90$^{\circ}$ grid, `Q' is in its stead.
532: The numbers after `2D' or `Q' indicate the resolution.  The
533:   lowest resolution, denoted by `1', extends the spherical grid from
534:   the outer edge of the butterfly mesh, 34 km, to 4000 km with 150 radial
535:   zones.  The middle resolution, `2', divides
536:   this same space into 300 radial zones, and the highest resolution,
537:   `3', divides the radii between 34 km and 1000 km into 300 zones.  For all resolutions, the space between the butterfly mesh and
538:   the outer boundaries are positioned logarithmically.
539: These simulations do show nontrivial differences in the SASI and
540:   post-shock flow.
541: However, the conclusions of the previous section are insensitive to
542:   changes in resolution and the range of $\theta$ simulated.
543: 
544: For these three resolutions, Fig. \ref{morphres} shows $\langle R_{\rm shock} \rangle$,
545: $a_1/a_0$, and $a_2/a_0$ vs. time
546: for a single luminosity, $L_{\nu_e} = 1.9$, and the 15-M$_{\sun}$ progenitor
547: model.
548: Their general evolution is similar to what was discussed in \S
549: \ref{section:shock} and shown in Fig. \ref{sphharm}.
550: However, there are some nontrivial differences.
551: 
552: After the Si/O interface is accreted, $\langle R_{\rm
553:   shock} \rangle$ of the lowest resolution model (15.0-2D1) secularly
554: creeps out to $\sim$220 km over 100 ms.  Then, outward
555: progression stalls again, and the model doesn't explode until 581 ms
556: after bounce.  Both the 15.0-2D2 and 15.0-2D3 models momentarily stall at
557: $\sim$200 km, a smaller radius. 150 ms later, the average shock radius
558: of the 15.0-2D2 model reaches
559: $\sim$220 km on its way to explosion.  At an even later time, 550 ms,
560: the shock radius of the 15.0-2D3 model reaches $\sim$220 km.  By this
561: measure, the trend with resolution is monotonic.  However, this does
562: not translate to monotonicity in explosion time with resolution.
563: From the lowest resolution (15.0-2D1) to the highest resolution
564: (15.0-2D3), the explosion times are 581, 359, and 649 ms.
565: Yet, considering another luminosity, $L_{\nu_e} = 2.0$ (see Table \ref{table:explosions}) does present a monotonic trend toward later
566: explosion times with higher resolution: from the lowest to the
567: highest resolution, $t_{\rm exp} = 325$, 364, and
568: 395 ms.  From this, one might conclude that
569: higher-resolution grids forestall explosion, yet only the
570: highest resolution shows an explosion for $L_{\nu_e} = 1.8$.
571: 
572: The $\ell = 1$ and $\ell = 2$ (the middle and bottom panels of Fig. \ref{morphres}) spherical
573: harmonic coefficients, like the monopole term, show some differences
574: with resolution, but present very few clear trends.  Any trends that do
575: appear are subtle.  
576: In the earliest phase, before accretion of the
577: Si/O interface, the highest resolution
578: simulations exhibit the greatest amplitude, though
579: all three resolutions have amplitudes in the $\ell =
580: 1$ coefficient that are less than 5\%.  
581: At the same time, the $\ell = 2$ component is comparable
582: for all three resolutions.  After accretion of the Si/O interface,
583: the nonradial components begin to rise for all three resolutions.
584: From $\sim$200 to $\sim$325 ms, the 15.0-2D1 model shows the
585: greatest increase in both nonradial components, while for 15.0-2D2 and
586: 15.0-2D3 they are comparable in growth.  At $\sim$325 ms, the 15.0-2D2
587: model shows a significant increase in the nonradial components that
588: coincides with explosion.  The nonradial components for the other two
589: simulations continue a secular increase in amplitude until their
590: explosions at later times.  All the while, the lowest resolution run
591: consistently maintains a slightly larger amplitude in the $\ell = 1$
592: component, but the $\ell = 2$ components are quite similar.
593: Generally, before accretion of the Si/O interface, oscillation
594: amplitudes for $a_1$ are mildly correlated with resolution, and
595: afterward it is mildly anti-correlated with resolution. 
596: 
597: Figure \ref{fullvs90} compares 180$^{\circ}$ (15.0-2D3, solid) and 90$^{\circ}$
598: (15.0-Q3, dashed) simulations.  $\langle R_{\rm shock} \rangle$ and
599: $a_2/a_0$ vs. time are plotted for two electron-neutrino luminosities $L_{\nu_e} = 1.8$ 
600: (purple curves) and $L_{\nu_e} = 1.9$ (green curves).  Once again dots
601: in the top panel mark $t_{\rm exp}$; interestingly, the 90$^{\circ}$
602: simulations consistently explode before the 180$^{\circ}$ simulations.
603: Other than for times near $\sim$300 to $\sim$400 ms, all models show similar evolution in the
604: amplitude of $a_2/a_0$.  The one exception appears around $\sim$300
605: to $\sim$400 ms, when the 90$^{\circ}$, $L_{\nu_e} = 1.9$ model starts
606: exploding and shows the largest amplitude.  Even though the 90$^{\circ}$
607: simulations explode earlier than the 180$^{\circ}$ simulations, the
608: critical luminosities for the 180$^{\circ}$ and 90$^{\circ}$ simulations
609: are very similar (see Table \ref{table:explosions} and \S \ref{section:critical}).
610: 
611: Using a ``ray-by-ray'' neutrino transport
612:     algorithm, \citet{buras06b} performed simulations of the 11.2-M$_{\sun}$ progenitor
613:     that teetered on the verge of explosion for a 90$^{\circ}$
614:     calculation, but exploded using 180$^{\circ}$.
615: They reported no
616: explosion with a 90$^{\circ}$ simulation but, with all else being equal,
617: obtained an explosion with a 180$^{\circ}$ simulation.  Differences between our simulations and theirs
618: include neutrino transport, EOS, and hydrodynamic solver.  While any
619: of these could account for the apparent discrepancy, we emphasize the
620: difference in the 90$^{\circ}$ grids.  The 90$^{\circ}$ simulations of
621: \citet{buras06b} included a wedge that extends plus and minus 45$^{\circ}$ of
622: the equator and suppressed $\ell = 1$ and $\ell = 2$ modes, the two
623: prominent modes of the SASI.  Our 90$^{\circ}$ simulations range from the pole to the equator
624: and inhibit only the $\ell = 1$ mode.  Regardless, \citet{buras06b} noted that
625: their 90$^{\circ}$ simulation was close to explosion and that extending
626: the grid to 180$^{\circ}$ simply tipped the scales toward explosion.
627: Using the same
628: hydrodynamics and neutrino transport as \citet{buras06b}, but a
629: 90$^{\circ}$ grid that goes from the pole to the equator,
630: \citet{marek07b} found that 90$^{\circ}$ simulations explode
631: insignificantly earlier than 180$^{\circ}$ simulations.
632: In either case, these results are consistent with our finding that the
633: critical luminosity for simulations using 90$^{\circ}$ and
634: 180$^{\circ}$ are very similar.
635: 
636: Resolution and the angular size of the grid non-trivially affect the shock
637: position and morphology in 2D simulations.
638: For one, the lowest resolution simulations
639: consistently maintain a slightly larger amplitude in the $\ell = 1$
640: component.  At early times, the lowest resolution simulations reach
641: $\sim$220 km before higher resolution simulations.  This does not lead to a
642: monotonic correlation in explosion times, though.
643: In changing the angular size of the grid, the 90$^{\circ}$ simulations
644: consistently explode at earlier times compared to the 180$^{\circ}$
645: simulations.
646: Despite differences that result from grid changes, the
647: conclusions of the previous section do not change.  Specifically, the
648: critical luminosity for the 90$^{\circ}$ and 180$^{\circ}$ simulations are
649: very similar.
650: 
651: \section{Heating and Advection Timescales}
652: \label{section:timescales}
653: 
654: Measures that are related to the critical $L_{\nu_e}$ and $\dot{M}$
655: condition \citep{burrows93} are the
656:   timescale for matter to traverse the gain region, the advection
657:   timescale, $\tau_{\rm adv}$, and the heating timescale of the matter
658:   in the gain region, $\tau_q$.
659: \citet{thompson00} argued that 1D core-collapse simulations do not
660: explode because $\tau_q$ is longer than $\tau_{\rm adv}$.  In other
661: words, core-collapse simulations fail because $\tau_{\rm adv} / \tau_{q}
662: < 1$.  He defined, $\tau_q \sim aT^4/\dot{Q}$, and
663:     the advection time as the flow time across a scale height.
664: \citet{janka01} explored similar integral concepts, and proposed
665: conditions for explosion that are akin to
666: \begin{equation}
667: \label{eq:taucondition}
668: \tau_{\rm adv} / \tau_q > 1 \, ,
669: \end{equation}
670: the opposite of the \citet{thompson00} failure condition.
671: \citet{thompson05} applied this
672:     condition to 1D core-collapse
673:     simulations and demonstrated that this ratio does indeed help to
674:     distinguish non-exploding models from exploding models.  They
675:     defined the advection timescale as $H/v_r$, where H is the
676:     pressure scale height, and the heating timescale as
677:     $(P/\rho)/\dot{q}$, where $\dot{q}$ is the local net
678:     heating rate.
679: Since, others have explored $\tau_{\rm adv}$ and $\tau_q$ as a condition
680: for explosion in 2D simulations \citep{buras06a,scheck08}.
681: While $\tau_{\rm adv}/\tau_{q} >
682: 1$ is to some extent useful in identifying explosions, our results suggest
683: that it is only useful as a rough diagnostic.
684: 
685: We define the heating time as a characteristic time for
686:   neutrinos to change the thermal energy of the matter in the gain
687:   region, where heating dominates cooling,
688:   \begin{equation}
689:     \label{eq:tauq}
690:     \tau_q = \frac{\int_{\rm gain} (\varepsilon - \varepsilon_0) dm }
691:     {\dot{Q}} \, ,
692:   \end{equation}
693:   where $\varepsilon_0$ is the zero point energy for
694:   the EOS and 
695:   \begin{equation}
696:     \label{eq:qdot}
697:     \dot{Q} = \int_{\rm gain} (\mathcal{H} - \mathcal{C}) dm \,
698:   \end{equation}
699:   is the net heating rate and is integrated over the gain region.
700: While \citet{scheck08} and \citet{buras06b} include $1/2v^2 + \Phi$ in the
701:     integrand of eq. (\ref{eq:tauq}), we suggest,
702:     like \citet{thompson00} and \citet{thompson05}, that comparing the
703:     heating rate to the thermal energy eq. (\ref{eq:tauq}) is a more natural definition of a heating
704:     timescale. To be clear though, these differences in definition
705:     alter the absolute scale, but not the general trends.
706: 
707: For spherically symmetric flow, the advection timescale is 
708:   \begin{equation}
709:     \label{eq:tauadv}
710:     \tau_{\rm adv} = \int_{\rm gain} \frac{dr}{v_r} \,
711:   \end{equation}
712:   where the the integral is integrated from the shock position to the
713: gain radius.
714: For 2D simulations, many incarnations of eq. (\ref{eq:tauadv}) have
715: appeared in the literature.  \citet{scheck08} use a solid-angle
716: averaged version of eq. (\ref{eq:tauadv}) for $\tau_{\rm adv}$.
717: \citet{buras06b} use eq. (\ref{eq:tauadv}) for the advection
718:     timescale in 1D, but for 2D simulations they redefine this
719:     timescale as the time for a mass shell to traverse the gain region.
720: We've explored three measures of the advection timescale.
721: The first replaces the denominator in eq. (\ref{eq:tauadv})
722:     with the solid-angle averaged velocity, $\langle v_r \rangle =
723:     (\int{v_r d \Omega})/4 \pi$.
724: For the second, if one assumes steady-state flow, then
725:     $\tau_m = M_{\rm gain}/\dot{M}$ is a useful measure of the
726:     accretion timescale, where $M_{\rm gain}$ is the mass in the gain
727:     region and $\dot{M}$ is the mass flux in or out of the gain
728:     region.  For simplicity, we use $\dot{M}$ just exterior to the
729:     shock.  The advantage of this definition is that it is relatively
730:     straightforward to measure $M_{\rm gain}$ and $\dot{M}$ for 2D simulations.
731: Prior to explosion, these two definitions give similar results, but
732: as the model explodes, the assumption of steady-state accretion is violated
733: and $\tau_m$ begins to underestimate $\tau_{\rm adv}$.
734: Neither of these definitions, however, fully capture the complex
735:     multi-dimensional nature of the post-shock flow.
736: 
737: We define a new timescale that better captures the multi-dimensional
738:     nature of the post-shock flow.
739: Post-processing the simulation data, we integrate the paths of tracer
740: particles for 150 ms, record their
741:     trajectories, and define a residence time, $\tau_r$, which is the
742:     duration of each particle's time spent in the gain region.
743: 50,000 particles are initiated at either 400 km, a
744:     radius well outside the shock, or 150 km, the middle of the gain
745:     region.
746: At either radius, the particles are randomly distributed in $\mu = \cos \theta$.
747: For most situations,
748: the particles traverse the gain region within the integration time,
749: 150 ms, and have
750:     accreted onto the PNS.
751: Every 10 ms, a new generation of 50,000 particles is generated,
752: providing a time-dependent distribution of $\tau_r$.
753: 
754: The ratio $\tau_{\rm adv}/\tau_q$ vs. time is shown in Fig. \ref{tscalesvst_1d} for the 15.0-1D sequence.  The top
755: panel shows the models that do not explode by 1.3 s after bounce, and
756: the bottom panel shows those that do explode.  Despite dramatic
757: oscillations, $\tau_{\rm adv}/\tau_{\rm q}$ rarely reaches a value of
758: one for the non-exploding models (top panel).  In all non-exploding
759: cases, except $L_{\nu_e} = 2.5$, the
760: oscillations in this ratio decay with a timescale commensurate with the decay times
761: of the shock radius oscillations (Fig. \ref{rvst_1d}).  For exploding
762: models (bottom panel), $\tau_{\rm adv}/\tau_{q}$ executes large
763: excursions from $\sim$0.1 to $\sim$100.  Note that $\tau_{\rm
764:   adv}/\tau_{\rm q}$ makes several excursions to values above order
765: unity and back again.  Smaller luminosities produce more excursions,
766: with one at $L_{\nu_e} = 2.9$ and four for
767: $L_{\nu_e} = 2.6$.
768: 
769: The ratio of these timescales for 2D simulations (15.0-2D3 sequence) are shown
770: in Fig. \ref{tscalesvst_2d}.  The models that don't explode, $L_{\nu_e} =
771: 1.5$, 1.6, and 1.7, show relatively constant
772: ratios for all time.  On the other hand, the exploding models,
773: $L_{\nu_e} = 1.8$, 1.9, and 2.0, start at a low ratio ($\sim$0.1)
774: and secularly increase to ratios of order $\sim$10.  Unlike the 1D
775: models, once this ratio reaches large values (in an average sense), it
776: remains there for the rest of the simulation.
777: 
778: Generally, the condition $\tau_{\rm
779:   adv}/\tau_{q} > 1$ is a useful diagnostic of explosion, but
780: given the ambiguities in defining
781:     $\tau_{\rm adv}$ and $\tau_q$, its accuracy is limited. 
782: For example, 1D simulations that explode make dramatic excursions above and
783:     below this line before explosion.
784: It
785: might be more sensible to state that this condition should be
786: satisfied for a finite time before explosion occurs.  
787: However, one shouldn't expect rigor when defining a heating timescale.
788: Note that this condition assumes a quasi-steady state, and the
789: large amplitude oscillations in 1D complicate this assumption.
790: Finally, $\tau_{\rm adv}/\tau_q$, in 2D exploding simulations, increases
791: monotonically for $\sim$100 ms, while this ratio remains low for 2D
792: non-exploding simulations.  This distinction implies that
793: 2D exploding models are well on their way to explosion before
794: $\tau_{\rm adv}/\tau_q > 1$ is satisfied.  Hence, $\tau_{\rm adv}/\tau_q$
795: is a useful diagnostic for explosion, but may not be a rigorous
796: condition for explosion.
797: 
798: In Fig. \ref{taures_mean_frac}, we plot for all generations of tracer
799: particles the mean residence times, $\langle
800: \tau_r \rangle$ (top panel), and the fraction ($f_r$) of particles with
801: $\tau_r > 40$ ms (bottom panel) as a function of
802: time for the $L_{\nu_e} = 1.9$ model of the 15.0-2D3 sequence.  The
803: particles with $R_{\rm start} = 400$ km (150 km) are shown with solid
804: (dashed) lines.  Both $\langle \tau_r \rangle$ and $f_r$ show that the particles with
805: $R_{\rm start} = 400$ km and 150 km track two fundamentally
806: different flows.  Generally, the particles with
807: $R_{\rm start} = 150$ km have longer average residence times and more
808: particles with $\tau_r > 40$ ms than the generation of particles with
809: $R_{\rm start} = 400$ km, which suggests that most of the mass
810: that accretes through the shock quickly finds its way to the PNS through narrow, low entropy plumes.  Despite the quick
811: accretion of new matter, the long $\tau_r$ of particles with $R_{\rm
812:   start} = 150$ km indicate that a significant amount of mass in the
813: gain region does linger, resulting in more heating and higher
814: entropies.
815: 
816: To verify this, we compute the total
817: amount of mass with high entropies and compare
818: this to the total mass in the gain region.  To define high entropies,
819: we choose a minimum entropy of 14 k$_{\rm B}$/baryon, which corresponds
820: to the maximum entropy for the 1D simulation of the same neutrino
821: luminosity (see Fig. \ref{entropy}) and gives a rough boundary between
822: the regions with short (low entropy) and long (high entropy) residence
823: times.  The ratio of mass with high entropy to the total mass in the gain region
824: grows from 0\% to 70\% between 0.2 s and 0.5 s after bounce.  For the
825: rest of the simulation the ratio remains at this level.  Therefore, even though
826: most of the accreted mass advects quickly through the gain region via the
827: down-flowing plumes, a significant fraction of the mass in the gain
828: region is better characterized with long residence times and high entropies.
829: 
830: In addition, $\langle \tau_r \rangle$ and
831: $f_r$ show four distinct phases, with each successive phase
832: having longer $\langle \tau_r \rangle$ and higher $f_r$.
833: Distributions of $\tau_r$ that represent these four phases are shown in
834: Fig. \ref{taures}, and spatial colormaps of entropy are shown in
835: Figs. \ref{stills1}, \ref{stills2}, and \ref{stills3} that
836: correspond to the beginning (Fig. \ref{stills1}), middle
837: (Fig. \ref{stills2}), and end (Fig. \ref{stills3}) of each generation's
838: integration time of 150 ms.    The top panel of
839: Fig. \ref{taures} shows the
840: distribution for four generations with $R_{\rm start} = 400$ km and at $t = 0.130$, 0.370, 0.530, and 0.6 s after
841: bounce.  The bottom panel shows similar residence-time distributions
842: for generations with $R_{\rm start} = 150$ km.
843: 
844: The entropy maps in the top-left panels of Figs. \ref{stills1}, \ref{stills2}, and \ref{stills3} show that there
845: is very little convection or SASI for the first phase ($t = 0.130$ s),
846: and the distributions of $\tau_r$ (Fig. \ref{taures}) show that all particles traverse
847: the gain region quickly, within $\sim$9 ms.  At $\sim$150 ms, when the Si/O interface
848: is accreted, both $\langle \tau_r \rangle$
849: and $f_r$ show a dramatic
850: increase.  The mean residence time increases from $\sim$10 ms to
851: $\sim$40 ms for $R_{\rm start}=400$ and 150 km, and $f_r$ increases
852: from 0\% to 35\% for $R_{\rm start} = 400$ km and $\sim$50\% for $R_{\rm
853:   start} = 150$ km.  The second phase, from
854: $\sim$200 ms to $\sim$450 ms, corresponds
855: to the growth and nonlinear saturation of the SASI (top-right panels of Fig. \ref{stills1}-\ref{stills3}).  During this phase ($t = 0.370$ s in Fig. \ref{taures}), $\langle \tau_r
856: \rangle$ for $R_{\rm start} = 400$ km (150 km) is $\sim$22 ($\sim$32), and $f_r$ is $\sim$10\% ($\sim$30\%).  The third phase, from $\sim$0.45 to $\sim$0.6 s,
857: shows a rise in both $\langle \tau_r \rangle$ and
858: $f_r$.  The distributions in Fig. \ref{taures} that represent this phase, $t= 0.530$ s, have mean residence times of $\sim$50 ms ($\sim$30 ms) for $R_{\rm start} = 150$ km (400 km) and $f_r$ of 50\% (20\%).  The fourth and final phase is marked by a significant increase of
859: $\langle \tau_{\rm res} \rangle$ and $f_{\rm res}$ for $R_{\rm start}=
860: 150$ km particles and continued rise for $R_{\rm start}= 400$ km.  In
861: Fig. \ref{taures}, the example distribution is shown at $t =
862: 0.6$ s.  For $R_{\rm start}= 400$ km, $\langle \tau_{\rm res} \rangle \sim 40$ ms
863: and $f_{\rm res} \sim 25$\%, while $\langle \tau_{\rm res} \rangle \sim 90$ ms
864: and $f_{\rm res} \sim 70$\%.  This phase is also characterized by a
865: large bump of particles with very long residence times in
866: Fig. \ref{taures}, which is much larger for $R_{\rm start}=150$ km than
867: for $R_{\rm start} = 400$ km.  The appearance of the bump at large $\tau_r$ is one of the most pronounced indicators that the star has reached an explosive situation.
868: 
869: In Fig. \ref{entropy}, we compare 1D entropy profiles of the 15.0-1D,
870: $L_{\nu_e} = 1.9$ simulation (solid red lines) with the entropy
871: distributions shown in Fig. \ref{stills1}.  For each zone of the 2D
872: simulation, points mark the entropy and radius from the center.
873: Generally, both 1D and 2D profiles show the shock as an abrupt jump in
874: entropy of the accreting matter, the gain region as
875: indicated by the highest entropies, and the
876: cooling region where entropy is significantly reduced as matter continues to accrete onto the PNS.  Where 1D
877: profiles show a negative entropy
878: gradient, 2D simulations are convective.  The most striking
879: differences, though, are that when the SASI dominates the post-shock flow, $t=
880: 0.370$, 0.530, and 0.600 s, the shock is asymmetric, and the gain region has a large range of
881: entropies, in which high entropies correspond
882: to regions of long residence times, and low entropies
883: correspond to plumes that funnel matter toward the PNS.
884: 
885: Other than passage through shocks, the entropy is determined by
886: neutrino heating and cooling.  Therefore, entropy is a good
887: measure of the integrated history of net heating.  Considering mass- and volume-weighted averages, the entropy
888: for 2D simulations is a couple $k_B$/baryon (less than 10\%) higher than 1D
889: simulations.  However, Fig. \ref{entropy} shows that 2D simulations have
890: regions with $\sim$30\% higher entropies than 1D simulations for the
891: same $L_{\nu_e}$ and $\dot{M}$.  In other words, the long residence
892: times that accompany the SASI in 2D simulations result in regions
893: with higher integrated net heating and offer an explanation for the
894: lower critical luminosity of 2D simulations.
895: 
896: \section{Conditions at Explosion}
897: \label{section:conditions}
898: 
899: As relevant indicators of the core-collapse mechanism, we
900: investigate the net heating rate ($\dot{Q}$), heating efficiency
901: ($\dot{Q}/L_{\nu_e \bar{\nu}_e}$), and the mass in the gain region ($M_{\rm gain}$).  In general, these quantities show distinctly
902: different evolutions for 1D and 2D simulations near their respective
903: critical luminosities.  Plots of the 1D simulations reflect the very
904:   large radial shock oscillations, while plots of the 2D simulations
905:   show relatively smooth evolution and bifurcations distinguishing
906:   explosive models from non-explosive models.  
907: \citet{scheck08} compared these quantities for 1D and 2D simulations
908:   with the same neutrino luminosity and found similar bifurcations in
909:   these integral measures.  We show that this bifurcation is generic
910:   when we include 1D and 2D non-exploding and 2D exploding models. 
911: In Table \ref{table:explosions}, we summarize the conditions at
912: explosion, but first we take care to define $t_{\rm exp}$, the time of
913: explosion.
914: 
915: The time of explosion for 1D
916: simulations is quite pronounced and occurs near the time when
917: the ratio $\tau_{\rm adv}/\tau_q$ becomes of order unity
918: or larger for the last time.  However, for 2D
919: simulations, all measures show a gradual trend toward explosion
920: over 50 to 100 ms timescales.  Therefore, to define $t_{\rm exp}$
921: we identify the last time that $\tau_{\rm adv}$ is less than 50 ms (65
922: ms) for 15-M$_{\sun}$ (11.2-M$_{\sun}$) models.
923: This definition establishes the time of explosion after there is a
924: clear trend toward explosion, but before post-shock material is
925: launched outward.  Acknowledging the ambiguity of such a definition, we assign an uncertainty to $t_{\rm exp}$ of $\pm 50$
926: ms.
927: 
928: Having defined $t_{\rm exp}$, we now define the conditions at explosion.
929: For 1D simulations, $\dot{M}_{\rm exp}$ is defined at 250 km
930: and at $t_{\rm exp}$.  For 2D simulations, $\dot{M}_{\rm exp}$ is
931: defined just exterior to the shock (at roughly 300 to 250 km) at the
932: time of explosion.  Corresponding to the uncertainty in $t_{\rm
933:   exp}$, we determine the uncertainty in $\dot{M}_{\rm exp}$.
934: Figs. \ref{qdotvstime}, \ref{effvstime}, and \ref{mass_gain}, show a
935: general upward trend of $\dot{Q}$, $\dot{Q}/L_{\nu \bar{\nu}_e}$, and
936: $M_{\rm gain}$ around
937: the time of explosion.  The values in Table \ref{table:explosions}
938: are obtained by averaging $\dot{Q}$, $\dot{Q}/L_{\nu \bar{\nu}_e}$, and
939: $M_{\rm gain}$ for a 50 ms range centered on $t_{\rm exp}$.  We
940: also experimented with a range of 10 and 100 ms and obtained similar
941: results.  For 1D simulations, we note the conditions at the time of
942: explosion, but due to large nonlinear oscillations the specific values
943: are sensitive to the definition of $t_{\rm exp}$.
944: 
945: Figure \ref{qdotvstime}
946: compares the time evolutions of the net heating rates, $\dot{Q}$, for
947: the 15.0-1D and 15.0-2D3 sequences.  
948: While the top panel shows $\dot{Q}$ for the
949: sequence of 1D simulations near explosion, the bottom panel shows $\dot{Q}$ for the sequence of 2D simulations (solid) near
950: explosion.  For comparison, 1D models (dotted) for the
951: same range of $L_{\nu_e}$ as the 2D sequence are included.  The dots show $\dot{Q}$ at the time of
952: explosion for the 2D simulations.
953: As expected, $\dot{Q}$ correlates with $L_{\nu_e}$
954: (Fig. \ref{qdotvstime}).  Consequently, the range of $\dot{Q}$ for the
955: 1D exploding models ($\sim$1 to $\sim$11 B/s, where 1 Bethe (B) $=
956: 10^{51}$ erg) is higher than the
957: range of $\dot{Q}$ for the 2D exploding models ($\sim$0.5 to $\sim$6
958: B/s).  The 1D sequence in the top panel shows pronounced oscillations
959: in $\dot{Q}$ that correspond to the oscillations in shock radius in
960: Fig. \ref{rvst_1d}.  The maxima in $\dot{Q}$ coincide with the minima
961: in the shock radius, which has been observed before for 1D simulations
962: \citep{buras06a}.
963: 
964: The 2D sequence, on the other hand, shows no
965: such oscillations in $\dot{Q}$.  Rather, the bottom panel of Fig. \ref{qdotvstime}
966: shows a relatively smooth evolution with time.  After bounce,
967: $\dot{Q}$ evolves secularly downward for non-exploding 1D and
968: 2D simulations.  In contrast, 50 ms to 100 ms before $t_{\rm exp}$,
969: $\dot{Q}$ for exploding 2D simulations remains flat with time or even
970: begins to rise slightly.  In either case, there is a clear bifurcation
971: in the evolution for exploding 2D simulations and non-exploding 1D and
972: 2D simulations.
973: 
974: Similar to Fig. \ref{qdotvstime}, in Fig. \ref{effvstime} we plot the heating efficiency,
975: $\dot{Q}/L_{\nu_e \bar{\nu}_e}$.  Even after dividing by $L_{\nu_e \bar{\nu}_e}$
976: the efficiency has a positive correlation with $L_{\nu_e}$.  This
977: suggests that the correlation between $\dot{Q}$ and $L_{\nu_e}$ is
978: steeper than linear.  The
979: efficiencies at $t_{\rm exp}$ for the 1D sequence range from 3\% to
980: 10\%, while the efficiencies of the 2D simulations at $t_{\rm exp}$ are
981: 3.4\%, 3.5\%, and 4.4\% for $L_{\nu_e} = 1.8$, 1.9, and 2.0, respectively. 
982: Naturally, the trends in $\dot{Q}/L_{\nu_e
983:   \bar{\nu}_e}$ are similar to the those of $\dot{Q}$, including the
984: bifurcation of $\dot{Q}/L_{\nu_e \bar{\nu}_e}$ for exploding and
985: non-exploding models.
986: 
987: Fig. \ref{mass_gain} shows the time evolution of $M_{\rm gain}$ for
988: the 15.0-1D and 15.0-2D3 sequences.  Once again, the oscillations in
989: $M_{\rm gain}$ correspond to the oscillations in shock radius, but
990: this time the minima in $M_{\rm gain }$ correspond to the minima
991: in shock radius.  In the exploding models of 15.0-2D3, there is a clear
992: secular increase in the amount of mass in the gain region leading up
993: to and past explosion.  This is consistent with the idea that the SASI
994: helps to increase the amount of mass in the gain region, tipping the
995: scales toward explosive solutions.  However, we note in
996: Fig. \ref{sphharm} that even the non-exploding models show some $\ell =
997: 1$ oscillations, and hence, the SASI is necessary but not
998: sufficient in increasing the mass in the gain region.  At
999: $t_{\rm exp}$, $M_{\rm gain} =$ 0.099, 0.0113, and 0.0152 M$_{\sun}$ for
1000: $L_{\nu_e} = 1.8$, 1.9, and 2.0, respectively.
1001: 
1002: We summarize the conditions at explosion in Table
1003: \ref{table:explosions}.  We list for each sequence in Table
1004: \ref{table:sequences}, the electron-neutrino
1005: luminosity, $L_{\nu_e}$ (column 1), time of explosion, $t_{\rm exp}$
1006: (column 2), accretion rate at the time of explosion near the
1007: shock, $\dot{M}_{\rm exp}$ (column 3), the net heating rate, $\dot{Q}$
1008: (column 4), the heating efficiency, $\dot{Q}/L_{\nu_e
1009:   \bar{\nu}_e}$ (column 5), and the mass in the gain region at the
1010: time of explosion, $M_{\rm gain}$ (column 6).  The mass accretion rate
1011: at $t_{\rm exp}$ is addressed in the next section.  The 11.2-1D sequence
1012: manifests a wide range of $\dot{Q}$, from 0.63 to 4.46 B/s, and the range
1013: of the corresponding heating efficiencies is similarly wide, from 2.4\% to 14\%.  In contrast,
1014: the 15.0-1D models have much larger $\dot{Q}$ at $t_{\rm exp}$, but
1015: with a smaller range.  Typically they are $\sim$3.5 B/s with one at
1016: 5.57 B/s.  The heating efficiencies are similarly high, $\sim$5.9 to
1017: $\sim$10\%.  On the other hand, 2D simulations have lower efficiencies
1018: at explosion.  All 2D
1019: models that use the 15-M$_{\sun}$ progenitor have $\dot{Q}$ that lies
1020: in the range $\sim$1 to $\sim$2 B/s, and have efficiencies that
1021: range from $\sim$3 to $\sim$4\%.  For most models, $M_{\rm gain}$ at $t_{\rm
1022:   exp}$ is $\sim$0.01 M$_{\sun}$, but can be as low as 0.0073 M$_{\sun}$ and
1023: as high as 0.0319 M$_{\sun}$. 
1024: 
1025: In Figs. \ref{qdotvstime}, \ref{effvstime}, and
1026: \ref{mass_gain}, the exploding 2D models show distinctly different
1027: values of these indicators
1028: compared to those of the non-exploding 1D and 2D simulations.  $\dot{Q}$,
1029: $\dot{Q}/L_{\nu_e \bar{\nu}_e}$, and $M_{\rm gain}$ trend downward for
1030: the non-exploding models.  Just before, during, and after explosion,
1031: the heating rate, heating efficiency, and gain mass increase with time
1032: for the 2D exploding models.
1033: \citet{scheck08} have shown a similar trend in comparing 1D and 2D
1034: models at the same neutrino luminosity, and we show that this
1035: bifurcation exists among the 2D exploding and non-exploding models as well.
1036: 
1037: \section{Critical Luminosity and Mass Accretion Rate}
1038: \label{section:critical}
1039: 
1040: Figure \ref{lumvsmdot} indicates that the concept of a
1041: critical luminosity versus accretion rate condition is relevant for time-dependent
1042: 1D and 2D simulations and quantifies the difference in critical
1043: luminosity between 1D and 2D simulations.
1044: It plots electron- plus anti-electron-neutrino luminosity, $L_{\nu_e \bar{\nu}_e}$ ,
1045:   vs. accretion rate, $\dot{M}$.
1046: The trajectory of each model evolves from right to left
1047: as the accretion rate decreases with time and the luminosity is held
1048: constant.  For models that explode, we plot the luminosities and accretion rates
1049: at explosion. These points strongly suggest $L_{\nu_e
1050:   \bar{\nu}_e}$-$\dot{M}$ curves that separate exploding from
1051: non-exploding models.  Indeed, models that do not explode remain in
1052: the low luminosity and high-accretion-rate realm that is delineated by the
1053: critical curves.
1054: With error bars, we show the range of
1055:   accretion rates encompassed by $t_{\rm exp} \pm 50$ ms.  In general, 1D simulations are represented by
1056:   orange, 2D simulations by green hues, and 2D-90$^{\circ}$ runs by blue
1057:   hues.  The three resolutions are represented by stars (1),
1058:   upside-down triangles (2), and squares (3).  The fit for $L_{\nu_e \bar{\nu}_e}$ as
1059:   a function of $\dot{M}$ from \citet{burrows93} is plotted (solid
1060:   line).  While their fit passes through the 15.0-1D
1061:   points, it overpredicts the critical luminosities for the
1062:   11.2-1D runs by $\sim$15\%.  The critical luminosity for all 2D
1063:   simulations is $\sim$70\% of the critical luminosity for 1D
1064:   calculations.
1065: 
1066: It is interesting that the points for the 15.0-1D sequence are
1067: consistent with the results of \citet{burrows93}.  However, we caution
1068: against interpreting this coincidence too literally.  While we do
1069: employ very similar heating and cooling terms, it is not a priori clear how
1070: the many differences between our more realistic time-dependent models
1071: should compare to the assumptions made in \citet{burrows93} to obtain
1072: steady-state solutions.
1073: 
1074: Furthermore, we note that the slope of our results differ
1075: from the slope of the \citet{burrows93} fit and the trend
1076: is discontinuous between the 11.2 and 15 M$_{\sun}$ results.
1077: The fit from \citet{burrows93} assumed a specific mass for the
1078: PNS, while our simulations have masses that naturally
1079: arise from the initial Fe-core mass and the subsequent accretion
1080: rates.  Just after the Si/O interface accretes onto the
1081: PNS, the mass of the PNS, $M_{\rm PNS}$, is $\sim$1.3 M$_{\sun}$ baryonic for
1082: the 11.2-M$_{\sun}$ progenitor and $\sim$1.45 M$_{\sun}$ baryonic for the
1083: 15-M$_{\sun}$ progenitor.  Afterwards, the progenitor-dependent
1084: accretion rates continue to increase the mass until
1085: explosion.  Of the models from the 15.0-2D3 suite that exploded, the PNS
1086: masses at $t_{\rm exp}$ are 1.63, 1.59, and 1.52 M$_{\sun}$ for
1087: $L_{\nu_e} = 1.8$, 1.9, and 2.0.  In contrast, the masses from the
1088: 11.2-2D3 suite are 1.4, 1.38, 1.35, 1.33, and 1.32 M$_{\sun}$ for
1089: $L_{\nu_e} = 0.8$, 0.9, 1.0, 1.1, and 1.2, respectively.  Assuming
1090: that the explosion mechanism can be described by the transition from
1091: steady-state accretion to a neutrino-driven wind (see
1092: \citet{burrows87} and \citet{burrows93}, and work in 
1093: preparation), we expect the critical luminosity to have
1094: a mass dependence of $L_{\nu_e} \propto M_{\rm PNS}^{4/5}$.
1095: Interestingly, this scaling with PNS mass resolves the
1096: slope discrepancy and the discontinuity between the 15 and
1097: 11.2-M$_{\sun}$ models in Fig. \ref{lumvsmdot}.
1098: 
1099: Although the concept of a critical $L_{\nu_e}$ and $\dot{M}$ condition was derived
1100: using steady-state accretion models \citep{burrows93},
1101: Fig. \ref{lumvsmdot} shows that the condition is appropriate for
1102: time-dependent 1D and 2D models.
1103: We find that the critical luminosity for explosion is a
1104: function of $\dot{M}$, and that it roughly follows the
1105: trend suggested by \citet{burrows93}.
1106: The critical luminosity for explosions in 2D is $\sim$70\% of
1107: the luminosity required for explosions in 1D simulations.
1108: While there are differences
1109: between the 90$^{\circ}$ and 180$^{\circ}$ simulations and between simulations
1110: with different resolutions, they roughly show
1111: the same reduction in critical luminosity when we go to 2D.
1112: 
1113: 
1114: \section{Conclusions}
1115: \label{section:conclusions}
1116: Since \citet{burrows93} published the critical luminosity condition for explosions by the neutrino mechanism, the
1117: relevance of this condition for 1D and 2D time-dependent
1118: simulations has remained unresolved.  Furthermore, recent simulations
1119: have hinted that 2D simulations teeter on the verge of explosion, when
1120: 1D simulations completely fail.  One wonders if the concept of a
1121: critical condition in 2D simulations can quantitatively explain the
1122: trend toward successful explosions by the neutrino mechanism.
1123: 
1124: With 95 simulations that parameterize $L_{\nu_e}$, $\dot{M}$, resolution,
1125: and dimensionality, we investigate the criteria for neutrino-driven
1126: explosions. The results of which indicate that:
1127: \begin{itemize}
1128: \item {\bf A critical luminosity and accretion rate condition is observed
1129:   in time-dependent 1D simulations.}  \citet{burrows93} used the success and failure of obtaining
1130:   steady-state accretion solutions to suggest a critical luminosity and
1131:   mass accretion rate condition for explosions.  Others have found a critical neutrino
1132:   luminosity condition for explosion in time-dependent 1D simulations
1133:   \citep{ohnishi06,buras06a}.  We note that a critical luminosity is applicable
1134:   for time-dependent 1D
1135:   simulations
1136:   and that it is dependent upon $\dot{M}$, as suggested by \citet{burrows93}.
1137: \item {\bf Radial oscillations are characteristic of the transition
1138:     between 1D accretion (non-exploding) and exploding simulations.}
1139:   Our results add to a growing body of literature that obtain radial
1140:   oscillations near the critical luminosity
1141:   \citep{mayle85,ohnishi06,buras06a}.
1142: \item {\bf A critical $L_{\nu_e}$  and $\dot{M}$ condition distinguishes
1143:     accretion and exploding models of time-dependent 2D simulations.}
1144: \item {\bf Unlike 1D simulations, radial oscillations are not prominent
1145:   near the critical luminosity of 2D simulations.  Rather non-radial
1146:   SASI oscillations characterize 2D simulations, even for
1147:   non-exploding models.}
1148: \item {\bf We suggest that $\tau_{\rm adv}/\tau_q$
1149: is a useful diagnostic for explosion, but may not be a rigorous
1150: condition for explosion.}  Specifically, we found that some 1D models
1151: strongly satisfied this condition several times before explosion, and
1152: 2D models that exploded exhibited trends toward explosion well before
1153: this condition was met.
1154: \item {\bf Several integral quantities such as heating rate, heating
1155:     efficiency, and mass in the gain region grow with time in 1D and 2D
1156:     exploding models, while they decline in 1D and 2D non-exploding
1157:     models.}
1158:   \citet{scheck08} compared these quantities for 1D and 2D simulations
1159:   with the same neutrino luminosity and found declining evolutions in
1160:   the 1D non-exploding model and increasing trends in the 2D
1161:   exploding model.  We find similar trends.
1162:   From these results, one might suggest that the SASI leads to
1163:   increases in gain region mass and net heating rate, which in turn
1164:   leads to explosion.  However, 2D simulations that do not explode
1165:   despite having pronounced SASI oscillations show declining
1166:   curves. Furthermore, 1D exploding models also show increases in
1167:   heating rate and gain region mass at
1168:   explosion.  Therefore, we conclude that the rise in these quantities
1169:   near explosion are not a unique result of the SASI.  Instead, the upward evolution is correlated with
1170:   explosion, whether it be in 1D or 2D simulations.  This, combined
1171:   with the fact that the integral measures change from declining to
1172:   increasing evolution near explosion, suggests that the upward trends
1173:   are a consequence of explosion rather than a cause of explosion.
1174: \item {\bf We suggest an improved measure of the residence time of matter in the gain region,
1175:   $\tau_r$, that better elucidates the conditions leading to explosions.} While $\tau_{\rm adv}$ is a convenient measure of the timescale
1176:   that a parcel of mass is exposed to net heating,
1177:   we find $\tau_r$ a more informative measure for diagnosing
1178:   the conditions and criteria leading to and during explosion in 2D simulations.  For
1179:   example, by following the trajectories of particles that
1180:   start outside and within the gain region, we note that much of the
1181:   mass that accretes through the shock receives very little heating as
1182:   it quickly advects onto the inner core.  On the other hand, there is
1183:   a significant amount of mass within the gain region that has long
1184:   residence times and receives a great deal of neutrino heating.  It
1185:   is this latter material that experiences longer and longer
1186:   residence times and prolonged heating that likely reverses accretion
1187:   into explosion.
1188: \item {\bf Resolution does affect the SASI and the post-shock flow in
1189:     2D simulations, but hardly
1190:   alters the critical luminosity and accretion rate condition.}  It
1191:   noticeably alters the explosion time, but the correspondence between
1192:   $t_{\rm exp}$ and resolution is not monotonic, nor is it easy to
1193:   disentangle the trends.  However, the 2D critical luminosity
1194:   vs. $\dot{M}$ curve is little affected by resolution.
1195: \item {\bf The 90$^{\circ}$ simulations explode earlier than
1196:   the 180$^{\circ}$ counterpart, but again the critical luminosity
1197:   condition is hardly affected.}  We contrast this
1198: with the results of \citet{buras06b} who found an explosion for
1199: a 180$^{\circ}$ simulation, while a 90$^{\circ}$ simulation did not explode.
1200: However, their 90$^{\circ}$ simulation is $\pm45^o$ of the
1201: equator and suppresses $\ell= 1$ and $\ell = 2$ nonradial flow, the
1202: dominant components of the SASI.  We simulated the region from the pole to the
1203: equator, which suppresses $\ell = 1$ but allows $\ell = 2$.
1204: In agreement with our results, \citet{marek07b} found earlier explosions for
1205: 90$^{\circ}$ compared to 180$^{\circ}$ simulations using the same
1206: hydrodynamics and neutrino transport as \citet{buras06b} but a
1207: 90$^{\circ}$ grid similar to ours.
1208: In any case, \citet{buras06b} speculated that
1209: their 90$^{\circ}$ simulation was close to explosion and not very
1210: different than the 180$^{\circ}$ simulation, and we show that the critical luminosity for
1211: simulations using 90$^{\circ}$ and 180$^{\circ}$ are very similar.
1212: \item {\bf The critical luminosity for explosions in 2D simulations is
1213:     $\sim$70\% of
1214:   the luminosity required for explosions in 1D simulations.}
1215:   Irrespective of resolution or the angular size of the simulation, this
1216:   conclusion holds.
1217: \end{itemize}
1218: 
1219: By employing local heating and cooling algorithms, we avoid expensive
1220: neutrino transport, but we also sacrifice accuracy.
1221: In our simulations, heating and cooling by neutrinos are
1222: decoupled from one another, which is not consistent with neutrino transport
1223: calculations.
1224: In addition, simulations using MGFLD \citep{dessart06} and
1225: ``ray-by-ray'' Boltzmann transport \citep{marek07} have neutrino
1226: luminosities that decrease by $\sim$50\% from 100 ms to 300 ms
1227: past bounce.
1228: Our constant $L_{\nu_e}$ certainly does not reflect
1229: this evolution in neutrino luminosities.
1230: Furthermore, a large fraction of $L_{\nu_e}$ is derived from accretion at
1231:   early times \citep{scheck06}.  In our simulations, $L_{\nu_e}$ is completely
1232:   decoupled from such effects, and, hence, we ignore any feedback
1233:   due to this coupling.  While our assumptions compromise the
1234:   accuracy of the results, they enable the parameter study that we
1235:   present, and we suspect that the general conclusions and trends regarding
1236:   the critical luminosity condition are still
1237:   relevant for simulations including detailed transport.
1238: 
1239: \citet{marek07} suggest that using a softer EOS in conjunction
1240: with general relativistic gravity leads to more compact PNSs and more
1241: favorable conditions for explosion; we include neither effect.
1242: The soft EOS that accompanies explosion has an incompressibility of
1243: nuclear matter, $K$, of 180 MeV, while the stiffer EOS, which does not
1244: explode, has $K = 263$ MeV.  While this suggests an
1245: interesting dependence of explosion on the EOS, we note that
1246: laboratory experiments indicate that $K = 240 \pm 20$ MeV
1247: \citep{shlomo06,lattimer07}, a value that is more consistent with the
1248: non-explosive model.
1249: Regardless, with a compact PNS, the post-shock flow is positioned deeper in the
1250: potential, making it harder to explode.  All else being equal, this
1251: should increase the critical luminosity.  However, \citet{marek07}
1252: argue that the more compact PNS is hotter, emitting more energetic
1253: neutrinos (though they are gravitationally redshifted) and at a higher
1254: luminosity, which might compensate for the higher critical luminosity.  
1255: Because we parameterize the luminosity and temperature in this study,
1256: these effects are not addressed here and must await a more consistent
1257: treatment.
1258: 
1259: Of the conclusions listed above, the most striking is that
1260: time-dependent 1D and 2D simulations show a critical neutrino luminosity and
1261: accretion rate condition for successful explosions.
1262: Equally significant is
1263: that the critical luminosity for 2D simulations is $\sim$70\% of the
1264: critical luminosity for 1D
1265: simulations.  This quantitatively supports
1266: suspicions that multi-dimensional effects increase the likelihood of
1267: successful explosions by the neutrino mechanism.  While none have explicitly explored these results,
1268: the simulations of the past decade are consistent with this
1269: conclusion, while also being intuitive and concise.  In this
1270: sense, these results are expected and reassuring.  Remarkably, this
1271: global condition for successful explosions, which was
1272: informed by 1D steady-state solutions, survives despite the
1273: complexities of time-dependent multi-dimensional simulations.
1274: 
1275: For all permutations of the core-collapse mechanism considered in this paper,
1276: oscillations are ubiquitous
1277: during the transition from accretion to explosion.
1278: Interestingly, the character of these oscillations for 1D and 2D
1279: simulations is quite different, and leads to distinctly different critical
1280: luminosities.  Specifically, we suggest that the extra degree of freedom in 2D
1281: simulations offers an easier route to
1282: explosion via the SASI.  This suggests that for other than the
1283: lightest progenitors, the core-collapse mechanism is inherently
1284: multi-dimensional, and an accurate theory for the core-collapse
1285: mechanism must be developed in the context of the multi-dimensional Universe.
1286: 
1287: Our results show that the critical luminosity for 2D
1288: simulations is lower than the critical luminosity in 1D simulations,
1289: but why is this the case?  The most obvious cause is
1290: that the residence time of matter in the gain region is different for
1291: 1D and 2D simulations, and the longer residence times in 2D
1292: simulations reduce the neutrino luminosity required for explosion.
1293: So, what accounts for the different residence times when the accretion
1294: rate is the same?  After all, we have shown that during the steady-state
1295: accretion phase, the advection timescales determined by the flow
1296: across the gain region ($\tau_{\rm adv}$, as calculated
1297: by eq. (\ref{eq:tauadv})) and by $\tau_m = M_{\rm gain}/ \dot{M}$ are
1298: similar.  Put another way, the solid-angle, average, radial velocity at radius $r$ is
1299: \begin{equation}
1300: v_r(r) = \frac{\dot{M}(r)}{4 \pi r^2 \rho(r)} \, ,
1301: \end{equation}
1302: and if $\rho(r)$ and $\dot{M}(r)$ are the same for 1D and 2D simulations,
1303: the flow across the gain region should be similar.  Yet, 2D
1304: simulations explode when 1D simulations do not.  This discrepancy can
1305: be resolved by noting that the flow in the gain region is by no means
1306: spherically symmetric in 2D simulations.  The plots, histograms,
1307: and entropy maps of Figs. \ref{taures_mean_frac}-\ref{stills3} show
1308: that much of the material that accretes through the shock is channeled
1309: to plumes and quickly accrete onto the PNS.  At the same time, there
1310: are large regions of high entropy material that have long residence
1311: times in the gain region.  When the residence times within
1312: these high entropy regions reach some threshold, it seems as though sufficient heating
1313: has occurred to launch an explosion, and because relevant regions behind the
1314: shock are in sonic contact, the shock radius expands globally.  Hence, the
1315: critical luminosities for 2D simulations are lower because some matter
1316: resides for longer times in the gain region for the same $\dot{M}$.
1317: 
1318: Since the concept of a critical luminosity condition applies for 1D and 2D
1319: simulations, we assume that it should also apply for 3D
1320: simulations.
1321: Increasing the dimensionality from 1D to 2D, has proven favorable for successful explosions by the
1322: neutrino mechanism.  Specifically, the
1323: extra degree of freedom in 2D simulations increases the dwell time
1324: of matter in the gain region.  Similarly, we suspect that differences between 2D
1325: and 3D simulations will continue the trend toward lower critical
1326: luminosities for higher dimensions.
1327: For example, convection in 2D simulations experiences a turbulent energy cascade from
1328: smaller to larger scales.
1329: In Nature and in 3D simulations, however, the turbulent energy cascade is described
1330: well by Kolmogorov theory, in which energy cascades from larger to
1331: smaller scales.  If we analyze the transport of matter through the
1332: convective gain region as a diffusive problem, then
1333: the time to diffuse through the gain region would roughly scale as
1334: \begin{equation}
1335: \label{eq:tdiff}
1336: t_{\rm diff} \sim d \left ( \frac{l^2}{\overline{v^2} \tau_{\rm coll} } \right ) \, ,
1337: \end{equation}
1338: where $d$ is the number of degrees of freedom, $l$ is the characteristic size, $\overline{v^2}$ is the mean-squared
1339: speed of the diffusing particle, and $\tau_{\rm coll}$ is the average time
1340: between collisions.
1341: The extra degree of freedom in 3D simulations alone
1342: could increase the dwell time by $\sim$50\% compared to 2D
1343: simulations.  In addition, we suspect that the smaller scales of 3D
1344: simulations would produce shorter $\tau_{\rm coll}$ compared to 2D
1345: simulations and, consequently, further increase the dwell times.
1346: Whether these scaling arguments bear out in realistic 3D simulations, in which
1347: advection and the SASI are key in the flow, will be
1348: determined by future 3D simulations.  If they are relevant, then the
1349: increased dwell time might result in a critical luminosity in 3D
1350: simulations that is even lower than the critical luminosity in 2D
1351: simulations.  In turn, this might lead to successful explosions in 3D
1352: radiation-hydrodynamic simulations.
1353: 
1354: 
1355: \acknowledgments
1356: We would like to thank Luc Dessart and Christian Ott for fruitful
1357: comments.
1358:  We acknowledge support for this work                                           
1359: from the Scientific Discovery through Advanced Computing                        
1360: (SciDAC) program of the DOE, under grant numbers DE-FC02-01ER41184              
1361: and DE-FC02-06ER41452, and from the NSF under grant number AST-0504947.         
1362: J.W.M. thanks the Joint Institute for Nuclear Astrophysics (JINA) for           
1363: support under NSF grant PHY0216783.  In addition, J.W.M. is supported by an NSF Astronomy and Astrophysics Postdoctoral Fellowship under award AST-0802315.
1364: We thank Jeff Fookson and Neal Lauver of the Steward Computer Support
1365: Group for their invaluable help with the local Beowulf cluster, Grendel.
1366: 
1367: \bibliographystyle{apj}
1368: %\bibliography{jeremiah}
1369: \begin{thebibliography}{54}
1370: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
1371: 
1372: \bibitem[{{Bethe} \& {Wilson}(1985)}]{bethe85}
1373: {Bethe}, H.~A., \& {Wilson}, J.~R. 1985, \apj, 295, 14
1374: 
1375: \bibitem[{{Bionta} {et~al.}(1987){Bionta}, {Blewitt}, {Bratton}, {Caspere}, \&
1376:   {Ciocio}}]{bionta87}
1377: {Bionta}, R.~M., {Blewitt}, G., {Bratton}, C.~B., {Caspere}, D., \& {Ciocio},
1378:   A. 1987, Physical Review Letters, 58, 1494
1379: 
1380: \bibitem[{{Blondin} \& {Mezzacappa}(2006)}]{blondin06}
1381: {Blondin}, J.~M., \& {Mezzacappa}, A. 2006, \apj, 642, 401
1382: 
1383: \bibitem[{{Blondin} {et~al.}(2003){Blondin}, {Mezzacappa}, \&
1384:   {DeMarino}}]{blondin03}
1385: {Blondin}, J.~M., {Mezzacappa}, A., \& {DeMarino}, C. 2003, \apj, 584, 971
1386: 
1387: \bibitem[{{Bruenn}(1985)}]{bruenn85}
1388: {Bruenn}, S.~W. 1985, \apjs, 58, 771
1389: 
1390: \bibitem[{{Bruenn}(1989)}]{bruenn89}
1391: ---. 1989, \apj, 340, 955
1392: 
1393: \bibitem[{{Buras} {et~al.}(2006{\natexlab{a}}){Buras}, {Janka}, {Rampp}, \&
1394:   {Kifonidis}}]{buras06b}
1395: {Buras}, R., {Janka}, H.-T., {Rampp}, M., \& {Kifonidis}, K.
1396:   2006{\natexlab{a}}, \aap, 457, 281
1397: 
1398: \bibitem[{{Buras} {et~al.}(2003){Buras}, {Rampp}, {Janka}, \&
1399:   {Kifonidis}}]{buras03}
1400: {Buras}, R., {Rampp}, M., {Janka}, H.-T., \& {Kifonidis}, K. 2003, Physical
1401:   Review Letters, 90, 241101
1402: 
1403: \bibitem[{{Buras} {et~al.}(2006{\natexlab{b}}){Buras}, {Rampp}, {Janka}, \&
1404:   {Kifonidis}}]{buras06a}
1405: ---. 2006{\natexlab{b}}, \aap, 447, 1049
1406: 
1407: \bibitem[{{Burrows} (1987)}]{burrows87}
1408: {Burrows}, A. 1987, \apjl, 318, L57-L61
1409: 
1410: \bibitem[{{Burrows} {et~al.}(2007{\natexlab{a}}){Burrows}, {Dessart}, \&
1411:   {Livne}}]{burrows07d}
1412: {Burrows}, A., {Dessart}, L., \& {Livne}, E. 2007{\natexlab{a}}, in American
1413:   Institute of Physics Conference Series, Vol. 937, Supernova 1987A: 20 Years
1414:   After: Supernovae and Gamma-Ray Bursters, ed. S.~{Immler} \& R.~{McCray},
1415:   370--380
1416: 
1417: \bibitem[{{Burrows} {et~al.}(2007{\natexlab{b}}){Burrows}, {Dessart}, {Livne},
1418:   {Ott}, \& {Murphy}}]{burrows07c}
1419: {Burrows}, A., {Dessart}, L., {Livne}, E., {Ott}, C.~D., \& {Murphy}, J.
1420:   2007{\natexlab{b}}, \apj, 664, 416
1421: 
1422: \bibitem[{{Burrows} {et~al.}(2007{\natexlab{c}}){Burrows}, {Dessart}, {Ott}, \&
1423:   {Livne}}]{burrows07b}
1424: {Burrows}, A., {Dessart}, L., {Ott}, C.~D., \& {Livne}, E. 2007{\natexlab{c}},
1425:   \physrep, 442, 23
1426: 
1427: \bibitem[{{Burrows} \& {Goshy}(1993)}]{burrows93}
1428: {Burrows}, A., \& {Goshy}, J. 1993, \apjl, 416, L75+
1429: 
1430: \bibitem[{{Burrows} {et~al.}(1995){Burrows}, {Hayes}, \& {Fryxell}}]{burrows95}
1431: {Burrows}, A., {Hayes}, J., \& {Fryxell}, B.~A. 1995, \apj, 450, 830
1432: 
1433: \bibitem[{{Burrows} {et~al.}(2006){Burrows}, {Livne}, {Dessart}, {Ott}, \&
1434:   {Murphy}}]{burrows06}
1435: {Burrows}, A., {Livne}, E., {Dessart}, L., {Ott}, C.~D., \& {Murphy}, J. 2006,
1436:   \apj, 640, 878
1437: 
1438: \bibitem[{{Burrows} {et~al.}(2007{\natexlab{d}}){Burrows}, {Livne}, {Dessart},
1439:   {Ott}, \& {Murphy}}]{burrows07a}
1440: ---. 2007{\natexlab{d}}, \apj, 655, 416
1441: 
1442: \bibitem[{{Burrows} {et~al.}(2000){Burrows}, {Young}, {Pinto}, {Eastman}, \&
1443:   {Thompson}}]{burrows00}
1444: {Burrows}, A., {Young}, T., {Pinto}, P., {Eastman}, R., \& {Thompson}, T.~A.
1445:   2000, \apj, 539, 865
1446: 
1447: \bibitem[{{Dessart} {et~al.}(2006){Dessart}, {Burrows}, {Livne}, \&
1448:   {Ott}}]{dessart06}
1449: {Dessart}, L., {Burrows}, A., {Livne}, E., \& {Ott}, C.~D. 2006, \apj, 645, 534
1450: 
1451: \bibitem[{{Foglizzo}(2002)}]{foglizzo02}
1452: {Foglizzo}, T. 2002, \aap, 392, 353
1453: 
1454: \bibitem[{{Foglizzo} {et~al.}(2007){Foglizzo}, {Galletti}, {Scheck}, \&
1455:   {Janka}}]{foglizzo07}
1456: {Foglizzo}, T., {Galletti}, P., {Scheck}, L., \& {Janka}, H.-T. 2007, \apj,
1457:   654, 1006
1458: 
1459: \bibitem[{{Foglizzo} \& {Tagger}(2000)}]{foglizzo00}
1460: {Foglizzo}, T., \& {Tagger}, M. 2000, \aap, 363, 174
1461: 
1462: \bibitem[{{Herant} {et~al.}(1994){Herant}, {Benz}, {Hix}, {Fryer}, \&
1463:   {Colgate}}]{herant94}
1464: {Herant}, M., {Benz}, W., {Hix}, W.~R., {Fryer}, C.~L., \& {Colgate}, S.~A.
1465:   1994, \apj, 435, 339
1466: 
1467: \bibitem[{{Hirata} {et~al.}(1987){Hirata}, {Kajita}, {Koshiba}, {Nakahata}, \&
1468:   {Oyama}}]{hirata87}
1469: {Hirata}, K., {Kajita}, T., {Koshiba}, M., {Nakahata}, M., \& {Oyama}, Y. 1987,
1470:   Physical Review Letters, 58, 1490
1471: 
1472: \bibitem[{{Janka}(2001)}]{janka01}
1473: {Janka}, H.-T. 2001, \aap, 368, 527
1474: 
1475: \bibitem[{{Janka} \& {M\"{u}ller}(1995)}]{janka95}
1476: {Janka}, H.-T., \& {M\"{u}ller}, E. 1995, \apjl, 448, L109
1477: 
1478: \bibitem[{{Janka} \& {M\"{u}ller}(1996)}]{janka96}
1479: ---. 1996, \aap, 306, 167
1480: 
1481: \bibitem[{{Kitaura} {et~al.}(2006){Kitaura}, {Janka}, \&
1482:   {Hillebrandt}}]{kitaura06}
1483: {Kitaura}, F.~S., {Janka}, H.-T., \& {Hillebrandt}, W. 2006, \aap, 450, 345
1484: 
1485: \bibitem[{{Lattimer} \& {Prakash}(2007)}]{lattimer07}
1486: {Lattimer}, J.~M., \& {Prakash}, M. 2007, \physrep, 442, 109-165
1487: 
1488: \bibitem[{{Liebend{\"o}rfer}(2005)}]{liebendorfer05a}
1489: {Liebend{\"o}rfer}, M. 2005, \apj, 633, 1042
1490: 
1491: \bibitem[{{Liebend{\"o}rfer} {et~al.}(2001{\natexlab{a}}){Liebend{\"o}rfer},
1492:   {Mezzacappa}, \& {Thielemann}}]{liebendorfer01b}
1493: {Liebend{\"o}rfer}, M., {Mezzacappa}, A., \& {Thielemann}, F.-K.
1494:   2001{\natexlab{a}}, \prd, 63, 104003
1495: 
1496: \bibitem[{{Liebend{\"o}rfer} {et~al.}(2001{\natexlab{b}}){Liebend{\"o}rfer},
1497:   {Mezzacappa}, {Thielemann}, {Messer}, {Hix}, \& {Bruenn}}]{liebendorfer01a}
1498: {Liebend{\"o}rfer}, M., {Mezzacappa}, A., {Thielemann}, F.-K., {Messer}, O.~E.,
1499:   {Hix}, W.~R., \& {Bruenn}, S.~W. 2001{\natexlab{b}}, \prd, 63, 103004
1500: 
1501: \bibitem[{{Liebend{\"o}rfer} {et~al.}(2005){Liebend{\"o}rfer}, {Rampp},
1502:   {Janka}, \& {Mezzacappa}}]{liebendorfer05b}
1503: {Liebend{\"o}rfer}, M., {Rampp}, M., {Janka}, H.-T., \& {Mezzacappa}, A. 2005,
1504:   \apj, 620, 840
1505: 
1506: \bibitem[{{Livne}(1993)}]{livne93}
1507: {Livne}, E. 1993, \apj, 412, 634
1508: 
1509: \bibitem[{{Livne} {et~al.}(2004){Livne}, {Burrows}, {Walder}, {Lichtenstadt},
1510:   \& {Thompson}}]{livne04}
1511: {Livne}, E., {Burrows}, A., {Walder}, R., {Lichtenstadt}, I., \& {Thompson},
1512:   T.~A. 2004, \apj, 609, 277
1513: 
1514: \bibitem[{{Marek}(2007)}]{marek07b}
1515: {Marek}, A. 2007, PhD thesis, Max Planck Institute for Astrophysics
1516: 
1517: \bibitem[{{Marek} \& {Janka}(2007)}]{marek07}
1518: {Marek}, A., \& {Janka}, H.~. 2007, ArXiv e-prints, 708
1519: 
1520: \bibitem[{{Mayle}(1985)}]{mayle85}
1521: {Mayle}, R.~W. 1985, PhD thesis, Lawrence Livermore National Laboratory,
1522:   University of California, Berkeley
1523: 
1524: \bibitem[{{Mazurek}(1982)}]{mazurek82}
1525: {Mazurek}, T.~J. 1982, \apjl, 259, L13
1526: 
1527: \bibitem[{{Murphy} \& {Burrows}(2008)}]{murphy08}
1528: {Murphy}, J.~W., \& {Burrows}, A. 2008, submitted to \apj
1529: 
1530: \bibitem[{{Ohnishi} {et~al.}(2006){Ohnishi}, {Kotake}, \& {Yamada}}]{ohnishi06}
1531: {Ohnishi}, N., {Kotake}, K., \& {Yamada}, S. 2006, \apj, 641, 1018
1532: 
1533: \bibitem[{{Ott} {et~al.}(2008){Ott}, {Burrows}, {Dessart}, \& {Livne}}]{ott08}
1534: {Ott}, C.~D., {Burrows}, A., {Dessart}, L., \& {Livne}, E. 2008, ArXiv
1535:   e-prints, 804
1536: 
1537: \bibitem[{{Rampp} \& {Janka}(2002)}]{rampp02}
1538: {Rampp}, M., \& {Janka}, H.-T. 2002, \aap, 396, 361
1539: 
1540: \bibitem[{{Scheck} {et~al.}(2008){Scheck}, {Janka}, {Foglizzo}, \&
1541:   {Kifonidis}}]{scheck08}
1542: {Scheck}, L., {Janka}, H.-T., {Foglizzo}, T., \& {Kifonidis}, K. 2008, \aap,
1543:   477, 931
1544: 
1545: \bibitem[{{Scheck} {et~al.}(2006){Scheck}, {Kifonidis}, {Janka}, \&
1546:   {M{\"u}ller}}]{scheck06}
1547: {Scheck}, L., {Kifonidis}, K., {Janka}, H.-T., \& {M{\"u}ller}, E. 2006, \aap,
1548:   457, 963
1549: 
1550: \bibitem[{{Shen} {et~al.}(1998){Shen}, {Toki}, {Oyamatsu}, \&
1551:   {Sumiyoshi}}]{shen98}
1552: {Shen}, H., {Toki}, H., {Oyamatsu}, K., \& {Sumiyoshi}, K. 1998, Nuclear
1553:   Physics A, 637, 435
1554: 
1555: \bibitem[{{Shlomo} {et~al.}(2006){Shlomo}, {Kolomietz}, \& {Col\`{o}}}]{shlomo06}
1556: {Shlomo}, S., {Kolomietz}, V.~M., \& {Col\`{o}}, G. 2006,
1557: {\it Eur. Phys. J.}, A30, 2
1558: 
1559: \bibitem[{{Thompson}(2000)}]{thompson00}
1560: {Thompson}, C. 2000, \apj, 534, 915
1561: 
1562: \bibitem[{{Thompson} {et~al.}(2003){Thompson}, {Burrows}, \&
1563:   {Pinto}}]{thompson03}
1564: {Thompson}, T.~A., {Burrows}, A., \& {Pinto}, P.~A. 2003, \apj, 592, 434
1565: 
1566: \bibitem[{{Thompson} {et~al.}(2005){Thompson}, {Quataert}, \&
1567:   {Burrows}}]{thompson05}
1568: {Thompson}, T.~A., {Quataert}, E., \& {Burrows}, A. 2005, \apj, 620, 861
1569: 
1570: \bibitem[{{Wilson}(1985)}]{wilson85}
1571: {Wilson}, J.~R. 1985 (Boston: Jones \& Bartlett), 422
1572: 
1573: \bibitem[{{Woosley} {et~al.}(2002){Woosley}, {Heger}, \& {Weaver}}]{woosley02}
1574: {Woosley}, S.~E., {Heger}, A., \& {Weaver}, T.~A. 2002, Reviews of Modern
1575:   Physics, 74, 1015
1576: 
1577: \bibitem[{{Woosley} \& {Weaver}(1995)}]{woosley95}
1578: {Woosley}, S.~E., \& {Weaver}, T.~A. 1995, \apjs, 101, 181
1579: 
1580: \bibitem[{{Yamasaki} \& {Yamada}(2005)}]{yamasaki05}
1581: {Yamasaki}, T., \& {Yamada}, S. 2005, \apj, 623, 1000
1582: 
1583: \bibitem[{{Yamasaki} \& {Yamada}(2006)}]{yamasaki06}
1584: ---. 2006, \apj, 650, 291
1585: 
1586: \end{thebibliography}
1587: 
1588: 
1589: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1590: %%%%  Tables  %%%%%%%%%%%%%%%%%%%%%%%%%%%
1591: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1592: 
1593: \clearpage
1594: 
1595: \begin{deluxetable}{cccccc}
1596: \tabletypesize{\scriptsize}
1597: \tablecaption{
1598: Model parameters.\tablenotemark{1}
1599: \label{table:sequences}
1600: }
1601: \tablewidth{0pt}
1602: \tablehead{
1603:   \colhead{Sequence Name} &
1604:   \colhead{Mass (M$_{\sun}$)\tablenotemark{2}} &
1605:   \colhead{Dimension\tablenotemark{3}} & 
1606:   \colhead{$N_{r}$\tablenotemark{4}} &
1607:   \colhead{$R_{\rm max}$ (km)\tablenotemark{5}}&
1608:   \colhead{$L_{\nu_e}$ ($10^{52}$ erg s$^{-1}$)\tablenotemark{6}}}
1609: \startdata
1610: 15.0-1D  & 15.0 & 1D & 700 & 4000 & 1.6-2.9 \\
1611: 15.0-2D1 & 15.0 & 2D & 250 & 4000 & 1.5-2.0 \\
1612: 15.0-2D2 & 15.0 & 2D & 400 & 4000 & 1.5-2.0 \\
1613: 15.0-2D3 & 15.0 & 2D & 400 & 1000 & 1.5-2.0 \\
1614: 15.0-Q1 & 15.0 & 2D-90$^{\circ}$ & 250 & 4000 & 1.5-2.0 \\
1615: 15.0-Q2 & 15.0 & 2D-90$^{\circ}$ & 400 & 4000 & 1.5-2.0 \\
1616: 15.0-Q3 & 15.0 & 2D-90$^{\circ}$ & 400 & 1000 & 1.5-2.0 \\
1617: 11.2-1D  & 11.2 & 1D & 700 & 4000 & 1.0-1.8 \\
1618: 11.2-2D1 & 11.2 & 2D & 250 & 4000 & 0.7-1.2 \\
1619: 11.2-2D2 & 11.2 & 2D & 400 & 4000 & 0.7-1.2 \\
1620: 11.2-2D3 & 11.2 & 2D & 400 & 1000 & 0.7-1.2 \\
1621: 11.2-Q1 & 11.2 & 2D-90$^{\circ}$ & 250 & 4000 & 0.7-1.2 \\
1622: 11.2-Q2 & 11.2 & 2D-90$^{\circ}$ & 400 & 4000 & 0.7-1.2 \\
1623: 11.2-Q3 & 11.2 & 2D-90$^{\circ}$ & 400 & 1000 & 0.7-1.2 \\
1624: \enddata
1625: \tablenotetext{1}{This table summarizes the 95 simulations presented
1626:   in this paper. These
1627: simulations represent a four-dimensional parameterization that
1628: investigates the dependence of the conditions for explosion on the
1629: accretion rate (column 2), dimensionality (column 3), resolution
1630: (columns 4 \& 5), and neutrino luminosity (column 6).}
1631: \tablenotetext{2}{Progenitor model.}
1632: \tablenotetext{3}{Dimensionality.}
1633: \tablenotetext{4}{Number of radial or effective radial zones.}
1634: \tablenotetext{5}{Radius of the outer boundary}
1635: \tablenotetext{6}{The range of neutrino luminosities investigated.}
1636: \end{deluxetable}
1637: 
1638: \clearpage
1639: 
1640: \LongTables
1641: \begin{deluxetable}{cccccc}
1642: \tabletypesize{\scriptsize}
1643: \tablecaption{
1644: Conditions at the time of explosion.
1645: \label{table:explosions}
1646: }
1647: \tablewidth{0pt}
1648: \tablehead{ 
1649:   \colhead{$L_{\nu_e}$ ($10^{52}$ erg s$^{-1}$)\tablenotemark{1}} &
1650:   \colhead{$t_{\rm exp}$ (ms)\tablenotemark{2}} &
1651:   \colhead{$\dot{M}_{\rm exp}$ (M$_{\sun}$/s)\tablenotemark{3}} &
1652:   \colhead{$\dot{Q}$ (B/s)\tablenotemark{4}} &
1653:   \colhead{$\dot{Q}/L_{\nu_e \bar{\nu}_e}$\tablenotemark{5}} &
1654:   \colhead{$M_{\rm gain}$ (M$_{\sun}$)\tablenotemark{6}}}
1655: \startdata
1656: \cutinhead{11.2-1D\tablenotemark{7}}
1657: 1.3 & 932 & $0.084^{+0.006}_{-0.003}$ & 0.63 & 0.0244 & 0.0123 \\
1658: 1.4 & 496 & $0.112^{+0.012}_{-0.010}$ & 1.50 & 0.0535 & 0.0121 \\
1659: 1.5 & 312 & $0.136^{+0.025}_{-0.011}$ & 2.00 & 0.0668 & 0.0153 \\
1660: 1.6 & 409 & $0.124^{+0.002}_{-0.003}$ & 4.46 & 0.1393 & 0.0228 \\
1661: \cutinhead{15.0-1D}
1662: 2.6 & 718 & $0.227^{+0.017}_{-0.015}$ & 3.69 & 0.0710 & 0.0105 \\
1663: 2.7 & 459 & $0.263^{+0.027}_{-0.013}$ & 3.79 & 0.0701 & 0.0129 \\
1664: 2.8 & 335 & $0.312^{+0.001}_{-0.006}$ & 5.57 & 0.0995 & 0.0177 \\
1665: 2.9 & 216 & $0.341^{+0.027}_{-0.027}$ & 3.42 & 0.0589 & 0.0319 \\
1666: \cutinhead{15.0-2D1}
1667: 1.9 & 581 & $0.247^{+0.001}_{-0.001}$ & 1.49 & 0.0392 & 0.0123 \\
1668: 2.0 & 325 & $0.310^{+0.002}_{-0.001}$ & 1.97 & 0.0494 & 0.0167 \\
1669: \cutinhead{15.0-2D2}
1670: 1.9 & 359 & $0.313^{+0.001}_{-0.023}$ & 1.51 & 0.0399 & 0.0120 \\
1671: 2.0 & 364 & $0.313^{+0.001}_{-0.027}$ & 1.60 & 0.0401 & 0.0133 \\
1672: \cutinhead{15.0-2D3}
1673: 1.8 & 778 & $0.210^{+0.014}_{-0.007}$ & 1.21 & 0.0337 & 0.0099 \\
1674: 1.9 & 649 & $0.249^{+0.001}_{-0.015}$ & 1.33 & 0.0351 & 0.0113 \\
1675: 2.0 & 395 & $0.299^{+0.014}_{-0.031}$ & 1.76 & 0.0440 & 0.0152 \\
1676: \cutinhead{15.0-Q1}
1677: 1.7 & 837 & $0.202^{+0.007}_{-0.005}$ & 0.91 & 0.0268 & 0.0074 \\
1678: 1.8 & 419 & $0.283^{+0.029}_{-0.024}$ & 1.19 & 0.0330 & 0.0098 \\
1679: 1.9 & 418 & $0.284^{+0.028}_{-0.024}$ & 1.31 & 0.0344 & 0.0109 \\
1680: 2.0 & 361 & $0.313^{+0.001}_{-0.025}$ & 1.71 & 0.0428 & 0.0152 \\
1681: \cutinhead{15.0-Q2}
1682: 1.7 & 748 & $0.217^{+0.018}_{-0.009}$ & 0.91 & 0.0267 & 0.0073 \\
1683: 1.8 & 517 & $0.249^{+0.011}_{-0.002}$ & 1.30 & 0.0362 & 0.0119 \\
1684: 1.9 & 471 & $0.258^{+0.023}_{-0.010}$ & 1.19 & 0.0314 & 0.0100 \\
1685: 2.0 & 299 & $0.310^{+0.008}_{-0.001}$ & 1.54 & 0.0385 & 0.0126 \\
1686: \cutinhead{15.0-Q3}
1687: 1.7 & 695 & $0.235^{+0.014}_{-0.018}$ & 1.14 & 0.0335 & 0.0096 \\
1688: 1.8 & 602 & $0.249^{+0.001}_{-0.001}$ & 1.08 & 0.0301 & 0.0089 \\
1689: 1.9 & 420 & $0.282^{+0.029}_{-0.024}$ & 1.42 & 0.0373 & 0.0122 \\
1690: 2.0 & 383 & $0.307^{+0.004}_{-0.033}$ & 1.66 & 0.0415 & 0.0135 \\
1691: \cutinhead{11.2-2D1}
1692: 0.7 & 808 & $0.091^{+0.000}_{-0.001}$ & 0.23 & 0.0165 & 0.0050 \\
1693: 0.8 & 585 & $0.098^{+0.005}_{-0.002}$ & 0.30 & 0.0190 & 0.0055 \\
1694: 0.9 & 471 & $0.118^{+0.006}_{-0.013}$ & 0.43 & 0.0242 & 0.0068 \\
1695: 1.0 & 335 & $0.129^{+0.020}_{-0.005}$ & 0.55 & 0.0276 & 0.0086 \\
1696: 1.1 & 288 & $0.147^{+0.016}_{-0.019}$ & 0.89 & 0.0404 & 0.0131 \\
1697: 1.2 & 288 & $0.147^{+0.016}_{-0.019}$ & 0.89 & 0.0370 & 0.0131 \\
1698: \cutinhead{11.2-2D2}
1699: 0.7 & 828 & $0.092^{+0.001}_{-0.002}$ & 0.24 & 0.0170 & 0.0054 \\
1700: 0.8 & 595 & $0.097^{+0.005}_{-0.002}$ & 0.31 & 0.0193 & 0.0053 \\
1701: 0.9 & 529 & $0.104^{+0.012}_{-0.005}$ & 0.40 & 0.0220 & 0.0070 \\
1702: 1.0 & 339 & $0.128^{+0.019}_{-0.004}$ & 0.54 & 0.0268 & 0.0077 \\
1703: 1.1 & 235 & $0.163^{+0.001}_{-0.014}$ & 0.85 & 0.0386 & 0.0115 \\
1704: 1.2 & 235 & $0.163^{+0.001}_{-0.014}$ & 0.85 & 0.0354 & 0.0115 \\
1705: \cutinhead{11.2-2D3}
1706: 0.8 & 791 & $0.091^{+0.000}_{-0.001}$ & 0.32 & 0.0200 & 0.0057 \\
1707: 0.9 & 553 & $0.101^{+0.009}_{-0.004}$ & 0.40 & 0.0225 & 0.0059 \\
1708: 1.0 & 372 & $0.125^{+0.008}_{-0.001}$ & 0.60 & 0.0300 & 0.0087 \\
1709: 1.1 & 257 & $0.162^{+0.001}_{-0.025}$ & 0.91 & 0.0413 & 0.0122 \\
1710: 1.2 & 257 & $0.162^{+0.001}_{-0.025}$ & 0.91 & 0.0378 & 0.0122 \\
1711: \cutinhead{11.2-Q1}
1712: 0.7 & 777 & $0.091^{+0.001}_{-0.001}$ & 0.23 & 0.0165 & 0.0048 \\
1713: 0.8 & 500 & $0.111^{+0.012}_{-0.009}$ & 0.31 & 0.0191 & 0.0060 \\
1714: 0.9 & 784 & $0.091^{+0.001}_{-0.001}$ & 0.78 & 0.0434 & 0.0133 \\
1715: 1.0 & 294 & $0.144^{+0.019}_{-0.017}$ & 0.52 & 0.0260 & 0.0076 \\
1716: 1.1 & 177 & $0.166^{+0.034}_{-0.004}$ & 0.91 & 0.0414 & 0.0124 \\
1717: 1.2 & 177 & $0.166^{+0.034}_{-0.004}$ & 0.91 & 0.0379 & 0.0124 \\
1718: \cutinhead{11.2-Q2}
1719: 0.7 & 631 & $0.096^{+0.003}_{-0.002}$ & 0.22 & 0.0158 & 0.0040 \\
1720: 0.8 & 525 & $0.105^{+0.013}_{-0.006}$ & 0.28 & 0.0174 & 0.0056 \\
1721: 0.9 & 413 & $0.124^{+0.001}_{-0.004}$ & 0.39 & 0.0218 & 0.0071 \\
1722: 1.0 & 293 & $0.145^{+0.018}_{-0.017}$ & 0.54 & 0.0269 & 0.0075 \\
1723: 1.1 & 270 & $0.158^{+0.004}_{-0.025}$ & 0.82 & 0.0371 & 0.0107 \\
1724: 1.2 & 270 & $0.158^{+0.004}_{-0.025}$ & 0.82 & 0.0340 & 0.0107 \\
1725: \cutinhead{11.2-Q3}
1726: 0.7 & 759 & $0.091^{+0.001}_{-0.001}$ & 0.22 & 0.0160 & 0.0047 \\
1727: 0.8 & 583 & $0.098^{+0.005}_{-0.003}$ & 0.30 & 0.0186 & 0.0056 \\
1728: 0.9 & 417 & $0.124^{+0.001}_{-0.005}$ & 0.39 & 0.0218 & 0.0065 \\
1729: 1.0 & 293 & $0.145^{+0.018}_{-0.017}$ & 0.56 & 0.0282 & 0.0083 \\
1730: 1.1 & 256 & $0.162^{+0.001}_{-0.024}$ & 0.84 & 0.0383 & 0.0125 \\
1731: 1.2 & 256 & $0.162^{+0.001}_{-0.024}$ & 0.84 & 0.0351 & 0.0125 \\
1732: \enddata
1733: \tablenotetext{1}{Electron-neutrino luminosity}
1734: \tablenotetext{2}{The time of explosion}
1735: \tablenotetext{3}{The mass accretion rate at explosion,}
1736: \tablenotetext{4}{The net heating rate}
1737: \tablenotetext{5}{The heating efficiency, where $L_{\nu_e \bar{\nu_e}}$ is the combined electron- and anti-electron-neutrino luminosity.}
1738: \tablenotetext{6}{The mass in the gain region.}
1739: \tablenotetext{7}{The table is divided into sections corresponding to the conditions at
1740: explosion for each sequence presented in Table \ref{table:sequences}.}
1741: \end{deluxetable}
1742: 
1743: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1744: %%%%  Figures  %%%%%%%%%%%%%%%%%%%%%%%%%%%
1745: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1746: 
1747: \clearpage
1748: 
1749: \begin{figure}
1750: \plotone{f1.ps}
1751: \caption{Accretion rate, $\dot{M}$, vs. postbounce time above the
1752:   stalled shock (250 km). The solid and dashed lines show the
1753:   time-dependent accretion rate for the 11.2- and 15.0-M$_{\sun}$ models,
1754:   respectively \citep{woosley02,woosley95}.  While the outer portions
1755:   of the Fe core accrete ($t = 0$-50 ms), $\dot{M}$ is as high as 10
1756:   M$_{\sun}$/s and decreases to 2 M$_{\sun}$/s.  After the Fe core fully
1757:   accretes, the accretion rates for the two models
1758:   diverge.  For the 11.2-M$_{\sun}$ (15-M$_{\sun}$) model, it takes 50 ms (100 ms) to accrete the
1759:   Si-burning shell.
1760: As the Si/O interface accretes, $\dot{M}$ plummets to 0.3 M$_{\sun}$/s
1761: (0.2 M$_{\sun}$/s) for the 15 M$_{\sun}$ (11.2 M$_{\sun}$) model.  Afterward,
1762: $\dot{M}$ declines slowly to 0.2 M$_{\sun}$/s (0.08 M$_{\sun}$/s).
1763: Together, these two models slowly sweep through a
1764:   range of accretion rates from 0.3 M$_{\sun}$/s to 0.08 M$_{\sun}$/s,
1765:   which enables a parameterization in $\dot{M}$. \label{mdotvstime}}
1766: \end{figure}
1767: 
1768: \clearpage
1769: 
1770: \begin{figure}
1771: \epsscale{0.7}
1772: \plotone{f2.ps}
1773: \caption{Shock radius, $R_{\rm
1774:   shock}$, vs. postbounce time for the 1D
1775: sequences.  The top panel displays the radii for 15.0-1D, and the
1776: bottom panel shows the radii for 11.2-1D.  Each line is labeled by the
1777: electron-neutrino luminosity in units of $10^{52}$ ergs s$^{-1}$.
1778: Comparing models within a sequence, all models show similar behavior prior to the accretion of
1779: the Si/O interface.  The accretion of the Si/O interface either initiates an
1780: explosion or excites radial oscillations in the shock radius.  For the
1781: 15.0-1D sequence, the oscillation periods range from $\sim$90 ms
1782: for $L_{\nu_e}=2.2$ to $\sim$170 ms for $L_{\nu_e} = 2.8$.  Of the
1783: models that oscillate, the lower luminosity simulations decay in
1784: the oscillation amplitude, while higher luminosity models oscillate
1785: and explode.  The timescales for decay range from $\sim$450 ms for
1786: $L_{\nu_e} = 2.2$ to 1.0 s for $L_{\nu_e} = 2.5$.
1787: \label{rvst_1d}}
1788: \end{figure}
1789: 
1790: \clearpage
1791: 
1792: \begin{figure}
1793: \epsscale{0.75}
1794: \plotone{f3.ps}
1795: \caption{Shock morphology vs. postbounce time for the 15.0-2D3
1796:   sequence. The top panel shows $a_0$ or $\langle R_{\rm shock} \rangle$, the average shock
1797: radius, and the middle and bottom panels plot the $\ell = 1$ and
1798: $\ell = 2$ components divided by to the average shock radius:
1799: $a_1/a_0$ and $a_2/a_0$, respectively.  The
1800: filled circles in the top panel indicate $t_{\rm exp}$, the time of explosion, and
1801: the error bars show our estimate of uncertainty ($\pm 50$ ms).
1802: Unlike the 1D models (Fig. \ref{rvst_1d}), the 2D simulations show
1803: very little radial oscillation near the critical luminosity, but
1804: significant nonradial oscillations.  In general, $a_1/a_0$ grows from
1805: 0\% to $\sim$10\% for all models, and $a_2/a_0$ grows to $\sim$5\%.  The $a_1/a_0$ amplitudes hint at a correlation with
1806: $L_{\nu_e}$ (see \S \ref{section:shock} for a discussion).  Near
1807: explosion the amplitude of $a_1/a_0$ can peak around $\sim$20\%. \label{sphharm}}
1808: \end{figure}
1809: 
1810: \clearpage
1811: 
1812: \begin{figure}
1813: \epsscale{0.8}
1814: \plotone{f4.ps}
1815: \caption{The shock morphology vs. time for three
1816: different resolutions of 2D simulations (15.0-2D1, 15.0-2D2, and
1817: 15.0-2D3).  The layout is similar to Fig. \ref{sphharm}.  The
1818: electron-neutrino luminosity for each is $1.9 \times 10^{52}$ ergs
1819: s$^{-1}$.  From lowest resolution (15.0-2D1) to the highest resolution
1820: (15.0-2D3), the explosion times are 581, 359, and 649 ms.
1821: Hence, $t_{\rm exp}$ is not monotonic with resolution.  See \S
1822: \ref{section:grid} for a discussion.
1823: \label{morphres}}
1824: \end{figure}
1825: 
1826: \clearpage
1827: 
1828: \begin{figure}
1829: \epsscale{0.75}
1830: \plotone{f5.ps}
1831: \caption{$\langle R_{\rm shock} \rangle$ and
1832: $a_2/a_0$ vs. time for two models of 15.0-Q3
1833: and 15.0-2D3.  This compares the differences among the 2D runs using
1834: 180$^{\circ}$ (solid) and 90$^{\circ}$ (dashed).  $L_{\nu_e} = 1.8$ corresponds to the
1835: purple curves and $L_{\nu_e} = 1.9$ corresponds to the green curves.
1836: Inspection of $\langle R_{\rm shock} \rangle$ (top panel of
1837: Fig. \ref{fullvs90}) shows that the 90$^{\circ}$ simulations consistently explode
1838: before the 180$^{\circ}$ simulations.  Other than the time near $\sim$300 to $\sim$400 ms, all models show similar evolution in the
1839: amplitude of $a_2/a_0$.  The one exception appears around $\sim$300
1840: to $\sim$400 ms, when the 90$^{\circ}$, $L_{\nu_e} = 1.9$ model explodes and shows the largest amplitude.
1841: \label{fullvs90}}
1842: \end{figure}
1843: 
1844: \clearpage
1845: 
1846: \begin{figure}
1847: \epsscale{0.7}
1848: \plotone{f6.ps}
1849: \caption{The ratio of advection and heating timescales, $\tau_{\rm adv}/\tau_{\rm
1850:   q}$, vs. postbounce time for the 15.0-1D sequence.  The top
1851: panel shows the models that do not explode by 1.3 s after bounce, and
1852: the bottom panel shows those that do explode.  Despite dramatic
1853: oscillations, $\tau_{\rm adv}/\tau_{\rm q}$ rarely reaches a value of
1854: above $\sim$1 for the non-exploding models (top panel).  In all non-exploding
1855: cases, except $L_{\nu_e} = 2.5$ (in units of $10^{52}$ erg s$^{-1}$), the
1856: oscillations decay (see text and the caption of Fig. \ref{rvst_1d} for
1857: the timescales).  For exploding
1858: models (bottom panel), $\tau_{\rm adv}/\tau_{q}$ makes large
1859: excursions from $\sim$0.1 to $\sim$100.  Larger luminosities produce more excursions,
1860: with one at $L_{\nu_e} = 2.9$ and four at
1861: $L_{\nu_e} = 2.9$.  We conclude that $\tau_{\rm
1862:   adv}/\tau_{q} > 1$ is a useful diagnostic, but not a rigorous condition for explosion.\label{tscalesvst_1d}}
1863: \end{figure}
1864: 
1865: \clearpage
1866: 
1867: \begin{figure}
1868: \epsscale{0.75}
1869: \plotone{f7.ps}
1870: \caption{Similar to Fig. \ref{tscalesvst_1d}, but we plot
1871: $\tau_{\rm adv}/\tau_{q}$ vs. postbounce time for the 15.0-2D3 sequence.  The models that don't explode, $L_{\nu_e} =
1872: 1.5$, 1.6, and 1.7 (in units of $10^{52}$ erg s$^{-1}$) show relatively constant
1873: ratios for all times.  On the other hand, the exploding models,
1874: $L_{\nu_e} = 1.8$, 1.9, and 2.0, start at low ratios ($\sim$0.1)
1875: and increase to ratios of order $\sim$10.  Unlike the 1D
1876: models, once this ratio reaches large values, it
1877: remains there for the rest of the simulation.  The times of explosion
1878: are 395, 648, and 778 ms for $L_{\nu_e} = 1.8$, 1.9, and 2.0,
1879: respectively.  Notice that the ratio of timescales begin to rise
1880: before $\tau_{\rm adv}/\tau_q = 1$, suggesting that these models are
1881: on their way to explosion before the condition $\tau_{\rm adv}/\tau_q
1882: > 1$ is met.\label{tscalesvst_2d}}
1883: \end{figure}
1884: 
1885: \clearpage
1886: 
1887: \begin{figure}
1888: \epsscale{0.68}
1889: \plotone{f8.ps}
1890: \caption{The mean residence time ($\langle \tau_r \rangle$) (top
1891: panel) and the fraction ($f_r$) of particles with $\tau_r > 40$ ms (bottom
1892: panel) as a function of time for $R_{\rm start} = 400$ km (solid line)
1893: and $R_{\rm start} = 150$ km (dashed line).  These residence times
1894: were calculated for the 15.0-2D3, $L_{\nu_e} = 1.9$ model.  50,000 tracer particles are
1895: initiated randomly in solid angle at either $R_{\rm start} = 150$ km
1896: or $R_{\rm start} = 400$ km.  The trajectories for these particles are
1897: integrated for 150 ms, and a new generation is initiated every 10 ms.
1898: The time on the horizontal axis corresponds to the start time for a
1899: generation.  From $\sim$200 ms and
1900: beyond, particles that start in the gain region ($R_{\rm start} = 150$
1901: km) consistently show larger $\langle \tau_r \rangle$ and more
1902: particles with longer residence times.  There are four phases in this
1903: plot that correspond to very short residence times, steady growth in
1904: the SASI, an increase in the vigor of the SASI, and the initial
1905: stages of explosion. See \S \ref{section:timescales} for a discussion
1906: and Figs. \ref{stills1}, \ref{stills2}, and \ref{stills3} for entropy maps.
1907: \label{taures_mean_frac}}
1908: \end{figure}
1909: 
1910: \clearpage
1911: 
1912: \begin{figure}
1913: \epsscale{0.6}
1914: \plotone{f9.ps}
1915: \caption{The distribution of residence times,
1916: $\tau_r$, for four generations of tracer particles corresponding to
1917: the four phases in Fig. \ref{taures_mean_frac}.
1918:   The times
1919: shown in the legend correspond to the starting time for each
1920: generation, and $R_{\rm start}$ indicates the starting radius for all
1921: 50,000 particles of each generation.  $R_{\rm start} = 400$ km is
1922: situated well outside the shock radius, and $R_{\rm start} = 150$ km is approximately the middle
1923: of the gain region.  The top panel shows that most of the mass that
1924: accretes through the shock traverses the gain region on relatively
1925: short timescales, $\sim$40 ms, while the bottom panel shows that
1926: in addition to rapidly accreting plumes, there are regions in
1927: the gain region in which mass has a large residence time and is
1928: subjected to prolonged heating.  
1929: \label{taures}}
1930: \end{figure}
1931: 
1932: \clearpage
1933: 
1934: \begin{figure}
1935: \epsscale{0.6}
1936: \plotone{f10.ps}
1937: \caption{Colormaps of entropy ($k_{\rm B}$/baryon) for 15.0-2D3,
1938:   $L_{\nu_e} = 1.9$ and for the four phases
1939:   shown in Fig. \ref{taures}.  Each panel is 400 km $\times$ 800 km.
1940:   These snapshots of the flow correspond to the times in
1941:   Fig. \ref{taures}, which label each histogram by the start time of
1942:   the four generations of particles shown.  The four phases are:  1)
1943:   there is very little convection or SASI and the particles accrete quickly ($\sim$9 ms) onto the
1944:   PNS (top left), 2) the SASI undergoes a steady growth and
1945:   the particles initiated at 400 km (200 km) have a mean residence time
1946:   of 22 ms (32 ms) (top right), 3) the SASI is becoming quite vigorous and the
1947:   residence times for both sets of particles are steadily increasing
1948:   (bottom left), and 4) The SASI is very vigorous and 150 ms later
1949:   (see Fig. \ref{stills3})
1950:   the model undergoes explosion (bottom right). \label{stills1}}
1951: \end{figure}
1952: 
1953: \clearpage
1954: 
1955: \begin{figure}
1956: \epsscale{0.6}
1957: \plotone{f11.ps}
1958: \caption{The same as Fig. \ref{stills1}, except the flow is shown
1959:   75 ms after the start of each generation of particles, which is the
1960:   midpoint of particle tracking.  Low entropy plumes that quickly
1961:   transport newly-accreted material to the PNS are quite pronounced in
1962:   the bottom-right panel. \label{stills2}}
1963: \end{figure}
1964: 
1965: \clearpage
1966: 
1967: \begin{figure}
1968: \epsscale{0.6}
1969: \plotone{f12.ps}
1970: \caption{The same as Fig. \ref{stills1}, except the flow is shown
1971:   150 ms after the start of each generation of particles, which is the
1972:   end of particle tracking. Note that the simulation is exploding in
1973:   the bottom-right panel.\label{stills3}}
1974: \end{figure}
1975: 
1976: \clearpage
1977: 
1978: \begin{figure}
1979: \epsscale{0.775}
1980: \plotone{f13.ps}
1981: \caption{Entropy vs. radius for the times shown in
1982:   Fig. \ref{stills1}.  Solid red lines are entropy profiles for the
1983:   15.0-1D, $L_{\nu_e} = 1.9$ simulation, and points show the entropy
1984:   and radius for each zone of the 15.0-2D3, $L_{\nu_e} =
1985:   1.9$ model.  These entropy profiles are coeval with the residence
1986:   time distributions in Fig. \ref{taures} and the
1987:   entropy maps in Fig. \ref{stills1}.  Features of the entropy
1988:   profiles highlight the shock ($\sim$150 km for $t
1989:   = 0.130$ s and $\sim$200 km for all other times), the gain region,
1990:   ($\sim$100 km to $\sim$ 200 km), and the cooling
1991:   region ($\sim$50 km to $\sim$100 km).  1D profiles show a negative entropy
1992:   gradient in the gain region that results in convection in 2D
1993:   simulations.  When the SASI dominates the post-shock flow, $t=
1994:   0.370$, 0.530, and 0.600 s, the shock is distinctly aspherical, and a
1995:   large range of entropies characterize the gain region.  Low
1996:   entropies correspond to plumes that funnel matter onto the PNS,
1997:   while high entropies correspond to regions of
1998:   long residence times and more integrated net heating.  This is the
1999:   likely cause for 2D simulations having lower critical luminosities
2000:   than 1D simulations.
2001:   \label{entropy}}
2002: \end{figure}
2003: 
2004: \clearpage
2005: 
2006: \begin{figure}
2007: \epsscale{0.7}
2008: \plotone{f14.ps}
2009: \caption{Time evolution of the net heating rate, $\dot{Q}$, for
2010: the 15.0-1D and 15.0-2D3 sequences.
2011: The top panel plots $\dot{Q}$ for the 1D sequence only for the range of $L_{\nu_e}$
2012: near the 1D critical luminosity.  The bottom panel highlights the range
2013: of $L_{\nu_e}$ near the critical luminosity of the 2D sequence, and
2014: plots both 1D and 2D simulations.  The dots show the $\dot{Q}$ at the time of
2015: explosion for the 2D simulations.  $\dot{Q}$ shows pronounced
2016: oscillations for the 1D simulations near explosion (top panel), with the
2017: peaks corresponding to dips in shock radius in Fig. \ref{rvst_1d}.
2018: For non-exploding 1D and 2D simulations
2019: $\dot{Q}$ evolves downward.
2020: Exploding 2D simulations, however, have $\dot{Q}$ that tends toward
2021: flat or upward evolutions.
2022: \label{qdotvstime}}
2023: \end{figure}
2024: 
2025: \clearpage
2026: 
2027: \begin{figure}
2028: \epsscale{0.8}
2029: \plotone{f15.ps}
2030: \caption{Similar to Fig. \ref{qdotvstime}, but here we plot the heating
2031:   efficiency, $\dot{Q}/L_{\nu_e \bar{\nu}_e}$, vs. postbounce time.  The
2032: efficiencies at $t_{\rm exp}$ for the 1D sequence near the critical
2033: luminosity range from 3\% to
2034: 10\%, while the efficiencies of the 2D simulations at $t_{\rm exp}$ are
2035: 3.4\%, 3.5\%, and 4.4\% for $L_{\nu_e} = 1.8$, 1.9, and 2.0,
2036: respectively.  For non-exploding 1D and 2D simulations
2037: $\dot{Q}/L_{\nu_e \bar{\nu}_e}$ evolves downward.
2038: Exploding 2D simulations, however, have efficiencies that tend to stay
2039: flat or increase.  \label{effvstime}}
2040: \end{figure}
2041: 
2042: \clearpage
2043: 
2044: \begin{figure}
2045: \epsscale{0.8}
2046: \plotone{f16.ps}
2047: \caption{Similar to Fig. \ref{qdotvstime}, but here we plot the mass in
2048:   the gain region, $M_{gain}$, vs. postbounce time.  The oscillations
2049:   of $M_{\rm gain}$ in the 1D models (top panel) correspond to the
2050:   oscillations in shock radius.  The minima in $M_{\rm gain }$ correspond to the minima
2051: in shock radius.  In the exploding models of 15.0-2D3, there is a
2052: secular increase in the amount of mass in the gain region leading up
2053: to and past explosion that is distinct from the non-exploding 1D and
2054: 2D models. \label{mass_gain}}
2055: \end{figure}
2056: 
2057: \clearpage
2058: 
2059: \begin{figure}
2060: \plotone{f17.ps}
2061: \caption{The electron- and anti-electron-neutrino luminosity, $L_{\nu_e \bar{\nu}_e}$ ,
2062:   vs. mass accretion rate, $\dot{M}$, of explosion.  The accretion rate is defined just
2063:   exterior to the shock at $t_{\rm exp}$.  Because the time of
2064:   explosion is difficult to define, we also show the range of
2065:   accretion rates 50 ms before and after $t_{\rm exp}$
2066:   as error bars.  In general, 1D simulations are represented by
2067:   orange, 2D simulations by green hues, and 2D-90$^{\circ}$ runs by blue
2068:   hues.  The points that lie in the $\dot{M}$ range from $\sim$0.8 to
2069:   $\sim$1.7 M$_{\sun}$/s correspond to simulations using the
2070:   11.2-M$_{\sun}$ progenitor, and the points in the range from $\sim$0.2 to
2071:   $\sim$3.5 M$_{\sun}$/s correspond to 15 M$_{\sun}$. The three resolutions are represented by stars (1),
2072:   upside-down triangles (2), and squares (3).  The fit for $L_{\nu_e \bar{\nu}_e}$ as
2073:   a function of $\dot{M}$ from \citet{burrows93} is plotted (solid
2074:   line).  While this analytic expression passes through the 15.0-1D
2075:   points it over predicts the critical luminosities for the
2076:   11.2-1D runs by $\sim$15\%.  These results show that time-dependent
2077:   1D and 2D simulations reproduce the critical $L_{\nu_e}$ and
2078:   $\dot{M}$ condition, the critical luminosity is indeed a function of
2079:   $\dot{M}$, and the critical luminosity for 2D
2080:   simulations is $\sim$70\% the critical luminosity for 1D calculations.  \label{lumvsmdot}}
2081: \end{figure}
2082: 
2083: \end{document}
2084: 
2085: