1: %\documentclass[12pt,preprint]{aastex}
2: \documentclass[10pt,preprint]{aastex}
3: %\documentclass{emulateapj}
4:
5: \shorttitle{Surface Detonations}
6: \shortauthors{Meakin et al.}
7:
8: \newcommand{\etal} { et~al.\ }
9: \newcommand{\sol}{$M_\odot$}
10: \def\nuc#1#2{\relax\ifmmode{}^{#1}{\protect\text{#2}}\else${}^{#1}$#2\fi}
11: \def\solrat#1#2{$[$#1/#2$]$}
12: \def\msol#1{\relax$#1\,M_\odot\/$ }
13: \def\mcol{\multicolumn}
14: \def\mult{$\times$}
15: \def\msun{M$_{\odot}$}
16: \def\lang{\langle}
17: \def\rang{\rangle}
18: \def\tnm{\tablenotemark}
19: \def\tnt{\tablenotetext}
20: \def\iso#1#2{$^{#2}${#1}}
21: \def\betp{$\beta^+$}
22: \def\betm{$\beta^-$}
23:
24: \begin{document}
25: \title{STUDY OF THE DETONATION PHASE IN THE GRAVITATIONALLY CONFINED DETONATION MODEL
26: OF TYPE Ia SUPERNOVAE}
27: \author{Casey A. Meakin\altaffilmark{1,2,3,5}, Ivo Seitenzahl\altaffilmark{3,4},
28: Dean Townsley\altaffilmark{1,2,3}, George C. Jordan IV\altaffilmark{1,2},
29: James Truran\altaffilmark{1,2,3}, Don Lamb\altaffilmark{1,2,4}}
30: \altaffiltext{1}{Center for Astrophysical Thermonuclear Flashes, University of Chicago, Chicago, IL}
31: \altaffiltext{2}{Department of Astronomy and Astrophysics, University of Chicago, Chicago, IL}
32: \altaffiltext{3}{Joint Institute for Nuclear Astrophysics, University of Chicago, Chicago, IL}
33: \altaffiltext{4}{Enrico Fermi Institute, University of Chicago, Chicago, IL}
34: \altaffiltext{5}{Steward Observatory, University of Arizona, Tucson, AZ}
35: \email{casey.meakin@gmail.com}
36:
37: \begin{abstract}
38: We study the gravitationally confined detonation (GCD) model of Type
39: Ia supernovae through the detonation phase and into homologous
40: expansion. In the GCD model, a detonation is triggered by the surface
41: flow due to single point, off-center flame ignition in carbon-oxygen
42: white dwarfs.
43: The simulations are unique in terms of the degree to which non-idealized physics is
44: used to treat the reactive flow, including weak reaction rates and a time dependent
45: treatment of material in nuclear statistical equilibrium (NSE).
46: Careful attention is paid to accurately calculating the final composition of material
47: which is burned to NSE and frozen out in the rapid expansion following the passage of a
48: detonation wave over the high density core of the white dwarf; and an efficient method
49: for nucleosynthesis post-processing is developed which obviates the need for costly network
50: calculations along tracer particle thermodynamic trajectories. Observational diagnostics are
51: presented for the explosion models, including abundance stratifications and integrated yields.
52: We find that for all of the ignition conditions studied here, a self regulating process comprised
53: of neutronization and stellar expansion results in final \iso{Ni}{56} masses of $\sim$1.1\msun.
54: But, more energetic models result in larger total NSE and stable Fe peak yields.
55: The total yield of intermediate mass elements is $\sim0.1$\msun and the explosion energies are
56: all around 1.5$\times$10$^{51}$ ergs. The explosion models are briefly compared
57: to the inferred properties of recent Type Ia supernova observations.
58: The potential for surface detonation models to produce lower
59: luminosity (lower \iso{Ni}{56} mass) supernovae is discussed.
60: \end{abstract}
61:
62: \keywords{stars: evolution - stars: nucleosynthesis - supernovae -
63: hydrodynamics }
64:
65:
66: \section{INTRODUCTION}
67:
68: \par The currently favored model for Type Ia supernovae (SNe Ia) is the
69: thermonuclear incineration of a white dwarf (WD) which has accreted mass
70: to near the Chandrasekhar limit from a binary companion
71: \citep[e.g.,][]{branch1995,hillebrandt2000}. The enormous luminosity
72: and homogeneity in the properties of the light curves of SNe Ia make
73: them exceptionally good standard candles and as such have shown that the
74: expansion rate of the universe is accelerating and provided intriguing evidence
75: for a cosmological constant \citep{riess1998}.
76:
77: \par Despite the success in using SNe Ia as cosmological probes and identifying
78: a plausible astrophysical progenitor site for the explosions, a detailed
79: understanding of the explosion mechanisms itself remains elusive.
80: Several uncertainties stand in the way of a definitive solution to the
81: SNe Ia problem. On the one hand, the conditions under which the
82: thermonuclear runaway commences remains poorly understood so that the
83: initial number and distribution of flamelets that seed the runaway
84: is still a free parameter. On the other hand, although significant progress has been
85: made in simulating flame fronts in multi-dimensional stellar models
86: \citep{gamezo2005,schmidt2006b,roepke2007a,townsley2007,jordan2007}, the challenge
87: associated with modeling an unresolved turbulent deflagration
88: \citep[e.g.][]{schmidt2006a} with limited computational resources injects
89: an additional degree of uncertainty into the outcome of a model for any
90: given choice of initial conditions.
91:
92: %\par A flamelet which ignites off center will rise buoyantly towards the
93: %stellar surface until burning is quenched at low densities.
94: %In order of decreasing energy release, or equivalent flame bubble
95: %ignition further from the stellar center, the following evolutionary paths
96: %can occur \citep[e.g.][]{livne2005}: (1) The nuclear energy released unbinds
97: %the star. (2) The nuclear energy release falls short of unbinding the star
98: %but leads to significant expansion of the star which quenches the nuclear
99: %burning and prevents a significant surface flow. (3) The nuclear energy
100: %released during the deflagration phase is not enough to significantly expand
101: %the star and a large fraction of the hot ash is confined to the star in the
102: %form of a surface flow which eventually converges at the pole opposite to
103: %breakout.
104:
105: %\par All three of these cases have been considered as possible scenarios for
106: %Type Ia supernovae. The first case is known as a {\em deflagration} or
107: %{\em pure deflagration} model and is distinct in the sense that no detonation
108: %forms in this model. The second case, with a marginally bound
109: %core following deflagration, may detonate upon recollapse \citep{khoklov}
110: %giving rise to a healthy Type Ia explosion. Most recently, it has been proposed
111: %that the surface flow in the third case initiates a detonation that consumes the
112: %unburnt portion of the C/O core \citep{plewa2004,plewa2007} and producing a
113: %luminous Ia explosion.
114:
115: %\par A caveate to the above scenarios is the possibility that deflagration to
116: %detonation transitions (DDT) occur at locations where the nuclear flame makes
117: %a transition to the distributed burning regime (REF) and the remainder of the
118: %unburnt stellar core is consumed by one or more detonation waves. The possibility
119: %of a DDT exists in any of the above scenarios and the evolution therefore depends
120: %strongly upon the initial flamelet distribution.
121:
122: %\par The overall problem requires mastery of at least three
123: %reactive-hydrodynamic
124: %subproblems, including: (1) The nature of the initial conditions and the birth
125: %of a thermonuclear flame, including the spatial distribution of that initial flame,
126: %its location within the star and whether or not it is composed of a single initial
127: %spark or a distribution of flame kernels. (2) The propagation of the nuclear flame
128: %through the star, which is best described as a Rayleigh-Taylor unstable deflagration.
129: %(3) The mode by which the (subsonic) deflagration makes a transition to a detonation.
130: %Each of these problems continues to be an active field of research and have only
131: %recently started to be investigated with multi-dimensional, multi-physics simulation
132: %codes with an interesting degree of realism.
133:
134:
135: %\par The GCD model can be conveniently broken down into four distinct stages of
136: %evolution, including: (1) The initial growth of the flame bubble and the subsequent
137: %erruption of hot ash onto the surface of the star. (2) The confinement of the hot
138: %ash to the stellar surface by gravity, which drives a flow around the outside of
139: %the star towards the anti-pode. (3) The collision of the surface flow at the
140: %anti-pode which leads to the formation of an inward and outwardly directed jet.
141: %(4) The inwardly directed jet eventually leads to conditions under which a
142: %detonation arises and sweeps over, the remaining unburned material in the stellar core
143: %producing a large fraction of a solar mass worth of iron peak and intermediate mass
144: %elements, disrupting the star and imparting $\sim$10$^{51}$ ergs of kinetic energy to
145: %the ejecta.
146:
147: \par In this paper we describe progress on our ongoing effort to improve
148: the simulation of SNe Ia in multi-dimensions, including methods
149: to perform detailed nucleosynthesis post-processing in a computationally efficient manner.
150: We extend the study of the GCD
151: model for a single ignition point slightly offset a range of distances from the center of the
152: star, as described in \citet{townsley2007}, through
153: the detonation phase and into homologous expansion.
154: In \S\ref{sec:numerics} we describe the treatment of the reactive-hydrodynamics
155: problem used in our simulation code. In \S\ref{sec:def} we review the relevant properties of
156: the deflagration phase for single point flame ignition. In \S\ref{sec:det} we examine
157: in some detail the initiation of the detonation, the properties of the detonation
158: wave which disrupts the star, and the resultant remnant morphology.
159: In \S\ref{sec:yields} we discuss in detail the nucleosynthetic yields for the explosions
160: studied and decribe the methodology used to efficiently calculate iron peak yields
161: from the simulation data. We conclude with a summary of the salient features of the
162: explosion models in light of observed Type Ia supernvoae.
163:
164:
165: \section{NUMERICAL METHODS: HYDRODYNAMICS AND NUCLEAR BURNING}
166: \label{sec:numerics}
167: \par In this section we review the computational tools used to model the hydrodynamic
168: and nuclear evolution of the stellar plasma, including the treatment of subsonic
169: (deflagration) and supersonic (detonation) burning fronts. The basic code framework is
170: FLASH \citep{fryxell2000}, a modular, block-structured adaptive mesh refinement (AMR),
171: Eulerian, reactive-hydrodynamics code. We use a directionally split PPM solver \citep{colella1984}
172: generalized to treat non-ideal gasses \citep{colella1985} to handle the hydrodynamic
173: evolution.
174:
175: \par The energetics scheme used to treat flames and detonation waves
176: in our simulations uses 3 progress variables to track carbon burning,
177: NSQE relaxation, and NSE relaxation. The rates connecting these burning
178: stages are calibrated using a large (200 nuclide) nuclear reaction network
179: for the conditions relevant to the Type Ia problem. Additionally, energy losses
180: (through neutrino emission) and changes in the electron mole fraction
181: $Y_e$ due to weak interactions taking place in material which has burned to NSE are
182: incorporated. Details can be found in \citet{calder2007,townsley2007,seitenzahl2008a}.
183: %\par Nuclear burning is coupled to the hydrodynamic evolution using an operator splitting
184: %formulation whereby hydrodynamic and nuclear burning modules are called succesively over
185: %the course of a single timestep \citep{fryxell2000}.
186:
187: \par Both detonation waves and flames are impossible to resolve in full star simulations
188: because they are characterized by length scales that are more than ten orders of magnitude
189: smaller than the radius of the white dwarf to be modeled, $R_{\rm wd}\sim10^8$ cm.
190: Therefore, these reaction fronts must be treated in a special manner. Subsonic burning fronts
191: (deflagrations) are advanced using an advection-diffusion-reaction (ADR) equation.
192: In short, a thickened flame front ($\sim$4 grid zones wide) is advanced at a
193: speed $v_f = \max(v_l, v_t)$, where $v_l$ is the laminar flame speed calculated
194: by \citet{timmes1992} and $v_t$ is a Rayleigh-Taylor driven
195: turbulent flame speed. Details concerning the implementation, calibration and noise
196: properties of the flame treatment can be found in \citet{townsley2007} and
197: Asida et al. (2008, in preparation) and references therein.
198:
199: \par Detonations are handled naturally by the reactive hydrodynamics solver
200: in FLASH without the need for a front tracker. This approach is possible because
201: unresolved Chapman-Jouguet (CJ) detonations retain the correct jump conditions and propagation speeds.
202: Numerical stability is maintained by preventing nuclear burning within the
203: shock. This is necessary because shocks are artificially spread out over a few zones
204: by the PPM hydrodynamics solver, which can lead to unphysical burning within
205: shocks that can destabilize the burning front \citep{fryxell1989}.
206: The energetics in the detonation differ from that in the deflagration front only
207: in how carbon burning proceeds, as represented by the first progress variable $\phi_1$
208: and an explicit carbon burning rate is used \citep{caughlan1988}. The additional burning stages
209: (NSQE and NSE relaxation) are tracked by $\phi_2$ and $\phi_3$ and are evolved in the same
210: manner as in the post flame ash \citep{calder2007,townsley2007}.
211: At densities above $\sim 10^7$ g cm$^{-3}$
212: detonations propagating through a mixture that is equal parts \nuc{12}{C} and \nuc{16}{O} have
213: Mach numbers that are larger than the CJ value, but by only a few percent
214: \citep{gamezo1999,sharpe2001}.
215: Cellular structure smaller than the grid scale will be suppressed in our simulations
216: but is free to form on resolved scales. The impact of cellular structure on the global evolution
217: of the model is still uncertain.
218: However, since cellular structure alters the detonation wave
219: speed by only a few percent for the conditions being modeled \citep{timmes2000b} the effect is likely to be small.
220: Additional details related to the treatment of detonation waves are discussed in \S\ref{sec:det}.
221:
222: %In order to treat detonations in the explicit manner described, a carbon
223: %burning reaction rate is needed \citep{caughlan1988} to advance the first
224: %progress variable in place of the flame model. The additional burning stages
225: %($\phi_2$, $\phi_3$) are advanced using the relaxation timescales in the same
226: %manner as in the flame.
227:
228: %\par This approximate treatment of the nuclear burning has the following shortcoming:
229: %For densities above $10^7$ g/cm$^3$ detonations propagating through a mixture that
230: %is equal parts \nuc{12}{C} and \nuc{16}{O} are in the ``pathological'' regime
231: %\citep{gamezo1999,sharpe2001}. [Basic features: detonation speeds vary less
232: %than X\% over the region of interest.]
233: %Detonation waves are unstable to perturbations transverse to the direction
234: %of propagation which give rise to a so-called ``cellular structures'' on length
235: %scales comparable to the induction length \citep{gamezo1999,timmes2000b}.
236: %Suppresed on smaller scales by resolution -- what might be the subgrid-scale
237: %effect of these instabilities? clumpiness, propagation speed differences.
238: %(Summarize. Conclude by reciting the magnitude of the errors expected. $\sim$ a few \%)
239:
240: \par Self gravity is calculated using a multi-pole solver with a maximum spherical harmonic
241: index l$_{max}$=10. The Helmholtz equation of state of \citet{timmes2000a} is used to
242: describe the thermodynamic properties of the stellar plasma including contributions from
243: blackbody radiation, ions, and electrons of an arbitrary degree of degeneracy.
244:
245: \par Passive tracer particles are included in our simulations which record
246: the time history of the flow properties along Lagrangian trajectories. These records
247: can be used to calculate detailed nucleosynthetic yields as well as to provide
248: additional diagnostic for complex flows. We use 10$^5$ tracer particles
249: for 2D models and 10$^6$ for 3D models. The particles used in this study are
250: initialized at the beginning of each simulation with a mass weighted distribution.
251: In \S\ref{subsec:freezeout-method} we present a novel method to calculate post-explosion
252: yields which does not require the prohibitively expensive post-processing of a large
253: number of tracer particle with a nuclear reaction network, but rather uses information
254: readily extracted from the tracers to calibrate an efficient table look-up scheme.
255:
256: \section{FLAME IGNITION AND DEFLAGRATION}
257: \label{sec:def}
258: \par In this paper we extend the study of the GCD model for the single point flame ignition models of
259: \citet{townsley2007} through the detonation phase and into homologous
260: expansion. The general simulation
261: setup is the same, and we review it here briefly, along with a description of
262: the basic progression of the evolution preceding detonation.
263: After carbon burning ignites at the center of the white dwarf, a
264: convective core is formed which expands as it heats. Our simulations begin
265: when the nuclear burning timescale becomes shorter than the eddy turnover
266: time, so that the first flamelet is ignited near (within a few hundred km of)
267: the center of the white dwarf. As discussed by \citet{townsley2007}, there is
268: still significant uncertainty
269: in the form which the nuclear flame will take at birth (also see
270: e.g.~\citealt{woosley2004}), relating to the number and location of what are
271: generally assumed to be relatively small ($<1$ km) ignition regions. For
272: reasons related to simplicity of setup and limitations of the imposed
273: cylindrical symmetry, we restrict our study to off-center, single point
274: ignitions in a quiescent background star.
275: The initial WD used in these simulations has a uniform temperature $4\times
276: 10^7$ K, a mass of \msol{1.365}\footnote[1]{This was erroneously given as
277: \msol{1.38} in \citet{townsley2007}, none of these parameters have
278: changed from that work to this.}, a central density of $2.2\times 10^9$ g
279: cm$^{-3}$, and is composed of equal mass fractions of \nuc{12}{C} and
280: \nuc{16}{O}. This progenitor is much colder than reality, but it is
281: expected, and we have confirmed by comparison, that this does not have a
282: significant effect on the structure of the white dwarf or the dynamics
283: of the explosion. At the beginning of the
284: simulation a spherical region of radius $r_{\rm bub}$=16 km placed on
285: the polar axis at a range of distances between
286: $r_{\rm off}=$20 and 100 km from the center of the star is converted to NSE
287: ash in pressure equilibrium with the remainder of the star. A summary of the
288: initial flame bubble parameters studied in this paper is given in
289: Table \ref{tab:models}.
290:
291: \par The basic stages of single bubble flame evolution can be described in
292: terms of two key length scales, the grid resolution, $\Delta$, which sets the
293: limit to which we can resolve flame structure, and the critical wavelength,
294: sometimes called the fire polishing length, $\lambda_c\equiv 6\pi s^2/Ag$
295: \citep{khoklov1995}, where $s$ is the front propagation speed (flame speed)
296: and $A$ is the Atwood number
297: $A=(\rho_{\rm fuel}-\rho_{\rm ash})/(\rho_{\rm fuel}+\rho_{\rm ash})$.
298: Perturbations in a
299: Rayleigh-Taylor (R-T) unstable flame front smaller than $\lambda_c$ are
300: ``polished out'' by the propagation of the flame, while those of larger
301: scale are enhanced by R-T growth, wrinkling the flame bubble.
302: \cite{townsley2007} distinguished three phases of flame evolution that occur
303: successively as the buoyant flame bubble grows in size and rises toward the
304: stellar surface: laminar bubble growth when $r_{\rm bub} \lesssim\lambda_c$,
305: resolved R-T unstable growth, and finally R-T unstable growth on the sub-grid
306: scale when $\lambda_c < \Delta$. One immediate consequence of this progression
307: and the increase of $g$ with radius, is that the resolution limits the largest
308: ignition offset position $r_{\rm off}$ which can be used and still start in
309: the laminar growth phase. This limit is roughly 100 km for $\Delta=4$ km, the
310: resolution of all the simulations here. At this resolution the critical wavelength
311: is less than the grid scale, $\lambda_c < \Delta$, outside roughly 400 km from the
312: center of the star.
313:
314: For even modest offsets, the hot ash is buoyant enough that it erupts from
315: the surface of the star before more than a few percent of the star is burned.
316: This creates a vigorous flow over the surface of the WD, which is still
317: relatively compact due to the small amount of burning. The progress of this
318: eruption and flow is shown in Figure \ref{fig:breakout} for the case with
319: $r_{\rm off} = 40$ km.
320: As mentioned previously, there is indication from comparisons of recent work
321: \citep{roepke2007a,townsley2007,jordan2007} that the eruption pattern arising
322: from a given offset is dependent on the choice of burning model, and therefore is
323: currently uncertain. This has important consequences at the collision
324: region because it sets the velocity and density structure of the
325: incoming flows as well as the surface gravity (via the degree of stellar
326: expansion) under which the collision occurs.
327:
328: We study the cases from \citet{townsley2007}, which demonstrated collisions
329: that created detonation conditions, along with several supplementary cases near the
330: minimum offset distance that led to detonation conditions. This gives the range $r_{\rm
331: off}=$20 to 100 km. The expansion which occurs during the
332: deflagration and surface flow stages is very nearly homologous and has only a
333: small degree of asymmetry, such that most of the asymmetry is created later
334: during the detonation phase (see below).
335: Figure \ref{fig:homologous-deflagration} compares the scaled density profile
336: for the initial model and the 25 km and 100 km offset cases at the time the
337: detonation initiates. Density is scaled by the central value and radius is
338: scaled by the distance from the center at which the density drops to $1/e$
339: times the central value. Profiles along the equator and along the symmetry
340: axis are both shown, demonstrating that the star remains very symmetric out
341: to approximately 2 density scale heights away from the center. This region
342: contains approximately 90\% of the stellar mass, and even the asymmetry
343: beyond this is fairly modest, but is likely to lead to some asymmetry in the
344: highest-velocity spectral features.
345:
346: Based on the variation in expansion found in the resolution study performed
347: by \citet{townsley2007}, and other cases generally, it appears that the
348: conditions at the collision region and the density structure at detonation
349: are not a single parameter family in the mass burned, or equivalently nuclear
350: energy release, in the deflagration, $E_{\rm n,def}$, prior to the collision.
351: It does seem that $E_{\rm n, def}$ is the primary parameter, but other
352: contributing factors include the time dependence of the energy release,
353: morphology of the burning region in the flame plume, and the possibility of
354: secondary or tertiary ignition sites. The sudden input of energy in the
355: deflagration puts the star in an oscillation, and the timing of the
356: detonation initiation with respect to this oscillation is important for
357: setting the density structure at detonation, and thereby the burning
358: products (see \S5). The timing and magnitude of the
359: nuclear energy release will change the magnitude and phase of this full-star
360: oscillation, but further investigation, including studies of 3-dimensional
361: deflagration morphologies, is needed to characterize these relationships.
362:
363: As an aside we note the impact of neutronization due to electron captures
364: during the deflagration. Neutronization influences the dynamics of a rising
365: flame bubble by changing the average binding energy of the final NSE state
366: obtained as well as the electron pressure available per gram of stellar plasma.
367: We assessed the impact of neutronization by recalculating several models through
368: bubble rise with weak rates suppressed, by enforcing $\dot Y_e = 0$. Flame bubbles
369: ignited closer to the stellar center, and hence at higher densities,
370: are more strongly affected. Models ignited at $r_{\rm off}$=40 km and 30 km
371: {\em burned $\sim$10\% and $\sim$40\% more mass, respectively, with weak rates
372: suppressed} while the model ignited at $r_{\rm off}=$80 km
373: was negligibly affected by the weak reactions during bubble rise and breakout.
374: The effect that the weak reactions have on the burned mass depends on the
375: developement of turbulence which is not well represented in the 2D simulations
376: presented here. Therefore, while we have demonstrated that weak reactions play
377: a non-negligible role in the present suite of models the impact that they have
378: on more realistic 3D flows remains an open question.
379:
380: %(Based on a comparison with Dean's earlier runs to my more recent ones
381: % w/ and w/out weak rates it appears that the sensitivity of the calculations
382: % to the development of turbulence is very strong and a simple explanation of the
383: % differences remains lacking.)
384: %
385: % It appears
386: % that the change in the binding energy of the final state dominates such that flames with
387: % weak reactions suppressed release less energy in burning to NSE thereby producing less
388: % buoyant ash which rises more slowly, burns to a larger lateral extent, and produces
389: % more total burned mass prior to breakout.
390: %
391: % \par (Dean to Casey: This is not consistent with the studies I did on this, in
392: % which the bubble rise was directly observed to be faster with neutronization
393: % turned off. I didn't check the burned mass dependence. It may be that the
394: % faster rise generates more turbulence...)
395:
396:
397: \section{DETONATION}
398: \label{sec:det}
399:
400: %\par Even slightly off-center flame ignition in the core of white dwarf appears
401: %to strongly diminish the chances that a deflagration by itself can unbind a
402: %white dwarf and lead to an explosion comparable to those observed as Ia SNe.
403:
404: %\par While the conditions of flame ignition, including the spatial distributions and formation
405: %rate, remains unknown an increasing number of studies have begun to explore the mapping
406: %between proposed distributions and final outcomes, incorporating
407: %ever more realistic physics and relying increasingly on 3D full star simulations
408: %\citep{gamezo2005,jordan2007,roepke2007a}.
409: %Although the correct treatment of a turbulent flame remains an active area of research,
410: %recent simulations which use state of the art turbulent flame models
411: %cannot reproduce the observed features of normal luminosity Ia SNe when burning takes
412: %place solely in a deflagration \citep{roepke2007b}. These results have renewed interest
413: %in scenarios which undergo a transition to detonation after a phase of subsonic burning.
414:
415: \par Single point off-center ignition results in a buoyant plume of hot ash which is brought
416: to the surface of the star before more than a few percent of the stellar core is consumed
417: by the flame (see M$_{\rm def}$ in Table~\ref{tab:models}). As the hot ash rises to the surface,
418: the nuclear energy that is released excites a stellar pulsation which initially expands the star.
419: Against this background pulsation,
420: the hot ash from the burning is expelled from the stellar interior. A large fraction of this ash
421: is confined to the star's surface by gravity. This ash sweeps over the surface of the
422: star together with a flow of unburned stellar material which is pushed ahead of it.
423: In all but the most expanded model in our parameter study (i.e., those with ignition points
424: r$_{\rm off} \ge$ 25 km), the resulting surface flows converge at a point opposite to breakout
425: which we refer to as the {\em collision region} (Fig.~\ref{fig:breakout}).
426: These converging surface flows result in a bi-directional, collimated jet-like
427: flow which both expels material away from the star's surface and drives a flow of high
428: temperature material into the stellar core.
429: The inward directed component of the collimated flow
430: reaches high enough densities and temperatures that a ``surface detonation'' inititiates which
431: sweeps over the core and completely disrupts the white dwarf, giving rise
432: to a luminous supernova explosion.
433:
434: \par In the following subsections we describe the
435: characteristics of the bi-directional jet which forms in the colliding surface
436: flow and initiates the detonation (\S\ref{subsec:jet}),
437: we discuss the characteristics of the ensuing detonation phase of burning (\S\ref{subsec:det}),
438: and we describe the final state of free expansion which results (\S\ref{subsec:remnant}).
439: A detailed analysis of the nucleosynthetic yields is presented in the next section, \S5.
440:
441: %Only the model with an initial flame bubble
442: %ignited 20 km off center fails to detonate within the time simulated. In this model the large
443: %amplitude of the stellar pulsation produced by the energy release in the deflagration
444: %stalls the surface flow and lowers the densities in the collision region to values
445: %well below the canonical value of $\rho\sim$10$^7$ g cm$^{-3}$ necessary to trigger a
446: %detonation \citep{niemeyer1997,seitenzahl2008b}.
447:
448:
449:
450:
451: \subsection{Jet Formation and Detonation Initiation}
452: \label{subsec:jet}
453:
454: %% FORMATION AND INIITAL DEVELOPMENT OF THE JET
455:
456: \par {\em Jet Formation and Characteristics.---} As the surface flow produced by the deflagration
457: converges, material accumulates in a small region on the hemisphere opposite to the
458: breakout location. The material which initially piles up, consists of unburned
459: carbon and oxygen rich surface material which is pushed ahead of the ash as it flows around
460: the stellar surface. As material accumulates in this region it is heated by compression until it reaches
461: temperatures sufficient to initiate carbon burning, which further heats the compressed material
462: and raises its pressure. Shortly after the initial collision, a conical shock forms which separates the
463: compressed material along the axis from the inflowing surface flow. The surface flow material
464: burns as it passes through this shock and ``accretes'' into the collision region.
465: The resultant pressure in the collision region roughly
466: balances the ram pressure of the accreting surface flow, $p_{\rm coll} \sim [\rho v^2]_{\rm surf}$ (see Fig.~\ref{fig:jet-slice}).
467: The pressure achieved in the compressed region soon exceeds the (nearly hydrostatic) background pressure
468: sufficiently that it redirects the accreting material and drives a bidirectional jet-like
469: flow which has components aligned along the polar axis. A closeup of the collision region
470: thus formed is shown in Figure~\ref{fig:jet} (which corresponds to the region outlined
471: by the dashed box in Figure~\ref{fig:breakout}). The velocity vectors reveal the
472: bidirectional nature of the flow. The ash from the deflagration is just approaching
473: the collision region at the time shown, well after the collimated jet has formed.
474: The width of the jet increases with time as material continues to accrete into
475: the region, but retains structure on scales $<$50 km, which are well resolved in our
476: simulations which have a grid resolution of 4 km.
477:
478:
479: %% JET PROFILES
480:
481: \par In Figure~\ref{fig:jet-slice} the flow properties along the jet axis are shown just prior to
482: the onset of detonation for two 2D models and one 3D model.
483: The 2D models shown span the conditions studied in this paper, including the model with the most
484: expanded core (left-panel) and the least expanded core (middle-panel) which
485: detonate in our study.
486: All of the collimated flows share the same overall structure with
487: the more expanded stars having shallower density gradients in the collision region.
488: The velocity profiles are roughly linear, decreasing from a maximum inwardly
489: directed velocity of $\sim$ (1 to 2)$\times 10^9$ cm/s to a comparable
490: velocity directed away from the stellar surface.
491: While the inward flow is attended by a great deal of small scale internal
492: substructure and turbulence (Fig.~\ref{fig:jet}), three distinct ``fronts'' are readily identifiable
493: along the axis of the jet: a leading subsonic compression wave, followed by a fuel-ash boundary layer,
494: and finally an internal shock. The fuel-ash boundary layer is marked in Figure~\ref{fig:jet-slice}
495: by the dashed vertical line. The material ahead of this line has not yet been compressed
496: to high enough densities that carbon burning can proceed.
497: The compression wave(s) which eminates from the head of the jet as it moves
498: into the star can be seen as perturbations preceeding the fuel-ash boundary in all of the variables
499: plotted in Figure~\ref{fig:jet-slice} and can also be seen as the pressure waves extending into
500: the star ahead of the jet in Figure~\ref{fig:jet} (left-panel). The compression
501: wave moves into the star at the sound speed, which is
502: $c_s\sim 3.5\times 10^8$ cm/s at this location. The head of the jet, as marked by
503: the location of the fuel-ash boundary, moves inward at a fraction of the sound speed so that
504: the size of the compressed region grows with time.
505: Trailing behind the fuel-ash boundary is a shock wave which
506: separates the low density, high velocity flow produced in the collision region from the compression
507: wave which moves ahead of it. It is the ram pressure of this high velocity flow which drives the
508: compression wave into the star. The ram pressure of this high velocity flow is balanced by the
509: gas pressure of the compressed, overlying material, as shown by the red line in the bottom panels of
510: Figure~\ref{fig:jet-slice}. In all of the models studied, the inward directed jet continues to
511: compress material, heating it to carbon burning conditions until a detonation arises
512: and distrupt the star.
513:
514: \par An important question in the context of the present study concerns the extent to which the jet-like
515: flows which develop depend upon the 2D geometry used.
516: Simulations of off-center ignition using 3D grids have been made \citep[e.g.][]{roepke2007a,jordan2007}
517: with the general conclusions that focusing of surface flow also occurs in 3D and is not strongly
518: diminished compared to 2D. As a point of direct comparison, we have simulated a 3D model from flame
519: ignition through detonation using the same methods as used in the 2D models presented here.
520: This 3D model used a finest resolution of $\Delta$=8 km and was ignited
521: by a 16 km flame bubble displaced 80 km from the stellar center. The development of the collision
522: region and the subsequent detonation in the 3D model is remarkably similar to that found in the 2D models.
523: For comparison, the profile of the jet formed in the 3D model just prior to detonation is included in
524: the right panel of Figure~\ref{fig:jet-slice}.
525:
526:
527:
528: %% SHAPED CHARGE JETS
529:
530: \par Jet formation within a converging flow and jet penetration are well studied phenomena.
531: For instance, engineers have designed shaped charge explosives which create jets by explosively
532: collapsing a convex ``liner'' material, most often cone-shaped, onto itself \citep[e.g.][]{birkhoff1948}.\footnote[2]{Mining engineers have employed similar methods as early as 1792.}
533: The jets thus formed have notoriously strong penetrative power and can
534: slice and penetrate sheets of steel which are several times thicker than
535: the shaped charge diameter and are applied in both military and industrial
536: capacities including metal perforation, armor penetration, and oil well drilling.
537: While there are many differences between shaped charge jet formation and
538: the collision region flows present in our calculations, the phenomena bear interesting
539: similarities which may provide insight into the depth to which a converging surface flow may
540: penetrate the underlying carbon-oxygen rich layers of a white dwarf.
541: The jet models which have been
542: made to interpret experimental results estimate penetration depth by balancing
543: the ram pressure of the jet material with that of the target material in a frame of reference
544: that is moving with the jet-target interface. The penetration depth under the simplifying
545: circumstances of constant density jet and target materials depend on only the density ratio and
546: the jet length. This picture is greatly complicated in the stellar surface flow
547: case where compressibility plays a central role and the pressure balance at the jet-star interface
548: is between the dynamical pressure of the jet and the gas pressure of the core.
549: While it is
550: beyond the scope of the present paper to fully analyze the problem of compressible jet formation and
551: penetration, we conclude by noting that the strong penetrative power observed in shaped charge jets
552: provides support for the deep penetration
553: seen in all of the simulations in our study which develop collision regions. In all of the cases
554: which we study, the jets which have formed penetrate into denser layers of the white dwarf until a
555: detonation occurs.
556:
557:
558:
559:
560: %% THE INITIATION OF THE DETONATION
561:
562: \par {\em Detonation Initiation.---} Once the density of the material undergoing carbon burning
563: in the jet exceeds $\rho_{\rm det}\sim 10^7$ g/cm$^3$ a detonation initiates
564: which then propagates away from the head of the jet at the Chapman-Jouguet speed,
565: $D_{CJ} \sim 1.2\times 10^9$ cm/s with Mach number $M = D_{CJ}/c_s \sim 3.4$.
566: The time sequence shown in Figure~\ref{fig:jet} captures the moment when the
567: detonation initiates at the head of the jet and begins to spread outward. Because
568: of the weak dependence of the detonation wave speed on the upstream density, the detonation
569: front radiates from its point of initiation nearly spherically.
570:
571: \par The initiation of the detonation, which takes place at the fuel-ash boundary, when
572: $\rho\sim 10^7$ g/cm$^{3}$ and T$\sim 3\times10^9$ K., resembles
573: a Zel'dovich gradient mechanism \citep{zeldovich1970,khokhlov1997}. Detonation initiation through this process
574: involves a complicated interplay between burning and hydrodynamic flow
575: that requires a coherent build up of acoustic energy by the nuclear energy release.
576: An often cited criteria for the initiation
577: of a detonation in the context of degenerate carbon-oxygen material is that a ``critical''
578: mass of material needs to be heated and compressed above a certain temperature and density
579: threshold \citep{arnett1994,niemeyer1997,roepke2007a}. While these studies indicate the general
580: conditions under which detonations might readily arise, thermodynamic
581: conditions and heated masses alone represent a gross oversimplification of the underlying initiation process which
582: depends sensitively on the gradients of thermodynamic variables within the heated region.
583: Since gradients play a central role, the resolution and the geometry of the flows being simulated,
584: such as those presented here, are important considerations when investigating the potential
585: for detonation. The suite of simulations studied in this paper use a finest zone size which is 4 km
586: and limits the steepness of temperature gradients which can be represented in our models.
587: And although detonations do arise in our simulations, drawing conclusions from the results of
588: simulations alone concerning the success or failure of detonation will require investigations at
589: significantly higher resolution than has been possible to date.
590:
591: \par We have made some efforts to address the robustness of initiation with a suite of simulation
592: models which employ a patch of mesh refinement over the collision region having zones as fine as
593: 125 m. One of the principal findings of this study, which is being prepared for publication
594: elsewhere (C. Meakin et al. in prep), is that the gradients at the head of the inward directed
595: jet component become steeper at higher resolution which at first appears to inhibit detonation.
596: However, the higher resolution flows develop turbulent structures within the shear layers that
597: form at the interface between the head of the jet and the background stellar material,
598: such as through the Kelvin-Helmholtz instability, which thicken the fuel-ash boundary to an extent
599: that induction time gradients conducive to the spontaneous initiation of a detonation may develop after all.
600:
601: %Although the initiation process is unresolved in the calculations presented here,
602: %this mechanism is being studied with a suite of models which employ a patch of mesh
603: %refinement over the jet region having zones as fine as $\sim 0.1$ km, a scale on which
604: %the gradient mechanism (critical radius) is resolved \citep{niemeyer1997,seitenzahl2008b}.
605: %One of the conclusions of this study, which is being prepared for publication
606: %\citep{seitenzahl2008b}, is that the initiation of the detonation is robust within the
607: %jet and forms through a gradient mechanism at the head of the jet.
608:
609:
610:
611: %
612: %CONDITIONS FOR FAILURE OF A COLLISION REGION TO FORM
613: %
614: %\par Only one model in our suite fails to detonate within the simulated time. In this model
615: %the flame was ignited in a 16 km radius bubble offset 20 km from the stellar center.
616: %The energy released in the deflagration leads to a large expansion of the
617: %stellar core which stalls the surface flow before collision and jet formation can
618: %take place. The total energy released in this model is not enough to unbind the
619: %star, however, and it will eventually collapse again, and remains a candidate for
620: %a pulsational delayed detonation model \citep{khoklov1991,arnett1994}.
621: %The boundary between failure and success in detonation does not depend on only one
622: %parameter, the burned mass, but also depends on the both the spatial and temporal release
623: %of the energy. For instance multi-point ignition, mixing caused by a turbulent background,
624: %and initial gradients in the composition of the core might all lead to different
625: %degrees of expansion and surface flow strengths for the same amount of net burned mass
626: %in the deflagration phase.
627:
628: %The high degree of nonlinearity and feedback inherent in this problem
629: %requires large scale, 3D simulations to study the mapping between
630: %ignition conditions and possible surface detonations in the context of colliding
631: %surface flows. In addition to the uncertainties inherent in turbulent flame models,
632: %viscosity dependent processes such as vortext shedding and advective instabilities
633: %during the early phases of flame bubble evolution, which take place at Reynolds
634: %numbers that are presently impossible to simulate, remain unclear.
635:
636:
637:
638: \subsection{Propagation of the Detonation Wave over the Stellar Core}
639: \label{subsec:det}
640:
641: \par Once the detonation wave forms it propogates outward from the spot
642: of initiation nearly spherically, and consumes the unburned carbon and oxygen remaining
643: in the core. The time sequence in Figure~\ref{fig:det-temp} shows the geometry of the detonation
644: wave as it propagates over the stellar core.
645: The detonation wave speed is a weak function of the upstream plasma density and varies by
646: only $\pm$5\% for the conditions present in the uburned core, where $10^7<\rho<10^9$ g cm$^{-3}$
647: \citep[see Figure 2 of][]{gamezo1999}. The detonation wave traverses the expanded white dwarf
648: in $t_{\rm cross}\sim 2 r_{\rm det}/{D_{CJ}}\sim 0.4$ s where the core size is roughly
649: $r_{\rm det}\sim 2\times10^8$ cm and the detonation wave speed is $D_{CJ}\sim10^9$ cm s$^{-1}$.
650:
651:
652: \par As the detonation wave propagates it compresses upstream material prior
653: to burning. Upstream material with a density greater than $\sim 10^7$ g cm$^{-3}$
654: is compressed and heated strongly enough by the shock that complete relaxation to nuclear
655: statistical equilibrium (NSE) occurs before the rarefaction wave behind the detonation expands
656: the material and it freezes-out (see \S5). At lower upstream densities
657: relaxation to NSE is incomplete and the ash is composed of intermediate
658: mass elements (IMEs) such as Si, S, Ca, and Ar, i.e., the products of incomplete
659: silicon burning \citep[e.g.][]{woosley1973,arnett1996}.
660:
661: \par Material which is compressed to densities exceeding $\sim10^8$ g cm$^{-3}$ in
662: the detonation wave develops a non-negligible neutron excess through electron capture
663: reactions. The strong density dependence of the weak reaction rates limit this
664: neutronization to the central-most regions of the star as evident in
665: Figure~\ref{fig:det-ye} which shows the spatial distribution of electron mole
666: fraction $Y_e$ as the detonation wave sweeps over the stellar core. As discussed in \S5, the
667: final composition of the material burned to NSE, including the fraction which is \nuc{56}{Ni},
668: depends on the degree of neutronization.
669:
670:
671: \par Detonation waves are subject to transverse instabilities which influence
672: the structure of the reaction zone and the reaction products and introduce
673: inhomogeneities in the downstream flow \citep[e.g.,][]{gamezo1999,timmes2000b,sharpe2001}.
674: Therefore, in order to faithfully capture in entirety the properties of the burning in a detonation
675: wave the reaction length scale must be resolved. An additional complication arises
676: in modeling detonations when the density scale height in the medium through which the detonation
677: propagates is comparable to or smaller than the reaction length. Under these
678: conditions steady detonation wave theory cannot be applied and the resulting reactive-
679: hydrodynamic flow remains an active field of research \citep{sharpe2001}.
680: In the context of a carbon-oxygen, near Chandrasekhar-mass white
681: dwarf (M$_{\rm Ch}$), such conditions arise when the upstream density is $\sim10^7$ g cm$^{-3}$.
682: Significant deviations from a Chapman-Jouguet detonation
683: may arise and influence the resulting intermediate mass element (IME) yield.
684: Since IMEs, such as Si and Ca, are
685: primary observational diagnostics of the explosion mechanism underlying SNe Ia
686: \citep[e.g.][]{wang2003,wang2007}, these uncertainties have important implications
687: for modeling all delayed detonation scenarios.
688:
689: \par In the models presented here, the stellar cores undergo only modest expansion
690: during the deflagration and detonation phases. Between 90\% and 97\% of the unburned mass in the core
691: has a density which exceeds $\sim10^7$ g cm$^{-3}$ at the time detonation initiates and
692: all of this material undergoes complete relaxation to NSE,
693: resulting in primarily \nuc{56}{Ni} and and a small fraction of stable Fe-peak elements
694: (\S5 and Table~\ref{tab:models}). Therefore, only a small amount of mass, which is confined to a thin
695: shell in the outer part of the core, is burned to IMEs by the detonation wave.
696: Within this narrow shell the length scales associated with transverse
697: instabilities exceed the grid scale used ($\Delta$ = 4 km) \citep{gamezo1999} and our
698: numerical methods are sufficient to capture them.
699: However, material in this narrow region undergoes rapid expansion
700: after the detonation wave passes and it quickly mixes with the turbulent layer of deflagration
701: ash which lies immediately above it so that it is difficult to discern the presence of cellular
702: structure if it did indeed arise. Significantly higher fidelity simulations are required
703: in order to study the impact that transverse instabilities have under these conditions.
704: While these affects are negligible in the present suite of models, more expanded, lower
705: density cores are likely to be much more strongly impacted by this uncertain physics.
706:
707:
708: \par Upon encountering the deflagration ash which enshrouds the star,
709: the detonation wave transitions into a shock wave which accelerates
710: the hot ash. After the detonation wave and the ensuing shock have propagated
711: off of the computational grid, what is left behind is a rapidly expanding
712: remnant consisting of a smoothly layered core of material burned to NSE with
713: a thin shell of IMEs outside of that, surrounded by a turbulent
714: layer of ash from the deflagration composed of both NSE and
715: IME material.
716:
717:
718:
719:
720:
721: \subsection{Transition to Free Expansion and Final Remnant Shape}
722: \label{subsec:remnant}
723:
724: \par As the detonation wave traverses the stellar core it shifts the density
725: distribution so that the peak in density is initially moved in the positive
726: $y$-direction. This can be seen in Figure~\ref{fig:det-propagation} which
727: presents a time series of velocity and density profiles spanning the time interval
728: over which the detonation wave traverses the stellar core. The initial shift in the
729: density peak towards positive $y$ is due to the strong rarefaction which follows the detonation wave
730: and expands the material behind it on a very short timescale ($\tau_{\rm expand}\sim 0.4$ s).
731: Within $\sim$1 s following the passage of the detonation over the core, the density
732: peak moves back in the negative $y$-direction and ends up south of the equator (negative $y$).
733: The binding energy released in the detonation wave is converted into the kinetic energy
734: associated with expansion within $\sim$1 s after the detonation wave completes its passage over
735: the star. The total energy budget is shown in Figure~\ref{fig:energy} for the model ignited
736: 25 km off center. By $t\sim4$ s the remnant is transitioning into a state of free expansion
737: and assumes a self-similar shape which is no longer changing with time and the radial velocity
738: is well described by a linear dependence on the distance from the center of the remnant.
739: Axial and equatorial profiles of density and radial velocity in the remnant are presented in
740: Figures~\ref{fig:det-profiles} for two models which span the explosion outcomes in our study.
741:
742:
743: \par Shown in Figure~\ref{fig:dens-contours} is the late time ($t>4$ s) remant shape presented as
744: a series of logarithmically spaced density contours for the same two models in Figure~\ref{fig:det-profiles}.
745: What can be seen in this figure is that the asymmetry imparted by the off-center ignition and surface detonation
746: manifests as a shift in the center of density contours even though each individual contour
747: is well described by a circle. The overall shape of the density distribution, therefore, can
748: be characterized by the radius $R_c$ and center $y_c$ of the circles which best describe each
749: density contour. This information is shown in Figure~\ref{fig:shape}. The degree to which these
750: two curves ($R_c$ and $y_c$) approximate the remnant is shown by the thin line in
751: Figure~\ref{fig:det-profiles}, which is the function $r(\rho) = y_c(\rho)\pm R_c(\rho)$.
752:
753: \par Superimposed over the relatively smooth overall shape of the final remnant are smaller scale density
754: inhomogeneities due to the turbulent flow associated with the deflagration and the surface flow which
755: preceeded detonation. These perturbations are quantified in the bottom panel of Figure~\ref{fig:shape}
756: as root mean square (rms) deviations in density taken along the best fit circle at each density.
757: The general trend is that more expanded stellar cores have larger
758: density perturbations in their surface layers at the time
759: of detonation. This can be accounted for partly by the fact that more expansion results from
760: a larger amount of energy liberated in the deflagration which goes into powering the surface flow.
761: Additionally, more expanded cores are less stably stratified at their surfaces (lower gravity and
762: shallower pressure gradients) and so are more easily perturbed by the surface flow which passes over
763: the star before detonation.
764:
765:
766:
767:
768: \section{NUCLEOSYNTHESIS}
769: \label{sec:yields}
770:
771:
772: \par The nucleosynthetic yield for the type of explosion model studied in this paper
773: consist of a mixture of ash due to both a deflagration and a detonation.
774: The total amount of the star burned in the deflagration amounts to less than $\sim$0.1 \msun
775: with nearly the entire remaining mass of the star consumed by the detonation wave which follows.
776: The progress variables described in \S\ref{sec:numerics} which are used to parameterize
777: the compositional evolution and energy release due to nuclear burning allow us to calculate the
778: bulk yield of IMEs and NSE material directly from the multi-dimensional
779: simulation data by performing simple sums \citep{calder2007,townsley2007}. In Table~\ref{tab:models}
780: the total mass burned in the deflagration M$_{\rm def}$ and the detonation M$_{\rm det}$ is summarized,
781: including the budget of IME and NSE material.
782: The deflagration, propagated with the ADR flame model (\S\ref{sec:numerics}), produces a total yield which
783: is approximately one third IMEs and two thirds NSE material, while more than 90\% of the material burned
784: in the detonation is completely relaxed to NSE.
785:
786: \par A general feature of these explosion models is that higher density cores at the time of detonation
787: produce a larger yield of NSE material and a smaller yield of IMEs. This trend is summarized in
788: Figure~\ref{fig:mnse-rhoc} which relates the final NSE yield to the central density of the white dwarf
789: at the time of detonation. The total mass of material having a density exceeding $\rho=10^7$ g/cm$^3$
790: is also shown as a function of central density, and provides a good measure of the mass of material that will
791: burn to NSE in the detonation. The relationship between the amount of mass above a certain density and
792: the central density is a property of the initial white dwarf density structure
793: and the wave form of the pulsation which is excited in the star
794: by the deflagration. The dashed line in Figure~\ref{fig:mnse-rhoc} shows the relationship expected if the
795: pulsation is described by the fundamental mode of the linear wave equation. This mode fits the simulation
796: data remarkably well considering how large (and non-linear) the pulsation amplitude is for the low central density end of the
797: figure. But this can be understood by the fact that the fundamental mode is close to homologous, i.e., the
798: displacement is nearly directly proportional to the radius of the white dwarf.
799:
800:
801: \par The data for the 3D model described in \S\ref{subsec:jet} has been included in
802: Figure~\ref{fig:mnse-rhoc} for comparison, and shows that the overall character of the
803: expansion driven by the deflagration is not dimensionality dependent, nor is the nucleosynthesis
804: that takes place in the detonation. Data from an additional 3D simulation which burned
805: significantly more mass in the deflagration is shown and labeled ``3D Multi''. This simulation
806: data, which is part of an extended suite of 3D models investigating the mapping of ignition conditions to
807: final explosion energies and nucleosynthetic yields (G. C. Jordan IV et al., in prep), was ignited
808: by a uniform distribution of 30 flame bubbles enclosed in an 80 km radius sphere having its center displaced 100 km
809: from the stellar center.
810: This distribution of ignition points is intended to be more representative
811: than single-point ignition of the non-axisymmetric conditions created by
812: the interaction of the growing bubble with the pre-ignition convective
813: core.
814: This model demonstrates that fundamental mode radial pulsation is a good description of
815: the core dynamics preceeding detonation even for significant degrees of expansion.
816: This model also demonstrates that off-center deflagration models are capable of producing a broad range of
817: NSE and \iso{Ni}{56} masses, and not just the most luminous SNe Ia events.
818:
819:
820: \subsection{Iron Peak Freeze-Out Yields: Method}
821: \label{subsec:freezeout-method}
822:
823: \par As NSE material expands and cools in the rarefaction that follows the detonation
824: wave, nuclear reactions eventually cease, the composition no longer changes and the
825: material is said to have gone through {\em freeze-out}.
826: In this section we describe a methodology to efficiently and accurately calculate
827: iron peak yields for material which burns to NSE and then freezes out in the expansion following the
828: detonation wave. In our hydrodynamic simulations $\sim10^5$ to $\sim10^6$ Lagrangian tracer
829: particles are passively advected through the computational domain with an initial distribution that
830: evenly samples the underlying mass distribution. Nucleosynthetic yields can then be calcualted by
831: integrating nuclear reaction networks along each of these particle trajectories and then summing the
832: yields. However, when the large number of particle trajectories required for accurate yield estimates is
833: multiplied by the number of simulation models desired for study, the computational cost of this
834: brute force method of post-processing becomes prohibitively expensive.
835: Therefore, we have developed an alternative approach to calculate the final composition of
836: material processed by the detonation, which takes advantage of the fact that the final nucleosynthetic
837: yield $X_{i,f}$ of material burned to NSE depends only on the final entropy
838: $S_f$, expansion timescale $\tau$, and degree of neutronization $\eta_f$ of the detonated material to a high
839: degree of precision with $X_{i,f} = X_{i,f}(S_f,\eta_f,\tau)$.
840:
841:
842:
843: % I. INDIVIDUAL TRAJECTORY
844: \subsubsection{Individual Trajectories}
845:
846: \par The temperature and density of a generic tracer particle processed by the detonation
847: is presented in Figure~\ref{fig:trajectory}. The evolution of $Y_e$ and the abundances of
848: nuclei along this trajectory have been calculated using a nuclear reaction network
849: initialized with the initial composition of the white dwarf material in the simulation,
850: equal mass fractions of \nuc{12}{C} and \nuc{16}{O}.
851: The network code used for the integrations is a version of the network used in
852: \citet{calder2007} expanded to 443 nuclear species (see Table~\ref{tab:network}). The
853: thermonuclear reaction rates are taken from an expanded version
854: ~\citetext{Schatz 2005, private communication} of the rate compilation
855: REACLIB~\citep{thielemann1986,rauscher2000}. We have also included the
856: temperature-dependent nuclear partition functions provided by~\citet{rauscher2000},
857: both in the determination of the rates of inverse reactions and in our determination
858: of NSE abundance patterns. Electron screening of thermonuclear
859: reaction rates is incorporated, adopting the relations for weak screening and strong
860: screening provided by~\citet{wallace1982} \citep[for additional details see the appendix of][]{calder2007}.
861: Contributions from weak reactions are included using the rates provided by \citet{langanke2001}.
862:
863: \par The time evolution during the rarefaction stage of several abundant
864: species along this trajectory, parameterized by plasma temperature, is shown in
865: Figure~\ref{fig:trajectory-burn} (solid lines).
866: The NSE composition corresponding to the same density and neutron excess for each temperature along the rarefaction part of the trajectory is also shown for comparison (dashed lines).
867: The NSE mass fractions were determined with the NSE solver described in \citet{calder2007} and \cite{seitenzahl2008a}, which uses the same nuclear physics as the reaction network code.
868: Figure~\ref{fig:trajectory-burn} illustrates the degree to which adopting an NSE state at a particular ``freeze out temperature''
869: is a poor approximation to the final freeze-out abundances. This is true because nuclei freeze-out
870: over a fairly large range in temperature and there is non-trivial evolution in a nuclide's abundance
871: after it falls out of NSE but before it reaches its asymptotic freeze-out value.
872:
873: \par The thermodynamic trajectories for Lagragian elements (i.e., tracer particles) processed
874: by the detonation wave are well characterized by an exponential temperature evolution
875:
876: \begin{equation}
877: \label{eqn:temp}
878: T(t) = T_0 \exp(-t/\tau)
879: \end{equation}
880:
881: \noindent and a corresponding density evolution found by assuming adiabaticity
882:
883: \begin{equation}
884: \label{eqn:entropy}
885: S(t) = S(T,\rho, \bar{A}, \bar{Z}) = S_f = (\hbox{constant})
886: \end{equation}
887:
888: \noindent where $S_f$ is the final entropy in the post-detonation state and $\bar{A}$ and $\bar{Z}$ are
889: the average atomic weight and charge of the plasma during the burn with $Y_e = \bar{Z}/\bar{A}$.
890: The density in equation \ref{eqn:entropy} is found using the same Helmholtz equation of state
891: \citep{timmes1999,timmes2000a,fryxell2000} used in the hydrodynamic
892: simulations. A parameterized trajectory, is shown in Figure~\ref{fig:trajectory}
893: for comparison to the particle trajectory.
894:
895: \par The abundance evolution for this parameterized trajectory is presented in
896: Figure~\ref{fig:trajectory-burn} (dotted line) and shows agreement to a high degree
897: of precision with the tracer particle trajectory.
898: Because the composition is well described by a NSE distribution at temperatures above $T_9 \sim 5.5$
899: the final yields are not dependent on the peak temperature reached by the particle trajectory
900: and depend only on the entropy, the total amount of neutronization, and the
901: expansion timescale. Final asympotic freeze-out yields are safely adopted when the plasma temperature
902: drops below $T\sim 10^9$ K with no evolution in the abundances taking place at lower temperatures.
903: Because the electron capture rates are a strong function of density, neutronization occurs in the short lived
904: high density region formed immediately behind the detonation front, while the material is still in NSE and
905: well before freeze-out begins.
906:
907: %\par The stability of the network integrations require that the abundances are initialized with
908: %an appropriate NSE distribution, and this is accomplished using the NSE solver described in
909: %\citet{calder2007} and \cite{seitenzahl2008a} which uses the same nuclear physics as the
910: %reaction network code.
911: %\par Electron capture rates are a strongly increasing function of density so that most
912: %neutronization in the plasma occurs in a narrow region of high density trailing the detonation
913: %front while the material is still in NSE, well before freeze out. This allows us to use a
914: %constant electron fraction for our network integrations chosen as the final value in the
915: %freeze out.
916:
917: % II. ENTROPY-NEUTRONIZATION CORRELATION AND LOOKUP TABLE
918:
919: \subsubsection{Systematic Properties of the Post-Detonation State and Generating a Lookup Table}
920: \par A tight correlation between the degree of neutronization $\eta_f$ and the final entropy
921: $S_f$ of the detonated material further simplifies the procedure and allows us to calculate
922: the final composition for each grid zone in our simulations using a one parameter
923: freeze-out abundance lookup table $X_{i,f} = X_{i,f}(\eta_f)$.
924: The asymptotic values of the degree of neutronization $\eta_f = (1-2 Y_{e,f})$
925: and the entropy $S_f$ for all of the stellar matter burned into NSE in the detonation wave
926: is found to lie along a narrow ridge in the $S_f$-$\eta_f$ plane as shown
927: in Figure \ref{fig:ye-s-table}.
928: This correlation arises from the monotonic dependence of the entropy deposition
929: and the neutronization rate $\dot{Y_e}(\rho)$ on the post-shock plasma density
930: in the detonation wave (which is itself a monotonic function of the pre-shocked,
931: upstream density). (Note, this tight correlation doesn't exist for the material which is
932: burned in the deflagration so that the method described here is not applied to that phase of burning.)
933:
934: A lookup table is constructed along the black line shown in
935: Figure \ref{fig:ye-s-table} and has been sampled at 50 locations logarithmically spaced in neutron excess $\eta$.
936: For every value of $\eta$ in the table, the freeze-out abundances are calculated by integrating the nuclear reaction network,
937: with weak interactions turned off, over an adiabatic analytic trajectories as described above for the corresponding value of
938: $S_f$. The network is initialized with NSE composition at $T_9 = 6.0$ and the integration constinued until $T_9<1.0$, at which
939: time freeze-out has occured. A summary of the stable iron peak yields for material along this locus of points is presented in
940: Figure~\ref{fig:yields-table}. The freeze-out abundances for a computational zone are then be found by using the table entry with
941: the corresponding value of $\eta$. The table lookup is computationally fast, and once the table is created no additional network
942: calculations are necessary.
943:
944: As described above, the final freezeout yield depends on the expansion timescale $\tau$.
945: The expansion timescale, defined by eq.[\ref{eqn:temp}] and found by fitting an exponential to the temperature histories
946: of the tracer particles, varies smoothly across the face of the white
947: dwarf in the narrow range $0.2 < \tau < 0.6$ s for all of the models simulated
948: (the range is narrower for an individual explosion model). If one wanted
949: to incorporate this information into the processing of the final yields, the range in expansion timescales would need
950: to be reflected in the network calculations.
951: In the work presented here, we adopt a central value of $\tau = 0.4$ s and
952: we discuss the sensitivity of the final yields to variations in this value in \S\ref{subsec:final-yields} below.
953:
954: % III. PRESENTATION OF FINAL YIELDS
955: \subsection{Iron Peak Freeze-Out Yields: Results}
956: \label{subsec:final-yields}
957:
958: \par We calculate the final yield of stable iron peak isotopes for all of our explosion models
959: using the procedure outlined in \S\ref{subsec:freezeout-method} above. In Figure~\ref{fig:yields}
960: we present yields for isotopes in the mass range A = 45 to 68 (\nuc{45}{Ti} to \nuc{68}{Zn}),
961: accounting for the decay of radioactive isotopes.
962: The yields ($X_i$) are scaled to the \iso{Fe}{56} abundance ($X_{Fe}$, from the decay chain
963: \nuc{56}{Ni}$\rightarrow$\nuc{56}{Co}$\rightarrow$\nuc{56}{Fe}) and the corresponding relative
964: solar system ratio ($X_{i,\odot}/X_{Fe,\odot}$) based on the abundances of \citet{lodders2003}.
965: In Table~\ref{tab:scaled-yields} we present the elemental abundances for the iron peak elements
966: from Ti to Zn scaled to Fe and solar system ratios. In all cases we highlight results for three
967: explosion models which bracket the range of initial conditions and final outcomes found in our
968: simulation suite. In addition, we provide a detailed list of the final integrated iron peak yields in units
969: of solar mass in Table~\ref{tab:yields}, including the abundances of radioactive isotopes
970: and their half lives. This table can be used to determine the absolute yield of a particular
971: isotope, or to examine the isotopic ratios of specific elements of interest.
972:
973: \par The iron peak yields are similar to pure deflagration
974: models such as the one-dimensional model of Nomoto \citep[W7 yields in][]{brachwitz2000} and the
975: three dimensional model presented in \citet{travaglio2004}.
976: The iron peak yield for the pure deflagration models are more neutron rich than our
977: models, however, because of the higher densities under which the deflagration burns material
978: to NSE. The highest density core to detonate in our model suite (with $r_{\rm off}$=100 km)
979: neutronized the most in the detonation and bears the most similarity to a pure deflagration
980: model in terms of integrated yields.
981: The most neutron-rich isotopes of each element (e.g. \iso{Ti}{50}, \iso{Cr}{54},
982: \iso{Fe}{58}) have no appreciable contribution from the NSE material created by the detonation. There
983: is likely some of these species in the small amount of deflagration material not included in our
984: post-processing. Notably, none of our models produce untoward overabundances ($\gtrsim 2$ times solar,
985: indicated by the dotted lines) of either \iso{Fe}{54} or
986: \iso{Ni}{58}. Overproduction of these nuclides continues to be a serious shortcoming of deflagration
987: models, both spherical and multi-dimensional, which process much of the stellar interior to NSE before
988: expansion can occur.
989: The spatial distribution of the material
990: in our detonation models, however, remains layered in space and velocity while the burning products
991: in the pure deflagration model are strongly mixed due to the turbulent nature under which burning proceeds
992: in a deflagration \citep[e.g.][]{roepke2007b,gamezo2003}.
993:
994: \par Interestingly, the total yield of \iso{Ni}{56} in all of the explosion models presented here is
995: $\sim$1.1\msun independent of the degree of expansion which takes place prior to detonation.
996: This is due to a self regulating process comprised of pre-expansion and neutronization which
997: counteract each other. While the total yield of material burned to NSE is larger for stars
998: which detonate at higher central densities, more neutronization takes place which shifts
999: the iron peak yield to more neutron rich isotopes and away from \iso{Ni}{56} \citep[see e.g.][]{timmes2003}.
1000: The highest density core at detonation produces overall more stable iron peak isotopes
1001: but approximately the same \iso{Ni}{56} yield as the core with the lowest density at detonation.
1002: It can be seen from Figure~\ref{fig:mnse-rhoc}, however, that this trend cannot hold for
1003: significantly more expanded cores since the total mass of high density material drops off
1004: precipitously as lower central densities are reached and will therefore result in SNe Ia
1005: explosions which have smaller \iso{Ni}{56} yields.
1006:
1007: \par {\em Dependence on progenitor neutronization. ---}
1008: The progenitor white dwarf model used in our explosion calculations
1009: is composed of equal parts \nuc{12}{C} and \nuc{16}{O} so that $Y_{e} = 0.5$ everywhere in the
1010: unburned fuel prior to detonation.
1011: However, the progenitor is expected to develop a neutron excess before flame ignition
1012: both during the CNO and He burning cycles and during a $\sim$1000 yr epoch of hydrostatic
1013: carbon burning which is sometimes referred to as ``simmering''.
1014: Recent studies of the ``simmering'' epoch indicate that when carbon burning runs away locally
1015: and a flame is born, $\eta^{\rm sim} \approx 10^{-3}$
1016: \citep{piro2008,chamulak2008}, while stars with an initial metallicity comparable to solar
1017: will develop a neutron excess of $\eta_{\odot} \approx 1.5\times 10^{-3}$ by the time core He burning
1018: commences.
1019:
1020: \par The neutronization which takes place during the detonation is restricted to the densest,
1021: central-most regions of the stellar core because of the strong density dependence of the electron
1022: capture rates (see Figure~\ref{fig:det-ye}). The resulting distribution of neutronization
1023: is presented in Figure~\ref{fig:ye-s-cdf}. This distribution
1024: extends to values lower than the minimum $\eta_{\rm min}^{\rm sim}$ expected in the progenitor
1025: prior to explosive burning. We explore the impact that such a neutronization
1026: floor will have on the yields by enforcing $\eta = \max(\eta_{\rm min},\eta)$ prior to
1027: calculating the iron peak yields using the lookup table method described in \S\ref{subsec:freezeout-method}.
1028: The results are presented in Figure~\ref{fig:yields} (right panel) which shows yields calculated after
1029: applying neutronization floors of $\eta_{\rm min} = 0$, $10^{-3}$, and $2\times 10^{-3}$ for the model
1030: which was least neutronized during the detonation (with $r_{\rm off} = $25 km),
1031: and therefore has the most mass affected by a floor in $\eta$.
1032: The iron peak elements which are primarily produced at low $\eta$ and are therefore most
1033: strongly affected by a neutronization floor are V, Cr, Mn, and Zn, although the isotopic
1034: ratios across the entire iron peak are affected.
1035:
1036: \par {\em Sensitivity to scatter in $S_f$ and $\tau$. ---}
1037: It is possible to construct a higher dimensional lookup table for calculating yields
1038: which accounts for the scatter about the $\eta$-$S_f$ curve used to generate the table
1039: (Figure~\ref{fig:ye-s-table}), but the total error associated with neglecting this scatter
1040: in final entropy $S_f$ is small. In Figure~\ref{fig:yields-texp-entropy} (right) we present the total
1041: variation in the yields due to shifting the $\eta$-$S_f$ curve in final entropy by $\pm$5\%, which is the range of the scatter.
1042: Similarly, the freezeout timescale that takes place in the wake of the detonation wave
1043: has some scatter about the fiducial value of $\tau=0.4$ s that has been used for the results
1044: presented above, spanning the range $0.2 <\tau < 0.6$ s.
1045: The total spread in yields adopting the extreme values for $\tau$ is shown in Figure~\ref{fig:yields-texp-entropy} (left).
1046: The variation in the yield will be significantly smaller than illustrated by these figures since there
1047: exists a smooth distribution between the extreme values of $S_f$ and $\tau$
1048: with the majority of the mass peaked about the central value.
1049: %While the effects of scatter in
1050: %$S_f$ and $\tau$ have a small impact on the final yields, it is important to investigate
1051: %the range of conditions present in any model before applying the freezeout technique described in
1052: %\S\ref{subsec:freezeout-method} above.
1053:
1054:
1055: \subsubsection{The velocity distribution of the yield.}
1056: \par {\em The yield of NSE material. ---}
1057: Inferring the abundance stratification in SNe Ia is possible by studying
1058: high fidelity, multi-epoch spectra; a technique which is proving to be a powerful new tool for
1059: constraining explosions models \citep{stehle2005,roepke2007b,mazzali2008}. The total
1060: number of objects which have had detailed internal abundance stratifications reconstructed
1061: to date is small (only 2 at the time of writing, including 2002bo and 2004eo). These two low
1062: luminosity SNe have inferred \iso{Ni}{56} masses in the range M[\iso{Ni}{56}]$\sim$0.43 - 0.52 \msun.
1063: Despite the very different \iso{Ni}{56} masses between these two observed SNe and the explosion models
1064: presented in this paper, it is interesting to compare the qualitative and quantitative properties of
1065: the abundance stratifications in an attempt to understand the nature of and the diversity inherent
1066: in the explosion mechanism.
1067:
1068: \par For the case of SN 2004eo, \citet{stehle2005} find M[\iso{Ni}{56}]$\sim$0.43 \msun. In their
1069: reconstruction, the \iso{Ni}{56} mass fraction drops below 0.5 at $v_{\rm exp}\sim$ 7,000 km/s,
1070: and drops below 0.1 at 12,000 km/s.
1071: For the case of SN 2002bo, \citet{mazzali2008} find M[\iso{Ni}{56}]$\sim$0.52 \msun. The \iso{Ni}{56}
1072: mass fraction drops below 0.5 at $v_{\rm exp}\sim$ 10,000 km/s and drops below 0.1 at 15,000 km/s.
1073: In both of these SNe, a high mass fraction of stable Fe (X$_{Fe}\sim 1$) is inferred at low velocities
1074: $v_{\rm exp} < $ 3,000 km/s.
1075:
1076: \par The distribution of the elemental abundances as a function of the
1077: radial velocity for our explosion models is presented in Figure~\ref{fig:yields-velocity}.
1078: In this figure, the elemental abundances are calculated by summing over isotopes and
1079: taking into account radioactive decays with half lives less than 1 day.
1080: Two models are shown which bracket the final outcome of all the explosions modeled in this paper.
1081: Both models produce $\sim$1.1\msun of \iso{Ni}{56}.
1082: The model ignited with $r_{\rm off}$=25 km is the most expanded at the time of detonation,
1083: neutronizes the least amount in the detonation, and has the lowest explosion energy,
1084: E$_{\rm tot} = 1.45\times 10^{51}$ erg.
1085: The contribution of stable Fe is the smallest in this model, having a mass fraction
1086: X$_{Fe}\sim$10$^{-3}$ out to $v_{\rm exp}\sim$ 2,000 km/s. The mass fraction of
1087: \iso{Ni}{56} drops below 0.5 at $v_{\rm exp}\sim$ 14,000 km/s, and drops
1088: below 0.1 at $\sim$ 16,000 km/s. The model ignited with $r_{\rm off}$=100 km is the least
1089: expanded at the time of detonation, is neutronized the most by the detonation wave, and has
1090: the largest explosion energy, E$_{\rm tot}=1.52\times10^{51}$ erg.
1091: Although the ejecta at low velocities is still dominated by \iso{Ni}{56},
1092: stable Fe with a mass fraction exceeding X$_{Fe}\sim$0.1 is present out to $v_{\rm exp}\sim$6,000 km/s.
1093: The mass fraction of \iso{Ni}{56} drops below 0.5 at a velocity of $v_{\rm exp}\sim$ 16,000 km/s and
1094: drops below 0.1 at 18,500 km/s.
1095:
1096: \par While the total yield of \iso{Ni}{56} is significantly larger in the explosion models presented
1097: here, the qualitative layered structure of the remnant and the near absence of unburned
1098: carbon and oxygen are in good agreement between the models and the observations. As discussed
1099: in \S\ref{sec:yields} and summarized in Figure~\ref{fig:mnse-rhoc}, lower \iso{Ni}{56}
1100: masses can be produced in surface detonation models which release more energy during the
1101: deflagration phase. However, detonations which take place at lower central densities
1102: and produce smaller \iso{Ni}{56} masses, undergo signifiantly less neutronization and will
1103: therefore fail to reproduce the stable Fe core which has been inferred in the two models
1104: discussed above. On the other hand, it is possible that the neutron rich region seen
1105: at low velocities in SNe Ia remnants are a vestige of the progenitor conditions at ignition.
1106: The nature of the progenitor at flame ignition, including the central density and $Y_e$
1107: distribution, are uncertain and will remain so until the evolution leading up to ignition
1108: is better understood, including the much debated and poorly understood Urca process
1109: \citep[see e.g.,][]{lesaffre2005,arnett1996}.
1110:
1111: \par {\em The yield of non-NSE material. ---}
1112: The detailed yield for material which has not completely relaxed to NSE prior to freezeout
1113: and is composed primarily of IMEs such as Si, S, and Ca is not presented here. In addition to
1114: producing IMEs, material which has begun
1115: silicon burning but has not yet reached an NSE state will contribute to the iron peak with isotopic
1116: ratios that are very different from that which reaches an NSE state.
1117: While the impact of this burning process is small for the suite of explosion models presented in this
1118: paper, which produce primarily NSE material, it is essential to accurately calculate the composition
1119: and distribution of this material for lower luminosity (lower \iso{Ni}{56} mass) SNe Ia explosion models
1120: which will have a significantly larger contribution of non-NSE material.
1121: Additionaly, although the total contribution of IMEs to the mass of the remnant is small in all of the
1122: explosions presented here this material plays a central role in modeling the observational
1123: signatures of these explosion models
1124: and is therefore crucial for comparing our calculations to observational data.
1125: Therefore, a procedure for determining incomplete silicon burning yields which is similar to the
1126: method described in \S\ref{subsec:freezeout-method} is being developed (C. Meakin et al., in preperation).
1127:
1128: %\section{DISCUSSION}
1129: %\par (ala Don Lamb: some words connecting the results of the models presented to the
1130: %observational data. in particular, how the inferred morphology and asymmetries detected
1131: %in spectra and light curves, and the distribution of composition throughout the remantn
1132: %compare with the models presented here.)
1133:
1134:
1135:
1136: \section{CONCLUSIONS}
1137:
1138: \par We have studied the final outcomes for a range of single point flame ignition models of
1139: thermonuclear supernovae within the computational framework developed at the FLASH center
1140: (\S\ref{sec:numerics} and \citet{fryxell2000,calder2007,townsley2007,seitenzahl2008a}).
1141: For the first time in this work, we have extended the 3-stage reactive ash model for nuclear burning
1142: described in \citet{townsley2007}
1143: to study the ignition and propagation of the detonation mode of burning.
1144: As a result, our explosion models
1145: are unique in terms of the degree to which non-idealized nuclear physics are employed, including
1146: a non-static, tabularized treatment of the nuclear statistical equilibrium (NSE) state and the
1147: inclusion of contemporary weak reaction rates \citep{seitenzahl2008a}.
1148: In addition, we have demonstrated here, by reaction network post-processing of
1149: recorded Lagrangian histories, that the 3-stage reactive ash model
1150: provides a suitable reproduction of fluid density and temperature histories to
1151: allow detailed nucleosynthesis, including self-consistent neutronization.
1152: Using these techniques, we have followed the progression of a thermonuclear flame (deflagration)
1153: from a single ignition point which is varied to successively larger distances from the center of a
1154: carbon oxygen white dwarf, and we have described in detail the resulting surface flows and
1155: detonation which ensue.
1156:
1157: \par Detonations arise within a colliding surface flow for all models which are
1158: ignited at a radial location which exceeds $\sim$ 20 km in our 2D simulations.
1159: Flames ignited closer to the stellar center release enough nuclear energy to signifianctly
1160: expand the stellar core to a degree that it stalls the surface
1161: flow, thus preventing a strong collision region and detonation. The nuclear binding energy released
1162: in these stalled surface flow models, however, is not enough to gravitationally unbind the star and
1163: they remain viable candidates for a pulsational detonation upon recollapse \citep{khoklov1991,arnett1994}.
1164: Models which detonate release $\sim 2\times10^{51}$ erg in nuclear binding energy,
1165: resulting in a supernova-like explosion with total energy $E_{\rm tot}\sim1.5\times 10^{51}$ erg.
1166:
1167: \par In all of the models in our parameter study which produce supernova-like explosions, detonation
1168: initiates within a jet-like flow which forms in the converging surface flow. This is in agreement with
1169: the results presented in \citet{kasen2007}. However, we do not find that the
1170: detonation initiates through a shock to detonation transition (SDT) as suggested by these authors, but
1171: instead find that the detonation occurs through a gradient mechanism. The initiation of the detonation
1172: takes place within the compressed gas which lies ahead of the high velocity jet, and ahead of the
1173: internal shock which forms within the jet (see \S\ref{sec:det} and Figure~\ref{fig:jet-slice}).
1174: The focusing of the surface flow and the formation of the jet is also present in 3D simulations
1175: (\S\ref{sec:det}, and \citealt{jordan2007,roepke2007a}),
1176: and is therefore not an artifact of the 2D axisymmetric geometry used.
1177:
1178: \par Within a few seconds after the detonation wave disrupts the stellar core, homologous expansion is
1179: beginning to be established. By t$\sim$ 4 s from flame ignition more than 90\% of the total energy
1180: is in the kinetic energy of expansion, and the expansion velocity has acquired a linear dependence on radius.
1181: The final remnant posesses both global and small scale asymmetries which will influence the observational signature.
1182: When the remnant enters the homologous expansion phase it is characterized by a smooth, layered
1183: inner core surrounded by a low density, flocculent layer of deflagration ash which was dumped onto
1184: the surface of the star prior to detonation. The smooth inner core of the remnant has a global north-south
1185: asymmetry due to off-center ignition and surface detonation, which is well characterized by circular isodensity
1186: contours which are progressively off-center at higher densities (see \S~\ref{subsec:remnant} and Figure~\ref{fig:shape}).
1187:
1188:
1189: \par We have analyzed in detail the nucleosynthesis of material burned to NSE in the detonation.
1190: These results have been generated from the multi-dimensional simulation data using a newly
1191: developed post-processing method which takes advantage of the uniqueness of the NSE state and
1192: systematic properties of detonation waves. The
1193: method, presented in \S\ref{sec:yields}, obviates the need for computationally prohibitive network calculations
1194: along each of the millions of particle trajectories which are necessary for good mass resolution in 3D explosion models.
1195: This work addresses only material which has relaxed to NSE, which forms, by mass, nearly all of the yield from the
1196: 2-dimensional explosion models of this study.
1197: Extending the method to include detailed isotopic yields for material incompletely relaxed to NSE (incomplete carbon,
1198: oxygen, and silicon burning) is being developed and will be described in a forthcoming publication (C. Meakin, in prep).
1199: Nucleosynthesis of material processed in a deflagration instead of detonation burning mode can be processed with a
1200: similarly parameterized method, though requiring more parameters, if it reaches NSE. This leaves only the relative
1201: minority of tracks in partially burned deflagration material to be processed directly (only a few percent of
1202: all the trajectories).
1203:
1204: \par Larger offsets of the ignition point lead to less stellar expansion prior
1205: to detonation and therefore the production of more NSE material. However, we find that
1206: the amount of \iso{Ni}{56} produced stays roughly fixed at $\sim 1.1 M_\odot$ for all
1207: of our 2-dimensional explosion models which extend down to a
1208: central density of $\rho_c\sim4\times 10^{8}$ g/cm$^{3}$ at the time of detonation.
1209: This regulation is due to the enhanced neutronization at the higher densities characteristic
1210: of the less-expanded cases.
1211: Higher density cores at the time of detonation result in more neutronization, and therefore
1212: a larger fractional yield of stable Fe-peak isotopes (e.g., \iso{Fe}{54} and \iso{Fe}{58}).
1213: The isotopic distribution we find in the Fe-peak is very similar to that found for pure deflagration
1214: models, but is characterized by a lower degree of neutronization. Less neutronization is a result of the
1215: lower densities under which the burning proceeds in our surface detonation models compared to pure
1216: deflagrations, due to the pre-detonation expansion.
1217: Between 0.06 and
1218: 0.14\msun\ of intermediate mass elements are produced at high velocities.
1219: Regions in which more than half of the mass is in the form of IMEs lie
1220: above an expansion velocity of 14,000 km/s for all of the 2-dimensional
1221: detonation models calculated.
1222:
1223: % probably should say more about IMEs compared with observation
1224:
1225: \par We successfully reproduced the relationship between the central density and mass-density distribution
1226: in the pre-detonation expanded star by superposing on the hydrostatic star the lowest order radial mode
1227: calculated in a linear approximation.
1228: We find that much smaller \iso{Ni}{56} yields are expected in cores which undergo more expansion prior
1229: to detonation (see Figure~\ref{fig:mnse-rhoc}). This degree of expansion appears to be achievable in 3-dimensional
1230: simulations which relax the constraints on axisymmetry of the ignition conditions necessary for 2-dimensional
1231: simulations. Thus it is expected that more realistic simulations, which include the pre-ignition convection
1232: field and its effect on the growing flame bubble, will be characterized by such larger expansions.
1233: However, further analysis of such simulations, which will be the subject of future papers, is required.
1234:
1235: % -- I think this is going a bit too far in speculation for the conclusions -- it is fine in the main text -Dean
1236: %In these models, almost no stable Fe will be synthesized
1237: %because detonation will occur at such low densities that the iron peak will be completely dominated by
1238: %\iso{Ni}{56}. However, the progenitor is thought to undergo
1239: %some degree of neutronization during the simmering phase prior to flame ignition which will result in
1240: %a contribution of stable iron.
1241:
1242: \par Future work on elucidating the SNe Ia explosion mechanism which is being pursued at the FLASH center
1243: involves the following. (1) We are extending our survey of the mapping between flame ignition conditions and final outcomes
1244: within the computational framework developed at the FLASH center, including multi-point ignition conditions and 3D models.
1245: (2) A simulation pipeline is being constructed to generate synthetic observational diagnostics
1246: for the explosion models, including light curves and spectra, which will allow a more direct comparison between
1247: the systematic properties of the single degenerate Type Ia model and observational data.
1248:
1249: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% AKNOWLEDGEMENTS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1250:
1251: \begin{acknowledgements}
1252: We thank Snezhana Abarzhi for bringing to our attention the literature on
1253: cummulative jets and shaped charges. We also thank Fang Peng for making her nuclear reaction network code available to us for this work.
1254: This work is supported in part at the University of Chicago by the Depart of Energy under Grant B523820 to the ASC/Alliances
1255: Center for Astrophysical Thermonuclear Flashes, and the National Science Foundation under Grant PHY 02-16783 for the
1256: Frontier Center ``Joint Institute for Nuclear Astrophysics'' (JINA), and at the Argonne National Laboratory by
1257: the U.S. Department of Energy, Office of Nuclear Physics, under contract DE-AC02-06CH11357.
1258: \end{acknowledgements}
1259:
1260:
1261:
1262:
1263: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% APPENDIX %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1264:
1265: \begin{appendix}
1266:
1267: \section{FREEZE OUT ABUNDANCES: DETAILED IRON PEAK YIELDS}
1268:
1269: \par The iron peak nucleosynthetic yields for three models spanning the range
1270: of ignition conditions simulated are summarized in Table~\ref{tab:yields}.
1271: The isotopes presented have been selected based on a limiting abundance
1272: ($M_{i;0,f} > 10^{-20}$ \msun). Two columns are shown for each model including the
1273: initial yield and the final yield after radioactive decays have been taken into
1274: account. The half lives and decay modes are presented for unstable isotopes.
1275:
1276: \end{appendix}
1277:
1278:
1279: %%%%%%%%%%%%%%%%%%%%%%%%%% REFERENCES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1280: %\input{biblio.tex}
1281:
1282: \begin{thebibliography}
1283:
1284: \bibitem[Arnett(1996)]{arnett1996} Arnett, D.\ 1996, Supernovae
1285: and Mucleosynthesis: An Investigation of the History of Matter, from the
1286: Big Bang to the Present, by D.~Arnett.~Princeton: Princeton University
1287: Press, 1996.,
1288:
1289: \bibitem[Arnett \& Livne(1994)]{arnett1994} Arnett, D., \& Livne,
1290: E.\ 1994, \apj, 427, 315
1291:
1292: %\bibitem[Asida et al.(2007)]{asida2007} Asida, S. et al, 2007, in preperation.
1293:
1294: \bibitem[Birkhoff et al.(1948)]{birkhoff1948} Birkhoff, G.,
1295: MacDougall, D.~P., Pugh, E.~M.,
1296: \& Taylor, G.\ 1948, Journal of Applied Physics, 19, 563
1297:
1298:
1299: \bibitem[Brachwitz et al.(2000)]{brachwitz2000} Brachwitz, F., et al., 2000, \apj, 536, 934
1300:
1301: \bibitem[Branch et al.(1995)]{branch1995} Branch, D., Livio, M.,
1302: Yungelson, L.~R., Boffi, F.~R., \& Baron, E.\ 1995, \pasp, 107, 1019
1303:
1304: \bibitem[Calder et al.(2007)]{calder2007} Calder, A.C., et al, 2007 \apj 656, 313C
1305:
1306: \bibitem[Caughlan \& Fowler(1988)]{caughlan1988} Caughlan, G.~R., \&
1307: Fowler, W.~A.\ 1988, Atomic Data and Nuclear Data Tables, 40, 283
1308:
1309: \bibitem[Chabrier \& Potekhin (1998)]{chabrier1998} Chabrier, G., Potekhin, A., 1998, \pre, 58, 4941
1310:
1311: \bibitem[Chamulak et al.(2008)]{chamulak2008} Chamulak, D.~A., Brown, E.~F., Timmes, F.~X.,
1312: \& Dupczak, K.\ 2008, ArXiv e-prints, 801, arXiv:0801.1643
1313:
1314:
1315: \bibitem[Colella \& Glaz(1985)]{colella1985} Colella, P., \& Glaz,
1316: H.~M.\ 1985, Journal of Computational Physics, 59, 264
1317:
1318: \bibitem[Colella \& Woodward(1984)]{colella1984} Colella, P., \&
1319: Woodward, P.~R.\ 1984, Journal of Computational Physics, 54, 174
1320:
1321: %\bibitem[Courant \& Friedrichs (1948)] {courant1948} Courant, R., \& Friedrichs, K.O., 1948,
1322: %\textit{Supersonic Flow and Shock Waves}, Interscience, New York
1323:
1324: \bibitem[Dursi \& Timmes(2006)]{dursi2006} Dursi, L.~J., \&
1325: Timmes, F.~X.\ 2006, \apj, 641, 1071
1326:
1327:
1328: \bibitem[Fryxell, Mueller \& Arnett(1989)]{fryxell1989} Fryxell, B., Mueller, W., \& Arnett, D.,
1329: MPIA technical report
1330:
1331: \bibitem[Fryxell et al.(2000)]{fryxell2000} Fryxell, B., et al.\
1332: 2000, \apjs, 131, 273
1333:
1334: \bibitem[Gamezo et al.(1999)]{gamezo1999} Gamezo, V.~N., Wheeler,
1335: J.~C., Khokhlov, A.~M., \& Oran, E.~S.\ 1999, \apj, 512, 827
1336:
1337: \bibitem[Gamezo et al.(2003)]{gamezo2003} Gamezo, V.~N., Khokhlov,
1338: A.~M., Oran, E.~S., Chtchelkanova, A.~Y.,
1339: \& Rosenberg, R.~O.\ 2003, Science, 299, 77
1340:
1341:
1342: \bibitem[Gamezo et al.(2005)]{gamezo2005} Gamezo, V.~N., Khokhlov,
1343: A.~M., \& Oran, E.~S.\ 2005, \apj, 623, 337
1344:
1345: \bibitem[Hillebrandt \& Niemeyer(2000)]{hillebrandt2000} Hillebrandt,
1346: W., \& Niemeyer, J.~C.\ 2000, \araa, 38, 191
1347:
1348:
1349: \bibitem[Jordan et al.(2008)]{jordan2007} Jordan, G.~I., Fisher,
1350: R., Townsley, D., Calder, A., Graziani, C, Asida, S, Lamb, D., \& Truran,
1351: J.\ 2007, \apj, 681, 1448
1352:
1353:
1354: \bibitem[Kasen \& Plewa(2007)]{kasen2007} Kasen, D., \& Plewa, T.\ 2007, \apj, 662, 459
1355:
1356:
1357: %\bibitem[Khokhlov(1989)] {khoklov1989} Khokhlov, A.M., 1989, \mnras, 239, 785
1358:
1359: \bibitem[Khokhlov(1991)]{khoklov1991} Khokhlov, A.~M.\ 1991, \aap, 245, 114
1360:
1361: \bibitem[Khokhlov(1995)]{khoklov1995} Khokhlov, A.~M.\ 1995, \apj,
1362: 449, 695
1363:
1364:
1365:
1366: \bibitem[Khokhlov et al.(1997)]{khokhlov1997} Khokhlov, A.~M., Oran,
1367: E.~S., \& Wheeler, J.~C.\ 1997, \apj, 478, 678
1368:
1369:
1370: \bibitem[Langanke \& Mart\'inez-Pinedo(2001)] {langanke2001} Langanke, K., Mart\'inez-Pinedo, G., 2001,
1371: ADNDT, 79, 1
1372:
1373: \bibitem[Lee \& Higgins(1999)]{lee1999} Lee, J.H.S., \& Higgins, A.J. 1999.
1374: Phil. Trans. R. Soc. Lond. A 357, 3503
1375:
1376:
1377: \bibitem[Lesaffre et al.(2005)]{lesaffre2005} Lesaffre, P.,
1378: Podsiadlowski, P., \& Tout, C.~A.\ 2005, Nuclear Physics A, 758, 463
1379:
1380: \bibitem[Lodders(2003)]{lodders2003} Lodders, K.\ 2003, \apj, 591, 1220
1381:
1382:
1383: \bibitem[Mazzali et al.(2008)]{mazzali2008} Mazzali, P.~A., Sauer,
1384: D.~N., Pastorello, A., Benetti, S., \& Hillebrandt, W.\ 2008, \mnras, 456
1385:
1386:
1387:
1388:
1389:
1390: \bibitem[Niemeyer \& Woosley(1997)]{niemeyer1997} Niemeyer, J.~C.,
1391: \& Woosley, S.~E.\ 1997, \apj, 475, 740
1392:
1393: \bibitem[Nomoto et al.(1984)]{nomoto1984} Nomoto, K., Thielemann, F.K., Yokoi, K, 1984, \apj, 286, 644
1394:
1395: \bibitem[Piro \& Bildsten(2008)]{piro2008} Piro, A.~L., \& Bildsten, L.\ 2008, \apj, 673, 1009
1396:
1397: \bibitem[Plewa(2007)]{plewa2007} Plewa, T.\ 2007, \apj, 657, 942
1398:
1399:
1400: \bibitem[Rauscher \& Thielemann(2000)]{rauscher2000} Rauscher, T., Thielmann, F.-K., 2000, ADNDT, 75, 1R
1401:
1402: \bibitem[Riess et al.(1998)]{riess1998} Riess, A.~G., et al.\
1403: 1998, \aj, 116, 1009
1404:
1405: \bibitem[R{\"o}pke et al.(2007a)]{roepke2007a} R{\"o}pke, F.~K.,
1406: Woosley, S.~E., \& Hillebrandt, W.\ 2007, \apj, 660, 1344
1407:
1408: \bibitem[R{\"o}pke et al.(2007b)]{roepke2007b} R{\"o}pke, F.~K.,
1409: Hillebrandt, W., Schmidt, W., Niemeyer, J.~C., Blinnikov, S.~I., \&
1410: Mazzali, P.~A.\ 2007, \apj, 668, 1132
1411:
1412: \bibitem[Schmidt et al.(2006a)]{schmidt2006a} Schmidt, W., Niemeyer,
1413: J.~C., \& Hillebrandt, W.\ 2006, \aap, 450, 265
1414:
1415: \bibitem[Schmidt et al.(2006b)]{schmidt2006b} Schmidt, W., Niemeyer,
1416: J.~C., Hillebrandt, W.,R{\"o}pke, F.~K.\ 2006, \aap, 450, 283
1417:
1418: \bibitem[Seitenzahl et al.(2008a)]{seitenzahl2008a} Seitenzahl, I., Townsley, D., Peng, F., Truran, J.,
1419: ADNDT, in press. %NSE table paper
1420:
1421: \bibitem[Seitenzahl et al.(2008b)]{seitenzahl2008b} Seitenzahl, I., et al., in preperation.
1422: %detonation initiation paper
1423:
1424: \bibitem[Sharpe(2001)]{sharpe2001} Sharpe, Gary J., 2001, \mnras, 322, 614
1425:
1426:
1427: \bibitem[Stehle et al.(2005)]{stehle2005} Stehle, M., Mazzali,
1428: P.~A., Benetti, S., \& Hillebrandt, W.\ 2005, \mnras, 360, 1231
1429:
1430:
1431: \bibitem[Thielemann et al.(1986)] {thielemann1986} Thielemann, F.-K., Truran, J., Arnould, M., 1986,
1432: Advances in Nuclear Astrophysics, ed. E. Vangioni-Flam et al. (Gif-sur-Yvette: Edidtions Fronti\`eres), 525
1433:
1434: \bibitem[Timmes \& Arnett (1999)] {timmes1999} Timmes, F.~X.,Arnett, D., 1999, \apjs, 125, 277
1435:
1436: \bibitem[Timmes et al.(2003)]{timmes2003} Timmes, F.~X., Brown,
1437: E.~F., \& Truran, J.~W.\ 2003, \apjl, 590, L83
1438:
1439: \bibitem[Timmes \& Swesty(2000a)]{timmes2000a} Timmes, F.~X., \&
1440: Swesty, F.~D.\ 2000, \apjs, 126, 501
1441:
1442: \bibitem[Timmes et al.(2000b)]{timmes2000b} Timmes, F.~X., et al.\
1443: 2000, \apj, 543, 938
1444:
1445: \bibitem[Timmes \& Woosley(1992)]{timmes1992} Timmes, F.~X., \&
1446: Woosley, S.~E.\ 1992, \apj, 396, 649
1447:
1448:
1449: \bibitem[Travaglio et al.(2004)] {travaglio2004} Travaglio, C., et al., 2004, \aap, 425, 1029
1450:
1451: \bibitem[Townsley et al.(2007)]{townsley2007} Townsley, D., et al., 2007, \apj, submitted
1452:
1453: \bibitem[Wallace et al.(1982)]{wallace1982} Wallace, R., Woosley, S., Weaver, T., 1982, \apj, 225, 1021
1454:
1455: \bibitem[Wang et al.(2003)]{wang2003} Wang, L., et al.\ 2003,
1456: \apj, 591, 1110
1457:
1458: \bibitem[Wang et al.(2007)]{wang2007} Wang, L., Baade, D.,
1459: \& Patat, F.\ 2007, Science, 315, 212
1460:
1461:
1462: \bibitem[Woosley et al.(1973)]{woosley1973} Woosley, S.~E., Arnett,
1463: W.~D., \& Clayton, D.~D.\ 1973, \apjs, 26, 231
1464:
1465: \bibitem[Woosley et al.(2004)]{woosley2004} Woosley, S.~E., Wunsch,
1466: S., \& Kuhlen, M.\ 2004, \apj, 607, 921
1467:
1468:
1469: \bibitem[Zel'dovich et al.(1970)]{zeldovich1970} Zel'dovich, Ya. B., Librovich, V. B.,
1470: Makhviladze, G. M., \& Sivashinsky, G. I. 1970, Acta Astron., 15, 313
1471:
1472: \end{thebibliography}
1473:
1474:
1475:
1476: %%%%%%%%%%%%%%%%%%%%%%%%%%%% FIGURES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1477: \clearpage
1478: %\input{figures.tex}
1479:
1480: \begin{figure}
1481: % \includegraphics{figs/rpv1_16o40.jpg}
1482: \includegraphics[scale=0.75]{f1.jpg}
1483: \caption{\label{fig:breakout}
1484: This time sequence of ash abundance (represented by the $\phi_1$
1485: progress variable) shows how a bubble ignited near the stellar core rises
1486: buoyantly, erupts from the star's surface and drives a flow which is largely
1487: confined to the surface of the star by gravity. The black contour line indicates
1488: a density of 10$^7$ g cm$^{-3}$. Eventually the surface flow converges at the
1489: opposite pole from the breakout location, compressing material in that region
1490: until it begins to burn carbon. The dashed box in the right panel indicates
1491: the region detailed in Figure \ref{fig:jet} below, where the converging
1492: flow produces a jet that initiates a detonation wave.}
1493: \end{figure}
1494:
1495:
1496: \begin{figure}
1497: % \includegraphics{figs/homologous_expansion.25_and_100.ps}
1498: \includegraphics{f2.jpg}
1499: \caption{The radial density profile scaled to the central density and the
1500: density e-folding height for the initial white dwarf (thick grey
1501: line), and at the time when a surface detonation initiates ($t_{\rm det}$,
1502: see Table \ref{tab:models}) for flame bubbles ignited at 25 (thin black lines)
1503: and 100 km (thick black lines) off-center.
1504: Equatorial (solid lines) and polar (dashed lines) profiles are shown
1505: for both of the pre-detonation models and are well described by homologous
1506: expansion with minimal asymmetry.
1507: \label{fig:homologous-deflagration}}
1508: \end{figure}
1509:
1510:
1511: \begin{figure}
1512: % \includegraphics{figs/jet_fig_16o40_press.jpg}%{figs/jet_initiation_final.ps}
1513: \includegraphics[scale=0.6]{f3.jpg}
1514: \caption{A time sequence showing the ``collision region'' and the bidirectional
1515: jet-like flow for the model ignited 40 km off-center.
1516: The pressure field and the velocity field are shown as the detonation wave initiates
1517: and begins to break away from the end of the inwardly moving jet component.
1518: The red and black contour lines indicate carbon depletion at the 1\% and 99\% levels,
1519: respectively. In the region where the detonation initiates the
1520: density and temperature are 10$^7$ g/cm$^3$ and 4$\times$10$^9$ K. The longest
1521: velocity vector indicates a flow speed of $v_{\rm vec} = 10^9$ cm/s, while other
1522: vectors have lengths linearly proportional to the flow speed.
1523: \label{fig:jet}}
1524: \end{figure}
1525:
1526:
1527: \begin{figure}
1528: \epsscale{1.0}
1529: %\includegraphics{figs/jet_slices_w3d_wpram.jpg}
1530: \includegraphics[scale=0.62]{f4.jpg}
1531: \caption{Flow properties along the jet axis prior to detonation for 2D and 3D models, including
1532: density, radial velocity, temperature, and gas pressure. The radially directed ram pressure,
1533: $p_{\rm ram} = \rho v_r^2$, is shown in the pressure figure for each model by the thin red line.
1534: The 2D models shown were ignited by 16 km radius flame bubbles with offset (left) $r_{\rm off} = 25$ km and
1535: (middle) $r_{\rm off} = 100$ km. A low resolution ($\Delta = 8$ km) 3D model is shown (right)
1536: which was ignited by a 16 km radius flame bubble with offset $r_{\rm off} = 80$ km for comparison.
1537: The dashed vertical lines mark the locations of the burning front for each model, taken to be
1538: where $\phi_1 = 0.5$. In the top panel, the horizontal dot-dashed line marks a density of
1539: $10^7$ g/cm$^3$, a value above which detonation readily arises in the simulations once it reaches
1540: a temperature of $T\sim 2\times 10^9$ K. The velocity zero point is marked by a dot-dashed horizontal line.
1541: \label{fig:jet-slice}}
1542: \end{figure}
1543:
1544:
1545: \begin{figure}
1546: %\includegraphics{figs/det_fig_16o40_temp.jpg}
1547: \includegraphics[scale=0.6]{f5.jpg}
1548: \caption{In this time sequence, the detonation wave breaks away from the jet in which it
1549: formed and sweeps across the stellar core. The black line is the 10$^7$ g/cm$^3$ iso-density
1550: contour which is roughly coincident with the stellar surface.
1551: \label{fig:det-temp}}
1552: \end{figure}
1553:
1554:
1555:
1556:
1557: \begin{figure}
1558: %\includegraphics{figs/ye_fig_16o40.jpg}
1559: \includegraphics[scale=0.6]{f6.jpg}
1560: \caption{The time evolution of the electron mole fraction ($Y_e$) is shown as the detonation
1561: wave passes over the stellar center. The dip in $Y_e$ reveals that neutronization is taking
1562: place in the high density core where material has burned to NSE. The expansion
1563: which follows the detonation freezes the $Y_e$ distribution as the material evolves into a
1564: supernova remnant. The neutron rich material (low $Y_e$) which surrounds the stellar
1565: core is the ash from the deflagration which had burned at high densities before it
1566: erupted from the star and spread out over the surface.
1567: The black line is the 10$^7$ g/cm$^3$ iso-density contour and the light
1568: blue line indicates the contour of carbon depletion at the 99\% level.\label{fig:det-ye}}
1569: \end{figure}
1570:
1571:
1572: \begin{figure}
1573: %\includegraphics{figs/vyslices.460_to_510.ps}{figs/ddslices.460_to_510.ps}
1574: \includegraphics[scale=0.45]{f7a.jpg}
1575: \includegraphics[scale=0.45]{f7b.jpg}
1576: \caption{The profile of the (left) velocity and (right) density along the
1577: polar axis is shown for several moments evenly spaced in time ($\delta t=$0.025 s)
1578: as the detonation wave passes across the stellar core for the model ignited
1579: with a flame bubble 40 km from the stellar center.
1580: \label{fig:det-propagation}}
1581: \end{figure}
1582:
1583:
1584:
1585:
1586: \begin{figure}
1587: \epsscale{1.0}
1588: %\includegraphics{figs/energy_25.ps}
1589: \includegraphics{f8.jpg}
1590: \epsscale{1.0}
1591: \caption{The time evolution of the kinetic, internal, and gravitational potential
1592: energy for a model with a flame bubble ignited 25 km from the stellar center.
1593: The nuclear energy released by burning in the deflagration ($t < t_{det}$, with $t_{det} = 2.45$s)
1594: and the detonation ($t > t_{det}$) can be seen as a change in the sum of
1595: these three energy components (blue).\label{fig:energy}}
1596: \end{figure}
1597:
1598:
1599:
1600:
1601: \begin{figure}
1602: %\includegraphics{figs/detprofiles_25and100.jpg}
1603: \includegraphics[scale=0.7]{f9.jpg}
1604: \caption{Late time ($t>4$ s) density and velocity profiles for post detonation state models
1605: having ignition:
1606: (left) 25 km and (right) 100 km off-center. The density
1607: is scaled by the peak value and the position is scaled by the density e-folding
1608: distance in the equatorial direction.
1609: % The spatial scales for the 25 km and 100 km offset models
1610: % are $r_{\rho} = 1.41\times 10^9$ cm and $r_{\rho} = 1.62\times 10^9$ cm,
1611: % respectively.
1612: The thick gray line shows the scaled density profile
1613: of the initial white dwarf model, while the post detonation state model is shown
1614: by thick black lines for profiles along the: (solid) equatorial, and (dashed) polar axes.
1615: The thin black line shows the reconstructed density profile along the polar
1616: axis based on the contour fits presented in Figure~\ref{fig:shape} (see text for more details).
1617: \label{fig:det-profiles}}
1618: \end{figure}
1619:
1620:
1621:
1622: %\begin{figure}
1623: % \epsscale{1.0}
1624: % \includegraphics{figs/line-density-25and100.ps}
1625: % \caption{Scaled line density, $\lambda/\lambda_{\rm max}$ (see eq.[\ref{eq:line-density}]), along
1626: % the symmetry axis at the following times: (black) just prior to detonation, and
1627: % (gray) at late times ($t>4$ s).
1628: % Shown are the models with flame bubbles ignited: (thick) 25 km, and
1629: % (thin) 100 km off-center.
1630: % The spatial scales are the same density e-folding lengths as in Figure~\ref{fig:det-profiles}.
1631: % \label{fig:line-density}}
1632: %\end{figure}
1633:
1634: \begin{figure}
1635: \epsscale{1.0}
1636: %\includegraphics{figs/dens_contours_25and100.jpg}
1637: \includegraphics[scale=0.7]{f10.jpg}
1638: \caption{Late time ($t>4$ s) density contours for the models ignited: (left) 25 km, and
1639: (right) 100 km off-center. The contours mark the locations at which
1640: $\ln (\rho/\rho_c) = -0.5, -1, -1.5, -2, -2.5$ where $\rho_c$ is the peak density.
1641: The magnitude of the largest velocity vectors are (left) $v = 2.4\times 10^9$ cm/s, and
1642: (right) $v = 2.5\times 10^9$ cm/s.
1643: The spatial scales are the same density e-folding lengths as in Figure~\ref{fig:det-profiles}.
1644: \label{fig:dens-contours}}
1645: \end{figure}
1646:
1647:
1648: \begin{figure}
1649: \epsscale{0.8}
1650: %\includegraphics{figs/shape_25_and_100.ps}
1651: \includegraphics{f11.jpg}
1652: \caption{The iso-density contours of the remnant during late times
1653: ($t>4$ s, Figure~\ref{fig:dens-contours}) are well described by
1654: circles of radius $R_c$ which have centers that are offset from the origin
1655: by an amount $y_c$ along the symmetry axis. The best fit radii and offsets are
1656: shown here as a function of density contour value scaled to the central density
1657: for the models ignited 25 km (thin-line), and 100 km (thick-line) off-center
1658: in the panels above: (top) radius, $R_c$; (middle) circle center, $y_c$.
1659: The spatial dimensions are scaled in terms of the e-folding density scale
1660: height in the equatorial direction of the remnant.
1661: (bottom) The degree of clumpiness is characterized by the ratio of the r.m.s.
1662: deviation in density along the best fit circle to the density contour value,
1663: denoted $\delta\rho/\rho$.
1664: The density perturbations at high density ($\ln\rho/\rho_c$>1.5) are due primarily
1665: to the narrow trail of ash left behind as the flame bubble rises our of the
1666: stellar core.
1667: \label{fig:shape}}
1668: \end{figure}
1669:
1670:
1671: \clearpage
1672:
1673:
1674: \begin{figure}
1675: \epsscale{1.0}
1676: %\includegraphics{figs/mnse_rhoc.ps}
1677: \includegraphics{f12.jpg}
1678: \caption{The total mass of NSE material and \nuc{56}{Ni} created in the detonation and the
1679: total mass of high density ($\rho>10^7$ g/cm$^3$) matter at detonation is plotted against
1680: the central density at detonation for all of the 2D models studied.
1681: The total mass of material having a density which exceeds
1682: $\rho=5\times 10^6$, $7.5\times 10^6$, and $10^7$ g/cm$^3$ during a 0.25 s time period
1683: preceeding detonation is shown by the curves which terminate at the detonation
1684: density for the 25 km and 100 km off-center ignition models.
1685: Data points for two 3D models are shown for comparison: the data points labeled
1686: ``3D Single'' show the \nuc{56}{Ni}, NSE, and high density material masses for a
1687: 3D model ignited 80 km off-center. The data point labeled ``3D Multi'' shows
1688: the central density and the high density material mass for a 3D multi-point ignition
1689: model which is described in the text (see \S\ref{sec:yields}). The dashed line shows
1690: the relationship between central density and high density material for the initial
1691: white dwarf model expanded by the (linear) fundamental pulsation mode.
1692: \label{fig:mnse-rhoc}}
1693: \end{figure}
1694:
1695:
1696:
1697: \begin{figure}
1698: %\includegraphics{figs/traject_exp_solid.jpg}
1699: \includegraphics[scale=0.6]{f13.jpg}
1700: \caption{The thermodynamic trajectory of a Lagrangian tracer particle having
1701: an expansion timescale $\tau = 0.42$s, and final entropy $s$ = 2.273 N$_A k$ and final
1702: electron mole fraction $Y_e$ = 0.49873. The dotted line shows the analytic
1703: adiabatic fit to this trajectory, parameterized by $\tau$, $Y_e$, and $s$.
1704: \label{fig:trajectory}}
1705: \end{figure}
1706:
1707:
1708:
1709:
1710: \begin{figure}
1711: \epsscale{0.9}
1712: %\includegraphics{figs/tj_exp_nse_many_highT.jpg}
1713: \includegraphics[scale=0.6]{f14.jpg}
1714: \caption{The time evolution is shown for various high abundance iron peak isotopes
1715: during the expansion which is here parameterized by the plasma temperature.
1716: For each species the abundances have been calculated with a network using the thermodynamic
1717: trajectory of the tracer particle (solid) and the analytic fit (dotted). For comparison, the
1718: NSE values are shown (dashed) using the thermodynamic conditions at each point along
1719: the particle trajectory. All three are in good agreement until $T\sim5.5\times10^9$ K,
1720: below which the NSE distribution begins to deviate from the network calculation at various
1721: temperatures. The tracer particle and the analytic fit agree to a high level of precision
1722: through freeze out. \label{fig:trajectory-burn}}
1723: \end{figure}
1724:
1725:
1726:
1727:
1728: \begin{figure}
1729: %\includegraphics{figs/ye_s.4_16_100.ps}
1730: \includegraphics[scale=0.6]{f15.jpg}
1731: \caption{The distribution of NSE mass in entropy and degree of neutronization
1732: for the model with 100 km off-center bubble ignition. The black line indicates the
1733: curve on which the freeze out yield table was calculated (\S\ref{subsec:freezeout-method}).
1734: %(right) Cumulative distribution function showing the fraction of NSE material below
1735: %a certain degree of neutonization.
1736: \label{fig:ye-s-table}}
1737: \end{figure}
1738:
1739:
1740:
1741: \begin{figure}
1742: %\includegraphics{figs/ye_s_cdf.ps}
1743: \includegraphics{f16.jpg}
1744: \caption{Cumulative distribution functions showing the fraction of NSE material below
1745: a certain degree of neutonization for the models ignited: 25 km, 40km, and 100 km
1746: off-center (from darkest to lightest, respectively).
1747: \label{fig:ye-s-cdf}}
1748: \end{figure}
1749:
1750:
1751: \begin{figure}
1752: %\includegraphics{figs/freezeout_table_unscaled.ps}{figs/freezeout_table_scaled.ps}
1753: \includegraphics[scale=0.45]{f17a.jpg}
1754: \includegraphics[scale=0.45]{f17b.jpg}
1755: \caption{Iron peak freezeout yields, accounting for
1756: radioactive decay, as a function of the degree of neutronization.
1757: The corresponding entropy of the material is related to the degree of neutronization,
1758: $S_f = S_f(\eta_f)$, by the black line shown in Figure~\ref{fig:ye-s-table}
1759: and an expansion timescale of $\tau = 0.4$ s.
1760: (left) Isotope mass fractions, X(\nuc{A}{Z}) for nuclide with atomic number A and
1761: proton number Z.
1762: (right) Isotope mass fractions scaled to X(\nuc{56}{Fe}) and normalized to the
1763: corresponding solar system abundance ratios of \citet{lodders2003} where
1764: [\nuc{A}{Z}/\nuc{56}{Fe}] =$\log_{10}$
1765: (X(\nuc{A}{Z})/X(\nuc{56}{Fe})) - $\log_{10}$(X(\nuc{A}{Z})/X(\nuc{56}{Fe}))$_{\odot}$.
1766: The dashed horizontal lines indicate where the scaled abundance ratio is equal to 0.5,
1767: 1.0, and 2.0 times the solar system value.
1768: \label{fig:yields-table}}
1769: \end{figure}
1770:
1771:
1772: \begin{figure}
1773: %\includegraphics{figs/final_iron_peak_yields.ps}{figs/final_iron_peak_yields_25km_etamin.ps}
1774: \includegraphics[scale=0.45]{f18a.jpg}
1775: \includegraphics[scale=0.45]{f18b.jpg}
1776: \caption{(left) The nucleosynthetic yields normalized by the solar system abundances are
1777: shown for three models which span the degree of pre-expansion and neutronization
1778: in our simulation suite.
1779: (right) The variation in the yields due to imposing a neutronization floor of
1780: $\eta_{\rm min}$ = 0, $10^{-3}$, and $2\times 10^{-3}$ for the model ignited
1781: 25 km off-center. Same notation as Figure~\ref{fig:yields-table} \label{fig:yields}}
1782: \end{figure}
1783:
1784:
1785: \begin{figure}
1786: %\includegraphics{figs/final_iron_peak_yields_100km_texp.ps}
1787: %{figs/final_iron_peak_yields_100km_entropy.ps}
1788: \includegraphics[scale=0.45]{f19a.jpg}
1789: \includegraphics[scale=0.45]{f19b.jpg}
1790: \caption{The variation in nucleosynthetic yields due to:
1791: (left) variations in expansion timescale, with $\tau_{exp} = $0.2, 0.4, and 0.6 s;
1792: and (right) entropy changes, where $\pm$5\% variations about the fiducial
1793: entropy-neutronization curve, $S_f = S_f(\eta_f)$, shown in Figure~\ref{fig:ye-s-table}
1794: have been used. Same notation
1795: as Figure~\ref{fig:yields-table}.
1796: \label{fig:yields-texp-entropy}}
1797: \end{figure}
1798:
1799:
1800:
1801: \begin{figure}
1802: %\includegraphics{figs/vyields.4_16_25.ps}{figs/vyields.4_16_100.ps}
1803: \includegraphics[scale=0.45]{f20a.jpg}
1804: \includegraphics[scale=0.45]{f20b.jpg}
1805: \caption{Distribution of elemental abundances are shown as a function of
1806: expansion velocity for models having initial flame bubbles ignited (left)
1807: 25 km and (right) 100 km from the stellar center. The elemental yields are
1808: calculated by taking into account radioactive decays with half lives less than
1809: 1 day. The dotted vertical line indicates the velocity above which less than
1810: 95\% of the material has been burned to NSE and therefore our nucleosynthesis
1811: post-processing method (\S\ref{subsec:freezeout-method}) is no longer reliable.
1812: At velocities where less than 95\% of the material burns to NSE we show only
1813: the total fraction of NSE and intermediate mass elements (IMEs). The red
1814: curve shows the fraction of the total stellar mass interior to the
1815: velocity. \label{fig:yields-velocity}}
1816: \end{figure}
1817:
1818:
1819:
1820: %%%%%%%%%%%%%%%%%%%%%%%%% TABLES %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1821: \clearpage
1822: \input{tab1.tex}
1823: \input{tab2.tex}
1824: \input{tab3.tex}
1825: \input{tab4.tex}
1826:
1827: %\input{tables.tex}
1828: %\input{tables.appendix.tex}
1829:
1830:
1831:
1832:
1833:
1834: \end{document}
1835: