0804.0239/ms.tex
1: \documentclass[10pt]{emulateapj}
2: \usepackage{apjfonts}
3: \usepackage{natbib}
4: \usepackage{amsfonts}
5: \usepackage{amsmath}
6: \usepackage[usenames]{color}
7: \citestyle{apj}
8: \bibliographystyle{apj}
9: 
10: 
11: \newcommand{\p}{\partial}    
12: \newcommand{\pr}{^\prime} 
13: 
14: \newcommand{\pav}{$\bar{\mathrm{P}}_{\rho\mathrm{12}}$\ }
15: \newcommand{\pavm}{$\bar{\mathrm{P}}_{\mathrm{M}}$\ }
16: \newcommand{\rtwelve}{10$^{12}$ g cm$^{-3}$}
17: 
18: \def\fnu{{\cal F}_\nu}
19: \def\etal{{et~al. \,}}
20: \def\sun{_\odot}
21: \def\mo{M$_\odot$}
22: \def\msun{M$_\odot$}
23: \def\mmsun{\mathrm{M}_\odot}
24: \def\mmsunm{$\mathrm{M}_\odot$}
25: \def\lo{L$_\odot\,$}
26: \def\teff{T$_{\rm eff}\,$}
27: \def\teffs{T$_{\rm eff}$s$\ $}
28: \def\Dwa{$\,$\uppercase\expandafter{\romannumeral5}$\,$}
29: \def\mic{$\mu$m$\,$}
30: \def\undertext#1{$\underline{\smash{\hbox{#1}}}$}
31: \def\sless{\lower2pt\hbox{$\buildrel {\scriptstyle <}
32:    \over {\scriptstyle\sim}$}}
33: \def\under{\undertext}
34: \def\sgreat{\lower2pt\hbox{$\buildrel {\scriptstyle >}
35:    \over {\scriptstyle\sim}$}}
36: \def\sharpnull#1{}
37: \def\aa{Astron. Astrophys.\ }
38: 
39: %\null\voffset=-4.0pc  %  +0.0pc
40: %\null\voffset=-1.0pc  %  +0.0pc
41: 
42: \setlength{\parskip}{5pt plus 1pt minus 1pt}    
43: \newcommand{\todo}[1]{{\color{blue}$\blacksquare$~\textsf{[TODO: #1]}}}
44: \newcommand{\todored}[1]{{\color{red}$\blacksquare$~\textsf{[TODO: #1]}}}
45: \newcommand{\sn}{$\mathrm{S}_n$}
46: 
47: %%% some hacks
48: %   Parameters for FLOAT pages (not text pages):
49: \renewcommand{\floatpagefraction}{0.7}	% require fuller float pages
50: % N.B.: floatpagefraction MUST be less than topfraction !!
51: \renewcommand{\dblfloatpagefraction}{0.7}	% require fuller float pages
52: \setcounter{topnumber}{2}
53: \setcounter{bottomnumber}{2}
54: \setcounter{totalnumber}{4}     % 2 may work better
55: \setcounter{dbltopnumber}{2}    % for 2-column pages
56: \renewcommand{\dbltopfraction}{0.9}	% fit big float above 2-col. text
57: \renewcommand{\textfraction}{0.07}	% allow minimal text w. figs
58: 
59: % make the argument stand out visually
60: %\newcommand{\todo}[1]{$\bullet\bullet\bullet$\textbf{[#1]}}
61: %\voffset=.0cm
62: 
63: \begin{document}
64: %-----------------------------------------------------
65: \slugcomment{Submitted to ApJ. March 31, 2008, accepted June 24, 2008.}
66: 
67: \title{2D Multi-Angle, Multi-Group Neutrino Radiation-Hydrodynamic 
68: Simulations of Postbounce Supernova Cores}
69: 
70: \author{Christian D. Ott\altaffilmark{1},
71: Adam Burrows\altaffilmark{2,1}, Luc Dessart\altaffilmark{2,1},
72: and Eli Livne\altaffilmark{3}}
73: \altaffiltext{1}{Department of Astronomy and Steward Observatory, 
74: The University of Arizona, 933 N. Cherry Ave., 
75: Tucson, AZ 85721; cott@as.arizona.edu, 
76: luc@as.arizona.edu}
77: \altaffiltext{2}{Department of Astrophysical Sciences,
78: Princeton University, Peyton Hall, Ivy Lane,
79: Princeton, NJ 08544, burrows@astro.princeton.edu} 
80: \altaffiltext{3}{Racah Institute of Physics, The Hebrew University,
81: Jerusalem, Israel; eli@phys.huji.ac.il}
82: 
83: \begin{abstract}
84: We perform axisymmetric (2D) multi-angle, multi-group neutrino
85: radiation-hydrodynamic calculations of the postbounce phase of
86: core-collapse supernovae using a genuinely 2D discrete-ordinate (\sn)
87: method. We follow the long-term postbounce evolution of the cores of
88: one nonrotating and one rapidly-rotating 20-M$_\odot$ stellar model
89: for $\sim$400 milliseconds from 160~ms to $\sim$550~ms after bounce.
90: We present a multi-D analysis of the multi-angle neutrino radiation
91: fields and compare in detail with counterpart simulations carried out
92: in the 2D multi-group flux-limited diffusion (MGFLD) approximation to
93: neutrino transport. We find that 2D multi-angle transport is superior
94: in capturing the global and local radiation-field variations
95: associated with rotation-induced and SASI-induced aspherical
96: hydrodynamic configurations. In the rotating model, multi-angle
97: transport predicts much larger asymptotic neutrino flux asymmetries
98: with pole to equator ratios of up to $\sim$2.5, while MGFLD tends to
99: sphericize the radiation fields already in the optically
100: semi-transparent postshock regions. Along the poles, the multi-angle
101: calculation predicts a dramatic enhancement of the neutrino heating by
102: up to a factor of 3, which alters the postbounce evolution and results
103: in greater polar shock radii and an earlier onset of the initially
104: rotationally weakened SASI. In the nonrotating model, differences
105: between multi-angle and MGFLD calculations remain small at early times
106: when the postshock region does not depart significantly from spherical
107: symmetry. At later times, however, the growing SASI leads to
108: large-scale asymmetries and the multi-angle calculation predicts up to
109: 30\% higher average integral neutrino energy deposition rates than
110: MGFLD.
111: \end{abstract}
112: 
113: \keywords{Hydrodynamics, Neutrinos, Radiative Transfer,
114: Stars: Evolution, Stars: Neutron, Stars: 
115: Supernovae: General}
116: 
117: \section{Introduction}
118: \label{section:intro}
119: 
120: Four decades after the first pioneering neutrino
121: radiation-hydrodynamic calculations of stellar collapse
122: (\citealt{colgate:66,arnett:66, leblanc:70,wilson:71}), the details of
123: the core-collapse supernova explosion mechanism remain
124: obscure. However, certain essentials are clear. The collapse of the
125: evolved stellar core to a protoneutron star (PNS) and its evolution to
126: a compact cold neutron star provides a gigantic reservoir of
127: gravitational energy, $\sim$3$\times$10$^{53}$~erg, a mass-energy
128: equivalent of $\sim$0.17~$\mmsun$. Any core-collapse supernova
129: mechanism must tap this energy and convert the fraction needed to
130: match Type-II supernova observations ($\sim$10$^{51}$~erg $\equiv$ 1
131: Bethe [B]) into kinetic and internal energy of the exploding stellar
132: envelope.
133: 
134: There is general agreement that the prompt hydrodynamic explosion
135: mechanism does not work and that the bounce shock always stalls,
136: falling short of blowing up the star (e.g.,
137: \citealt{bethe:90,janka:07}), and must be re-energized to lead to a
138: supernova. However, there is no agreement on the detailed mechanism
139: that revives and endows the shock with sufficient energy to make a
140: canonical $\sim$1-Bethe supernova. For decades, the
141: ``neutrino-driven'' mechanism, first proposed in its direct form by
142: \cite{colgate:66}, and in its delayed form by \cite{wilson:85} and
143: \cite{bethewilson:85}, seemed compelling.  It relies on a subtle
144: imbalance of neutrino heating and cooling that leads to a net energy
145: deposition behind the stalled shock, sufficient to revive it and drive
146: the explosion on a timescale of hundreds of milliseconds. While
147: appealing, it has been shown to fail for regular massive stars in
148: spherical symmetry (1D) when the best neutrino physics and transport
149: are used \citep{ramppjanka:00,liebendoerfer:01a,
150: liebendoerfer:05,thompson:03}. Yet, weak explosions may be obtained in
151: 1D for the lowest mass progenitors, O-Ne-Mg cores
152: (\citealt{kitaura:06,burrows:07c}).
153: 
154: It is now almost certain that the canonical explosion mechanism must
155: be multi-dimensional (2D/3D) in nature.  The multi-D dynamics
156: associated with convective overturn in the postshock region (e.g.,
157: \citealt{herant:94, bhf:95,jankamueller:96,buras:06b}) and the
158: recently identified standing accretion shock instability (SASI,
159: e.g., \citealt{foglizzo:00,foglizzo:07,
160: scheck:08,blondin:03,burrows:07a,iwakami:08}) lead to a
161: dwell time of accreting outer core material in the postshock region
162: that is larger on average than in the 1D case. This results in a
163: greater neutrino energy deposition efficiency behind the shock and,
164: thus, creates more favorable conditions for explosion
165: \citep{burrows:93, janka:01,thompson:05,marek:07}.
166: 
167: 
168: The first generation of multi-dimensional supernova calculations,
169: still employing gray flux-limited diffusion (or yet simpler schemes)
170: for neutrino transport, indeed found that neutrino-driven convective
171: overturn in the region between the stalled shock and the PNS
172: sufficiently increased the neutrino energy deposition rate to lead to
173: a delayed explosion (\citealt{herant:94,bhf:95, jankamueller:96,fh:00,
174: fryerwarren:02, fryerwarren:04}). The more sophisticated studies that
175: followed changed this picture. Recent long-term axisymmetric (2D)
176: supernova calculations with multi-group, multi-species neutrino
177: physics and transport find it difficult to explode garden-variety
178: massive stars via the neutrino mechanism. \cite{buras:06b} report
179: explosion only for the low-mass (11.2~$\mmsun$) progenitor of
180: \cite{wwh:02}, while \cite{marek:07} report the onset of explosion in
181: a 15-$\mmsun$ model of \cite{ww:95}, given moderately fast rotation
182: and the use of the Lattimer-Swesty equation of state (EOS;
183: \citealt{lseos:91}) with a nuclear compressibility modulus $K_0$ of
184: 180~MeV, which is significantly softer than the current best
185: experimental values ($K_0 = 240 \pm 20$~MeV; \citealt{shlomo:06}).  On
186: the other hand, \cite{bruenn:06} obtain explosions for 11-$\mmsun$ and
187: 15-$\mmsun$ progenitors from \cite{ww:95} only when they take silicon
188: and oxygen burning into account and due to a synergy between nuclear
189: burning, the SASI, and neutrino heating.
190: 
191: \cite{burrows:06,burrows:07a} do not obtain neutrino-driven explosions
192: (except in the case of O-Ne-Mg cores and accretion-induced collapse;
193: \citealt{dessart:06b}), but observe the excitation of PNS core
194: $g$-modes. In their calculations, the PNS core oscillations reach
195: non-linear amplitudes and damp via the emission of strong sound waves
196: that propagate through the postshock region and efficiently deposit
197: energy into the shock, eventually leading to late explosions at
198: $\sim$1~second after bounce. This \emph{acoustic} mechanism appears to
199: be robust enough to blow up even the most massive and extended
200: progenitors \citep{burrows:07a,ott:06b}, but remains controversial and
201: needs to be confirmed by other groups (see, e.g., 
202: \citealt{yoshida:07,weinberg:08}).
203: 
204: 
205: In the context of rapid progenitor rotation, \cite{burrows:07b},
206: \cite{dessart:08a}, and \cite{dessart:07} (the latter for the
207: accretion-induced collapse scenario) have shown that energetic
208: MHD-driven explosions may be obtained if field-amplification by the
209: magneto-rotational instability \citep{bh:91} is as efficient in the
210: core-collapse context as suggested \citep{akiyama:03}. Whether 
211: rotation alone and without strong magnetic fields favors or disfavors
212: a neutrino-driven explosion remains to be seen \citep{walder:05,
213: dessart:06b,ott:06spin}, but rapid rotation has been shown to damp
214: convection \citep{fh:00} and weaken the SASI \citep{burrows:07b}.
215: 
216: 
217: \subsection{Core-Collapse Supernova Theory and Neutrino
218: Radiation Transport}
219: 
220: Neutrinos, their creation, propagation, and interactions with
221: supernova matter, are of paramount importance to the core-collapse
222: supernova problem. They carry away $\sim$99\% of the final neutron
223: star's gravitational binding energy and $\sim$1\% of this energy would
224: be sufficient to blow up the star. Depending on progenitor
225: characteristics that set the postbounce rate of mass accretion onto
226: the PNS, a successful supernova explosion should occur within
227: $\sim$1--1.5~s after bounce to match observational and theoretical
228: neutron star upper mass limits around $\sim$2--2.5~$\mmsun$
229: (\citealt{lattimer:07}, and references therein).  Consequently, the
230: explosion mechanism must deliver canonical 1-B explosions on this
231: timescale and, if the explosion is neutrino-driven, the neutrino
232: heating efficiency\footnote{We define the heating efficiency as the
233: ratio of the energy deposition rate and the summed electron-neutrino
234: and anti-electron neutrino luminosities.
235: The $\mu$ and $\tau$ neutrinos and their
236: anti-particles do not contribute much to the heating.} must be on the
237: order of 10\% to yield an explosion that achieves an energy of 1~B
238: within $\sim$1~s.
239: 
240: 
241: The neutrinos travelling through the postshock region in a postbounce
242: supernova core are not in thermal equilibrium with the baryonic
243: matter. They should ideally be treated with full kinetic theory,
244: describing the neutrino distributions and their temporal distribution
245: with the Boltzmann equation (\citealt{mihalas:99}). Boltzmann
246: transport is in its most general form a 7-dimensional problem. The 6D
247: neutrino phase space (usually split up into 3D spatial coordinates,
248: neutrino energy, and 2 angular degrees of freedom) 
249: and time. In addition, there are up to 6 neutrino
250: types (3 particle species, and their anti-particles) to deal
251: with. Spherically-symmetric Boltzmann transport schemes have been
252: devised and implemented in the core-collapse context
253: \citep{mezzacappa:93b,messer:98,burrows:00,
254: yamada:99,mezzacappa:99,ramppjanka:02,liebendoerfer:04,hb:07}, but
255: general Boltzmann transport in multiple spatial dimensions is
256: computationally challenging and will remain so in the intermediate
257: term.  Hence, approximations must be made in devising computationally
258: tractable neutrino transport schemes for multi-D simulations.
259: 
260: A highly sophisticated approximation that arguably comes close to full
261: Boltzmann transport in the case of quasi-spherical configurations in
262: 2D is that presented in \cite{buras:06a}, and based on earlier work by
263: \cite{ramppjanka:02}. These authors solve equations for the zeroth and
264: first angular moments of spherically-symmetric radiation fields along
265: multiple radial rays (ray-by-ray approach; \citealt{bhf:95}) and
266: perform a variable Eddington factor closure \citep{mihalas:99} via a
267: single spherically symmetric Boltzmann solution on an averaged 1D
268: profile of the 2D hydrodynamics data.  Neighboring rays are coupled to
269: provide for limited treatment of latitudinal transport. Their
270: multi-group (multi-energy and multi-neutrino species) scheme includes
271: inelastic neutrino-electron scattering, aberration, gravitational
272: redshift, and frame effects to ${O}(\mathrm{v}/c)$.
273: 
274: \cite{livne:04} implemented a genuinely 2D direct solution of a
275: reduced Boltzmann equation via the method of discrete ordinates
276: (S$_n$: see, e.g.,
277: \citealt{yueh:77,mezzacappa:93b,adamslarsen:02,castor:04}, and
278: references therein) in the code VULCAN/2D, neglecting energy
279: redistribution and fluid-velocity dependence.
280: 
281: 
282: A common, more approximate way to handle neutrino transport that has a
283: long pedigree in 1D core-collapse studies is multi-group
284: (energy/neutrino species) non-equilibrium flux-limited diffusion
285: (MGFLD; \citealt{mihalas:99}, \citealt{arnett:66},
286: \citealt{bowerswilson:82}, \citealt{bruenn:85},
287: \citealt{myra:87,myra:90,baron:89,cooperstein:92}). FLD schemes solve a
288: diffusion equation for the mean radiation intensity, the zeroth
289: angular moment of the specific radiation intensity. Hence, they drop
290: all local angular dependence of the radiation field, while, in the
291: MGFLD case, retaining the spectral neutrino distribution. MGFLD
292: accurately describes the radiation field at high optical depth where
293: the diffusion approximation is exact. In the free-streaming limit, the
294: flux must be limited to maintain causality and an interpolation must
295: be performed between diffusion and free-streaming regimes by an
296: ad-hoc prescription (using a flux limiter).
297: 
298: 2D FLD schemes were pioneered in the core-collapse context by
299: \cite{leblanc:70} and modern MGFLD implementations can be found in
300: \cite{swesty:06} and in \cite{burrows:07a}.  It is not a priori clear
301: whether MGFLD is an accurate enough prescription to yield postbounce
302: supernova dynamics in qualitative and quantitative agreement with a
303: more accurate multi-angle treatment.
304: Since net energy deposition by neutrinos is favored only in the
305: semi-transparent gain layer, the quality of a MGFLD scheme may
306: sensitively depend on the flux limiter chosen \citep{burrows:00}.
307: The fact that 2D gray FLD schemes have in the past led to
308: neutrino-driven explosions
309: \citep{herant:94,bhf:95,fh:00,fryerwarren:02,fryerwarren:04}, while
310: MGFLD schemes appear not to \citep{walder:05,burrows:06,burrows:07a},
311: emphasizes the importance of a spectral treatment of neutrino
312: transport.
313: 
314: In 1D, MGFLD and Boltzmann neutrino transport were compared on static
315: hydrodynamic postbounce backgrounds by \cite{janka:92},
316: \cite{yamada:99}, \cite{messer:98}, and \cite{burrows:00}.  Also in
317: 1D, \cite{mezzacappa:93a} compared Boltzmann transport and MGFLD
318: evolutions in the collapse phase, while \cite{liebendoerfer:04}
319: performed the only comparison to date of 1D long-term Boltzmann and MGFLD
320: supernova evolutions.  The static studies all agree that
321: Boltzmann transport yields larger instantaneous neutrino heating rates
322: in the gain region, mostly because of a more slowly decreasing inverse
323: flux factor ($c$ over the ratio of flux to neutrino energy density),
324: a quantity that can be related to the rate of energy absorption. On the other
325: hand, \cite{liebendoerfer:04} find no significant dynamical
326: differences between MGFLD and Boltzmann transport evolutions in their
327: long-term comparison study with the 13-$\mmsun$ progenitor
328: model of~\cite{nomoto:88}.
329: 
330: 
331: In this paper, we present 2D \emph{multi-angle}, multi-group neutrino
332: transport supernova calculations using the Newtonian axisymmetric
333: VULCAN/2D code \citep{livne:93,livne:04,burrows:07a}.  Comparing
334: multi-D Boltzmann and MGFLD treatments, we perform postbounce
335: simulations with VULCAN/2D and compare 2D steady-state snapshots, as
336: well as fully-coupled dynamical 2D radiation-hydrodynamics evolutions,
337: for non- and rapidly-rotating 20-$\mmsun$ models whose precollapse
338: profiles are taken from \cite{wwh:02}.  We analyze our angle-dependent
339: neutrino radiation fields and provide for the first time local 2D map
340: projections of the specific intensity $I_\nu$.
341: 
342: In \S\ref{section:methods}, we describe our hydrodynamic and
343: radiation-transport schemes and the microphysics that we use in this
344: postbounce core-collapse supernova study. In \S
345: \ref{section:initialmodels}, we introduce the presupernova models and
346: the postbounce configurations, the setup, and the methodology of our
347: Boltzmann-transport--MGFLD comparisons. In \S\ref{section:snapshots},
348: we present results of snapshot Boltzmann transport calculations and
349: compare them with their MGFLD counterparts. In
350: \S\ref{section:evolution}, we then discuss time-dependent calculations,
351: the dynamical differences between Boltzmann transport and MGFLD
352: runs, and the consequences for postbounce supernova model
353: evolution. We wrap up in
354: \S\ref{section:summary} with a summary and critical discussion of the
355: work presented in this paper.
356: 
357: \section{Methods}
358: \label{section:methods}
359: 
360: \subsection{Hydrodynamics}
361: 
362: 
363: We employ the arbitrary Lagrangian-Eulerian (ALE, with second-order
364: total-variation-diminishing [TVD] remap) radiation-hydrodynamics code
365: VULCAN/2D. The hydrodynamics module was first described by
366: \citet{livne:93}\footnote{For details and an extension to
367: magneto-hydrodynamics not employed here, see \cite{livne:07}.}.
368: The 2D time-explicit hydrodynamics scheme is second-order accurate (in
369: smooth parts of the flow), unsplit, and implements a finite-difference
370: representation of the Newtonian Euler equations with artificial
371: viscosity on arbitrarily structured grids and in cylindrical
372: coordinates.  The computational grid employed here is set up to
373: resemble a spherical-polar grid at radii greater than 20~km and
374: gradually transitions to a Cartesian structure at smaller radii
375: \citep{ott:04}.  This (a) avoids hydrodynamic timestep restrictions
376: due to focussing of angular grid lines and (b) liberates the PNS core,
377: thus allowing mass motion along the axis of symmetry.
378: 
379: Self-gravity is implemented via direct grid-based solution
380: of the Newtonian Poisson equation, as described in
381: \citet{burrows:07a}, and we employ the finite-temperature
382: nuclear equation of state of \citet{shen:98a,shen:98b}.
383: The calculations are run with 230 logarithmically-spaced 
384: radial and 120 angular zones (including the
385: inner, quasi-Cartesian region). The grid encompasses
386: a radial extent of 4000~km and the full 180$^\circ$
387: of the axisymmetric domain.
388: 
389: 
390: \subsection{Neutrino Transport and Microphysics}
391: 
392: VULCAN/2D contains two multi-group, multi-species neutrino
393: radiation-transport options.  As we discuss below
394: (\S\ref{section:initialmodels}), both modules are used in this study.
395: The module implementing 2D transport in the MGFLD approximation, 
396: evolving the zeroth moment of the radiation
397: field, is discussed in \cite{burrows:07a}.  The angle-dependent
398: transport module that evolves the specific neutrino radiation
399: intensity,
400: $I(\mathbf{r},\mathbf{\Omega},\varepsilon_\nu,\mathrm{species},t)$, via
401: the method of discrete ordinates (S$_n$), was first discussed by
402: \citet{livne:04} (see also \citealt{morel:96,adamslarsen:02,castor:04}).
403: 
404: 
405: For convenience and future reference,
406: we define the zeroth, first, and second moments of the
407: radiation field,
408: \begin{eqnarray}
409: J_{\nu} &\equiv& \frac{1}{4\pi}
410: \oint_{4\pi} d\Omega\,\, I_{\nu}\,,\label{eq:J}\\ \vec{H}_{\nu}
411: &\equiv& \frac{1}{4\pi} \oint_{4\pi} d\Omega\,\, \vec{n}\,
412: I_{\nu}\,,\label{eq:H}\\ {\textsf
413: K}_{\nu} &\equiv& \frac{1}{4\pi} 
414: \oint_{4\pi} d\Omega\,\, \vec{n}\,\vec{n}\,I_{\nu}\,\label{eq:eddy}.
415: \end{eqnarray}
416: Note the vector and tensor natures of $\vec{H}_\nu$ and
417: ${\textsf K}_\nu$, respectively. $\vec{n}$ is the
418: radiation field unit vector whose coordinate-dependent
419: components are given in \cite{hb:07} for various common
420: coordinate systems. Here we employ cylindrical coordinates
421: (see Fig.~\ref{fig:coord}). The radiation-pressure tensor
422: ${\textsf K_\nu}$ obeys the trace condition $J_\nu =
423: \mathrm{Tr}[K_\nu]$~\citep{mihalas:99}. The spectral
424: neutrino flux is defined as $\vec{F}_\nu = 4\pi \vec{H}_\nu$.
425: 
426: 
427: As explained in \citet{livne:04}, the
428: time-implicit S$_n$ solver in VULCAN/2D updates the
429: specific intensity in the laboratory
430: frame via the Boltzmann transport equation
431: (\citealt{castor:72}) without
432: fluid-velocity dependence,
433: \begin{equation}
434: \label{eq:trans_master}
435: \frac{1}{c}\frac{\partial I}{\partial t} + \vec{n}\cdot \vec{\nabla} I
436: + \sigma I = S\,\,, 
437: \end{equation}
438: where we have dropped the neutrino group index $\nu$.
439:  $\sigma = \sigma^a + \sigma^s$, where
440: $\sigma^a(\mathbf{r},\varepsilon_\nu, \mathrm{species})$ 
441: is the inverse absorption mean-free path and 
442: $\sigma^s(\mathbf{r},\varepsilon_\nu,\mathrm{species})$
443: is the inverse scattering mean-free path (both equivalent to the
444: corresponding cross section multiplied by the number density).
445:  We assume scattering to be isotropic and
446: employ the transport cross section $\sigma^s = (1 - \langle \cos
447: \vartheta\rangle)\sigma^s_T$ instead of the total scattering cross
448: section $\sigma^s_T$. This approach has been shown to work well in
449: spherically-symmetric core-collapse supernova calculations
450: \citep{burrows:00,thompson:03}.  The right-hand side source term $S$
451: equals $S_\mathrm{em}(\mathbf{r},\varepsilon_\nu,\mathrm{species}) +
452: \sigma^s J$, where $S_\mathrm{em}$ is the emissivity. The transport
453: grid is identical to the hydrodynamics grid.  The specific intensity
454: and its moments are defined at cell centers, facilitating
455: spatially-consistent coupling with the scalar hydrodynamics variables,
456: as discussed in \citet{livne:04}.  Radiation stress at cell corners is
457: computed via linear interpolation employing cell-centered values of
458: the radiation flux.
459: 
460: As a consequence of the neglect of $O(\mathrm{v}/c)$ terms in our transport
461: formulation, neutrino advection, Doppler shifts and aberration effects
462: are not considered. This greatly limits the computational complexity
463: of the problem, but its impact on the transport solution depends on
464: the particular choice of reference frame and was examined in
465: \cite{hb:07}. It is clear that around core bounce and neutrino
466: breakout, during the non-linear phase of the SASI hundreds of
467: milliseconds after bounce, and in the case of rapid rotation,
468: including $O(\mathrm{v}/c)$ terms is advisable. We leave them out here in order
469: to make long-term multi-angle radiation-hydrodynamics simulations
470: feasible and allow direct comparison with the MGFLD variant of
471: VULCAN/2D. Full $O(\mathrm{v}/c)$ Boltzmann transport with energy
472: redistribution will be addressed using the code BETHE currently under
473: development by a subset of our group (\citealt{hb:07,murphy:08}).
474: 
475: \begin{figure}
476: \centering
477: \includegraphics[width=7cm]{f1.eps}
478: \caption{Coordinates used in the axisymmetric
479: \sn\ transport scheme implemented in VULCAN/2D.  The radiation
480: direction vector $\vec{n}$ is defined in terms of $\vartheta$ and
481: $\varphi$. $\vartheta$ is the angle with respect to the
482: coordinate-grid $z$-axis at all spatial positions ($z$,$\varpi$).  At each
483: ($z$,$\varpi$), the local momentum-space unit sphere is covered by $n$
484: zones in $\vartheta$ and at each $\vartheta$ location by a number
485: $m(\vartheta)$ of $\varphi$-zones, so that each zone in
486: $(\vartheta,\varphi)$ covers roughly the same solid angle.
487: \label{fig:coord}} 
488: \end{figure}
489: 
490: 
491: We discretize the angular radiation distribution
492: evenly in $\cos{\vartheta}$ from -1 to 1 and in
493: $\varphi$ evenly from 0 to $\pi$ (treating only
494: one hemisphere because of axial symmetry).
495: We make the number of $\varphi$-bins a function of
496: $\cos{\vartheta}$ to tile the hemisphere more
497: or less uniformly in solid angle. 
498: In our time-dependent runs we employ
499: 8 $\cos \vartheta$ bins, resulting in a total
500: of 40 angular zones. Steady-state radiation fields
501: are computed with 8 $\cos \vartheta$ bins, 
502: 12 $\cos \vartheta$ bins (92 total angular zones) 
503: and 16 $\cos \vartheta$ bins (162 total angular zones)
504: at each spatial grid point.
505: 
506: 
507: The standard set of neutrino-matter interactions
508: listed in \cite{thompson:03} is included and all
509: computations are performed with 16 discrete neutrino 
510: energy bins, approximately 
511: logarithmically spaced from 2.5~MeV to 220~MeV.
512: Electron neutrinos ($\nu_e$) and electron-antineutrinos
513: ($\bar{\nu}_e$) are treated independently while
514: we lump together the heavy-lepton $\mu$, $\bar{\mu}$, $\tau$, and 
515: $\bar{\tau}$ neutrinos into one 
516: group (``$\nu_{\mu}$'').
517: The code is very efficiently parallelized via MPI 
518: in energy groups and species.
519: As an additional simplification, we do not include 
520: energy redistribution by inelastic neutrino-electron
521: scattering. Such energy redistribution and
522: scattering are of modest ($\sim 10\%$) relevance
523: for the trapped electron fraction ($Y_e$) and
524: entropy of the inner core at core bounce, but otherwise
525: arguably quite subdominant \citep{thompson:03}.
526: 
527: 
528: \subsection{A Hybrid Approach -- Combining
529: {\rm S}$_n$ and MGFLD Neutrino Transport}
530: 
531: The time-implicit S$_n$ scheme in VULCAN/2D
532: is iterative and suffers convergence problems in
533: regions where the transport problem is 
534: scattering-dominated and the optical depth
535: is high $(\tau \sgreat 5)$. As a consequence, 
536: \citet{livne:04} limited the timestep
537: at postbounce times to $\sim$0.1--0.3~$\mu$s 
538: to ensure accuracy and stability. In the present
539: study, we take a different approach and
540: introduce a hybrid S$_n$--MGFLD transport scheme
541: that treats the quasi-isotropic transport problem 
542: in the optically-thick PNS interior 
543: in the diffusion approximation and transitions to full
544: multi-angle S$_n$ transport in a region
545: of moderate optical depth ($\tau\,\, \sgreat\, 2$), but
546: that is still significantly interior to the neutrinospheres
547: ($\tau \sim 2/3$) where the neutrinos decouple from
548: matter and begin to stream. 
549: 
550: \begin{figure*}[t]
551: \centering
552: \includegraphics[width=8.8cm]{f2a.ps}
553: \includegraphics[width=8.8cm]{f2b.ps}
554: \caption{Entropy colormaps of the nonrotating model s20.nr (left) and the
555: rotating model s20.$\pi$ (right) at 160~ms into their postbounce evolution
556: computed with MGFLD. Velocity vectors are superposed with vector
557: lengths saturated at 1.0$\times$$10^{9}$~cm~s$^{-1}$. Model s20.nr has
558: a practically spherical PNS and shows features of violent overturn in
559: the convectively unstable postshock region. The shock radius in this
560: model is $\sim$175~km at this point and the onset of the SASI is
561: apparent from the slightly deformed shock. Model s20.$\pi$, on the
562: other hand, has a strongly rotationally-flattened PNS and convective
563: overturn is confined to polar regions. These regions exhibit
564: the globally highest entropies and greatest entropy gradients, since
565: the polar velocity divergence at the shock is the highest. The shock
566: radius at this time in model s20.$\pi$ is $\sim$230~km and no SASI
567: features are visible.
568: \label{fig:intro2D}}
569: \end{figure*}
570: 
571: 
572: We chose a radius of 20~km in our calculations for the transition from
573: MGFLD to S$_n$. This is a sensible choice, (1) because the
574: neutrinosphere radii of all groups (energies/species) remain larger
575: than 20~km throughout the postbounce period our simulations cover and,
576: (2) because 20~km also marks the radius at which the transition from
577: the inner irregular quasi-Cartesian grid to the outer regular grid is
578: complete. This boundary is smooth and the S$_n$--MGFLD transition does
579: not suffer from Cartesian cornerstone effects.
580: 
581: The transition is implemented by setting up for each energy group and
582: species an approximate specific intensity $I_\nu$ at the centers of the
583: zones below the S$_n$--MGFLD interface using the information available
584: from MGFLD. This approximate $I_\nu$ is obtained via its angular expansion
585: to first order in $\vec{n}$ (the Eddington approximation): 
586: \begin{equation}
587: I_\nu = I_0 + 3 (\vec{n}\cdot\vec{H})\,\,.
588: \label{eq:approxI}
589: \end{equation}
590: Here, $I_0 = J_\mathrm{MGFLD}$ and 
591: $\vec{H} = \vec{F}_\mathrm{MGFLD}/4\pi$,
592: where $\vec{F}_\mathrm{MGFLD}$ is the flux, 
593: and $\vec{H}$ is the first moment of
594: $I_\nu$. In MGFLD, $\vec{F}_\mathrm{MGFLD}$ is computed via
595: \begin{equation}
596: \vec{F}_\mathrm{MGFLD} = - \mathrm{FL}[D] \vec{\nabla} J\,\,,
597: \end{equation} where \begin{equation} D = \frac{1}{3\sigma}\,\,, \end{equation}
598: with Bruenn's flux limiter\footnote{We use Bruenn's flux limiter in
599: VULCAN/2D, because \cite{burrows:00} found it to perform best in their
600: comparison of flux limiters with angle-dependent transport.}
601: \citep{bruenn:85}, 
602: \begin{equation} \mathrm{FL}\big[D\big] =
603: \frac{D}{1+D|\vec{\nabla} J|/J}\,\,.  
604: \end{equation} 
605: The first angular moment of eq.~(\ref{eq:approxI}), $\vec{F}_\mathrm{S_n} = \int
606: \vec{n}\, I d\Omega$, is then equal\footnote{Given the limited number
607: of angular zones of $I$ and the fact that we are not using
608: Gaussian-quadrature-type angular zoning, the integrals of $I$ are only
609: accurate to $\sim$5\% when 8 $\vartheta$-zones are used and accurate
610: to $\sim$1\% when 12 and more $\vartheta$-zones are employed. To
611: ensure conservation of energy in the S$_n$--MGFLD matching, we employ
612: purely geometrical and temporally constant correction factors to
613: enforce $\vec{F}_\mathrm{S_n} = \vec{F}_\mathrm{MGFLD}$ at the
614: interface.} to $\vec{F}_\mathrm{MGFLD}$ and the S$_n$--MGFLD matching
615: is consistent and provides a representation of the specific intensity
616: $I$ that is accurate to first order in $\vec{n}$.  Given the
617: essentially isotropic neutrino radiation field deep inside the PNS,
618: this approximation yields excellent results. We note that the scheme
619: makes the implicit assumption that the radial gradient of the mean
620: intensity at the transition radius is always negative or zero. This
621: condition is generally fulfilled in PNSs.
622: 
623: 
624: \section{Initial Models and Setup}
625: \label{section:initialmodels}
626: 
627: We employ the spherically-symmetric solar-metallicity
628: $\mathrm{20-}\mmsun$ (at ZAMS) model s20.0 from the stellar
629: evolutionary study of \cite{wwh:02}, who evolved it to the onset of
630: core collapse. At that moment, its iron core mass\footnote{Determined
631: by the discontinuity in the electron fraction, $Y_e$, at the outer
632: edge of the iron core where $Y_e$$\sim$0.5.} is $\sim$1.46~$\mmsun$
633: and its central density has reached
634: $\sim$8.4$\times$10$^{9}$~g~cm$^{-3}$. A graph of the progenitor's
635: precollapse density stratification as a function of enclosed mass can
636: be found in Fig.~1 of \cite{burrows:07a}. Note that in the study of
637: \cite{wwh:02}, iron core mass and extent vary non-monotonically in the
638: 10--20~$\mmsun$ ZAMS mass range and that their solar-metallicity
639: 20-$\mmsun$ model has, in fact, a more compact central configuration
640: than the corresponding 15-$\mmsun$ model.  Stellar evolution theory of
641: massive stars has yet to converge and studies by different groups do
642: not presently yield the same presupernova structures.
643: 
644: We set up two initial models in VULCAN/2D: s20.nr and s20.$\pi$. Both
645: models are mapped from 1D onto our 2D hydrodynamic grid under
646: the assumption of spherical symmetry. Model s20.nr is kept
647: nonrotating, while we impose an initial angular velocity profile
648: in model s20.$\pi$ according to the rotation law,
649: \begin{equation}
650: \Omega(\varpi) = \Omega_0 \frac{1}{1+(\varpi/A)^2}\,\,,
651: \label{eq:rotlaw}
652: \end{equation} where $\varpi$ is the distance from the rotation axis and $A$ is a
653: parameter governing precollapse differential rotation.  This rotation
654: law enforces constant angular velocity on cylindrical shells and, for
655: sensible choices of $A$, reproduces qualitatively
656: (\citealt{ott:06spin}) predictions from presupernova models that
657: include rotation in a 1D fashion~\citep{heger:00,heger:05}. Since the
658: computational complexity of this study inhibits us from performing a
659: sweep of the $\Omega_0$--$A$ parameter space, we chose $A = 1000$~km
660: and $\Omega_0=\pi$~rad~s$^{-1}$. Hence, the initial central period is
661: 2~s -- this is an identical rotational setup to the fiducial model in
662: \cite{burrows:07b}.  As discussed in \cite{ott:06spin}, 2~s is rather
663: short, leads to a rapidly rotating postbounce configuration with a
664: millisecond-period PNS, and, unless significant postbounce spin-down
665: (e.g. via MHD torques) occurs, is inconsistent with average pulsar
666: birth spin estimates. We chose such rapid rotation simply because we
667: wish to study a postbounce supernova core with significant
668: rotationally-induced asymmetry. Key model parameters and
669: characteristics are summarized in Table~\ref{table:models}.
670: 
671: 
672: We collapse both models with the MGFLD variant of VULCAN/2D and evolve
673: them to $\sim$160~ms after core bounce.
674: Then, we transition to \sn\ Boltzmann transport and solve for
675: the stationary neutrino radiation field based on the artificially
676: frozen hydrodynamics data at this postbounce time. Once we have
677: obtained a converged angle-dependent radiation field, we activate
678: neutrino-matter coupling and hydrodynamics and evolve in time the coupled
679: radiation-hydrodynamics equations. For direct comparison, we also
680: continue the MGFLD simulations to later times. 
681: All steady-state snapshots are computed in three momentum-space 
682: angular resolutions, S$_{16}$, S$_{12}$, and S$_8$, while the
683: long-term evolution calculations could only be performed with
684: S$_8$, due to computational constraints.
685: 
686: In Fig.~\ref{fig:intro2D}, we show entropy colormaps of both models at
687: 160~ms after bounce. Fluid velocity vectors are superposed, providing
688: a snapshot of the flow. By 160~ms after bounce, in the nonrotating
689: model s20.nr convection in the high-entropy ($O$(10)
690: $\mathrm{k}_\mathrm{B}$/baryon) gain layer has developed fully. The
691: shock sits at $\sim$175~km and is slightly deformed by the onset of
692: the SASI. Not visible on the scale of this figure is the
693: lepton-gradient-driven convective region deep inside the PNS,
694: which was extensively  discussed in \cite{dessart:06a}.
695: 
696: The PNS in the rapidly rotating model s20.$\pi$ is rotationally
697: flattened, with unshocked low-entropy inner-core
698: pole/equator asymmetry ratios below $\sim$0.5. The shock
699: is slightly prolate and has attained an average radius of 
700: $\sim$230~km.  The moment-of-inertia-weighted 
701: mean period of the unshocked (specific entropy
702: $s \le$~3~k$_\mathrm{B}$) inner core is $\sim$2.0~ms. Differential
703: rotation between $\sim$20 and 200~km is very large, with the
704: angular velocity  $\Omega$ dropping from $\sim$1600~rad~s$^{-1}$
705: to a mere $\sim$15~rad~s$^{-1}$ over this radial equatorial interval.
706: Yet, the specific angular momentum $j$ is still monotonically and rapidly 
707: increasing. It flattens, but does not decrease, only at radii greater
708: than $\sim$100~km.  This positive gradient in $j$ stabilizes the postbounce
709: core against convective instability at low latitudes \citep{fh:00}, 
710: confining overturn to the polar regions and large equatorial radii where
711: the $j$ gradient is less steep.
712: 
713: 
714: 
715: \begin{deluxetable}{llcllll}
716: \tabletypesize{\scriptsize}
717: \tablecaption{Model Summary\label{table:models}}
718: \tablehead{
719: \colhead{Model Name}&
720: \colhead{Progenitor}&
721: \colhead{$\Omega_0$}&
722: \colhead{$A$}&
723: \colhead{$t_b$}&
724: \colhead{$t_\mathrm{snap}$}&
725: \colhead{$t_f - t_b$}\\
726: \colhead{}&
727: \colhead{}&
728: \colhead{(rad s$^{-1}$)}&
729: \colhead{(km)}&
730: \colhead{(ms)}&
731: \colhead{(ms)}&
732: \colhead{(ms)}
733: }
734: \startdata
735: s20.nr&s20.0&0.0&---&179.2&160.0&500.0\\
736: s20.$\pi$&s20.0&$\pi$&1000&193.7&160.0&550.0
737: \enddata
738: \tablecomments{Summary of model parameters. The progenitors
739: are taken from \cite{wwh:02}. $\Omega_0$ is the initial
740: central angular velocity, $A$ is the differential rotation
741: parameter of the rotation law (eq. \ref{eq:rotlaw}). $t_b$ is
742: the time of core bounce, $t_\mathrm{snap}$ is the time after
743: $t_b$ at which the postbounce snapshots are taken, and $t_f - t_b$
744: is the point at which we stop our simulations.}
745: \end{deluxetable}
746: 
747: 
748: \begin{figure}[t]
749: \centering
750: \includegraphics[width=8.5cm]{f3.ps}
751: \caption{Polar plot of the normalized specific intensity
752: $I_\nu(r,\vartheta,\varphi)/\mathrm{max}[I_\nu(r,\vartheta,\varphi)]$
753: in model s20.nr at 160~ms after core bounce, at selected equatorial
754: radii, and for $\nu_e$ neutrinos at $\varepsilon_\nu = 12.6$~MeV.
755: At each radius, we normalize the specific intensity by its local
756: maximum. Shown is the variation with $\varphi$ at fixed $\vartheta =
757: \pi/2$. The graphs are based on a S$_{16}$ calculation.  At $r
758: =$~30~km, the radiation field is practically isotropic, but is already
759: appreciably forward-peaked at the neutrinosphere ($r_\nu = 55$~km;
760: optical depth $\tau = 2/3$) and thereafter smoothly transitions over
761: $\sim$200--300~km to the free-streaming limit.
762:  \label{fig:s20nr_psi1}}
763: \end{figure}
764: 
765: \vspace*{.8cm}
766: 
767: 
768: \section{Results: Snapshots}
769: \label{section:snapshots}
770: 
771: In this section, we present our \sn\ multi-angle transport results for
772: steady-state model snapshots at 160~ms after core bounce.  We diagnose
773: the angle-dependent neutrino radiation fields and carry out a 
774: comparison between multi-angle and MGFLD transport results based on
775: local and global radiation-field variables.
776: 
777: \begin{figure*}[t!]
778: \centering
779: \includegraphics[width=17cm]{f4.ps}
780: \caption{ Hammer-type interpolated (smoothed) map projections of the
781: normalized specific intensity $I_\nu(\vartheta,\varphi) / J_\nu$ in
782: model s20.nr at 160 ms after bounce. The colormap is logarithmic and in
783: each individual projection is setup to range from
784: $\mathrm{max}(I_\nu(\vartheta,\varphi) / J_\nu)$ (red) to
785: 10$^{-4}\mathrm{max}(I_\nu(\vartheta,\varphi) / J_\nu)$ (black).
786: Shown is the specific intensity of $\nu_e$, $\bar{\nu}_e$, and
787: ``$\nu_\mu$'' neutrinos at $\varepsilon_\nu = 12.6$~MeV (rows) on the
788: equator ($\theta$=90$^\circ$, measured from the pole) and at radii of
789: 60, 120, and 240 km (columns). The Hammer projection is set up in such
790: a way that $\vartheta$ varies in the vertical from $0^\circ$ (top) to
791: $180^\circ$ (bottom) and $\varphi$ varies horizontally from
792: $-180^\circ$ (left) to $+180^\circ$ (right). Grid lines are drawn in
793: $\vartheta$- and $\varphi$-intervals of 30$^\circ$. Note (a) the
794: increasing forward-peaking of $I_\nu$ with increasing radius (and
795: decreasing optical depth) and (b) that at any given radius $I_\nu$ of
796: ``$\nu_\mu$'' is more forward-peaked than that of the $\bar{\nu}_e$
797: component, which, in turn, is always more forward-peaked than the
798: $\nu_e$ component. This fact is a consequence of a transport mean-free
799: path that varies with species (and energy; not shown here) and is
800: smallest for the electron neutrinos.
801: \label{fig:hammer1}}
802: \end{figure*}
803: 
804: 
805: \subsection{Angular Distributions}
806: 
807: The quintessential problem in treating neutrino radiation transport in
808: core-collapse supernova cores is the fact that the neutrino transport
809: mean-free path, the average distance a neutrino can travel without
810: experiencing scattering or absorption, changes by orders of magnitude
811: from inside to outside.  Moreover, the neutrino transport
812: mean-free-path $\lambda_\nu$ varies locally strongly with neutrino
813: energy ($\propto \varepsilon_\nu^2$) and matter
814: density.  As a consequence, gray transport schemes are problematic,
815: since neutrino-energy averages can be defined only locally and the
816: mean neutrino energy varies significantly throughout the supernova
817: core.
818: 
819: From a more geometric point of view, the radiation field in momentum
820: space goes from being completely isotropic (net flux $\sim$zero) to
821: being focussed into the radial direction (``forward-peaked'') in the
822: free-streaming regime. In the MGFLD approximation, the mean intensity
823: $J_\nu$ is evolved in time and the angular information, in particular
824: the information on the degree of forward-peaking, is captured only
825: by computing spatial gradients in $J_\nu$ and employing a flux
826: limiter to interpolate between diffusion and free streaming.
827: 
828: 
829: \begin{figure*}[t!]
830: \centering
831: \includegraphics[width=10cm]{f5.ps}
832: \caption{ Hammer map projections of the interpolated (smoothed)
833: normalized specific intensity $I_\nu(\vartheta,\varphi) / J_\nu$ at
834: 160~ms postbounce in model s20.nr. The projections are set up in
835: identical fashion to Fig.~\ref{fig:hammer1}. Shown here is the
836: variation of the angular distribution with energy group (columns) and
837: angular position (rows) for electron neutrinos. The radius is fixed to
838: 150~km.  As expected in the coordinates used for the S$_n$ transport
839: in VULCAN/2D (see Fig.~\ref{fig:coord}), $I_\nu$ becomes
840: forward-peaked into $\vartheta=0^\circ$ and degenerate in $\varphi$
841: along the pole ($\theta=0^\circ$), forward-peaked into
842: $\vartheta=45^\circ,\varphi=0^\circ$ on the diagonal
843: ($\theta=45^\circ$), and forward-peaked
844: $\vartheta=90^\circ,\varphi=0^\circ$ on the equator
845: ($\theta=90^\circ$). The degree of the radiation anisotropy and its
846: variation from forward-peaked at $\varepsilon_\nu = 12.6$~MeV to less
847: forward-peaked at $\varepsilon_\nu = 37.5$~MeV is apparent.
848: \label{fig:hammer2}} 
849: \end{figure*}
850: 
851: 
852: The \sn\ Boltzmann solver in VULCAN/2D is able to 
853: self-consistently handle the transition from isotropic to
854: forward-peaked radiation.  Figure~\ref{fig:s20nr_psi1} depicts the
855: angular distribution in the azimuthal angle $\varphi$ (see
856: Fig.~\ref{fig:coord}) of the normalized specific spectral neutrino
857: intensity $I_\nu$ for electron neutrinos at 12.6~MeV. In
858: Fig.~\ref{fig:s20nr_psi1}, the polar angle $\vartheta$ is set equal to
859: $\pi/2$ and the $\varphi$-distribution is given at various radii in
860: the equatorial plane of model s20.nr.  At 30~km from the center, the
861: radiation field of $\nu_e$'s at $\varepsilon_\nu$~=~12.6~MeV is nearly
862: isotropic,which corresponds to a circle in
863: Fig.~\ref{fig:s20nr_psi1}. With increasing radius (and, of course,
864: decreasing matter density) the transport mean-free path at fixed
865: $\varepsilon_\nu$ increases and the radiation field gradually departs
866: from isotropy and becomes more and more forward-peaked. We define the
867: neutrinosphere as the surface at which the optical depth $\tau_\nu$,
868: given by \begin{equation} \tau_\nu = \int_\infty^R \frac{dr}{\lambda_\nu}\,\,, \end{equation}
869: is equal to 2/3. At around this $\tau_\nu$, the neutrinos decouple
870: from matter and begin to stream freely. At 160~ms after bounce in
871: model s20.nr, the 12.6~MeV $\nu_e$ neutrinosphere is located at
872: $r\sim$55~km. As is obvious from Fig.~\ref{fig:s20nr_psi1}, the
873: radiation field at the neutrinosphere is not yet dramatically
874: forward-peaked, but becomes so with increasing radius. However,
875: complete forward-peaking only obtains at radii $\sgreat$~250--300~km,
876: beyond which the angular resolution of our \sn\ scheme becomes
877: suboptimal, even with $n=16$. However, calculations with varying
878: number of $\vartheta$ (and, hence, $\varphi$) angles reveal that the
879: transition from isotropy to moderate and large anisotropy is
880: adequately reproduced at small and intermediate radii (out to
881: $\sim$200~km) even in the case of S$_8$.
882: 
883: For the purpose of displaying and studying the local neutrino
884: radiation field, we provide equal-area Hammer-type map projections
885: \citep{hammer:1892}.  Such map projections are new to the field of
886: neutrino radiation transport and beautifully reveal the multi-D
887: angular-dependence of the radiation field.  In Fig.~\ref{fig:hammer1},
888: we present such Hammer projections on the equator (spatial $\theta =
889: 90^\circ$) of model s20.nr at radii of 60, 120, and 240~km for the
890: three neutrino species included in our simulations at $\varepsilon_\nu
891: = 12.6$~MeV . In each plot, we normalize the specific intensity to the
892: mean intensity to set a common scale. The colormap is logarithmic and
893: chosen to have regions on the sphere with high intensity appear red
894: and regions of low intensity appear black.
895: 
896: 
897: For neutrinos on the equator, the momentum-space forward direction is
898: $(\vartheta = 90^\circ, \varphi = 0)$. Electron neutrinos generally
899: have the shortest transport mean-free path of all species in the
900: core-collapse context and decouple from matter at the lowest
901: densities. The Hammer projection in the top-left corner of
902: Fig.~\ref{fig:hammer1} of the $\varepsilon_\nu$ = 12.6~MeV equatorial
903: radiation field at 60~km corresponds roughly to the blue line graph in
904: Fig.~\ref{fig:s20nr_psi1}, which portrays only its variation with
905: $\varphi$.  At fixed neutrino energy group $\varepsilon_\nu$, electron
906: anti-neutrinos and ``$\nu_\mu$'' neutrinos decouple at smaller
907: radii. Hence, as Fig.~\ref{fig:hammer1} shows, at 60~km, they already
908: manifest greater local anisotropy than the $\nu_e$s. This trend
909: continues at all considered radii in Fig.~\ref{fig:hammer1}.
910: 
911: In Fig.~\ref{fig:hammer2}, we again present Hammer projections of the
912: normalized specific intensity, but this time consider only $\nu_e$s,
913: keep the radius fixed at 150~km, and vary the neutrino energy and the
914: angular position on the grid. The bottom row of Fig.~\ref{fig:hammer2}
915: shows the normalized $I_\nu$ at the equator ($\theta=90^\circ$) and
916: for the 12.6-MeV and 35.7-MeV $\nu_e$ energy groups. The center and
917: top rows show the same groups at $\theta=45^\circ$ and at
918: $\theta=0^\circ$, respectively.  From the discussion of
919: Fig.~\ref{fig:hammer1}, we are already familiar with the overall
920: radiation field geometry.  The transport mean-free path scales roughly
921: inversely with $\varepsilon_\nu^2$. Hence, at any given position in
922: the postbounce supernova core, more energetic neutrinos should be
923: locally more isotropically distributed in momentum space than less
924: energetic ones. The less forward-peaked angular $I_\nu$ distribution
925: of the higher-energy neutrinos reflects this.
926: 
927: The degree of forward-peaking in $\vartheta$ and $\varphi$ of the
928: radiation field in the quasi-spherically symmetric nonrotating model
929: s20.nr is essentially independent of the angular position on the grid
930: and the radiation fields at any given radius can be transformed into
931: one another by simple rotation. Because of the aspherical and oblate
932: distribution of matter in the rotating model s20.$\pi$, the
933: forward-peaking is also a function of polar angle. Due to its PNS's
934: oblateness (see Fig.~\ref{fig:intro2D}), the neutrinos generally
935: decouple at significantly smaller radii near the pole than near the
936: equator, in turn leading to more strongly forward-peaked radiation
937: fields in the polar than in the equatorial regions
938: \citep{janka:89a,janka:89b,walder:05,dessart:06b,dessart:07}.
939: 
940: 
941: 
942: \begin{figure*}[t]
943: \centering
944: \includegraphics[width=5.95cm]{f6a.ps}
945: \includegraphics[width=5.95cm]{f6b.ps}
946: \includegraphics[width=5.95cm]{f6c.ps}
947: \caption{Normalized Eddington tensor ${\textsf k}$ components in
948: spherical coordinates as a function of neutrino energy $\varepsilon_\nu$
949: and spherical radius $r$. {\bf Left}: Angular-averaged $k_{rr}$ and
950: $k_{\vartheta\vartheta}$ for electron neutrinos in model
951: s20.nr. $k_{\varphi\varphi}$ is not shown, but has essentially
952: identical behavior to $k_{\vartheta\vartheta}$.  The diagonal
953: components start out with $1/3$ at small radii, as expected for the
954: prevailing isotropic radiation fields. With increasing radius
955: (and decreasing density), the local radiation field becomes more
956: anisotropic and forward-peaked. This occurs at progressively larger
957: radii with increasing $\varepsilon_\nu$ and is reflected by the
958: increasing $k_{rr}$ and the decreasing $k_{\vartheta\vartheta}$ in the
959: plot. The off-diagonal component $k_{r\vartheta}$ is not shown, does
960: not exhibit clear systematics, and is generally a factor of 10--100
961: smaller than the diagonal components.  {\bf Center}: Same as left,
962: but showing profiles extracted from regions near the pole in the
963: rapidly-rotating model s20.$\pi$. Interior to $\sim$100~km, $k_{rr}$,
964: and $k_{\vartheta\vartheta}$ show the same systematics with
965: $\varepsilon_\nu$ as in the nonrotating model. However, at larger radii
966: they are reversed, $k_{rr}$ and $k_{\vartheta\vartheta}$ exhibiting
967: greater isotropy for lower $\varepsilon_\nu$. See text for discussion.
968: {\bf Right}: Equatorial profiles of $k_{rr}$, $k_{\vartheta\vartheta}$
969: and $k_{\varphi\varphi}$ for electron neutrinos in model s20.nr. Due to
970: rotational flattening of the PNS, the transition to free streaming
971: occurs over a much larger range of radii near the equator.
972: $k_{\vartheta\vartheta}$ shows a significantly 
973: larger variation as a function of energy than $k_{\varphi\varphi}$.
974: \label{fig:eddy}} 
975: \end{figure*}
976: 
977: 
978: \subsection{Eddington Factors}
979: \label{section:Eddy}
980: 
981: The radiation-pressure tensor ${\textsf K}_\nu$, also known as
982: the Eddington tensor, represents the second angular moment of the
983: specific intensity and is defined by eq.~(\ref{eq:eddy}).
984: In the following, we use its normalized variant 
985: ${\textsf k}_\nu = {\textsf K}_\nu / J_\nu$.
986: 
987: In spherical symmetry, ${\textsf k}_\nu$ is diagonal and has a single
988: independent component, the Eddington factor $k_\nu$.  For isotropic
989: radiation, $k_\nu = 1/3$ and ${\textsf k}_\nu =
990: \mathrm{diag}(1/3,1/3,1/3)$, while in the streaming regime, $k_\nu =
991: 1$ and ${\textsf k}_\nu = \mathrm{diag}(1,0,0)$.  In the transition
992: from isotropy to free streaming, $k_\nu$ generally varies from $1/3$
993: to $1$, but in special cases, e.g., enhanced radiation perpendicular
994: to the radial direction, may assume values below $1/3$.  Note that one
995: of the common assumptions of MGFLD is the Eddington closure, setting
996: $k_\nu = \frac{1}{3}$ everywhere.
997: 
998: In axisymmetry and ignoring velocity-dependent terms, the Eddington tensor
999: has four independent components whose individual meaning depends on
1000: the coordinates chosen\footnote{Off-diagonal components of the
1001: Eddington tensor can be related to radiation shear viscosity
1002: \citep{mihalas:99}, which we do not consider here.}.  We assume and
1003: transform to spherical coordinates for our discussion, since they make
1004: the interpretation of the components most straightforward.
1005: 
1006: In Fig.~\ref{fig:eddy}, we present radial profiles of normalized
1007: Eddington tensor components at selected electron-neutrino energies
1008: $\varepsilon_\nu$ in models s20.nr and s20.$\pi$.  The nonrotating model
1009: can be considered nearly spherically symmetric, and, hence, should and
1010: does exhibit the expected Eddington-factor systematics. At small
1011: radii and high densities, where neutrinos and matter are in
1012: equilibrium, $k_{rr} = k_{\vartheta\vartheta} = k_{\varphi\varphi} =
1013: \frac{1}{3}$ and with increasing radius, $k_{rr} \rightarrow 1$ and
1014: $\{k_{\vartheta\vartheta},k_{\varphi\varphi}\} \rightarrow 0$.  As
1015: expected from the basic decoupling hierarchy, the value of the
1016: Eddington tensor components is a strong function of $\varepsilon_\nu$.
1017: Lower-$\varepsilon_\nu$ neutrinos decouple at higher densities, and, hence,
1018: have Eddington tensor components which depart from $\frac{1}{3}$ at
1019: smaller radii than $\nu_e$s of higher energy.
1020: This systematics applies, of course, to $\bar{\nu}_e$s 
1021: and ``$\nu_\mu$''s as well.
1022: The off-diagonal component
1023: $k_{r\vartheta}$ is zero in the isotropic region, does not exhibit clear
1024: systematics, and stays an order-of-magnitude smaller than the diagonal
1025: components for all $\varepsilon_\nu$ and species.
1026: 
1027: The rotating model s20.$\pi$ has a postshock configuration that is far
1028: from spherically symmetric (Fig.~\ref{fig:intro2D}).  We present in
1029: Fig.~\ref{fig:eddy} separate plots for its Eddington tensor components
1030: in regions near the pole and near the equator.  In the polar regions
1031: and at small radii ($r\,\sless\,$100~km), the Eddington tensor
1032: components show the same qualitative behavior as in model s20.nr. At
1033: larger radii, however, the systematics are reversed and
1034: lower-$\varepsilon_\nu$ electron neutrinos have more isotropic radiation
1035: fields (smaller $k_{rr}$) than their higher-$\varepsilon_\nu$
1036: counterparts. Analyzing their radiation fields and matter coupling in
1037: detail, we find that this surprising feature is a consequence of
1038: electron capture and the polar compactness (large density gradient due
1039: to rotation) of the supernova core.  Electron capture near the shock
1040: leads to isotropic neutrino emission that can locally isotropize the
1041: radiation field in semi-transparent regions.  With decreasing density
1042: and temperature, the mean energy of neutrinos emitted by capture
1043: processes shifts to lower $\varepsilon_\nu$.  This leads to greater local
1044: isotropization of lower-$\varepsilon_\nu$ neutrinos, which in turn is
1045: reflected in the more slowly increasing $k_{rr}$ of these
1046: neutrinos. This interpretation is confirmed by the fact that we do not
1047: find any such feature in the Eddington tensor components of the
1048: ``$\nu_\mu$'' neutrinos that are not produced in 
1049: capture processes. We also do not observe significant isotropization
1050: in the $\bar{\nu}_e$ radiation fields, since the emission of
1051: $\bar{\nu}_e$s by positron capture on neutrons is weaker due to
1052: the lower positron abundance.
1053: 
1054: 
1055: In regions of model s20.$\pi$ near the equator where the PNS is most
1056: extended, the neutrino radiation fields stay isotropic to large radii
1057: and decouple from matter only slowly with radius. Since the matter
1058: densities in the equatorial plane stay roughly a factor of four larger
1059: than in the polar regions, the cross-over feature in $\{k_{rr},
1060: k_{\vartheta\vartheta}, k_{\varphi\varphi}\}$ does not appear and
1061: these components follow the standard decoupling
1062: hierarchy. Interestingly, and different from in the nonrotating model,
1063: $k_{\vartheta\vartheta}$ and $k_{\varphi\varphi}$ show quantitatively
1064: distinct variation with $\varepsilon_\nu$, the latter exhibiting
1065: significantly less variation with $\varepsilon_\nu$ at any given radius.
1066: The interpretation of this observation is not straightforward, but we
1067: suggest that it can be attributed to the fact that in model s20.$\pi$ the
1068: radiation field at any given point on the equator of the
1069: rotationally-flattened core and for any $\varepsilon_\nu$ and neutrino
1070: species varies locally less in the $\vartheta$ direction than in the
1071: $\varphi$ direction. This, in combination with the fact that on the
1072: equator the radiation field asymptotically peaks into the 
1073: ($\vartheta=0$,$\varphi=0$) direction, results on average in smaller
1074: $k_{\varphi\varphi}$ with less spread in energy than exhibited by
1075: $k_{\vartheta\vartheta}$.
1076: The off-diagonal component $k_{r\vartheta}$ (not shown in
1077: Fig.~\ref{fig:eddy}) vanishes in $k_{rr} = k_{\vartheta\vartheta} =
1078: k_{\varphi\varphi} = \frac{1}{3}$ regions, but can become relatively
1079: large at greater radii (up to $\sim$0.2 in magnitude; increasing with
1080: $\varepsilon_\nu$ and radius) and flips sign at the equator. The
1081: interpretation of $k_{r\vartheta}$ is not straightforward, since its
1082: magnitude depends on the choice of coordinates. We do not
1083: attempt to study it, nor its implications for neutrino shear
1084: viscosity, in any detail.
1085: 
1086: 
1087: \begin{figure}[t]
1088: \centering
1089: \includegraphics[width=8.5cm]{f7.ps}
1090: \caption{ S$_n$--MGFLD comparison for the nonrotating model s20.nr at
1091: 160~ms after bounce. All S$_n$ results where obtained with a 16
1092: $\vartheta$-angle calculation.  See text for details and discussion.
1093: Top panel: Neutrino luminosity as a function of radius and broken down
1094: into the three neutrino species considered. The ``$\nu_\mu$'' neutrinos
1095: dominate in luminosity and their luminosity profiles are scaled by
1096: a factor of 1/4 to preserve the overall scale of the plot. 
1097: Center panel: Angle-averaged energy-mean inverse
1098: neutrino flux factor profiles.  Bottom panel: RMS neutrino energy
1099: profiles.
1100: \label{fig:s20nr_line}}
1101: \end{figure}
1102: 
1103: 
1104: \subsection{Global Radiation Field Diagnostics: Luminosities,
1105: Spectra, Flux Factors, and Neutrino Energy Deposition}
1106: \label{section:stills}
1107: 
1108: So far we have studied aspects of neutrino transport inaccessible to
1109: MGFLD. We now go on to discuss radiation field diagnostics that
1110: facilitate a \sn--MGFLD comparison. For further reference
1111: and comparison with previous studies \citep{janka:92,messer:98,
1112: burrows:00}, we define the neutrino luminosity per species 
1113: $L_{\nu_i}$ at spherical radius $r$, 
1114: \begin{equation}
1115: L_{\nu_i}(r) = \oint d\omega \int d\varepsilon_\nu\, F_r(r,\varepsilon_\nu,\nu_i) r^2\,, 
1116: \end{equation} 
1117: where $F_r$ is the spectral radial neutrino flux in species $\nu_i$ 
1118: at energy $\varepsilon_\nu$. $d\omega$ is the
1119: spatial solid-angle element, $d\omega = 2\pi \sin\theta d\theta$ in
1120: axisymmetry. Furthermore, we define the
1121: mean inverse flux factor $\langle 1 / {\textsf F}_{\nu_i}\rangle$,
1122: \begin{equation}
1123: \bigg\langle \frac{1}{{\textsf F}_{\nu_i}} \bigg\rangle = 
1124: \frac{c \int d\varepsilon_\nu E(\varepsilon_\nu,\nu_i)}
1125: {\int d\varepsilon_\nu F_r(\varepsilon_\nu, \nu_i)}\,,
1126: \end{equation}
1127: where $E(\varepsilon_\nu,\nu_i) = 4\pi c^{-1} J(\varepsilon_\nu,\nu_i)$
1128: is the spectral neutrino energy density,
1129: and the neutrino RMS energies are
1130: \begin{equation}
1131: E_{\mathrm{RMS},\nu_i} = \sqrt{{\frac{\int d\varepsilon_{\nu_i}
1132: \varepsilon_{\nu_i}^2 J(\varepsilon_{\nu_i})}
1133: {\int d\varepsilon_{\nu_i} J(\varepsilon_{\nu_i})}}}\,.
1134: \end{equation}
1135: 
1136: The above three quantities are particularly useful diagnostics, since the
1137: $\varepsilon_\nu$-averaged energy deposition rate by charged-current
1138: absorption of $\nu_e$ and $\bar{\nu}_e$ on neutrons and protons scales
1139: linearly with their product (\citealt{messer:98}).
1140: 
1141: \vspace*{.5cm}
1142: 
1143: \subsubsection{Model s20.nr}
1144: \label{section:s20nr}
1145: 
1146: In Fig.~\ref{fig:s20nr_line}, we plot neutrino luminosities
1147: $L_{\nu_i}$, angle-averaged mean inverse flux factors, and the
1148: angle-averaged $E_\mathrm{RMS}$ for the postbounce snapshot at 160~ms
1149: of the nonrotating model s20.nr. The asymptotic total luminosity at
1150: this time is $\sim$150~B~s$^{-1}$ and is already dominated by the
1151: thermally-produced ``$\nu_\mu$''s that cool the PNS, but contribute
1152: little to the heating in the gain region, since they cannot take part
1153: in charged-current absorption processes. In this quasi-spherically
1154: symmetric model, we define a spherical gain radius $r_\mathrm{gain}$
1155: as the radial position beyond which net neutrino energy deposition
1156: occurs. At 160~ms after bounce, $r_\mathrm{gain} \simeq $~90~km and
1157: the gain region extends almost out to the shock at $\sim$175~km.  The
1158: MGFLD luminosities in Fig.~\ref{fig:s20nr_line} are systematically
1159: lower by $\sim$5\% for $\nu_e$s, $\sim$3.5\% for $\bar{\nu}_e$s, and
1160: $\sim$4\% for ``$\nu_\mu$''s, but qualitatively resemble the \sn\
1161: luminosity profiles in the gain region. At around the shock position,
1162: all MGFLD luminosities increase by $\sim$5\%. This is a due to the
1163: combination of the artificially spread-out shock (over $\sim$4--5
1164: zones), the rapid change of the inverse neutrino mean-free path in the
1165: spread-out shock and the implementation of the flux limiter in
1166: VULCAN/2D.  Since this MGFLD artefact occurs right at the shock, it
1167: can have only little influence on the heating in the gain region, but
1168: leads to somewhat overestimated asymptotic luminosities in the MGFLD
1169: case.
1170: 
1171: The center panel of Fig.~\ref{fig:s20nr_line} shows the
1172: $\varepsilon_\nu$-averaged inverse flux factors for the three neutrino
1173: species in the MGFLD and \sn\ steady-state calculations of model
1174: s20.nr.  For isotropic radiation $\langle 1 / {\textsf
1175: F}_{\nu_i}\rangle$ tends to infinity, while it approaches one when the
1176: radiation field becomes forward-peaked at low optical depth.
1177: Focussing on the gain region between $r_\mathrm{gain}$ and the shock
1178: position, we find that MGFLD yields mean inverse flux factors that are
1179: up to $\sim$5\% larger for $\bar{\nu}_e$s (less for the other
1180: species) in the inner gain region. At radii $\sgreat\,$150~km, the
1181: MGFLD $\langle 1 / {\textsf F}_{\nu_i}\rangle$ quickly drops to 1
1182: (free streaming), becoming up to 8\% lower than the \sn\ values in the
1183: outer gain region. We note that ``$\nu_\mu$'' interact only via
1184: neutral-current weak interactions, hence, decouple from matter at
1185: higher densities and temperatures. Next in the decoupling hierarchy
1186: are electron anti-neutrinos followed by electron neutrinos.  Both \sn\
1187: and MGFLD realize this hierarchy at radii below $\sim$150~km, beyond
1188: which MGFLD rapidly transitions to free streaming irrespective of
1189: neutrino species.
1190: 
1191: 
1192: \begin{figure}[t]
1193: \centering
1194: \includegraphics[width=8.5cm]{f8.ps}
1195: \caption{ Neutrino luminosity spectra extracted at a radius of 500~km for
1196: $\nu_e$, $\bar{\nu}_e$, and ``$\nu_\mu$'' neutrinos at 160~ms after bounce in
1197: model s20.nr. Solid lines correspond to S$_n$ results, while dashed
1198: lines are obtained using MGFLD. The spectra have the canonical shape
1199: and the quantitative behavior found in nonrotating intermediate-time
1200: postbounce supernova calculations~(e.g., \citealt{thompson:03}) with
1201: the ``$\nu_\mu$''-neutrinos peaking at the highest energies, since they
1202: decouple from the fluid at the smallest radii. MGFLD and S$_n$ spectra
1203: agree closely in shape, but  MGFLD is overestimating slightly the
1204: total asymptotic luminosity (cf. Fig~\ref{fig:s20nr_line}).
1205: \label{fig:s20nr_spec160}}
1206: \end{figure}
1207: 
1208: \begin{figure}[t]
1209: \centering
1210: \includegraphics[width=8.5cm]{f9.ps}
1211: \caption{Angle-averaged specific neutrino net gain profile in the 
1212: s20.nr model at 160~ms after core bounce. Shown are the MGFLD results, as
1213: well as results from steady-state \sn\ calculations with 8, 12, and 
1214: 16 $\vartheta$-angles, corresponding to a total number of angular zones of 40,
1215: 92, and 162. The gain region extends from $\sim$90~km to the shock position
1216: at $\sim$175~km. The three different \sn\ resolutions yield net gain profiles
1217: that agree very well (relative differences below 1\% even for S$_8$). The MGFLD
1218: calculation underestimates the total net gain in the outer gain region by 
1219: at most 10\% locally and $\sless\,$5\% on average.
1220: \label{fig:s20nr_gainprof}}
1221: \end{figure}
1222: 
1223: \begin{figure}[h!]
1224: \centering
1225: \includegraphics[width=8.8cm]{f10.ps}
1226: \caption{ 2D colormap of the specific (per gram) net gain
1227: distribution in model s20.nr at 160~ms after core bounce. The left half
1228: of the plot depicts the MGFLD result, \sn\ is shown on the right.
1229: The differences between S$_n$ and MGFLD are marginal at this time in this
1230: model and are practically indiscernible by eye. As a consequence of
1231: convection in the gain region and the onset of the SASI, even this
1232: nonrotating model exhibits significant angular and radial variations in the
1233: neutrino energy deposition not captured by the average profiles
1234: in Fig.~\ref{fig:s20nr_gainprof}.
1235: \label{fig:s20nr_net_gain160}}
1236: \end{figure}
1237: 
1238: \begin{figure*}[t!]
1239: \includegraphics[width=6.0cm]{f11a.ps}
1240: \includegraphics[width=6.0cm]{f11b.ps}
1241: \includegraphics[width=6.0cm]{f11c.ps}
1242: \caption{Colormaps of the radial spectral flux at $\varepsilon_\nu
1243: =$~12.6~MeV of $\nu_e$ (left), $\bar{\nu}_e$ (center), and
1244: ``$\nu_\mu$'' (right) neutrinos in the rapidly-rotating model
1245: s20.$\pi$ at 160~ms after bounce.  Isoenergy density contours ($4\pi
1246: c^{-1} J_\nu$, vertical color legend) are superposed.  The left half
1247: of each panel displays the MGFLD result -- \sn\ is shown in the right
1248: half. The radiation fields are oblate in the PNS core and deform to a
1249: prolate shape further out. Note that \sn\ predicts a prolateness of
1250: the radiation field to much greater radii than MGFLD does. The latter
1251: leads to nearly spherically symmetric radiation fields at radii
1252: greater than $\sgreat\,$150--200~km independent of neutrino species.
1253: \label{fig:s20pi_fluxes}}
1254: \end{figure*}
1255: 
1256: 
1257: \begin{figure}[t]
1258: \centering
1259: \includegraphics[width=8.5cm]{f12.ps}
1260: \caption{ Radial neutrino ``luminosity'' profiles ($4\pi r^2 F_r$) as
1261: seen by observers near the pole (solid lines) and near the equator (dashed
1262: lines) in model s20.$\pi$ at 160~ms after bounce. Red
1263: graphs correspond to \sn\ results, black graphs depict MGFLD results.
1264: Top, center, and bottom panels show $L_\nu$ for $\nu_e$, $\bar{\nu}_e$,
1265: and ``$\nu_\mu$,'' respectively. All \sn\ results were obtained with
1266: $n = 16$, but for comparison we also plot in the top panel polar
1267: profiles that were obtained with S$_{8}$ and S$_{12}$ and find that
1268: both S$_{16}$ and S$_{12}$ are very well converged, while S$_{8}$
1269: has troubles at radii greater than $\sim$200~km. However, it agrees very well
1270: at smaller radii with the higher-resolution \sn\ calculations.
1271: \label{fig:s20pi_lum160}}
1272: \end{figure}
1273: 
1274: 
1275: The behavior we observe with radius of the luminosity and mean-inverse
1276: flux factor agrees with the general findings of \cite{messer:98}.  In
1277: particular, we agree with their assessment that the artificially
1278: accelerated transition to free streaming in MGFLD occurs not at the
1279: neutrinospheres (which are generally below the gain region), but at
1280: relatively large radii within which most of the neutrino source is
1281: enclosed.
1282: 
1283: In the bottom panel of Fig.~\ref{fig:s20nr_line} 
1284: we present profiles of the RMS neutrino energy  
1285: for all species in MGFLD and \sn\ snapshots of model s20.nr. The
1286: corresponding luminosity spectra (extracted at 500~km) are shown 
1287: in Fig.~\ref{fig:s20nr_spec160}.
1288: Both MGFLD and \sn\ capture the energy systematics that 
1289: is set essentially by the matter temperature in the decoupling
1290: region. Neutrino species that decouple at smaller radii (higher
1291: densities and temperatures) have higher RMS energies and
1292: harder spectra than neutrinos decoupling at larger radii.
1293: Quantitative differences in RMS energies and in the spectra
1294: between MGFLD and \sn\ are small, the slightly higher MGFLD
1295: spectral luminosities being mostly a result 
1296: of the artificially enhanced MGFLD luminosities
1297: near and beyond the shock.
1298: 
1299: 
1300: We now conclude our discussion of the 160~ms-postbounce snapshot of
1301: model s20.nr by considering the instantaneous neutrino energy
1302: deposition rates.  Figure~\ref{fig:s20nr_gainprof} depicts
1303: angle-averaged radial profiles of the specific neutrino
1304: heating/cooling rates in units of erg~(g s)$^{-1}$. The region of net
1305: gain extends from $\sim$90~km to the shock radius and the chief
1306: contribution to the heating comes from charged-current
1307: $\bar{\nu}_e$-capture processes on protons, exceeding the
1308: corresponding $\nu_e$-capture on neutrons by a factor of two and more
1309: in the narrow radial interval from 145 to 175~km.  MGFLD
1310: underestimates the specific net gain in the angle-averaged
1311: radial profile by at most 10\% locally and by $\sim$5\% on average at
1312: radii greater than $\sim$110~km.  The integral total net gain
1313: predicted by S$_{16}$ is 2.13~B~s$^{-1}$. This is only 3\% larger than
1314: the MGFLD value of 2.07~B~s$^{-1}$. We note in passing that S$_8$
1315: overestimates the integrated gain rate by at most $\sim$1.6\% while S$_{12}$
1316: agrees with S$_{16}$ to better than $\sim$0.3\%.
1317: 
1318: Figure~\ref{fig:s20nr_net_gain160} depicts the 2D distribution of
1319: neutrino heating and cooling in the snapshot of model s20.nr
1320: considered here. Regions of net gain range from green to red, cooling
1321: regions are blue to black. The colormap demonstrates the somewhat
1322: misleading character of angle-averaged profiles. While we find that
1323: there is little spatial angular variation in the neutrino radiation
1324: field, the neutrino--matter coupling depends strongly on 
1325: angular position, and energy deposition is generally greatest in
1326: regions of high entropy (cf. Fig~\ref{fig:intro2D}).
1327: 
1328: 
1329: \subsubsection{Model s20.$\pi$}
1330: \label{section:s20pi}
1331: 
1332: As we discussed in the context of the Eddington tensor in
1333: \S\ref{section:Eddy} and as may be guessed from the significant
1334: rotational deformation of the core in model s20.$\pi$
1335: (Fig.~\ref{fig:intro2D}), the radiation field in this model exhibits a
1336: strong rotationally-induced asymmetry between pole and equator.  In
1337: Fig.~\ref{fig:s20pi_fluxes}, we present 2D colormaps of the radial
1338: spectral flux component (in erg s$^{-1}$ cm$^{-2}$ MeV$^{-1}$) and
1339: isoenergy-density contours ($4\pi J_\nu / c$ in erg cm$^{-3}$
1340: MeV$^{-1}$) at a representative $\varepsilon_\nu$ of 12.6~MeV and for all
1341: species. Numbers for both \sn\ and MGFLD are compared side by side.
1342: The global radiation-field anisotropy systematics are qualitatively
1343: similar to what was found in the previous MGFLD rotating core-collapse
1344: study of \cite{walder:05}. At small radii, the radiation field (energy
1345: density) follows the density distribution and is oblate, but in the
1346: snapshot at 160~ms after bounce shown in Fig.~\ref{fig:s20pi_fluxes}
1347: has a pole--equator ratio of only 1:2. This ratio increases as the PNS
1348: cools and contracts. The polar compactness of the PNS core leads to a
1349: decoupling of matter and neutrinos at smaller radii in regions near
1350: the pole, resulting there in greater spectral fluxes at higher
1351: neutrino energies and in a prolate distribution of neutrino fluxes
1352: and isoenergy-density contours.
1353: 
1354: 
1355: 
1356: \begin{figure}[t]
1357: \centering
1358: \includegraphics[width=8.5cm]{f13.ps}
1359: \caption{{\bf Top}: Neutrino ``luminosity'' spectra ($4\pi r^2
1360: F_r(\varepsilon_\nu)$) in \sn\ (red) and MGFLD (black) variants of model
1361: s20.$\pi$ as seen by observers near the pole.
1362:  $\nu_e$ spectra have solid lines, $\bar{\nu}_e$ spectra are
1363: shown in dashed lines, and ``$\nu_\mu$''s have dashed-dotted
1364: spectra. The spectra are taken from a S$_{16}$ calculation at a
1365: radius of 300~km at 160~ms after core bounce.
1366: {\bf Bottom:}  ``Luminosity'' spectra seen by equatorial observers.
1367: \label{fig:s20pi_spec160}}
1368: \end{figure}
1369: 
1370: 
1371: The most striking difference between the \sn\ and MGFLD radiation
1372: fields presented in Fig.~\ref{fig:s20pi_fluxes} is the former's much
1373: greater prolateness at large radii for all species (and all energies,
1374: though we show only $\varepsilon_\nu =$~12.6~MeV).  With the MGFLD
1375: prescription, the prolateness of the flux is muted and does not extend
1376: to large radii. Though the radiation fields are smoothed out at radii
1377: $\sgreat\,$150~km by MGFLD, the \sn\ fluxes and energy densities
1378: remain prolate through the entire postshock region and beyond. At
1379: radii outside $\sim$200~km, the typical striping pattern of
1380: \sn\ (\citealt{castor:04}) becomes visible, though not 
1381: yet dominant.
1382: 
1383: In Fig.~\ref{fig:s20pi_lum160}, we plot line profiles of the polar and
1384: equatorial ``luminosities'' ($4\pi r^2 F_r$) of each neutrino
1385: species. Profiles obtained with \sn\ and MGFLD are shown. The
1386: asymptotic ``luminosities'' obtained with S$_n$ have pole-to-equator
1387: ratios of 2.2 ($\nu_e$), 1.8 ($\bar{\nu_e}$), and 2.4 (``$\nu_\mu$'').
1388: MGFLD smoothes out these large asymmetries, yielding higher equatorial
1389: and significantly lower polar ``luminosities'' at radii greater than
1390: $\sim$100~km. This is consistent with the more qualitative findings
1391: based on Fig.~\ref{fig:s20pi_fluxes}.  We note that the MGFLD variant
1392: of VULCAN/2D still conserves total flux and energy. For the \sn\
1393: calculation, we find total asymptotic luminosities of
1394: 21.1~B~s$^{-1}$ for $\nu_e$ neutrinos (MGFLD:
1395: 20.4~B~s$^{-1}$),
1396: 22.7~B~s$^{-1}$ for $\bar{\nu}_e$ neutrinos (MGFLD:
1397: 22.6~B~s$^{-1}$), and
1398: 53.0~B~s$^{-1}$ for ``$\nu_\mu$'' neutrinos (MGFLD:
1399: 52.3~B~s$^{-1}$). Hence, \sn\ and MGFLD total
1400: luminosities per species agree very well (and differ at most by
1401: $\sim$3.5\% in the $\nu_e$ case), while their flux distributions
1402: disagree significantly.
1403: 
1404: Figure~\ref{fig:s20pi_spec160}, depicting polar and equatorial
1405: ``luminosity'' spectra ($4\pi r^2 F_r(\varepsilon_\nu)$), reveals that in the
1406: \sn\ calculation (polar: black graphs, equatorial: red graphs) the
1407: neutrino radiation emerging from the PNS and postshock environments
1408: through the polar region not only has greater fluence, but also a
1409: significantly different and -- in the $\nu_e$ case -- a significantly
1410: harder spectrum.  $\nu_e$ neutrinos decouple at the largest
1411: radii. Their ``luminosity'' spectrum observed by a polar observer peaks at
1412: $\varepsilon_\nu \sim$9.5~MeV, while for an observer near the equator it
1413: peaks at $\sim$7.5~MeV. Both $\bar{\nu}_e$ and ``$\nu_\mu$'' neutrinos
1414: (which decouple further in) exhibit a smaller variation in peak
1415: energy from pole to equator. The MGFLD calculation, on the other hand,
1416: shows much smaller variations in neutrino energy and flux between
1417: pole and equator (green and blue graphs, respectively).
1418:  We note in passing that the emerging
1419: neutrino spectra of model s20.$\pi$ are systematically softer by up to
1420: $\sim$10\% in each species than those of the nonrotating model s20.nr
1421: presented in Fig.~\ref{fig:s20nr_spec160}. This is a direct
1422: consequence of the rotationally-induced lower overall compactness of
1423: the PNS in model s20.$\pi$.
1424: 
1425: The RMS neutrino energies in model s20.$\pi$ show the same overall
1426: qualitative behavior and decoupling hierarchy discussed in the context
1427: of model s20.nr. Hence, we do not show them here, but rather state
1428: quantitative results. They do, of course, trace the strong
1429: pole--equator asymmetry that we observe in the radiation field.  The
1430: RMS energies in the 160~ms \sn\ snapshot are 11.5~MeV (pole) and
1431: 10.5~MeV (equator) for $\nu_e$, 16.7~MeV (pole) and 15.2~MeV (equator)
1432: for $\bar{\nu}_e$, and 25.8~MeV (pole) and 24.8~MeV (equator) for
1433: ``$\nu_\mu$'' neutrinos. The MGFLD values converge at pole and equator
1434: to 10.5~MeV ($\nu_e$), 15.2~MeV ($\bar{\nu}_e$), and 25.0~MeV
1435: (``$\nu_\mu$'').
1436: 
1437: \begin{figure}[t]
1438: \centering
1439: \includegraphics[width=8.5cm]{f14.ps}
1440: \caption{ 
1441: Mean inverse flux factors in model s20.$\pi$
1442: at 160~ms after bounce in polar regions (solid lines) and equatorial
1443: regions (dashed lines). Shown are profiles for $\nu_e$ neutrinos
1444: obtained with \sn\ (red) and MGFLD (black), as well as profiles
1445: for $\bar{\nu}_e$ neutrinos (\sn\ blue, MGFLD green). \sn\ 
1446: and MGFLD graphs agree well inside $\sim$50~km at the pole
1447: and inside $\sim$80~km in equatorial regions. For $\nu_e$ neutrinos,
1448: \sn\ yields systematically larger mean inverse flux factors in
1449: polar and equatorial regions. For $\bar{\nu}_e$, however, \sn\ predicts
1450: larger mean inverse flux factors in polar regions, yet transitions
1451: slightly faster than MGFLD to free streaming in equatorial
1452: regions.\label{fig:s20pi_FF160}}
1453: \end{figure}
1454: 
1455: 
1456: \begin{figure}[t]
1457: \centering
1458: \includegraphics[width=8.5cm]{f15.ps}
1459: \caption{ Colormaps of energy- and species-integrated specific
1460: neutrino energy deposition and loss rates in the rotating model
1461: s20.$\pi$ at 160~ms after core bounce (in units of
1462: erg~s$^{-1}$~g$^{-1}$).  The left section of the plot depicts the
1463: MGFLD result and the right shows the result of the \sn\
1464: calculation. Note the distinctively enlarged polar gain regions and
1465: greater specific gain of the S$_n$ result compared to the MGFLD
1466: calculation. This is in part a consequence of the larger polar
1467: neutrino fluxes and overall greater flux asymmetry in the S$_n$ model
1468: (see Fig.~\ref{fig:s20pi_fluxes}).  A feature prevalent in both S$_n$
1469: and MGFLD versions of this rapidly rotating core is an extended loss
1470: region between the shock and the small gain region at low latitudes
1471: (cf. Fig.~\ref{fig:s20pi_netgain2}). The material in the loss region
1472: is still proton rich ($Y_e $\sgreat0.4) and efficiently captures
1473: electrons as it advects in, radiating away a significant flux of
1474: neutrinos (see, e.g., the increase in the equatorial ``luminosity''
1475: between 120 and 150~km in the \sn\ variant of this model, visible in
1476: the top panel of Fig.~\ref{fig:s20pi_lum160}). Note that both MGFLD
1477: and \sn\ exhibit a very small artefact (lower gain/loss) at the
1478: symmetry axis associated with imperfect numerics/regularization.
1479: \label{fig:s20pi_netgain}}
1480: \end{figure}
1481: 
1482: \begin{figure}[t]
1483: \centering
1484: \includegraphics[width=8.5cm]{f16a.ps}
1485: \includegraphics[width=8.5cm]{f16b.ps}
1486: \caption{{\bf Top}: Averaged specific radial neutrino gain and loss
1487: profiles in model s20.$\pi$ at 160~ms after core bounce. Shown are
1488: results from the S$_n$ (red) and MGFLD (black) calculations. Both
1489: polar and equatorial radial profiles are obtained by averaging over
1490: 20$^\circ$ wedges. As is already clear from
1491: Fig.~\ref{fig:s20pi_netgain}, S$_n$ yields significantly greater polar
1492: specific neutrino energy gain than MGFLD. The S$_n$ gain region
1493: extends further in by $\sim$10~km and the gain is more than a factor
1494: of two larger in the interval from $\sim$90 to 200~km.  Given the
1495: larger flux asymmetry in the S$_n$ calculation
1496: (Fig.~\ref{fig:s20pi_fluxes}), less neutrino flux is going through
1497: regions of low latitude, resulting in the \emph{lower} specific gain at low
1498: latitudes predicted by S$_n$.  {\bf Bottom}: Neutrino gain density
1499: (density-weighted specific gain). Due to rapid rotation higher
1500: densities obtain out to larger radii at low latitudes. This results in
1501: a partial reversal of the picture presented by the upper panel;
1502: weighted by density, the neutrino gain (now per unit volume) in the
1503: equatorial wedge becomes comparable to that near the poles.
1504: Furthermore, equatorial regions, since they subtend the largest solid
1505: angles, contribute most to the volume integral. The integral numbers
1506: for the net gain in the polar wedge (counting both poles) for S$_n$
1507: (MGFLD) are 0.17~B~s$^{-1}$
1508: (0.047~B~s$^{-1}$) and in the equatorial wedge
1509: are 0.35~B~s$^{-1}$
1510: (0.47~B~s$^{-1}$). The total integrated net gain is
1511: 1.603~B~s$^{-1}$ and
1512: 1.637~B~s$^{-1}$ for S$_n$ and MGFLD,
1513: respectively. These numbers are surprisingly close given the large
1514: qualitative and quantitative \emph{local} differences in the neutrino
1515: gain distribution.
1516: \label{fig:s20pi_netgain2}}
1517: \end{figure}
1518: 
1519: 
1520: In Fig.~\ref{fig:s20pi_FF160}, we plot polar and equatorial mean
1521: inverse flux factor profiles for $\nu_e$ and $\bar{\nu}_e$ neutrinos
1522: in our steady-state snapshot for model s20.$\pi$. Results from MGFLD
1523: and S$_{16}$ runs are shown. A free-streaming radiation field has an
1524: inverse flux factor of one. Due to the steeper density gradient in
1525: polar regions, neutrinos decouple from matter at smaller radii than at
1526: the equator. While MGFLD must handle the decoupling and increased
1527: forward-peaking of the radiation field via the flux limiter, \sn\
1528: can track it self-consistently.  For $\nu_e$ neutrinos and along the
1529: poles, \sn\ predicts significantly greater mean inverse flux factors
1530: with shallower slopes than MGFLD, indicating a more gradual transition
1531: to free streaming than predicted by the flux limiter.  In the radial
1532: interval of $\sim$60--100~km, the relative difference is
1533: $\sim$12--19\%, decreasing to $\sim$6--12\% out to 200~km.  In
1534: equatorial regions, the $\nu_e$ radiation field is somewhat more
1535: forward-peaked in the \sn\ calculation at radii below $\sim$120~km,
1536: beyond which MGFLD transitions quickly to free streaming while \sn\
1537: approaches it more gradually, exhibiting $\sim$6--8\% larger mean
1538: inverse flux factors in the outer postshock region. For $\bar{\nu}_e$
1539: neutrinos, the behavior of the mean inverse flux factors in polar
1540: regions essentially mirrors that observed for the $\nu_e$s.  In
1541: equatorial regions, the \sn\ mean inverse flux factor of the
1542: $\bar{\nu}_e$s stays below that using MGFLD out to 165~km, beyond which
1543: the MGFLD $\bar{\nu}_e$ radiation field rapidly transitions
1544: to free streaming. At 180~km, the MGFLD $\bar{\nu_e}$ mean inverse 
1545: flux factor is $\sim$1\% smaller than that predicted by \sn. At 220~km,
1546: this difference has grown to $\sim$5\%.
1547: 
1548: 
1549: Having established the overall neutrino radiation-field
1550: characteristics in the 160-ms postbounce snapshot of model s20.$\pi$,
1551: we now turn our focus to the neutrino cooling and heating rates in
1552: this model. We have found little difference in the net neutrino
1553: heating between \sn\ and MGFLD variants in the 160-ms postbounce
1554: snapshot of the nonrotating model s20.nr. However, based on the
1555: differences between \sn\ and MGFLD in neutrino fluxes, RMS energies, and
1556: flux factors we have highlighted in this section, we may expect
1557: to find significant differences in the neutrino heating rates for
1558: model s20.$\pi$.
1559: 
1560: 
1561: Figure~\ref{fig:s20pi_netgain} depicts 2D colormaps of the neutrino
1562: energy gain and loss rate per unit mass (accounting for all energies
1563: and species), computed for the 160-ms postbounce snapshot of model
1564: s20.$\pi$ using \sn\ (left panel) and MGFLD (right panel). At low
1565: latitudes near the equator, \sn\ and MGFLD agree very well to the eye.
1566: In regions near the pole, both MGFLD and \sn\ show a pronounced region
1567: of net loss at $z$-coordinates between $\sim$40 and $\sim$80~km,
1568: beyond which a region of net gain (colors light blue and green to red)
1569: prevails out to the shock position at $\sim$230~km. While the
1570: gain region has roughly the same physical extent in MGFLD and \sn, the
1571: latter yields significantly higher energy deposition rates.  This is
1572: particularly the case in the lower gain region at polar angles below
1573: $\sim$20$^\circ$ and at radii between $\sim$80 and 150~km, where the
1574: \sn\ gain rate is larger by a factor of two and more. The top panel in
1575: Fig.~\ref{fig:s20pi_netgain2} provides a more quantitative comparison
1576: of \sn\ and MGFLD gain/loss rates, since it contrasts average specific
1577: gain/loss profiles obtained from polar and equatorial 20$^\circ$
1578: wedges.  In the polar region, the \sn\ gain region begins at a radius
1579: of $\sim$80~km (MGFLD: $\sim$88~km) and the \sn\ specific gain rate
1580: magnitude exceeds the MGFLD numbers by a factor of 2.6 at 100~km,
1581: increasing to 3.2 at 200~km.  Near the equator, net energy deposition
1582: occurs only in a small radial interval of $\sim$90--120~km and the
1583: MGFLD specific gain rate is larger by 80\% at 95~km, 41\% at 100~km,
1584: and 26\% at 110~km. The net energy loss between $\sim$120--210~km
1585: (captured by both \sn\ and MGFLD) results from strong electron capture
1586: that dominates energy deposition by neutrino absorption.
1587: 
1588: \begin{figure*}[t]
1589: \centering
1590: \includegraphics[width=8.5cm]{f17a.ps}
1591: \includegraphics[width=8.5cm]{f17b.ps}
1592: \caption{{\bf Left:} Evolution of the total neutrino net gain rate as
1593: a function of postbounce time in the \sn\ and MGFLD variants of models
1594: s20.nr and s20.$\pi$. At postbounce times before $\sim$300~ms in model
1595: s20.nr, \sn\ yields a net gain rate that is larger by (on average)
1596: $\sim$10--15\% than that predicted by MGFLD. As the SASI becomes more
1597: pronounced at postbounce times $\sgreat$~300~ms, the \sn\ net gain
1598: begins to more significantly exceed that of MGFLD, averaging out at
1599: $\sim$20-30\% larger values than the MGFLD net gain rate.
1600: %%
1601: In model s20.$\pi$, \sn\ and MGFLD net gain rates stay very
1602: close in the first $\sim$30~ms of evolution, yet depart when
1603: the \sn\ variant approaches its new dynamical equilibrium
1604: (see Fig.~\ref{fig:shockrad}) and provides for a larger gain region
1605: (mass and volume). This leads to a net gain rate that is larger
1606: by $\sim$20-25\% (on average in the postbounce interval
1607: from 200--350~ms). At later times, the MGFLD calculation, approaching
1608: the \sn\ variant's postshock extent (Fig.~\ref{fig:shockrad})
1609: produces larger net gain rates due to its larger equatorial neutrino
1610: fluxes at similar hydrodynamic configuration.
1611: %%
1612: %%
1613: %%
1614: %%
1615: {\bf Right:} Heating efficiency evolution in the two models with their
1616: \sn\ and MGFLD variants. We define the heating efficiency as the 
1617: ratio of total neutrino net gain rate and the sum of electron and
1618: anti-electron neutrino luminosities. 
1619: \label{fig:s20_gain}}
1620: \end{figure*}
1621: 
1622: 
1623: 
1624: The observed local differences in neutrino energy deposition between
1625: \sn\ and MGFLD are due primarily to the vastly different degree to
1626: which the two schemes capture the global pole-equator asymmetry of the
1627: radiation field in the rapidly-rotating postbounce supernova core of
1628: model s20.$\pi$. \sn\ yields much larger fluxes in the polar direction
1629: than MGFLD, but predicts lower neutrino fluxes in equatorial
1630: regions~(cf.\ Fig.~\ref{fig:s20pi_lum160}). Differences in the radial
1631: mean-inverse flux factors and RMS energies are much smaller, and,
1632: hence, are of only secondary importance.  The \sn\ steady-state
1633: snapshot yields an integrated gain rate of 1.603~B~s$^{-1}$ while
1634: MGFLD predicts 1.637~B~s$^{-1}$ for the s20.$\pi$ snapshot under
1635: consideration.  This corresponds to $\sim$2.1\% \emph{more} energy
1636: deposition per unit time in the MGFLD calculation.  Given the above
1637: discussion, the reader may be surprised by these numbers.  The
1638: explanation consists of two factors.  Owing to rotation, the amount of
1639: mass per unit volume (i.e., the rest-mass density) is higher at any
1640: given equatorial radius than at the same radius in the polar
1641: direction. Plotting the neutrino gain/loss rate per unit volume
1642: instead of per unit gram, the bottom panel of
1643: Fig.~\ref{fig:s20pi_netgain2} clearly shows the rotation-induced
1644: enhancement of the energy deposition (per unit volume) near the
1645: equator and the larger gain rate per unit volume predicted by MGFLD at
1646: small to intermediate radii.  The second factor is the simple fact
1647: that the volume of the equatorial gain regions is much larger than
1648: that of the polar gain regions.
1649: 
1650: 
1651: As we shall discuss in the following section, the large local
1652: differences in neutrino heating between the \sn\ and MGFLD snapshots
1653: have a dynamical consequence for the rapidly rotating model and lead
1654: to a significant polar expansion of the shock in the \sn\ postbounce
1655: evolution calculation.
1656: 
1657: 
1658: 
1659: \section{Results: Evolution Calculations}
1660: \label{section:evolution}
1661: 
1662: In order to study differences between \sn\ and MGFLD in a
1663: time-dependent postbounce setting, we follow our relaxed 160-ms
1664: \sn\ models in fully coupled radiation-hydrodynamics
1665: fashion for $\sim$340~ms (model s20.nr) and 390~ms
1666: (model s20.$\pi$) of postbounce time. In parallel with the \sn\ runs, we
1667: continue their MGFLD counterparts for the same time span.
1668: 
1669: %\vspace*{.5cm}
1670: \subsection{Model s20.nr}
1671: 
1672: \begin{figure*}[t]
1673: \centering
1674: \includegraphics[width=8.5cm]{f18a.ps}
1675: \includegraphics[width=8.3cm]{f18b.ps}
1676: \caption{{\bf Left:} $\nu_e$ ``luminosities'' ($4\pi r^2 F_r$) as a
1677: function of postbounce time as seen by observers located at a
1678: spherical radius of 250~km along the north pole (black lines), south
1679: pole (green lines), and in the equatorial plane (red lines) in \sn\
1680: (solid lines) and MGFLD (dashed lines) variants of model s20.nr (thin
1681: lines) and s20.$\pi$ (thick lines). Note that the south pole, north
1682: pole, and equator MGFLD ``luminosities'' in model s20.nr (thin dashed
1683: lines) are very similar. Their lines are indistinguishable.  The same
1684: holds for the south and north pole MGFLD ``luminosities'' in model
1685: s20.$\pi$ (thick black and green dashed lines).
1686: %%
1687: %%
1688: %%
1689: %%
1690: {\bf Right:} Angle-averaged RMS energies of $\nu_e$ (solid lines) and
1691: $\bar{\nu}_e$ (dashed lines) neutrinos as a function of postbounce
1692: time in the \sn\ and MGFLD simulations the two 
1693: models. \sn\ predicts systematically higher RMS neutrino
1694: energies in both models.
1695: \label{fig:s20_lum_rms}}
1696: \end{figure*}
1697: 
1698: 
1699: \begin{figure*}
1700: \centering
1701: \includegraphics[width=8.5cm]{f19a.ps}
1702: \includegraphics[width=8.5cm]{f19b.ps}
1703: \caption{{\bf Left}: Average shock radii as a function of postbounce
1704: time in \sn\ (red) and MGFLD (black) variants of the nonrotating model
1705: s20.nr.  Also shown are the overall average shock radius, the average
1706: of south-pole and north-pole shock radii, and the equatorial shock
1707: radius for the rapidly spinning model s20.$\pi$, again for \sn\ (blue)
1708: and MGFLD (green). In model s20.nr, MGFLD and \sn\ show little
1709: quantitative deviation from each other. In the s20.$\pi$ evolution,
1710: however, a significant increase in the various shock radii is
1711: noticable right at the beginning of the time-dependent \sn\
1712: calculation. At later times MGFLD catches up and the average shock
1713: radii approach each other.  The \sn\ variant exhibits larger
1714: variations, indicating stronger SASI-like shock excursions. 
1715: %%
1716: %%
1717: %%
1718: {\bf Right}: Evolution of the north-pole (positive) and south-pole
1719: (negative) shock radii for the \sn\ and MGFLD variants of the two
1720: models. Since the lowest-order and dominant mode of the 2D SASI is the
1721: $\ell$=1 polar sloshing mode, the polar shock radii are good
1722: indicators of its strength and periodicity.  Note the initial
1723: suppression, but late-time development of SASI-like polar shock
1724: excursions in the rotating model. 
1725: \label{fig:shockrad}}
1726: \end{figure*}
1727: 
1728: 
1729: \begin{figure*}[t]
1730: \centering
1731: \includegraphics[width=5.8cm]{f20a.ps}
1732: \includegraphics[width=5.8cm]{f20b.ps}
1733: \includegraphics[width=5.8cm]{f20c.ps}
1734: 
1735: \includegraphics[width=5.8cm]{f20d.ps}
1736: \includegraphics[width=5.8cm]{f20e.ps}
1737: \includegraphics[width=5.8cm]{f20f.ps}
1738: \caption{2D entropy colormaps portraying the postbounce evolution of
1739: model s20.nr between 160~ms (top-left panel) and 500~ms (bottom-right
1740: panel) after core bounce. Fluid-velocity vectors are superposed to
1741: provide an impression of the flow. Each panel's left-hand side
1742: corresponds to the MGFLD calculation and each panel's right-hand side
1743: shows the \sn\ result. The time of each panel is given relative to the
1744: time of core bounce. The sequence of panels portrays the canonical
1745: development of the SASI in the nonrotating axisymmetric context. \sn\
1746: and MGFLD evolution agree very well in the early SASI phases, but
1747: deviate in detail at later times, while still
1748: exhibiting the same overall SASI dynamics. \label{fig:s20nr_evol2D}}
1749: \end{figure*}
1750: 
1751: 
1752: Since we begin the MGFLD and \sn\ calculations from an identical
1753: hydrodynamic configuration at 160~ms after bounce, any qualitative
1754: or quantitative differences in their evolutions must ultimately be due
1755: to differences in the neutrino heating and cooling between \sn\ and
1756: MGFLD.
1757: 
1758: In the left panel of Fig.~\ref{fig:s20_gain}, we display the time
1759: evolution of the integral neutrino energy deposition (net gain) in the
1760: gain region of model s20.nr. The net gain systematically declines at
1761: early postbounce times, due (a) to the declining neutrino luminosity
1762: and (b) to the rapid settling of accreting material into the net loss
1763: region near the PNS core (cf. Fig.\ 7 of \citealt{marek:07}).  
1764: At later times, SASI-modulated convection increases
1765: the dwell time of accreting outer core material in the gain layer and
1766: the slope of the net gain evolution flattens. Both \sn\ and MGFLD
1767: track these systematics without qualitative difference. The \sn\
1768: calculation predicts on average $\sim$5--10\% higher net gain in the
1769: postbounce interval from $\sim$160~ms to $\sim$220~ms.  Between
1770: $\sim$220~ms and $\sim$280~ms, MGFLD and \sn\ net gain rates agree to
1771: within a few percent. Towards the end of this interval, the net gain
1772: of the \sn\ calculation grows and settles at values that are on
1773: average 20--30\% higher than those of the MGFLD run. This trend is
1774: confirmed by the right panel of Fig.~\ref{fig:s20_gain}, which portrays
1775: the heating efficiency, defined as the ratio of net gain to the
1776: sum of $\nu_e$ and $\bar{\nu}_e$ luminosities. 
1777: 
1778: 
1779: The left panel of Fig.~\ref{fig:s20_lum_rms} depicts the temporal
1780: evolution of the $\nu_e$ ``luminosities'' ($4\pi r^2 F_r$) as seen by
1781: observers situated at 250~km along the north pole and south pole as
1782: well as in the equatorial plane of models s20.nr and s20.$\pi$. Here
1783: we focus on model s20.nr and note for the \sn\ variant that north pole
1784: (thin solid black lines) and south pole (thin solid green lines)
1785: ``luminosities'' agree (on average) in magnitude, but exhibit
1786: oscillations about their temporal average that are roughly out of
1787: phase by half a cycle. The MGFLD calculation (thin dashed lines), on
1788: the other hand, does exhibit some short-period ``luminosity''
1789: variations, yet shows no appreciable difference between poles and
1790: equator.
1791: 
1792: 
1793: The time at which \sn\ begins to yield systematically larger neutrino
1794: heating rates (Fig.~\ref{fig:s20_gain}) coincides with the growth of
1795: the SASI-related shock excursions to large amplitudes 
1796: (Fig.~\ref{fig:shockrad}). This suggests that the increased
1797: heating is related at least in part to the \sn variant's ability to
1798: better capture radiation field asymmetries (see also the discussion
1799: in \S\ref{section:s20pi}), induced at late times by the rapidly varying
1800: shock and postshock hydrodynamics in this model. Other factors that
1801: contribute to the increased heating in the \sn\ calculation are the
1802: higher RMS neutrino energies (by $\sim$5\%; shown in the right panel
1803: of Fig.~\ref{fig:s20_lum_rms}) and the more gradual transition of the
1804: \sn\ neutrino radiation field to free streaming in the postshock
1805: region (see \S\ref{section:s20nr}).
1806: 
1807: 
1808: Figure~\ref{fig:s20nr_evol2D} contrasts \sn\ and MGFLD simulations of
1809: model s20.nr by means of colormaps depicting the specific entropy
1810: distributions in the two variants. To visualize the hydrodynamic flow,
1811: we superpose fluid velocity vectors.  Each panel of this figure
1812: corresponds to a specific postbounce time and each panel's
1813: left-hand-side depicts the state of the MGFLD calculation, while the
1814: right-hand-side depicts the corresponding \sn\ calculation. The figure
1815: covers a postbounce interval from 160~ms (top left) to 500~ms (bottom
1816: right). At the beginning of the runs, the SASI-driven deviation from
1817: sphericity of the stalled shock is mild, but grows with time, showing
1818: $\ell=1$ excursions now generally recognized as characteristic of the
1819: SASI\footnote{At least in detailed 2D models. \cite{iwakami:08}
1820: carried out an exploratory 3D numerical study with nonrotating 
1821: progenitors that suggests that in
1822: the 3D case the $\ell$=1 dominance still obtains, yet reaches smaller
1823: relative amplitudes, since not only higher $\ell$ modes, but also $m$
1824: modes, may now contain power. However, \cite{yamasaki:08}, who
1825: performed a perturbative study without symmetry constraints, argued
1826: that in the 3D case with rotation, a dominant $m=1$ ($m=2$) mode is
1827: likely to emerge in the case of slow (rapid) rotation.}
1828: \citep{scheck:08, marek:07,bruenn:06,burrows:07a}.
1829: 
1830: As expected from the discussion of the s20.nr 160-ms postbounce
1831: steady-state snapshot in \S\ref{section:snapshots}, \sn\ and MGFLD
1832: variants of this model do not differ significantly in the early SASI
1833: phase. However, at later SASI stages, in particular at postbounce
1834: times $\sgreat$~300--350~ms, the simulations diverge, showing different
1835: local qualitative and quantitative behavior within the overall SASI
1836: theme. This is also reflected in Fig.~\ref{fig:shockrad}, which
1837: depicts the evolution of the average shock radius, as well as the
1838: shock radii along north pole and south pole.  The shock positions in
1839: the \sn\ and MGFLD simulations remain close and the SASI stays
1840: practically in phase (right panel of Fig.~\ref{fig:shockrad}) until
1841: $\sim$~350~ms after bounce. Only then do they begin to show
1842: significant departures from each other. The SASI in the \sn\
1843: calculation appears more pronounced at later times, exhibiting larger
1844: local (in time) shock excursions.  Yet, quite surprisingly, given the
1845: significant increase in neutrino energy deposition, the \sn\
1846: calculation does not exhibit any increase in the average shock radius,
1847: nor does it appear to be any closer to explosion than its
1848: MGFLD counterpart.
1849: 
1850: 
1851: 
1852: \subsection{Model s20.$\pi$}
1853: 
1854: 
1855: \begin{figure*}[t]
1856: \centering
1857: \includegraphics[width=5.8cm]{f21a.ps}
1858: \includegraphics[width=5.8cm]{f21b.ps}
1859: \includegraphics[width=5.8cm]{f21c.ps}
1860: 
1861: \includegraphics[width=5.8cm]{f21d.ps}
1862: \includegraphics[width=5.8cm]{f21e.ps}
1863: \includegraphics[width=5.8cm]{f21f.ps}
1864: \caption{2D entropy colormaps portraying the postbounce evolution of
1865: the rapidly-rotating model s20.$\pi$ between 160~ms (top-left panel)
1866: and 550~ms (bottom-right panel) after core bounce. Fluid-velocity
1867: vectors are superposed to relay an impression of the flow and convey
1868: the partial suppression of convective overturn in regions of
1869: positive specific angular momentum gradient. As in
1870: Fig.~\ref{fig:s20nr_evol2D}, we plot the MGFLD result on the left-hand
1871: side and the \sn\ result on the right-hand side of each panel. Easily
1872: discernible is the immediate increase in the polar shock radius
1873: in the \sn\ calculation.  This is a direct consequence of the
1874: increased polar neutrino heating in this variant
1875: (Figs.~\ref{fig:s20pi_netgain} and \ref{fig:s20pi_netgain2}). At
1876: intermediate times, \sn\ and MGFLD shock positions grow closer, but 
1877: later on in the postbounce evolution, the \sn\ variant begins to
1878: develop larger top-bottom SASI-like asymmetry and polar shock excursions at
1879: earlier time than its MGFLD counterpart.\label{fig:s20pi_evol2D}}
1880: \end{figure*}
1881: 
1882: The diagnosis of the radiation-hydrodynamic evolution of the rapidly
1883: spinning model s20.$\pi$ is less straightforward than for the
1884: nonrotating model s20.nr. As discussed in \S\ref{section:s20pi}, rotation
1885: creates a global pole-equator asymmetry in the hydrodynamics of this
1886: model. MGFLD and \sn\ track the effect of globally asymmetric
1887: matter distributions on the neutrino radiation field to different
1888: degrees. In the steady-state snapshot at 160~ms, \sn\ predicts
1889: stronger neutrino heating in polar regions, yet weaker heating 
1890: in the higher-density, larger-volume equatorial regions.
1891: 
1892: The polar, equatorial, and angle-averaged shock positions portrayed by
1893: Fig.~\ref{fig:shockrad} show that the hydrodynamics responds
1894: immediately to the increased polar heating in the \sn\ calculation by
1895: a pronounced expansion of the shock along the poles. This
1896: expansion lasts for $\sim$40~ms, after which the shock has
1897: expanded by $\sim$20\% from $\sim$230~km to $\sim$275~km on both
1898: poles. It stagnates at this radius and subsequently contracts
1899: again when feedback of the hydrodynamics to the neutrino microphysics
1900: leads to increased cooling (cf.\ the increased polar neutrino emission
1901: shown in Fig.~\ref{fig:s20_lum_rms}). The increased
1902: postshock volume also results in a larger gain region and increased
1903: (compared to MGFLD) total neutrino energy deposition and heating
1904: efficiency. However, this increased heating is not able to sustain the
1905: large postshock volume. The shock slowly recontracts in the
1906: postbounce interval from $\sim$250~ms to $\sim$380~ms and eventually
1907: settles at radii similar to those obtained by the MGFLD shock.
1908: 
1909: In Fig.~21, we present a sequence of 2D entropy colormaps with
1910: superposed velocity vectors, portraying the postbounce evolution of
1911: model s20.$\pi$ from 160~ms on. The rapid rotation in this model not
1912: only partially stabilizes convection, but also weakens and delays the
1913: growth of the characteristic $\ell=1$ SASI\footnote{But see the work
1914: of \cite{yamasaki:08}, who find via perturbative analysis that in 3D,
1915: rotation enhances the development of azimuthal $m=1$ and $m=2$
1916: SASI-related spiral structure.}. Since larger shock radii are
1917: associated with an increased growth rate of the SASI (e.g.,
1918: \citealt{foglizzo:07,scheck:08}), the \sn\ variant begins to develop
1919: periodic shock excursions along the symmetry axis at much earlier
1920: times than the MGFLD simulation (Fig.~\ref{fig:shockrad}).  However,
1921: at times later than $\sim$400~ms, the MGFLD model picks up the
1922: large-amplitude SASI as well and both calculations exhibit large-scale
1923: radial shock excursions beyond $\sim$400~km along the pole
1924: (Fig.~\ref{fig:shockrad}).  The average shock radius increases in both
1925: calculations in this late postbounce phase.  We observe neither such
1926: large shock excursions nor a systematic late-time increase of the
1927: average shock radius in the nonrotating model. The observed behavior
1928: is most likely due to the rapid rotation and the resulting rarefaction
1929: of the polar regions that reduces, in particular at late times, the
1930: ram pressure of accretion and allows for the more pronounced SASI.
1931: 
1932: In the left panel of Fig.~\ref{fig:s20_lum_rms}, we contrast the
1933: $\nu_e$ ``luminosities'' ($4\pi r^2 F_r$) seen by observers located at
1934: a radius of 250~km above the north pole, the south pole, and in the
1935: equatorial plane of model s20.$\pi$. The MGFLD variant predicts a
1936: pole-equator flux asymmetry of $\sless$10\% that is roughly
1937: constant with time. The \sn\ calculation yields a very different
1938: picture. Polar and equatorial ``luminosities'' at 250~km (i.e., near the
1939: shock) are vastly different (cf. \S\ref{section:s20pi}).  Over time,
1940: the equatorial ``luminosity'' decreases while the ``luminosity'' along the
1941: poles is enhanced.  At $\sim$200~ms, polar and equatorial ``luminosities''
1942: differ by a factor of $\sim$3. By $\sim$500~ms, this factor has grown
1943: to 4.  In addition, the \sn\ simulation shows SASI-induced variations
1944: in north and south-pole ``luminosities'' that grow to $\sim$3--5\% at late
1945: times and are not tracked in the MGFLD variant. These variations are
1946: akin those reported for the nonrotating model s20.nr, yet have longer
1947: periods, since the large shock excursions in model s20.$\pi$ occur on
1948: longer timescales.
1949: 
1950: As in the nonrotating model, we also find in model s20.$\pi$ that \sn\
1951: yields systematically higher RMS neutrino energies for all species
1952: and at all times. However, as shown in the right panel of
1953: Fig.~\ref{fig:s20_lum_rms}, the angle-averaged RMS energies do not
1954: exhibit a significant increase in the time interval covered by our
1955: simulations. This, again, is due to rapid rotation which slows down the
1956: PNS's contraction. Not shown in Fig.~\ref{fig:s20_lum_rms}, but
1957: present in the \sn\ variant throughout its postbounce evolution, are 
1958: $\sim$10--20\% (roughly constant in time and independent of
1959: species) higher RMS energies for neutrinos emitted from polar regions
1960: compared to those emitted from the PNS equator. This is consistent
1961: with our analysis of the neutrino spectra and RMS neutrino energies
1962: for the 160~ms postbounce steady-state snapshot presented in
1963: \S\ref{section:s20pi}.
1964: 
1965: We end our postbounce simulations of model s20.$\pi$ with \sn\ and
1966: MGFLD at 550~ms after bounce. Though within roughly the same
1967: qualitative picture, the two approaches to neutrino transport yield
1968: appreciable differences in the postbounce radiation-hydrodynamics
1969: evolutions.  Importantly, and in contrast to our findings for the
1970: nonrotating model, \sn\ in model s20.$\pi$ does not lead to
1971: systematically higher integral neutrino energy deposition, and at late
1972: postbounce times, shows a volume-integrated heating rate that is even
1973: $\sim$30\% lower (on average) than in its MGFLD counterpart.
1974: 
1975: 
1976: \section{Summary and Discussion}
1977: \label{section:summary}
1978: 
1979: Using the code VULCAN/2D \citep{livne:04,burrows:07a,livne:07}, we
1980: perform long-term full-2D multi-angle, multi-group neutrino
1981: radiation-hydrodynamic calculations in the core-collapse supernova
1982: context. Based on postbounce hydrodynamic configurations from MGFLD
1983: simulations, we first compute 2D angle-dependent (\sn) steady-state
1984: solutions for models without precollapse rotation and with rapid
1985: rotation ($\Omega_0 = \pi$~rad~s$^{-1}$). From these snapshots, we
1986: numerically follow the radiation-hydrodynamics evolution with \sn\
1987: neutrino transport, tracking the nonrotating model to 500~ms and the
1988: rotating model to 550~ms after bounce.
1989: 
1990: Done for the first time in 2D, we investigate in
1991: detail the angle-dependent specific intensities and neutrino radiation
1992: fields. We compute angular moments of the specific
1993: intensity, including the Eddington tensor, and introduce Hammer-type
1994: map projections to visualize the angle dependence of the specific
1995: intensity. These we employ to demonstrate the decoupling systematics
1996: of the neutrinos and the gradual transition to free-streaming of the
1997: radiation fields with decreasing optical depth.
1998: 
1999: We compare our \sn\ simulations with MGFLD counterparts.  We find for
2000: both models and at all times that the \sn\ specific intensity
2001: distributions transition less rapidly from isotropy to free-streaming
2002: in the semi-transparent outer postshock regions.  \sn\ yields mean
2003: inverse flux factors and RMS neutrino energies in these regions that
2004: are $\sim$10\% larger than those obtained with MGFLD.  In the context
2005: of the neutrino mechanism of core-collapse supernova explosions,
2006: differences in the net neutrino energy deposition rates between MGFLD
2007: and multi-angle \sn\ transport are of greatest interest.  In the
2008: quasi-spherical early postbounce phase of the nonrotating model, we
2009: find that \sn\ predicts a 5--10\% greater neutrino energy deposition
2010: rate than MGFLD.  At later times, when the SASI has reached large
2011: amplitudes and globally deforms the postshock region, we find that
2012: \sn\ yields consistently larger (up to 30\% on average) 
2013: energy depositions and leads to significantly larger temporary
2014: shock excursions around average shock radii that do not depart
2015: much from those in the MGFLD calculation.
2016: 
2017: Convection on small and intermediate scales and SASI on large
2018: scales, are the key agents of the breaking of spherical symmetry in
2019: nonrotating (or slowly rotating) core-collapse supernovae. While we
2020: observe no large qualitative differences in the growth and dynamical
2021: evolutions of convection and SASI between nonrotating \sn\ and MGFLD
2022: models, we find that the imprint of the asymmetric hydrodynamics on
2023: the neutrino radiation fields is captured with greater detail by the
2024: multi-angle transport scheme. For the late-time, heavily
2025: SASI-distorted postbounce core, \sn\ predicts asymptotic neutrino
2026: fluxes that have variations with time and angle of 5--10\% in
2027: magnitude. MGFLD is able to capture the temporal variations of the
2028: neutrino luminosity, but smoothes out the angular flux variations at
2029: large radii/low optical depths.
2030: 
2031: 
2032: Rapid rotation leads to large deviations from spherical symmetry and a
2033: rotationally-deformed PNS emits, by von Zeipel's law of gravity
2034: darkening, a greater neutrino flux along its rotational axis than
2035: through its equatorial regions
2036: \citep{janka:89a,janka:89b,kotake:03,walder:05,buras:06b,
2037: dessart:06b}.  We find that both 2D MGFLD and \sn\ yield similar
2038: radiation fields and pole-equator flux ratios at radii smaller than
2039: $\sim$100~km. At larger radii, the MGFLD radiation fields sphericize
2040: and show little pole-equator asymmetry in their asymptotic
2041: variables. \sn, on the other hand, captures large pole-equator flux
2042: ratios of up to 4:1 at late times and predicts polar neutrino spectra
2043: that are harder in peak energy (RMS energy) than on the equator by up
2044: to 30\% (10--15\%) for $\nu_e$ neutrinos, and somewhat less for the
2045: other species. All this results in a neutrino energy deposition rate
2046: per unit mass in polar regions that is locally up to $\sim$2.5--3
2047: times higher when multi-angle transport is used. This increased polar
2048: neutrino heating has a dynamical effect on the postbounce evolution,
2049: leading to rapid shock expansion in the polar regions and an earlier
2050: onset of the (initially) rotationally-weakened SASI. However, at late
2051: times, the SASI in the MGFLD calculation catches up and yields shock
2052: excursions of a similar magnitude.
2053: 
2054: 
2055: In summary, our results show that 2D multi-angle neutrino transport
2056: manifests interesting differences with 2D MGFLD when addressing local
2057: and global radiation field asymmetries associated with rapid rotation
2058: and the non-linear SASI at late postbounce times. In addition,
2059: multi-angle transport results in enhanced neutrino energy
2060: deposition. The latter is most significant in the polar regions of
2061: rapidly rotating postbounce configurations and affects dynamically the
2062: postbounce evolution, including the growth of the SASI. However, in
2063: the large postbounce interval covered by our simulations, the local
2064: and global differences between multi-angle transport and MGFLD
2065: calculations do not appear large enough to alter the overall
2066: simulation outcome. Importantly, the multi-angle models do not appear
2067: to be closer to explosion than their MGFLD counterparts.
2068: 
2069: Although we neglect velocity-dependent transport terms and coupling of
2070: neutrino energy bins, we do not expect our conclusions to be altered
2071: by their inclusion, since they are not likely to affect significantly
2072: the differences between multi-angle transport and MGFLD. Further
2073: significant limitations of our present study are the neglect of
2074: general relativistic and MHD effects, the restriction to only one
2075: finite-temperature nuclear EOS, the limited resolution in
2076: momentum-space imposed by the computational cost of multi-angle
2077: calculations, and the use of two spatial dimensions, plus
2078: rotation. In the future, we will investigate
2079: the dependence of our results (e.g., heating rates, radiation-field
2080: asymmetries etc.) on the choice of flux limiter and will consider
2081: different progenitor models.
2082: 
2083: The core-collapse supernova problem is one of many feedbacks.  Larger
2084: heating rates and heating efficiencies than found in our models appear
2085: to be necessary to break the feedback cycle between neutrino radiation
2086: fields and hydrodynamics, revive the stalled shock, and unbind the
2087: supernova envelope -- if the neutrino mechanism is to obtain in the
2088: way presently envisioned. Future work will have to go beyond the
2089: limitation of axisymmetry and must address in detail the entire
2090: ensemble of possible factors relevant in the supernova problem,
2091: including, but not limited to, 3D dynamics, multi-angle neutrino
2092: transport with velocity dependence and inelastic $\nu_e$-$e^-$
2093: scattering, progenitor structure, rotational configuration,
2094: magnetohydrodynamics, convection, the SASI, PNS g-modes, general
2095: relativity, the nuclear EOS, and neutrino-matter interactions.
2096: 
2097: 
2098: 
2099: \section*{Acknowledgements}
2100: 
2101: We acknowledge helpful discussions with and input from Jeremiah
2102: Murphy, Ivan Hubeny, Casey Meakin, Jim Lattimer, Alan Calder, Stan
2103: Woosley, Ed Seidel, Harry Dimmelmeier, H.-Thomas Janka, 
2104: Kei Kotake, Thierry Foglizzo, Ewald
2105: M\"uller, Bernhard M\"uller, Martin Obergaulinger,
2106: Benjamin~D.~Oppenheimer, Thomas Marquart, and Erik Schnetter.  This
2107: work was partially supported by the Scientific Discovery through
2108: Advanced Computing (SciDAC) program of the US Department of Energy
2109: under grant numbers DE-FC02-01ER41184 and DE-FC02-06ER41452.
2110: C.D.O. acknowledges support through a Joint Institute for Nuclear
2111: Astrophysics postdoctoral fellowship, sub-award no.~61-5292UA of NFS
2112: award no.~86-6004791. E.L. acknowledges support by the Israel Science
2113: Foundation (grant 805/04).  The computations were performed at the
2114: local Arizona Beowulf cluster, on the Columbia SGI Altix machine at
2115: the Ames center of the NASA High End Computing Program, at the
2116: National Center for Supercomputing Applications (NCSA) under Teragrid
2117: computer time grant TG-MCA02N014, at the Center for Computation and
2118: Technology at Louisiana State University, and at the National Energy
2119: Research Scientific Computing Center (NERSC), which is supported by
2120: the Office of Science of the US Department of Energy under contract
2121: DE-AC03-76SF00098.
2122: 
2123: \newpage
2124: 
2125: \begin{thebibliography}{89}
2126: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
2127: 
2128: \bibitem[{Adams \& Larsen(2002)}]{adamslarsen:02}
2129: Adams, M.~L., \& Larsen, E.~W. 2002, Prog. Nuc. En., 40, 3
2130: 
2131: \bibitem[{{Akiyama} {et~al.}(2003){Akiyama}, {Wheeler}, {Meier}, \&
2132:   {Lichtenstadt}}]{akiyama:03}
2133: {Akiyama}, S., {Wheeler}, J.~C., {Meier}, D.~L., \& {Lichtenstadt}, I. 2003,
2134:   \apj, 584, 954
2135: 
2136: \bibitem[{{Arnett}(1966)}]{arnett:66}
2137: {Arnett}, W.~D. 1966, Canadian Journal of Physics, 44, 2553
2138: 
2139: \bibitem[{{Balbus} \& {Hawley}(1991)}]{bh:91}
2140: {Balbus}, S.~A., \& {Hawley}, J.~F. 1991, \apj, 376, 214
2141: 
2142: \bibitem[{{Baron} {et~al.}(1989){Baron}, {Myra}, {Cooperstein}, \& {van den
2143:   Horn}}]{baron:89}
2144: {Baron}, E., {Myra}, E.~S., {Cooperstein}, J., \& {van den Horn}, L.~J. 1989,
2145:   \apj, 339, 978
2146: 
2147: \bibitem[{{Bethe}(1990)}]{bethe:90}
2148: {Bethe}, H.~A. 1990, Rev. Mod. Phys., 62, 801
2149: 
2150: \bibitem[{{Bethe} \& {Wilson}(1985)}]{bethewilson:85}
2151: {Bethe}, H.~A., \& {Wilson}, J.~R. 1985, \apj, 295, 14
2152: 
2153: \bibitem[{{Blondin} {et~al.}(2003){Blondin}, {Mezzacappa}, \&
2154:   {DeMarino}}]{blondin:03}
2155: {Blondin}, J.~M., {Mezzacappa}, A., \& {DeMarino}, C. 2003, \apj, 584, 971
2156: 
2157: \bibitem[{{Bowers} \& {Wilson}(1982)}]{bowerswilson:82}
2158: {Bowers}, R.~L., \& {Wilson}, J.~R. 1982, \apjs, 50, 115
2159: 
2160: \bibitem[{{Bruenn}(1985)}]{bruenn:85}
2161: {Bruenn}, S.~W. 1985, \apjs, 58, 771
2162: 
2163: \bibitem[{{Bruenn} {et~al.}(2006){Bruenn}, {Dirk}, {Mezzacappa}, {Hayes},
2164:   {Blondin}, {Hix}, \& {Messer}}]{bruenn:06}
2165: {Bruenn}, S.~W., {Dirk}, C.~J., {Mezzacappa}, A., {Hayes}, J.~C., {Blondin},
2166:   J.~M., {Hix}, W.~R., \& {Messer}, O.~E.~B. 2006, J. Phys. Conf. Ser., 46, 393
2167: 
2168: \bibitem[{{Buras} {et~al.}(2006{\natexlab{a}}){Buras}, {Janka}, {Rampp}, \&
2169:   {Kifonidis}}]{buras:06b}
2170: {Buras}, R., {Janka}, H.-T., {Rampp}, M., \& {Kifonidis}, K.
2171:   2006{\natexlab{a}}, \aap, 457, 281
2172: 
2173: \bibitem[{{Buras} {et~al.}(2006{\natexlab{b}}){Buras}, {Rampp}, {Janka}, \&
2174:   {Kifonidis}}]{buras:06a}
2175: {Buras}, R., {Rampp}, M., {Janka}, H.-T., \& {Kifonidis}, K.
2176:   2006{\natexlab{b}}, \aap, 447, 1049
2177: 
2178: \bibitem[{{Burrows} {et~al.}(2007{\natexlab{a}}){Burrows}, {Dessart}, \&
2179:   {Livne}}]{burrows:07c}
2180: {Burrows}, A., {Dessart}, L., \& {Livne}, E. 2007{\natexlab{a}}, in {AIP
2181:   Conference Series}, ed. S.~{Immler} \& R.~{McCray}, Vol. 937, 370
2182: 
2183: \bibitem[{{Burrows} {et~al.}(2007{\natexlab{b}}){Burrows}, {Dessart}, {Livne},
2184:   {Ott}, \& {Murphy}}]{burrows:07b}
2185: {Burrows}, A., {Dessart}, L., {Livne}, E., {Ott}, C.~D., \& {Murphy}, J.
2186:   2007{\natexlab{b}}, \apj, 664, 416
2187: 
2188: \bibitem[{{Burrows} \& {Goshy}(1993)}]{burrows:93}
2189: {Burrows}, A., \& {Goshy}, J. 1993, \apjl, 416, L75
2190: 
2191: \bibitem[{{Burrows} {et~al.}(1995){Burrows}, {Hayes}, \& {Fryxell}}]{bhf:95}
2192: {Burrows}, A., {Hayes}, J., \& {Fryxell}, B.~A. 1995, \apj, 450, 830
2193: 
2194: \bibitem[{{Burrows} {et~al.}(2006){Burrows}, {Livne}, {Dessart}, {Ott}, \&
2195:   {Murphy}}]{burrows:06}
2196: {Burrows}, A., {Livne}, E., {Dessart}, L., {Ott}, C.~D., \& {Murphy}, J. 2006,
2197:   \apj, 640, 878
2198: 
2199: \bibitem[{{Burrows} {et~al.}(2007{\natexlab{c}}){Burrows}, {Livne}, {Dessart},
2200:   {Ott}, \& {Murphy}}]{burrows:07a}
2201: ---. 2007{\natexlab{c}}, \apj, 655, 416
2202: 
2203: \bibitem[{{Burrows} {et~al.}(2000){Burrows}, {Young}, {Pinto}, {Eastman}, \&
2204:   {Thompson}}]{burrows:00}
2205: {Burrows}, A., {Young}, T., {Pinto}, P., {Eastman}, R., \& {Thompson}, T.~A.
2206:   2000, \apj, 539, 865
2207: 
2208: \bibitem[{{Castor}(1972)}]{castor:72}
2209: {Castor}, J.~I. 1972, \apj, 178, 779
2210: 
2211: \bibitem[{{Castor}(2004)}]{castor:04}
2212: ---. 2004, {Radiation Hydrodynamics} (Radiation Hydrodynamics, by John
2213:   I.~Castor, pp.~368.~ISBN 0521833094.~Cambridge, UK: Cambridge University
2214:   Press, November 2004.)
2215: 
2216: \bibitem[{{Colgate} \& {White}(1966)}]{colgate:66}
2217: {Colgate}, S.~A., \& {White}, R.~H. 1966, \apj, 143, 626
2218: 
2219: \bibitem[{{Cooperstein} \& {Baron}(1992)}]{cooperstein:92}
2220: {Cooperstein}, J., \& {Baron}, E. 1992, \apj, 398, 531
2221: 
2222: \bibitem[{{Dessart} {et~al.}(2006{\natexlab{a}}){Dessart}, {Burrows}, {Livne},
2223:   \& {Ott}}]{dessart:06a}
2224: {Dessart}, L., {Burrows}, A., {Livne}, E., \& {Ott}, C.~D. 2006{\natexlab{a}},
2225:   \apj, 645, 534
2226: 
2227: \bibitem[{{Dessart} {et~al.}(2007){Dessart}, {Burrows}, {Livne}, \&
2228:   {Ott}}]{dessart:07}
2229: ---. 2007, \apj, 669, 585
2230: 
2231: \bibitem[{{Dessart} {et~al.}(2008){Dessart}, {Burrows}, {Livne}, \&
2232:   {Ott}}]{dessart:08a}
2233: ---. 2008, \apjl, 673, L43
2234: 
2235: \bibitem[{{Dessart} {et~al.}(2006{\natexlab{b}}){Dessart}, {Burrows}, {Ott},
2236:   {Livne}, {Yoon}, \& {Langer}}]{dessart:06b}
2237: {Dessart}, L., {Burrows}, A., {Ott}, C.~D., {Livne}, E., {Yoon}, S.-Y., \&
2238:   {Langer}, N. 2006{\natexlab{b}}, \apj, 644, 1063
2239: 
2240: \bibitem[{{Foglizzo} {et~al.}(2007){Foglizzo}, {Galletti}, {Scheck}, \&
2241:   {Janka}}]{foglizzo:07}
2242: {Foglizzo}, T., {Galletti}, P., {Scheck}, L., \& {Janka}, H.-T. 2007, \apj,
2243:   654, 1006
2244: 
2245: \bibitem[{{Foglizzo} \& {Tagger}(2000)}]{foglizzo:00}
2246: {Foglizzo}, T., \& {Tagger}, M. 2000, \aap, 363, 174
2247: 
2248: \bibitem[{{Fryer} \& {Heger}(2000)}]{fh:00}
2249: {Fryer}, C.~L., \& {Heger}, A. 2000, \apj, 541, 1033
2250: 
2251: \bibitem[{{Fryer} \& {Warren}(2002)}]{fryerwarren:02}
2252: {Fryer}, C.~L., \& {Warren}, M.~S. 2002, \apjl, 574, L65
2253: 
2254: \bibitem[{{Fryer} \& {Warren}(2004)}]{fryerwarren:04}
2255: ---. 2004, \apj, 601, 391
2256: 
2257: \bibitem[{{Hammer}(1892)}]{hammer:1892}
2258: {Hammer}, E. 1892, Petermanns Mitt., 38, 85
2259: 
2260: \bibitem[{{Heger} {et~al.}(2000){Heger}, {Langer}, \& {Woosley}}]{heger:00}
2261: {Heger}, A., {Langer}, N., \& {Woosley}, S.~E. 2000, \apj, 528, 368
2262: 
2263: \bibitem[{{Heger} {et~al.}(2005){Heger}, {Woosley}, \& {Spruit}}]{heger:05}
2264: {Heger}, A., {Woosley}, S.~E., \& {Spruit}, H.~C. 2005, \apj, 626, 350
2265: 
2266: \bibitem[{{Herant} {et~al.}(1994){Herant}, {Benz}, {Hix}, {Fryer}, \&
2267:   {Colgate}}]{herant:94}
2268: {Herant}, M., {Benz}, W., {Hix}, W.~R., {Fryer}, C.~L., \& {Colgate}, S.~A.
2269:   1994, \apj, 435, 339
2270: 
2271: \bibitem[{{Hubeny} \& {Burrows}(2007)}]{hb:07}
2272: {Hubeny}, I., \& {Burrows}, A. 2007, \apj, 659, 1458
2273: 
2274: \bibitem[{{Iwakami} {et~al.}(2008){Iwakami}, {Kotake}, {Ohnishi}, {Yamada}, \&
2275:   {Sawada}}]{iwakami:08}
2276: {Iwakami}, W., {Kotake}, K., {Ohnishi}, N., {Yamada}, S., \& {Sawada}, K. 2008,
2277:   \apj, 678, 1207
2278: 
2279: \bibitem[{{Janka}(1992)}]{janka:92}
2280: {Janka}, H.-T. 1992, \aap, 256, 452
2281: 
2282: \bibitem[{{Janka}(2001)}]{janka:01}
2283: ---. 2001, \aap, 368, 527
2284: 
2285: \bibitem[{{Janka} {et~al.}(2007){Janka}, {Langanke}, {Marek},
2286:   {Mart{\'{\i}}nez-Pinedo}, \& {M{\"u}ller}}]{janka:07}
2287: {Janka}, H.-T., {Langanke}, K., {Marek}, A., {Mart{\'{\i}}nez-Pinedo}, G., \&
2288:   {M{\"u}ller}, B. 2007, \physrep, 442, 38
2289: 
2290: \bibitem[{{Janka} \& {Moenchmeyer}(1989{\natexlab{a}})}]{janka:89a}
2291: {Janka}, H.-T., \& {Moenchmeyer}, R. 1989{\natexlab{a}}, \aap, 209, L5
2292: 
2293: \bibitem[{{Janka} \& {Moenchmeyer}(1989{\natexlab{b}})}]{janka:89b}
2294: ---. 1989{\natexlab{b}}, \aap, 226, 69
2295: 
2296: \bibitem[{{Janka} \& {M\"uller}(1996)}]{jankamueller:96}
2297: {Janka}, H.-T., \& {M\"uller}, E. 1996, \aap, 306, 167
2298: 
2299: \bibitem[{{Kitaura} {et~al.}(2006){Kitaura}, {Janka}, \&
2300:   {Hillebrandt}}]{kitaura:06}
2301: {Kitaura}, F.~S., {Janka}, H.-T., \& {Hillebrandt}, W. 2006, \aap, 450, 345
2302: 
2303: \bibitem[{{Kotake} {et~al.}(2003){Kotake}, {Yamada}, \& {Sato}}]{kotake:03}
2304: {Kotake}, K., {Yamada}, S., \& {Sato}, K. 2003, \apj, 595, 304
2305: 
2306: \bibitem[{{Lattimer} \& {Prakash}(2007)}]{lattimer:07}
2307: {Lattimer}, J.~M., \& {Prakash}, M. 2007, \physrep, 442, 109
2308: 
2309: \bibitem[{Lattimer \& Swesty(1991)}]{lseos:91}
2310: Lattimer, J.~M., \& Swesty, F.~D. 1991, {Nucl. Phys. A}, 535, 331
2311: 
2312: \bibitem[{{LeBlanc} \& {Wilson}(1970)}]{leblanc:70}
2313: {LeBlanc}, J.~M., \& {Wilson}, J.~R. 1970, \apj, 161, 541
2314: 
2315: \bibitem[{{Liebend{\" o}rfer} {et~al.}(2001){Liebend{\" o}rfer}, {Mezzacappa},
2316:   {Thielemann}, {Messer}, {Hix}, \& {Bruenn}}]{liebendoerfer:01a}
2317: {Liebend{\" o}rfer}, M., {Mezzacappa}, A., {Thielemann}, F., {Messer}, O.~E.,
2318:   {Hix}, W.~R., \& {Bruenn}, S.~W. 2001, \prd, 63, 103004
2319: 
2320: \bibitem[{{Liebend{\"o}rfer} {et~al.}(2004){Liebend{\"o}rfer}, {Messer},
2321:   {Mezzacappa}, {Bruenn}, {Cardall}, \& {Thielemann}}]{liebendoerfer:04}
2322: {Liebend{\"o}rfer}, M., {Messer}, O.~E.~B., {Mezzacappa}, A., {Bruenn}, S.~W.,
2323:   {Cardall}, C.~Y., \& {Thielemann}, F.-K. 2004, \apjs, 150, 263
2324: 
2325: \bibitem[{{Liebend{\"o}rfer} {et~al.}(2005){Liebend{\"o}rfer}, {Rampp},
2326:   {Janka}, \& {Mezzacappa}}]{liebendoerfer:05}
2327: {Liebend{\"o}rfer}, M., {Rampp}, M., {Janka}, H.-T., \& {Mezzacappa}, A. 2005,
2328:   \apj, 620, 840
2329: 
2330: \bibitem[{{Livne}(1993)}]{livne:93}
2331: {Livne}, E. 1993, \apj, 412, 634
2332: 
2333: \bibitem[{{Livne} {et~al.}(2004){Livne}, {Burrows}, {Walder}, {Lichtenstadt},
2334:   \& {Thompson}}]{livne:04}
2335: {Livne}, E., {Burrows}, A., {Walder}, R., {Lichtenstadt}, I., \& {Thompson},
2336:   T.~A. 2004, \apj, 609, 277
2337: 
2338: \bibitem[{{Livne} {et~al.}(2007){Livne}, {Dessart}, {Burrows}, \&
2339:   {Meakin}}]{livne:07}
2340: {Livne}, E., {Dessart}, L., {Burrows}, A., \& {Meakin}, C.~A. 2007, \apjs, 170,
2341:   187
2342: 
2343: \bibitem[{{Marek} \& {Janka}(2007)}]{marek:07}
2344: {Marek}, A., \& {Janka}, H.-T. 2007, Submitted to \apj. ArXiv e-prints,
2345:   0708.3372 [astro-ph]
2346: 
2347: \bibitem[{{Messer} {et~al.}(1998){Messer}, {Mezzacappa}, {Bruenn}, \&
2348:   {Guidry}}]{messer:98}
2349: {Messer}, O.~E.~B., {Mezzacappa}, A., {Bruenn}, S.~W., \& {Guidry}, M.~W. 1998,
2350:   \apj, 507, 353
2351: 
2352: \bibitem[{{Mezzacappa} \& {Bruenn}(1993{\natexlab{a}})}]{mezzacappa:93b}
2353: {Mezzacappa}, A., \& {Bruenn}, S.~W. 1993{\natexlab{a}}, \apj, 405, 669
2354: 
2355: \bibitem[{{Mezzacappa} \& {Bruenn}(1993{\natexlab{b}})}]{mezzacappa:93a}
2356: ---. 1993{\natexlab{b}}, \apj, 405, 637
2357: 
2358: \bibitem[{{Mezzacappa} \& {Messer}(1999)}]{mezzacappa:99}
2359: {Mezzacappa}, A., \& {Messer}, B. 1999, {J. Comp. Appl. Math}, 109, 281
2360: 
2361: \bibitem[{Mihalas \& Mihalas(1984)}]{mihalas:99}
2362: Mihalas, D., \& Mihalas, B. 1984, {Foundations of Radiation Hydrodynamics}
2363:   (Mineola, NY, USA: Dover Publications)
2364: 
2365: \bibitem[{Morel {et~al.}(1996)Morel, Wareing, \& Smith}]{morel:96}
2366: Morel, J.~E., Wareing, T.~A., \& Smith, K. 1996, J. Comput. Phys., 128, 445
2367: 
2368: \bibitem[{{Murphy} \& {Burrows}(2008)}]{murphy:08}
2369: {Murphy}, J.~W., \& {Burrows}, A. 2008, \apj in press, xxx
2370: 
2371: \bibitem[{{Myra} {et~al.}(1987){Myra}, {Bludman}, {Hoffman}, {Lichenstadt},
2372:   {Sack}, \& {van Riper}}]{myra:87}
2373: {Myra}, E.~S., {Bludman}, S.~A., {Hoffman}, Y., {Lichenstadt}, I., {Sack}, N.,
2374:   \& {van Riper}, K.~A. 1987, \apj, 318, 744
2375: 
2376: \bibitem[{{Myra} \& {Burrows}(1990)}]{myra:90}
2377: {Myra}, E.~S., \& {Burrows}, A. 1990, \apj, 364, 222
2378: 
2379: \bibitem[{{Nomoto} \& {Hashimoto}(1988)}]{nomoto:88}
2380: {Nomoto}, K., \& {Hashimoto}, M. 1988, \physrep, 163, 13
2381: 
2382: \bibitem[{{Ott} {et~al.}(2006{\natexlab{a}}){Ott}, {Burrows}, {Dessart}, \&
2383:   {Livne}}]{ott:06b}
2384: {Ott}, C.~D., {Burrows}, A., {Dessart}, L., \& {Livne}, E. 2006{\natexlab{a}},
2385:   Phys. Rev. Lett., 96, 201102
2386: 
2387: \bibitem[{Ott {et~al.}(2004)Ott, Burrows, Livne, \& Walder}]{ott:04}
2388: Ott, C.~D., Burrows, A., Livne, E., \& Walder, R. 2004, \apj, 600, 834
2389: 
2390: \bibitem[{{Ott} {et~al.}(2006{\natexlab{b}}){Ott}, {Burrows}, {Thompson},
2391:   {Livne}, \& {Walder}}]{ott:06spin}
2392: {Ott}, C.~D., {Burrows}, A., {Thompson}, T.~A., {Livne}, E., \& {Walder}, R.
2393:   2006{\natexlab{b}}, \apjs, 164, 130
2394: 
2395: \bibitem[{{Rampp} \& {Janka}(2000)}]{ramppjanka:00}
2396: {Rampp}, M., \& {Janka}, H.-T. 2000, \apjl, 539, L33
2397: 
2398: \bibitem[{{Rampp} \& {Janka}(2002)}]{ramppjanka:02}
2399: ---. 2002, \aap, 396, 361
2400: 
2401: \bibitem[{{Scheck} {et~al.}(2008){Scheck}, {Janka}, {Foglizzo}, \&
2402:   {Kifonidis}}]{scheck:08}
2403: {Scheck}, L., {Janka}, H.-T., {Foglizzo}, T., \& {Kifonidis}, K. 2008, \aap,
2404:   477, 931
2405: 
2406: \bibitem[{{Shen} {et~al.}(1998{\natexlab{a}}){Shen}, {Toki}, {Oyamatsu}, \&
2407:   {Sumiyoshi}}]{shen:98a}
2408: {Shen}, H., {Toki}, H., {Oyamatsu}, K., \& {Sumiyoshi}, K. 1998{\natexlab{a}},
2409:   Nucl. Phys. A, 637, 435
2410: 
2411: \bibitem[{{Shen} {et~al.}(1998{\natexlab{b}}){Shen}, {Toki}, {Oyamatsu}, \&
2412:   {Sumiyoshi}}]{shen:98b}
2413: ---. 1998{\natexlab{b}}, Progress of Theoretical Physics, 100, 1013
2414: 
2415: \bibitem[{{Shlomo} {et~al.}(2006){Shlomo}, {Kolomietz}, \&
2416:   {Col{\`o}}}]{shlomo:06}
2417: {Shlomo}, S., {Kolomietz}, V.~M., \& {Col{\`o}}, G. 2006, European Physical
2418:   Journal A, 30, 23
2419: 
2420: \bibitem[{{Swesty} \& {Myra}(2006)}]{swesty:06}
2421: {Swesty}, F.~D., \& {Myra}, E.~S. 2006, submitted to the Astrophys. J.,
2422:   astro-ph/0607281
2423: 
2424: \bibitem[{{Thompson} {et~al.}(2003){Thompson}, {Burrows}, \&
2425:   {Pinto}}]{thompson:03}
2426: {Thompson}, T.~A., {Burrows}, A., \& {Pinto}, P.~A. 2003, \apj, 592, 434
2427: 
2428: \bibitem[{{Thompson} {et~al.}(2005){Thompson}, {Quataert}, \&
2429:   {Burrows}}]{thompson:05}
2430: {Thompson}, T.~A., {Quataert}, E., \& {Burrows}, A. 2005, \apj, 620, 861
2431: 
2432: \bibitem[{{Walder} {et~al.}(2005){Walder}, {Burrows}, {Ott}, {Livne},
2433:   {Lichtenstadt}, \& {Jarrah}}]{walder:05}
2434: {Walder}, R., {Burrows}, A., {Ott}, C.~D., {Livne}, E., {Lichtenstadt}, I., \&
2435:   {Jarrah}, M. 2005, \apj, 626, 317
2436: 
2437: \bibitem[{{Weinberg} \& {Quataert}(2008)}]{weinberg:08}
2438: {Weinberg}, N.~N., \& {Quataert}, E. 2008, ArXiv 0802.1522 (astro-ph)
2439: 
2440: \bibitem[{{Wilson}(1971)}]{wilson:71}
2441: {Wilson}, J.~R. 1971, \apj, 163, 209
2442: 
2443: \bibitem[{{Wilson}(1985)}]{wilson:85}
2444: {Wilson}, J.~R. 1985, in Numerical Astrophysics, ed. J.~M. {Centrella}, J.~M.
2445:   {Leblanc}, \& R.~L. {Bowers}, 422
2446: 
2447: \bibitem[{{Woosley} {et~al.}(2002){Woosley}, {Heger}, \& {Weaver}}]{wwh:02}
2448: {Woosley}, S.~E., {Heger}, A., \& {Weaver}, T.~A. 2002, Rev. Mod. Phys., 74,
2449:   1015
2450: 
2451: \bibitem[{{Woosley} \& {Weaver}(1995)}]{ww:95}
2452: {Woosley}, S.~E., \& {Weaver}, T.~A. 1995, \apjs, 101, 181
2453: 
2454: \bibitem[{{Yamada} {et~al.}(1999){Yamada}, {Janka}, \& {Suzuki}}]{yamada:99}
2455: {Yamada}, S., {Janka}, H.-T., \& {Suzuki}, H. 1999, \aap, 344, 533
2456: 
2457: \bibitem[{{Yamasaki} \& {Foglizzo}(2008)}]{yamasaki:08}
2458: {Yamasaki}, T., \& {Foglizzo}, T. 2008, \apj, 679, 607
2459: 
2460: \bibitem[{{Yoshida} {et~al.}(2007){Yoshida}, {Ohnishi}, \&
2461:   {Yamada}}]{yoshida:07}
2462: {Yoshida}, S., {Ohnishi}, N., \& {Yamada}, S. 2007, \apj submitted
2463: 
2464: \bibitem[{{Yueh} \& {Buchler}(1977)}]{yueh:77}
2465: {Yueh}, W.~R., \& {Buchler}, J.~R. 1977, \apj, 217, 565
2466: 
2467: \end{thebibliography}
2468: 
2469: \end{document}
2470: 
2471: 
2472: