0811.3610/ms.tex
1: %\documentclass[12pt,preprint]{aastex}
2: \documentclass{emulateapj}
3: 
4: % PACKAGES TO USE
5: \usepackage{psfig}
6: \usepackage{amsmath}
7: \usepackage{amssymb}
8: \usepackage{graphicx}
9: \usepackage{appendix}
10: 
11: % NEW COMMANDS
12: \newcommand{\lSect}[1]{{\label{sec:#1}}}
13: \newcommand{\lFig}[1]{{\label{fig:#1}}}
14: \newcommand{\lEq}[1]{{\label{eq:#1}}}
15: \newcommand{\lTab}[1]{{\label{tab:#1}}}
16: % MACROS
17: \def\gtaprx {\lower .1ex\hbox{\rlap{\raise .6ex\hbox{\hskip .3ex
18: 	{\ifmmode{\scriptscriptstyle >}\else
19: 		{$\scriptscriptstyle >$}\fi}}}
20: 	\kern -.4ex{\ifmmode{\scriptscriptstyle \sim}\else
21: 		{$\scriptscriptstyle\sim$}\fi}}}
22: \def\ltaprx {\lower .1ex\hbox{\rlap{\raise .6ex\hbox{\hskip .3ex
23: 	{\ifmmode{\scriptscriptstyle <}\else
24: 		{$\scriptscriptstyle <$}\fi}}}
25: 	\kern -.4ex{\ifmmode{\scriptscriptstyle \sim}\else
26: 		{$\scriptscriptstyle\sim$}\fi}}}
27: %\newcommand{\FIGFF}[2]{{\ref{fig:#2}\PAN{#1}}}
28: \newcommand{\FIGFF}[2]{{\ref{fig:#2}{#1}}}
29: \newcommand{\Figff}[1]{{\FIGFF{}{#1}}}
30: \newcommand{\FIG}[2]{{Fig.~\FIGFF{#1}{#2}}}
31: \newcommand{\Fig}[1]{{\FIG{}{#1}}}
32: \newcommand{\FIGS}[2]{{Figs.~\FIGFF{#1}{#2}}}
33: \newcommand{\Figs}[1]{{\FIGS{}{#1}}}
34: \newcommand{\Figure}[1]{{Figure~\FIGFF{}{#1}}}
35: \newcommand{\Figures}[1]{{Figures~\FIGFF{}{#1}}}
36: \newcommand{\Sectff}[1]{{\ref{sec:#1}}}
37: \newcommand{\Sect}[1]{{\S~\Sectff{#1}}}
38: \newcommand{\Sects}[1]{{\S\S~\Sectff{#1}}}
39: \newcommand{\Section}[1]{{Section~\Sectff{#1}}}
40: \newcommand{\Sections}[1]{{Sections~\Sectff{#1}}}
41: \newcommand{\Eqref}[1]{{\ref{eq:#1}}}
42: \newcommand{\Eqff}[1]{{(\Eqref{#1})}}
43: \newcommand{\eqff}[1]{{\Eqref{#1}}}
44: \newcommand{\EQ}[1]{{Equation~\Eqff{#1}}}
45: \newcommand{\Equation}[1]{{Equation~\Eqff{#1}}}
46: \newcommand{\Eq}[1]{{eq.~\Eqff{#1}}}
47: \newcommand{\eq}[1]{{eq.~\eqff{#1}}}
48: \newcommand{\Eqs}[1]{{eqs.~\Eqff{#1}}}
49: \newcommand{\Leg}[1]{{\textit{#1}}}
50: 
51: 
52: \begin{document}
53: 
54: %\shorttitle{SN Ia Burning and Detonation}
55: %\shortauthors{Woosley}
56: 
57: \title{Type~Ia Supernova: Calculations of Turbulent Flames Using the
58:   Linear Eddy Model}
59: 
60: \author{S. E. Woosley\altaffilmark{1}, A. R. Kerstein\altaffilmark{2},
61:   V. Sankaran\altaffilmark{2}, and F. R\"opke\altaffilmark{3}}
62: 
63: \altaffiltext{1}{Department of Astronomy and Astrophysics, University
64:   of California, Santa Cruz, CA 95064; woosley@ucolick.org}
65: \altaffiltext{2}{Combustion Research Facility, Sandia National Laboratory,
66:   Livermore, CA; arkerst@sandia.gov} 
67: \altaffiltext{3}{Max Planck Institut f\"ur Astrophysik, Garching,
68:   Germany; fritz@mpa-Garching.mpg.de}
69: 
70: 
71: \begin{abstract} 
72: The nature of carbon burning flames in Type Ia supernovae is explored
73: as they interact with Kolmogorov turbulence. One-dimensional
74: calculations using the Linear Eddy Model of \citet{Ker91}
75: elucidate three regimes of turbulent burning. In the simplest case,
76: large scale turbulence folds and deforms thin laminar flamelets to
77: produce a flame brush with a total burning rate given approximately by
78: the speed of turbulent fluctuations on the integral scale, $U_L$. This
79: is the regime where the supernova explosion begins and where most of
80: its pre-detonation burning occurs. As the density declines,
81: turbulence starts to tear the individual flamelets, making broader
82: structures that move faster. For a brief time, these turbulent
83: flamelets are still narrow compared to their spacing and the concept
84: of a flame brush moving with an overall speed of $U_L$ remains
85: valid. However, the typical width of the individual flamelets, which
86: is given by the condition that their turnover time equals their
87: burning time, continues to increase as the density
88: declines. Eventually, mixed regions almost as large as the integral
89: scale itself are transiently formed. At that point, a transition to
90: detonation can occur. The conditions for such a transition are
91: explored numerically and it is estimated that the transition will
92: occur for densities near $1 \times 10^7$ g cm$^{-3}$, provided the
93: turbulent speed on the integral scale exceeds about 15\% sonic. An
94: example calculation shows the details of a detonation actually
95: developing.
96: \end{abstract}
97: 
98: 
99: \keywords{supernovae: general; hydrodynamics, shock waves, turbulence}
100: 
101: \section{INTRODUCTION}
102: \lSect{intro}
103: 
104: \citet{Dam40} first discussed multiple regimes of turbulent chemical
105: combustion and gave scaling relations for each. In modern terms, the
106: two regimes can be distinguished by their Karlovitz number, Ka,
107: \citep[e.g.,][]{Pet00},
108: \begin{equation}
109: \frac{U_L}{S_{\rm lam}} = {\rm Ka}^{2/3} \left( \frac{L}{\delta_{\rm
110:     lam}} \right)^{1/3},
111: \end{equation}
112: or equivalently,
113: \begin{equation}
114: {\rm Ka} = \left( \frac{\delta_{\rm lam}}{l_{\rm G}} \right)^{1/2}.
115: \lEq{Karlovitz}
116: \end{equation}
117: Here $U_L$ is the {\sl rms} velocity of turbulent fluctuations on an
118: integral scale, $L$; $S_{\rm lam}$ is the laminar conductive speed;
119: $\delta_{\rm lam}$ is the width of the laminar flame; and $l_G$ is the
120: Gibson length.  For isotropic Kolmogorov turbulence (assumed
121: throughout this paper), the turbulent speed on the scale of the flame
122: thickness is
123: \begin{equation}
124: v_{\rm turb}(\delta_{\rm lam}) = \left(\frac{\delta_{\rm lam}}{L}\right)^{1/3}
125: U_L,
126: \end{equation}
127: and the Gibson scale, the size of the eddy that turns over in a
128: (laminar) flame crossing time, is
129: \begin{equation}
130: l_G = \left( \frac{S_{\rm lam}}{U_L} \right)^3 L.
131: \lEq{Gibson}
132: \end{equation}
133: 
134: For Ka $\ltaprx 1$, individual laminar flames are moved around by the
135: largest turbulent eddies while smaller eddies have little effect.  The
136: overall burning progresses at a speed determined by turbulence
137: properties and is independent of the burning rate on small scales.
138: This regime has been extensively explored in the astrophysical context
139: \citep{Nie95a,Nie95b,Nie97} and its properties are reflected in the
140: Munich group's subgrid model for flame propagation
141: \citep{Sch06a,Sch06b}.
142: 
143: The condition Ka $\gg 1$, on the other hand, implies that turbulence
144: can penetrate into the flame and transport heat, and possibly fuel,
145: faster than laminar burning crosses a flame width. An equivalent
146: condition is that the Gibson scale is much less than the flame
147: thickness.  This regime too has been discussed in the astrophysical
148: literature \citep{Nie97,Nie97a,Kho97,Lis00b}. It is generally agreed
149: that if spontaneous detonation is to occur, it requires Ka $> 1$ and
150: probably Ka $> 10$ so that the burning region itself is disrupted, not
151: just the preheat zone.
152: 
153: It is also known in the chemical combustion community
154: \citep{Pet86,Pet00,Ker01} that the region Ka $\gg 1$ can be further
155: divided based on the value of the Damk\"ohler number, Da = $L/(U_L
156: \tau_{\rm nuc})$. Here $\tau_{nuc}$ is the characteristic burning time
157: scale, appropriately modified by turbulence. For Da $<$ 1, the eddy
158: turnover time on the integral scale is short compared with the nuclear
159: time; for Da $>$ 1, it is longer.  In the literature, the term
160: ``distributed reaction zone(s)'' has been used with reference to the
161: Da $<$ 1 regime, the Da $>$ 1 regime, or both lumped together (i.e.,
162: all flames with Ka $\gg$ 1).  Therefore we avoid this terminology,
163: choosing instead to follow \citet{Pet86} in referring to Da $<$ 1 as
164: the ``well-stirred reactor regime'' (WSR regime), and to follow
165: \citet{Ker01} in referring to Da $>$ 1 as the ``stirred flame regime''
166: (SF regime).
167: 
168: In the WSR regime, there is only one flame. It has a width broader
169: than the integral scale and a speed slower than the turbulent speed on
170: the integral scale. This sort of flame is similar to the usual laminar
171: flame, except that the turbulent diffusion coefficient ($D_{\rm turb}
172: \sim U_L L$) now substitutes for conduction. Within the SF regime, on
173: the other hand, there can be multiple burning regions, but the idea of
174: a flame brush composed of individual flamelets with well-defined local
175: properties is no longer valid.  The overall burning continues with an
176: average rate given by the turbulent speed on the integral scale, but
177: the flamelets do not have a uniform width and their number and
178: individual speeds are quite variable.
179: 
180: In this paper, we explore these three regimes of turbulent nuclear
181: combustion, Ka $< 1$, WSR, and SF, in the context of a Type Ia
182: supernova using a numerical tool, the Linear Eddy Model
183: (\Sect{LEM}). If a transition to detonation is to occur, we conclude
184: that it must happen in the SF regime, specifically where Da $\sim 10$
185: and in the presence of a high degree of turbulence, $U_L \gtaprx 0.15
186: \, c_{\rm sound}$. An example of a successful spontaneous transition to
187: detonation is given.
188: 
189: 
190: \section{Special Conditions in a Type Ia Supernova}
191: \lSect{special}
192: 
193: The conditions characterizing turbulent (nuclear) combustion in a Type
194: Ia supernova are novel and have no direct analogue on earth. This
195: makes the supernova an interesting environment for testing new
196: physics, but it also means that our terrestrial intuition regarding
197: flames can be misleading. For example, it is thought that laboratory
198: flames in the Ka $\gg 1$ regime simply go out \citep{Pet00} because
199: they are unable to maintain their heat in the presence of so much
200: turbulence. But the flame in a supernovae can never ``go out'' until
201: the star comes apart and, in terms, of local flame variables, that
202: takes a very long time. The relevant time scale is the hydrodynamic
203: time scale for the whole star, $\sim$1 s, not the shorter local
204: turbulent time scale, $U_L/L$. Also the star is very large, $\gtaprx
205: 10^8$ laminar flame thicknesses and 100 integral scales.  Rare events
206: have many opportunities for realization. There are also other
207: conditions that are vastly different and, in some cases, make
208: detonation more easily achievable.
209: 
210: 
211: \subsection{High Reynolds Number}
212: 
213: The Reynolds number in a Type Ia supernova is orders of magnitude
214: greater than achieved thus far in any terrestrial experiment or
215: in any numerical simulation. The source of viscosity is electron-ion
216: interactions in a fully ionized plasma \citep{Nan84}
217: \begin{equation}
218: \begin{split}
219: \nu &= \frac{1.9 \times 10^6}{Z} \left(\frac{(\rho_6/\mu_e)^{5/3}}{1 \ +
220:   \ 1.02 \, (\rho_6/\mu_e)^{2/3}}\right) \ \frac{1}{I_2} \\ &\sim 2 \times
221: 10^6 \ {\rm gm \ cm^{-1} \ s^{-1}}
222: \end{split}
223: \end{equation}
224: for Z = 7, $\mu_e$ = 2, $\rho_6 = 10$, and $I_2$ = 0.5. The Reynolds
225: number is then
226: \begin{equation}
227: Re = \frac{\rho U_L L}{\nu} \sim 5 \times 10^{13},
228: \end{equation}
229: for $\rho =$ 10$^7$ g cm$^{-3}$, $U_L$ = 100 km s$^{-1}$, and $L$ = 10
230: km.  This large value of Re implies a tiny Kolmogorov scale, much
231: smaller than any laminar flame thickness under consideration here, and
232: irresolvably small in most numerical simulations,
233: \begin{equation}
234: \eta \ = \ Re^{-3/4} L \ \sim \  10^{-4} \ {\rm cm}.
235: \end{equation}
236: Note also that this implies ten orders of magnitude in length scale
237: where the turbulence is (assumed to be) Kolmogorov and isotropic with
238: constant energy dissipation, $U_L^3/L$.
239: 
240: On earth, the Kolmogorov scale is usually not so small compared with
241: the flame thickness. This makes achieving the SF and WSR regimes more
242: difficult on earth, though certainly not impossible.
243: 
244: \subsection{Temperature-Dependent Heat Capacity}
245: \lSect{cp}
246: 
247: The heat capacity in the supernova at a relevant density, $\rho \sim
248: 10^7$ g cm$^{-3}$, is due to a combination of semi-degenerate
249: electrons, ions and radiation. Radiation is an important component of
250: the heat capacity at these low densities and hence the heat capacity
251: is a rapidly increasing function of the temperature. At a density of
252: $10^7$ g cm$^{-3}$, for 50\% C and 50\% O, the heat capacity at
253: constant pressure is 3.5, 6.0, 9.0, 12.7, 18.2,, and 27.3 $\times
254: 10^7$ erg gm$^{-1}$ K$^{-1}$ for a temperature of 0.5, 1.0, 1.5, 2.0,
255: 2.5, and 3.0 $\times 10^9$ K respectively. The power of T upon which
256: $C_p$ depends varies from 0.70 to 2.70 in the same temperature range.
257: 
258: This means that when the fuel is cold, a small amount of burning
259: raises the temperature a lot. Given the high power of temperature upon
260: which the burning depends, burning just a little fuel dramatically
261: shortens the nuclear time scale.
262: 
263: \subsection{Temperature-Dependent Reaction Rate}
264: \lSect{enuct}
265: 
266: The most important reaction rate in the regime where detonation might
267: occur is $^{12}$C + $^{12}$C. The rate for this reaction is
268: proportional to $\rho X^2(^{12}C) T^n$ with $n \sim 19 - 27$ for
269: temperatures in the range 1 - 3 $\times 10^9$ K (more sensitive at
270: lower temperature).  This high temperature sensitivity coupled with
271: the temperature-dependent heat capacity means that once about half of
272: the carbon has burned, the remaining increase in the temperature
273: happens very rapidly. As we shall see later this leads to small but
274: very rapid increases in the local pressure that can help initiate a
275: detonation.
276: 
277: \subsection{Strong Turbulence}
278: 
279: As burning plumes of ash float due to the Rayleigh-Taylor instability,
280: they create shear and turbulence on their boundaries. Since the plumes
281: are large, the speed at which buoyancy balances drag is high. A speed of
282: order 10 - 30\% sonic is necessary to burn a large fraction of the
283: star before it comes apart.  Typical turbulent speeds on an integral
284: scale of 10 km are about 150 km s$^{-1}$ \citep{Roe07a}, but speeds as
285: great as 1000 km s$^{-1}$ may occasionally occur. The sound speed in
286: the star at a density of 10$^7$ g cm$^{-3}$ is 3500 km s$^{-1}$, so
287: the fastest turbulence is not terribly subsonic. Fluctuations in
288: burning rate do not have to accelerate the burning by orders of
289: magnitude in order to make a detonation happen.
290: 
291: \subsection{Large Lewis Number}
292: \lSect{turb}
293: 
294: The Lewis number, which is the ratio of thermal diffusivity to mass
295: diffusivity, is very large in the supernova. The ionic diffusion
296: coefficient for a carbon-oxygen plasma is \citep{Han75,Bil01}
297: \begin{equation}
298: D_{\rm ion} \ = \ 3 \omega_p \alpha^2 \Gamma^{-4/3} \sim 0.1 \ {\rm cm^2 \ s^{-1}}, 
299: \end{equation}
300: where $\omega_p = (4 \pi n_i (Ze)^2/A m_p)^{1/2}$ is the plasma frequency,
301: $\alpha = (3/(4 \pi n_i))^{1/3}$, with $n_i$, the ion density, and
302: $\Gamma = (Ze)^2/\alpha k T$. The thermal diffusion coefficient is
303: \begin{equation}
304: D_{\rm rad} \ = \ \frac{4 a c T^3}{3 \rho^2 C_P \kappa} \ \sim \ 10^4.
305: \lEq{drad}
306: \end{equation}
307: Here representative conditions and opacities have been assumed: $T
308: \approx 1 - 2 \times 10^9$ K, $\rho = 10^7$ g cm$^{-3}$, and $\kappa
309: \approx 0.02$ cm$^2$ gm$^{-1}$ \citep{Tim00a}. Combining these two
310: equations, $Le = D_{\rm rad}/D_{\rm ion}\sim 10^5$.
311: 
312: Terrestrial Lewis numbers are close to unity. The large Lewis number
313: in the star has an effect on the laminar speed (\Sect{laminar}), but
314: this dependence is greatly mitigated in the turbulent regime
315: (\Sect{distrib}) where the turbulent diffusion coefficient exceeds
316: $D_{\rm rad}$. In that case, the effective Lewis number approaches
317: unity since both ions and heat are transported with equal efficiency by
318: the turbulent eddies \citep{Asp08}.
319: 
320: \section{The Linear Eddy Model}
321: 
322: \subsection{The LEM Code}
323: \lSect{LEM}
324: 
325: The range of scales that must be resolved to address the flame
326: propagation problem in a Type Ia supernova is very large, $\sim
327: 10^{-1} - 10^6$ cm even if the Kolmogorov scale is not resolved. This
328: range exceeds the current or anticipated capabilities of 3D
329: simulations, or even 2D simulations.  This conundrum arises for many
330: turbulent flow environments, and motivated the development of the
331: Linear Eddy Model (LEM), a 1D simulation tool \citep{Ker91}.
332: 
333: LEM simulates the evolution of scalar properties on a 1D spatial
334: domain.  This can be interpreted as property profile evolution along a
335: 1D line of sight through 3D turbulent flow.  The 1D domain is treated
336: as a closed system with respect to enforcement of conservation laws.
337: In this and other respects, LEM is not fully consistent with the
338: evolution observed along a line of sight, yet it captures the salient
339: features and provides useful results.
340: 
341: The physical processes that are time advanced on the LEM domain are
342: diffusive (e.g., Fickian) transport (in the present context
343: representing species transport, radiation transport or subgrid
344: turbulent transport), chemical (or nuclear) reactions, and turbulent
345: eddy motions.  For combustion applications, including the present 
346: application, LEM includes an equation of state, an energy equation, 
347: and thermal expansion, using a zero-Mach-number (constant-pressure) 
348: formulation \citep{Smi97}.  The novelty of the LEM approach is the 
349: representation of eddy motions.
350: 
351: On a 1D domain, advection in the usual sense cannot reorder the fluid
352: elements along the domain, hence cannot emulate the folding of
353: material surfaces that is an essential feature of turbulent stirring.
354: A model of eddy motions is introduced that is formulated to capture
355: this folding effect and other properties of turbulent stirring.
356: Namely, a turbulent eddy is represented as an instantaneous map,
357: called the triplet map, that is applied to a designated interval of
358: the 1D domain.
359: 
360: The triplet map is applied by first compressing all property profiles
361: in the interval by a factor of three.  The property profiles in the
362: interval are then replaced by three side-by-side copies (`images') of
363: the compressed profiles.  The middle copy is then flipped.
364: 
365: This procedure preserves the continuity of spatial property profiles.
366: It also leaves the total linear measure (model analog of volume) of
367: fluid corresponding to a given state or set of states (based on the
368: property values) unchanged.  This is the 1D analog of the solenoidal
369: condition.  Flipping of the middle copy introduces fluid-element
370: reordering analogous to the folding of material surfaces.  The
371: three-fold compression emulates gradient amplification by compressive
372: strain.  The dispersive effect of extensive strain is represented in
373: LEM by the mapping of neighboring fluid elements to different images.
374: 
375: Computationally, the triplet map is implemented as a permutation of
376: the cells of a uniform discretization of the 1D domain.  This and
377: other details of LEM are explained elsewhere (Kerstein 1991).  Here,
378: features relevant to what follows are described.  Some symbols that
379: are used have different meanings than in later sections.
380: 
381: An LEM simulation time advances processes other than advection until
382: it is time to implement a triplet map.  After the map, advancement of
383: the other processes resumes until it is time for the next map.
384: 
385: Model inputs are the initial and boundary conditions, the transport
386: coefficients and rate constants governing diffusive and chemical
387: advancement, and parameters controlling the time sequence of triplet
388: maps.  Map size $l$ is sampled from a probability density function
389: (pdf) $f(l)$ that is designed to reproduce relevant features of the
390: inertial-range turbulent cascade.  One key feature is the turbulent
391: diffusivity $D_{\rm turb}$ associated with maps of size $l<S$.  The 
392: inertial-range scaling $D_{\rm turb} \propto S^{4/3}$ is enforced.  
393: In \citet{Ker91}, it is shown that this implies $f(l)=Al^{-8/3}$, 
394: where $A$ is a normalization factor.
395: 
396: Map size $l$ is restricted to the range $[\eta, L]$, where $\eta$ and
397: $L$ are the model analogs of the Kolmogorov scale and the integral
398: scale, respectively.  Then as shown in the Appendix, the total
399: turbulent diffusivity is given by $D_{\rm turb}= \frac{1}{18} \Lambda
400: A(L^{4/3}-\eta^{4/3})$, where $\Lambda$ is the total frequency of maps
401: of all sizes per unit domain length.  Here, homogeneous turbulence is
402: assumed, so none of the model parameters depend on location, and map
403: location is sampled uniformly within the notionally infinite 1D domain.
404: 
405: The parameters $\eta$, $L$, and $D_{\rm turb}$ are model inputs and 
406: $\Lambda$ is determined from them.  $D_{\rm turb}$ is not
407: typically known for turbulent flows, but rather is inferred from the
408: relation $D_{\rm turb} = U_LL/C$, where $C$ is an empirical coefficient.  Here,
409: LEM results are compared to 3D simulations of turbulent premixed
410: combustion for which $U_L$ and $L$ are known, but $D_{\rm turb}$ is not, so $C$
411: must be specified in order to evaluate the LEM input parameter $D_{\rm turb}$.
412: \citet{Smi97} calibrated $C$ by comparing LEM results for turbulent
413: premixed methane-air combustion (using a simplified chemical
414: mechanism) to turbulent burning velocity measurements.  They chose the
415: value $C=15$ for some cases and $C=3.5$ for others, reflecting the
416: wide variation of turbulent burning velocity results obtained in
417: different experiments.
418: 
419: The experimental set-ups did not correspond to the idealized case of
420: flames freely propagating through stationary, homogeneous turbulence.
421: Here, LEM results are compared to 3D simulation results more closely
422: analogous to the LEM flow configuration.  The $C$ value inferred from
423: this comparison is close to the lower of the two $C$ values reported
424: by Smith and Menon.
425: 
426: % fig 1 - laminar flame conditions
427: \begin{figure*}
428: \begin{center}
429: \includegraphics[width=0.475\textwidth]{fig1a.eps}
430: \hfill
431: \includegraphics[width=0.475\textwidth]{fig1b.eps}
432: \caption{Laminar flame structure for initial carbon mass fractions of
433:   0.50 (left frame) and 0.75 (right frame) and a density of 1.0
434:   $\times 10^7$ g cm$^{-3}$. Carbon mass fraction is given in red and
435:   the temperature, in units of 10$^9$ K, is in black. The energy
436:   generation (green) has been normalized to a maximum value of 7.47
437:   $\times 10^{20}$ erg g$^{-1}$ s$^{-1}$ ($X^0_{12} = 0.50$) and $8.40
438:   \times 10^{21}$ erg g$^{-1}$ s$^{-1}$ ($X^0_{12} = 0.75$).  The
439:   flame is propagating to the left with a speed of $3.2 \times 10^3$
440:   cm s$^{-1}$ ($X^0_{12} = 0.50$) and $1.1 \times 10^4$ cm s$^{-1}$
441:   ($X^0_{12} = 0.75$). The assumed Lewis Number in both cases was
442:   1000. These calculations with the LEM code used 2048 zones and show
443:   its ability to resolve and physically calculate laminar flames where
444:   radiation transport is dominant. \lFig{flame}}
445: \end{center}
446: \end{figure*}
447: 
448: \subsection{Modifications for the SN Ia Problem}
449: \lSect{modifications}
450: 
451: The standard (i.e., terrestrial chemical combustion) version of LEM
452: was modified to use an equation of state, opacities, and a nuclear
453: reaction network appropriate to the supernova. The equation of state
454: \citep[the ``Helmholtz EOS'';][]{Tim00b} included pressure and energy
455: contributions from radiation, ions (treated as an ideal gas), and
456: electrons and pairs of arbitrary speed and degeneracy.  The opacity
457: routine \citep{Tim00a} included heat transport by radiative diffusion
458: and conduction. Energy generation and composition changes were
459: followed using a nuclear reaction network with 7 species: $^{4}$He,
460: $^{12}$C, $^{16}$O, $^{20}$Ne, $^{24}$Mg, $^{28}$Si, and $^{56}$Ni
461: (though no appreciable $^{56}$Ni was ever produced in this study).
462: These species were connected together by a chain of $(\alpha,\gamma)$
463: reactions as well as the heavy ion reactions 3$\alpha$, $^{12}$C +
464: $^{12}$C, $^{16}$O + $^{16}$O, and $^{12}$C + $^{16}$O.  Though it is
465: an unpublished version written by the authors, the network was similar to
466: that of \citet{Tim00c}. The network did not include electron screening
467: corrections to the reaction rates which were, in general,
468: small at the low densities considered here. Testing the fast 7 isotope
469: network against a larger 19 isotope network that included screening
470: and about 85 reactions \citep{Wea78} showed that the small network
471: gave an energy generation that was typically 20 - 40\% smaller. Since
472: this is equivalent to a small error in the temperature, the results of
473: the 7 isotope, unscreened network were deemed sufficiently accurate.
474: 
475: The code was then checked and calibrated against several other studies
476: of flame propagation. For simple laminar flames (\Sect{laminar}), the
477: turbulence parameter, $C$ (\Sect{LEM}), doesn't enter.  For turbulent
478: flames in the flamelet regime, the default setting $C = 15$ was used
479: (\Sect{flamelet}). However, for flames in the WSR regime
480: (\Sect{distrib}), good agreement with prior 3D studies required a
481: smaller value of $C = 5$. This value is within the range anticipated
482: from terrestrial experiments (\Sect{LEM}) and was used for all
483: calculations in the WSR and SF regimes (i.e., all studies with Ka $>
484: 1$) unless otherwise noted. Future 3D studies, especially of the SF
485: regime, are encouraged in order to gain better confidence and
486: understanding of $C$ for flames of this sort.
487: 
488: 
489: \section{Single Laminar Flames}
490: \lSect{laminar}
491: 
492: In the absence of turbulence, a flame has a simple structure in
493: which a self-similar profile of temperature and fuel concentration
494: propagates into the fuel with a well-defined width and speed. Heat is
495: transported by a combination of conduction and radiative diffusion,
496: and this heat raises the temperature to the point where fuel burns
497: on a diffusion time scale.  Such flames are well understood
498: \citep{Lan59}, even in the supernova context \citep{Tim92}. In
499: multiple dimensions, laminar flames may become deformed and take on a
500: cusp-like appearance due to the Landau-Darrieus instability
501: \citep{Nie95b}, but this instability does not lead to a major
502: destabilization of the overall burning \citep{Roe04}, nor does it
503: greatly affect the propagation speed \citep{Bel04a}.
504: 
505: As a test of the LEM implementation and to provide a calibration point
506: for more complex studies to follow, the one-dimensional laminar speed
507: was calculated for carbon-burning flames at a variety of densities for
508: two initial carbon mass fractions, $X^0_{12}$ = 0.50 and 0.75. The
509: results are given in Table 1 and \Fig{flame}. The speed of the flame
510: with carbon mass fraction 0.50 and density 10$^7$ g cm$^{-3}$ was $3.2
511: \times 10^3$ cm s$^{-1}$, in good agreement with the $3.5 \times 10^3$
512: cm s$^{-1}$ found by \citet{Asp08}, who used somewhat different
513: nuclear physics.  The ash temperatures for the two compositions
514: considered here were also in good agreement with the values calculated
515: by \citet{Woo07}. The flame speeds for other densities were in
516: agreement with previous studies by \citet{Bel04b}, and \citet{Tim92}.
517: 
518: In these calculations, heat was transported by conduction and
519: radiation and the laminar scale was well resolved. Zoning for the case
520: of 10$^7$ g cm$^{-3}$, $X^0_{12}$ = 0.50 was 0.0244 cm and the flame
521: width, several cm.  A recalculation of the flame speed with coarser
522: grids gave essentially the same answer. For resolutions of 0.196 cm
523: and 0.0978 cm the calculated flame speed was 3.2 and 3.1 $\times 10^3$
524: cm s$^{-1}$, respectively, indicating that our calculation and that of
525: \citet{Asp08} were both converged and did not differ because of
526: resolution.
527: 
528: Some dependence on Lewis Number was noted. Two calculations of the
529: speed for $X^0_{12}$ = 0.50 and Le = 100 and Le = 1000 gave speeds of
530: $3.0 \times 10^4$ cm s$^{-1}$ and $3.2 \times 10^4$ cm s$^{-1}$
531: respectively.
532: 
533: 
534: % fig 2 - multiple laminar flames
535: \begin{figure*}
536: \begin{center}
537: \includegraphics[width=0.475\textwidth]{fig2a.eps}
538: \hfill
539: \includegraphics[width=0.475\textwidth]{fig2b.eps}
540: \caption{Multiple laminar flames for a carbon mass fraction of 0.75,
541:   density, 1.0 $\times 10^7$ g cm$^{-3}$, integral length scale, 10 m,
542:   and turbulent speed on that scale, 400 m $^{-1}$. The left frame
543:   shows the entire flame brush, while the right frame shows greater
544:   detail for a few of the flamelets.  The entire collection of
545:   flamelets is moving to the left at an average rate of 500 m s$^{-1}$
546:   (see \Fig{flameletspd}).The time here is 0.015 s into the
547:   calculation and the instantaneous rate of burning corresponds to a
548:   speed of 980 m s$^{-1}$.  This calculation with the LEM code used
549:   40000 zones and Le = 100.  \lFig{flamelet}}
550: \end{center}
551: \end{figure*}
552: 
553: \section{Three Regimes of Turbulent Burning}
554: \lSect{regime}
555: 
556: \subsection {Large Scale Turbulence and Flame Brushes (Flamelet Regime)}
557: \lSect{flamelet}
558: 
559: The interior of a Type Ia supernova is turbulent, first because of the
560: convection that precedes the explosion, which gives a lower bound for
561: $U_L \sim$ 50 - 100 km s$^{-1}$ \citep{Woo04,Kuh06}, and second
562: because of the Rayleigh-Taylor and Kelvin-Helmholtz instabilities
563: associated with the flame itself \citep{Hil00}.  Because these speeds
564: are so much greater than the laminar speed (Table 1), especially at
565: low density, it is expected that turbulence has a major effect on
566: flame propagation. Initially, however, the flame is very thin compared
567: to the Gibson length and is just carried around by the eddies
568: \citep{Dam40}. Overall the burning rate is independent of the speed of
569: each little flamelet, and is governed instead by the speed of the
570: largest turbulent eddies. This is Damk\"ohler's ``large scale
571: turbulence'' limit.
572: 
573: To illustrate burning in this regime, we considered a laminar flame
574: similar to that in \Fig{flame} with a carbon mass fraction of 0.75 and
575: a fuel density $1.0 \times 10^7$ g cm$^{-3}$, but embedded in a
576: turbulent background with characteristic speed 400 m s$^{-1}$ on an
577: integral scale of 10 m. On this length scale in a supernova, one would
578: actually expect much larger turbulent speeds, 10 to 100 km s$^{-1}$,
579: cascading down from still higher values on larger scales. For the time
580: being such large, highly turbulent regimes with narrow flames remain
581: out of computational reach, even in 1D.  Conditions chosen here were
582: thus artificial, but illustrative.  The Gibson scale here
583: (\eq{Gibson}) is 21 cm, considerably larger than the laminar flame
584: width, which is $\sim$1 cm (\Fig{flame}), but still about still 50
585: times less than the integral scale.  The results are given in
586: \Fig{flamelet}.  Sometimes only a single flame was observed, but at
587: the particular moment sampled here, there were eleven (counting the
588: small blip of entrained ash at 800 cm).
589: 
590: \Fig{flameletspd} shows the overall propagation speed of the flame
591: brush. Here, as in the rest of the paper, the burning ``speed'' is
592: defined by an integral over the grid of the burning rate, 
593: \begin{equation}
594: v_f \ = \ \frac{\int S_{\rm nuc} dx}{q},
595: \end{equation}
596: where $S_{\rm nuc}$ is the nuclear energy generation rate on the grid
597: (erg g$^{-1}$ s$^{-1}$), and $q$ is the energy released by burning a
598: gram of fuel. For fuel with 50\% carbon, $q$ is $3.2 \times 10^{17}$
599: erg g$^{-1}$ and for 75\% carbon it is $4.5 \times 10^{17}$ erg
600: g$^{-1}$.
601: 
602: For the case shown in \Fig{flameletspd}, the average burning speed is
603: 500 m s$^{-1}$, close to the turbulent {\sl rms} speed on the integral
604: scale, 400 m s$^{-1}$. In LEM, one expects for the flamelet regime that
605: $v_{\rm turb}$ = 18 $D_{\rm turb}/L$ (see Appendix). Combining this
606: with $D_{\rm turb} = U_L L/C$ (\Sect{LEM}) and using the same value
607: for $C$ as the simulation ($C = 15$), one has $v_{\rm turb} = (18/15) \,
608: U_L$, which is excellent agreement.
609: 
610: 
611: \Fig{flameletspd} also shows that the burning rate is
612: far from regular. Speeds over three times the average are sometimes
613: seen. At other times, the flame briefly almost goes out.
614: A similar behavior would be expected in the actual supernova early on
615: when the density is higher and the turbulent speed is not that much
616: greater than the laminar one. Then, as here, there would be just a few
617: flames within the integral scale and the burning rate would be highly
618: irregular. For example, at a density of $10^9$ g cm$^{-3}$ with a
619: carbon mass fraction of 0.50, the laminar flame speed is 36 km
620: s$^{-1}$ \citep{Tim92} and the ratio $L/l_{\rm Gib} \sim (U_L/S_{\rm
621:   lam})^3$ is, for $U_L = 130$ km s$^{-1}$, also about 50.  The
622: difference at these high densities is that the flame would be
623: irresolvably small, even in the present version of LEM, for a
624: calculation that carried the entire integral scale.
625: 
626: \subsection{Transition to Stirred Flames}
627: 
628: The flames at such high density are individually very thin and the
629: laminar speed far below sonic, so detonation at early times is
630: avoided. As the density declines, the Gibson length shrinks and the
631: flame brush contains an increasingly large number of flamelets. Just
632: before entering the SF regime in the supernova, there
633: are roughly a thousand flamelets within the flame brush ($U_L \sim
634: 100$ km s$^{-1}$; $S_{\rm lam} \sim 100$ m s$^{-1}$).  The statistical
635: variations in overall speed are therefore small.
636: 
637: 
638: % fig 3 - turbulent laminar flame speed 
639: \begin{figure}
640: \includegraphics[angle=90,width=0.475\textwidth]{fig3.ps}
641: \caption{Speed of the flame brush as a function of time for the same
642:   calculation shown in \Fig{flamelet}. The speed, as measured by the
643:   overall energy generation on the grid, is highly variable ranging
644:   from a lower bound given by the laminar speed of a single flamelet
645:   ($1.1 \times 10^4$ cm s$^{-1}$) to several times the mean value. The
646:   dashed line indicates the average during the time interval studied,
647:   $5.5 \times 10^4$ cm s$^{-1}$, which is close to the {\sl rms}
648:   turbulent speed on the integral scale, $4 \times 10^4$ cm s$^{-1}$.
649:   The turnover time on the integral scale is 0.025 s and one sees the
650:   effect of the occasional large eddy in accelerating the burning.
651:   \lFig{flameletspd}}
652: \end{figure}
653: 
654: 
655: As the density in the star declines below several times 10$^7$ g
656: cm$^{-3}$, the Gibson length becomes less than the laminar flame width
657: and the transport of fuel and heat by turbulent advection starts to
658: become important on the scale of the flame. By the time the density
659: reaches $1.0 \times 10^7$ g cm$^{-3}$, for typical turbulent speeds,
660: not only is the preheat region of the flame being mixed by turbulence,
661: but the burning zone itself has been disrupted and combustion has a
662: qualitatively different character \citep{Asp08,Pet00}. In the star,
663: this transition is brought about by the slowing and thickening of the
664: laminar flame with decreasing density in a background of turbulence
665: with nearly constant energy dissipation rate and integral length
666: scale.
667: 
668: 
669: \subsection{The Well-Stirred Reactor (WSR Regime)}
670: \lSect{distrib}
671: 
672: Before discussing the stirred flame regime further though
673: (\Sect{stirred}), it is helpful to consider a limiting case where the
674: turbulence thickens a single flame to a dimension greater than the
675: integral scale, i.e., $\delta_{\rm turb} > l$, where $l$ is some
676: integral scale to be varied subject to the condition $v'^3(l)/l =
677: \epsilon$ = constant, where $v'$ is the velocity on the integral
678: scale, $l$.  While probably not realized in the supernova because of
679: the large integral scale, this is the regime most easily accessible to
680: multi-dimensional simulations and corresponds to what is commonly
681: meant by the ``well-stirred reactor'' \citep[e.g., Fig. 7
682:   of][]{Pet86}. The solutions here also obey scaling relations
683: predicted by \citet{Dam40}.
684: 
685: 
686: % fig 4 - comparison of flame properties andy lem t8 fuel = 1 vs lem
687: % t8 = 6
688: \begin{figure}
689: \includegraphics[width=0.475\textwidth]{fig4.eps}
690: \caption{Comparison of carbon mass fraction and temperature for two
691:   flames calculated at a density of 10$^{7}$ g cm$^{-3}$, integral
692:   scale, $l = 15$ cm, and turbulent speed, $2.45 \times 10^5$ cm
693:   s$^{-1}$. The purple and blue curves are the average temperature and
694:   carbon mass fraction from the 3D study of \citet{Asp08}. The black
695:   and red curves are from the corresponding study using LEM and $C =
696:   5$ (Table 2). The speed of the flame calculated in 3D was $1.8
697:   \times 10^4$ cm s$^{-1}$. With LEM it was $1.5 \times 10^4$ cm
698:   s$^{-1}$, about five times the laminar flame speed for these
699:   conditions. The flame widths calculated in the two studies are quite
700:   similar. Differences in the temperature are attributable, in part,
701:   to the different fuel temperature employed in the two studies - $1
702:   \times 10^8$ K for the 3D study and $6 \times 10^8$ K for the LEM
703:   study and the smaller network used for the 3D study (see text).
704:   \lFig{andywidth}}
705: \end{figure}
706: 
707: \subsubsection{An example compared with previous 3D simulations}
708: \lSect{Aspden}
709: 
710: \citet{Asp08} recently carried out 3D simulations of flame propagation
711: in the WSR regime for a density and composition appropriate to Type Ia
712: supernovae. We focus here on their calculation at 10$^{7}$ g
713: cm$^{-3}$, $X_{12}^0$ = 0.50, and $\epsilon = U_L^3/L = 10^{15}$ erg g
714: s$^{-1}$.  \Figures{andywidth}, \ref{fig:andy}, \ref{fig:andyspd}, and
715: Table 2 show the results for an LEM calculation at the same values of
716: integral scale, $l$ = 15 cm, and $v'(l)$ = 2.47 km s$^{-1}$ that they
717: employed. A single broad flame propagates at a steady speed
718: about 5 times faster than the laminar value. The width
719: of the flame, as measured by the full width at half maximum of the
720: energy generation curve is about 50 cm, or 20 times the laminar width
721: (\Fig{flame}). 
722: 
723: The most noticeable differences with the 3D calculation
724: (\Fig{andywidth}) are a consequence of differing assumptions regarding
725: the background (fuel) temperature in the supernova. Aspden et al. used
726: $1 \times 10^8$ K; here we use $6 \times 10^8$ K. The lower
727: temperature would characterize a region of low turbulence following
728: some expansion prior to a detonation transition. The latter is
729: characteristic of the convection zone in the supernova at the time of
730: runaway. For a non-turbulent medium, the former is more
731: physical. However, anticipating that we will be interested here in
732: turbulent energies $U_L^3/L \sim 10^{17} - 10^{18}$ erg g$^{-1}$
733: s$^{-1}$ and that at least a few turnovers on the integral scale are
734: expected prior to a detonation, the latter is also a reasonable
735: assumption. For $U_L \sim 10^8$ cm s$^{-1}$ and $L = 10$ km, 10$^{16}$
736: erg g$^{-1}$ would be deposited in one turnover time. At a density of
737: 10$^7$ g cm$^{-3}$, this corresponds to a temperature increase from $1
738: \times 10^8$ K (or essentially zero) to $6 \times 10^8$ K. Because of
739: the temperature sensitivity of the heat capacity (\Sect{cp}), we
740: expect our results to be insensitive to the assumed fuel
741: temperature. However we do note that, if a fuel temperature of $1
742: \times 10^8$ is assumed in LEM, better agreement with the width is
743: obtained than in \Fig{andywidth}, but the best value of $C$ is lowered
744: to $C = 2$.
745: 
746: While lower values of $C$ are actually more favorable to detonation, we
747: prefer to use the larger value because it is more consistent with the
748: range seen in terrestrial experiments (\Sect{LEM}) and because a
749: hotter fuel temperature is both physically justifiable and more stable
750: numerically.
751: 
752: 
753: % fig 5 - distributed flame - andy's conditions
754: \begin{figure}
755: \begin{center}
756: \includegraphics[width=0.475\textwidth]{fig5a.eps}
757: \hfill
758: \includegraphics[width=0.475\textwidth]{fig5b.eps}
759: \caption{Two recomputations of the turbulently broadened flame in
760:   Fig. 4.  Both calculations used the LEM code with a zoning of 2048
761:   zones ($\Delta x$ = 0.147 cm). The maximum energy generation in both
762:   plots is normalized to $2.3 \times 10^{20}$ erg g$^{-1}$
763:   s$^{-1}$. The calculation in the first frame used a diffusive energy
764:   transport coefficient based on radiation and conduction. The one in
765:   the second frame, virtually identical in result, used the subgrid
766:   model discussed in the text.
767:   \lFig{andy}}
768: \end{center}
769: \end{figure}
770: 
771: 
772: % fig 6 - flame speed -  andy's conditions
773: \begin{figure}
774: \includegraphics[angle=90,width=0.475\textwidth]{fig6.ps}
775: \caption{Speed of the turbulently broadened flame shown in \Fig{andy}
776:   as a function of time. The turnover time on the integral scale is
777:   $6.1 \times 10^{-5}$ s. It takes many turnover times to reach the
778:   terminal speed, but then that speed is maintained rather precisely.
779:   The dashed line shows the average speed, $1.5 \times 10^4$ cm
780:   s$^{-1}$, or 5 times the laminar value. \lFig{andyspd}}
781: \end{figure}
782: 
783: 
784: % fig 7 - carbon as a function of temperature
785: \begin{figure*}
786: \begin{center}
787: \includegraphics[width=0.475\textwidth]{fig7a.eps}
788: \hfill
789: \includegraphics[width=0.475\textwidth]{fig7b.eps}
790: \caption{Carbon mass fraction as a function of temperature for all
791:   grid points for the calculation of the laminar flame (left; see also
792:   \Fig{flame}) and the turbulently broadened flame (right; see also
793:   \Fig{andy}). The initial density and carbon mass fraction in both
794:   cases were $X^0_{12} = 0.5$ and $1.0 \times 10^7$ g
795:   cm$^{-3}$. Burning in the laminar flame occurs chiefly at a higher
796:   carbon mass fraction due to preheating by radiation. Burning in the
797:   turbulent flame occurs as hot fuel and cold ash are mixed with
798:   negligible radiation transport. Because burning a certain fraction
799:   of carbon releases a certain amount of energy and, at constant
800:   pressure, gives a unique temperature, all points lie on a
801:   well-defined line. A mixture of radiation transport and advection
802:   would have given points in between these two lines (see also
803:   \citet{Asp08}).  \lFig{c12oft}}
804: \end{center}
805: \end{figure*}
806: 
807: 
808: For the calculation shown in the first frame of \Fig{andy}, energy
809: transport by conduction and radiative diffusion was included, just as
810: in the laminar flame cases, and zoning was sufficient to resolve a
811: laminar flame had one been present. However, the transport here is
812: really dominated by the turbulence. The radiative diffusion
813: coefficient (\eq{drad}) varies from 730 cm$^2$ s$^{-1}$ in the fuel to
814: $1.8 \times 10^4$ cm$^2$ s$^{-1}$ in the ash while the turbulent
815: diffusion coefficient  
816: (\Sect{LEM}; $D_{\rm turb} = v' l /C$ with $v' = 2.47$ km s$^{-1}$, $l$ =
817: 15 cm, and $C = 5$) is $7.4 \times 10^5$ cm$^2$ s$^{-1}$.
818: 
819: 
820: \Fig{c12oft} shows the distribution of carbon with temperature in the
821: case of a laminar flame (\Fig{flame}) and the turbulent flame
822: (\Fig{andy}). Both plots show a single-valued, monotonic relation. For
823: the laminar flame this is not surprising. At each temperature in the
824: self-similar front there is a unique carbon abundance. When eddies,
825: coupled with diffusion move heat and carbon around in a way that, in
826: LEM at least, is discontinuous, the result is perhaps more
827: surprising. It is also important that the curves are substantially
828: different in the two cases.
829: 
830: The curve for the turbulent case in \Fig{andy} is what one would
831: result were carbon to burn at constant pressure {\sl with no transport
832:   whatsoever} \citep{Asp08}. This means the process that is
833: transporting heat is transporting composition with equal efficiency,
834: i.e., the effective Lewis number is one. One expects this sort of
835: behavior when turbulence is dominating in the transport of both. The
836: Lewis number in the calculation was still 1000, but small scale eddies
837: kept the mixture so well stirred that even a small amount of ionic
838: diffusion at the grid scale gave the same result as if Le = 1. In
839: fact, the actual turbulent transport would have been even more
840: efficient had the resolution been higher. Recall that the Kolmogorov
841: scale is not resolved here.
842: 
843: \subsubsection{A subgrid model for turbulent transport}
844: 
845: If the turbulent diffusion coefficient, $D_{\rm turb} \sim v' l$, is so much
846: greater than conduction and radiation, then the results should be
847: independent of $D_{\rm rad}$. The second frame of \Fig{andy} shows
848: that this is indeed the case. Here the calculation sets conduction and
849: radiative transport to zero, but instead uses a diffusion coefficient
850: coupling individual zones of
851: \begin{equation}
852: D_{\rm SG} = v(\Delta x) \, \Delta x \, \frac{15}{C}
853: \end{equation}
854: where $\Delta x$ is the grid resolution (constant in this paper), and 
855: \begin{equation}
856: v(\Delta x) = \left(\frac{\Delta x}{l}\right)^3 v'.
857: \end{equation}
858: The subgrid diffusion coefficient $D_{\rm SG}$ represents the mixing
859: effects of unresolved eddies smaller than the grid scale.  The results
860: for $C = 5$ are identical.
861: 
862: The exact value of $D_{\rm SG}$ is not very important so long as it is
863: derived from a length scale that is very much less than the
864: (turbulently broadened) flame thickness. There is a characteristic
865: eddy size that is chiefly responsible for the diffusion (Appendix A)
866: and so long as that scale is well resolved, the results are
867: similar. However there is a choice for $D_{\rm SG}$ that works best as
868: the resolution becomes coarser. \Fig{subgrid} shows that a diffusion
869: coefficient derived from the turbulent speed on the grid scale has
870: much better scaling properties than one based on e.g., $6 \Delta x$,
871: even though the smallest eddy in LEM is $6 \Delta x$. The figure also
872: shows that the diffusive properties of a temperature discontinuity
873: remain unaltered as the resolution is decreased by a factor of 50.
874: 
875: % fig 8 - resolution test of sub-grid model
876: \begin{figure*}
877: \begin{center}
878: \includegraphics[width=0.475\textwidth]{fig8a.eps}
879: \hfill
880: \includegraphics[width=0.475\textwidth]{fig8b.eps}
881: \hfill
882: \includegraphics[width=0.475\textwidth]{fig8c.eps}
883: \hfill
884: \includegraphics[width=0.475\textwidth]{fig8d.eps}
885: \caption{Resolution study of the subgrid model. Nuclear burning and
886:   conduction were turned off so that all transport was by turbulent
887:   eddies. Turbulence with a speed of 10 km s$^{-1}$ was assumed on an
888:   integral scale of 5 cm. An island of ash 20 cm wide and initially
889:   centered at 50 cm was inserted in a background of fuel. The ensuing
890:   diffusive spreading was calculated using a grid of 10,000 (first
891:   frame), 1000 (second frame), and 200 (third frame) zones. The extent
892:   to which the fuel diffuses after $5 \times 10^{-5}$ s is very
893:   similar in all three runs. Larger subgrid diffusion coefficients on
894:   the other hand gave a resolution-dependent spreading that was more
895:   extensive for lower resolution. The fourth frame shows the result
896:   for $D_{\rm turb} = v(6 \Delta x) \, 6 \Delta x$ instead of $D_{\rm turb} = v(\Delta
897:   x) \, \Delta x$ (see text). In that case, the image for 10000 zones
898:   (not shown) is virtually unchanged. (Note that the grid moves to
899:   keep the point $2 \times 10^9$ K centered.) \lFig{subgrid}}
900: \end{center}
901: \end{figure*}
902: 
903: 
904: 
905: % fig 9 - larger integral scale but still distributed regime
906: \begin{figure}
907: \begin{center}
908: \includegraphics[width=0.475\textwidth]{fig9a.eps}
909: \hfill
910: \includegraphics[width=0.475\textwidth]{fig9b.eps}
911: \caption{Flames for larger integral scales than \Fig{andy}, but still
912:   in the WSR regime. The integral scales are 120 cm (first frame) and
913:   960 cm (second frame), and the corresponding {\sl rms} turbulent
914:   speeds on those scales, 4.93 km s$^{-1}$ and 9.86 km s$^{-1}$. The
915:   average propagation speeds are 0.61 km s$^{-1}$ and and 2.3 km
916:   s$^{-1}$. Note that both the flame speeds and widths scale as
917:   $l^{2/3}$ in this regime, both in this figure and as compared with
918:   \Fig{andy}. In the flame with the larger integral scale (right
919:   frame), one begins to see some structure in the flame and small
920:   ``ledges'' of mixture with nearly constant fuel abundance and
921:   temperature (\Sect{ledges}).  \lFig{andy2}}
922: \end{center}
923: \end{figure}
924: 
925: \subsubsection{The effect of increasing the integral scale}
926: \lSect{scaling}
927: 
928: Encouraged by these results, which imply that one need not resolve the
929: laminar scale in this regime in order to obtain accurate results, we
930: proceeded to explore flame properties for a large range of $l$
931: consistent with the condition $\epsilon_t = v'^3/l = 10^{15}$ erg
932: g$^{-1}$ s$^{-1}$ (Table 2).
933: 
934: In general, the speed of a flame propagated by diffusion should obey
935: the scaling relation
936: \begin{equation}
937: v_f \ \sim \ \sqrt{\frac{D}{\tau_{\rm nuc}}}.
938: \lEq{scale}
939: \end{equation}
940: where $D_{\rm turb} \sim v' \, l$.  \Fig{andy2} shows the effect on
941: the flame of increasing $l$ at constant turbulent energy dissipation
942: $\epsilon$, mass density, and carbon mass fraction. For awhile, the
943: profiles follow the scaling implied by \eq{scale}.  If the burning time
944: is independent of $l$,
945: \begin{equation}
946: \begin{split}
947: v_{\rm turb} \ \propto (v' l)^{1/2} \ \propto \ l^{2/3}, \\
948: \delta_{\rm turb} \ \propto \ v_{\rm turb} \ \propto \ l^{2/3}.
949: \end{split}
950: \end{equation}
951: For given, $l$, the flame speed also remains roughly steady.
952: 
953: However, this scaling also predicts its own breakdown.  Since
954: $v_{turb}$ increases as $l^{2/3}$ while $v'$ only increases as
955: $l^{1/3}$ there must be a critical integral scale $\lambda$ for which 
956: the two become equal.  $\lambda$ is determined by the relation 
957: $\lambda/v' = \tau_{\rm nuc}$, where $v'$ is the velocity fluctuation 
958: at scale $\lambda$.  Rewriting this as 
959: $\lambda = \left( v'^3 \tau_{\rm nuc}^3/\lambda \right)^{1/2}$ gives
960: \begin{equation}
961: \lambda = \left(\epsilon \tau_{\rm nuc}^3\right)^{1/2}.
962: \lEq{lambda}
963: \end{equation}
964: 
965: At $l = \lambda$, the flame width and integral scale are the same and
966: the burning speed is approximately equal to the turbulent speed on
967: that scale \citep{Ker01,Woo07}.  This is also the size of an eddy that
968: burns in one turnover time. The quantity, $\epsilon = U_L^3/L$, is a
969: constant by assumption.  Approximate values of $\lambda$ have been
970: tabulated for various densities, turbulent energies and carbon mass
971: fractions by \citet{Woo07}.  Better values for the cases considered
972: here can be inferred from the data in Table 2 which uses our reaction
973: network. Once $l > \lambda$, the flamelets no longer has the 
974: appearance of a diffusively broadened structure, though the 
975: individual flamelets can occasionally coalesce into
976: larger structures. This is the SF regime of the next section.  As
977: \Fig{andy3} and \Fig{andyvofl} show, the flame structure and speed
978: both become highly variable as the integral scale approaches this
979: value.
980: 
981: 
982: % fig 10 - larger integral scale approaching lambda
983: \begin{figure*}
984: \begin{center}
985: \includegraphics[width=0.475\textwidth]{fig10a.eps}
986: \hfill
987: \includegraphics[width=0.475\textwidth]{fig10b.eps}
988: \includegraphics[width=0.475\textwidth]{fig10c.eps}
989: \hfill
990: \includegraphics[width=0.475\textwidth]{fig10d.eps}
991: \caption{A turbulent flame calculated for a still larger integral
992:   scale than in \Fig{andy2} and sampled at four different times.  The
993:   density and carbon mass fraction continue to be $1.0 \times 10^7$ g
994:   cm$^{-3}$ and $X^0_{12}$ = 0.50, and $\epsilon = 10^{15}$ erg
995:   g$^{-1}$ s$^{-1}$. The integral scale here is 76.8 m and the flame's
996:   average speed is 8.3 km s$^{-1}$ compared with an {\rm rms}
997:   turbulent speed on the integral scale of 19.7 km s$^{-1}$.  No
998:   well-organized, self-similar solution is visible at the four
999:   different times (contrast to \Fig{andy} and \Fig{andy2}). At some
1000:   times the flame resembles the multiple structures in the flamelet
1001:   regime (compare the lower right with \Fig{flamelet}). At others,
1002:   sizable islands of well-mixed fuel and ash have nearly constant
1003:   temperature. Some of these (lower left) are almost as large as the
1004:   integral scale itself. The average width of the flame is still
1005:   somewhat broader than the integral scale by about a factor of two.
1006:   \lFig{andy3}}
1007: \end{center}
1008: \end{figure*}
1009: 
1010: % fig 11 - flame speeds for different integral scales
1011: \begin{figure*}
1012: \begin{center}
1013: \includegraphics[angle=90,width=0.475\textwidth]{fig11a.ps}
1014: \hfill
1015: \includegraphics[angle=90,width=0.475\textwidth]{fig11b.ps}
1016: \includegraphics[angle=90,width=0.475\textwidth]{fig11c.ps}
1017: \hfill
1018: \includegraphics[angle=90,width=0.475\textwidth]{fig11d.ps}
1019: \caption{Instantaneous flame speed divided by average speed for four
1020:   different choices of $l$ and $v'$ at a density of 10$^7$ g
1021:   cm$^{-3}$, fuel carbon concentration 50\%, and turbulent energy
1022:   dissipation rate 10$^{15}$ erg g$^{-1}$ s$^{-1}$. The four choices
1023:   correspond to the four integral scales in Table 2, $l$ = 120 cm, 960
1024:   cm, 614 m, and 4.92 km. For the smallest scale (top left), $l <
1025:   \lambda = 64$ m, and the flame speed (and width) are nearly
1026:   constant. Going to larger length scales one sees the effect of
1027:   individual large eddies and unsteady burning. In the last frame
1028:   (lower right), the flame width and speed are approximately the same
1029:   as turbulence on the integral scale and large burning rates are seen
1030:   approximately every $l/v'$ seconds. The speed plotted here is the
1031:   rate at which a single flame would move into the fuel with an
1032:   burning rate equivalent to that on the entire grid. The actual
1033:   burning here actually happens mostly in regions with much smaller
1034:   carbon concentrations than in the unmixed fuel.  \lFig{andyvofl}}
1035: \end{center}
1036: \end{figure*}
1037: 
1038: 
1039: % fig 12 nuclear time scale
1040: \begin{figure*}
1041: \begin{center}
1042: \includegraphics[angle=90,width=0.475\textwidth]{fig12a.ps}
1043: \hfill
1044: \includegraphics[angle=90,width=0.475\textwidth]{fig12b.ps}
1045: \caption{Nuclear time scale and energy generation as a function of
1046:   carbon mass fraction for carbon burning under isobaric conditions.
1047:   The initial carbon mass fraction in the first frame was 0.50, though
1048:   only the evolution below 0.30 is shown. Similarly in the second
1049:   frame the initial carbon mass fraction was 0.75. Solid lines show
1050:   the induction time scale (the time remaining until most of the
1051:   carbon is consumed) and the nuclear energy generation rate divided
1052:   by the maximum value achieved in the evolution. Dashed lines show
1053:   the nuclear time scale, $[(1/X_{12})(dX_{12}/dt)]^{-1}$, and the
1054:   temperature divided by the ash temperature. For the 50\% carbon run
1055:   the maximum energy generation and ash temperature were $3.89 \times
1056:   10^{19}$ erg g$^{-1}$ s$^{-1}$, and $2.59 \times 10^9$ K. For the
1057:   75\% run, the corresponding values were $1.29 \times 10^{21}$ erg
1058:   g$^{-1}$ s$^{-1}$, and $2.95 \times 10^9$ K. The induction time was
1059:   arbitrarily defined to reach zero when the remaining carbon fraction
1060:   was 0.03 for the 50\% carbon case and 0.05 for the 75\% carbon
1061:   case. Energy released beyond these points is negligible and the time
1062:   for the carbon to go to precisely zero is arbitrarily long. For high
1063:   remaining carbon abundance, the induction time scale is much shorter
1064:   than the instantaneous nuclear time scale because much of the
1065:   remaining carbon will burn at a higher temperature. Note the
1066:   relatively small change in induction time over an interesting range
1067:   of carbon mass fraction, 0.2 to 0.1 (left) and 0.4 to 0.1
1068:   (right). \lFig{taunuc}}
1069: \end{center}
1070: \end{figure*}
1071: 
1072: \subsubsection{The nuclear time scale in the WSR regime}
1073: 
1074: Burning in the WSR, $l < \lambda$, produces flames that, for a given
1075: constant density and fuel composition, have a well-defined nuclear
1076: time scale, $\tau_{\rm nuc}$.  \Fig{taunuc} shows the temperature, and
1077: nuclear time scale as function of carbon mass fraction in the fuel for
1078: two initial carbon concentrations (50\% and 75\% by mass) and fuel
1079: density 10$^7$ g cm$^{-3}$. The burning is assumed to be isobaric and
1080: two time scales are computed. One time scale reflects the
1081: instantaneous rate of carbon consumption, $\tau = X_{12}/(dX_{12}/dt)$
1082: (the dash-dotted line). More relevant to the flame speed, however, is
1083: the {\sl induction} time scale, which is the time required to consume
1084: most of the remaining fuel. In a situation where burning increases the
1085: temperature, and therefore the reaction rate, the induction time can
1086: be much shorter than the instantaneous burning time scale (defined by the 
1087: current abundance divided by the current burning rate). Dividing the
1088: turbulent flame width by the turbulent flame speed for those cases in
1089: Table 2 where both quantities are well defined gives an approximately
1090: constant value $\tau_{\rm nuc} \approx 0.003$ s for $\epsilon =
1091: 10^{15}$ erg g$^{-1}$ s$^{-1}$ and $X_{12}^0$ = 0.50.  This
1092: corresponds to a carbon mass fraction in \Fig{taunuc} of $X_{12}
1093: \approx 0.20$ where the energy generation is about 1/$e$ of its
1094: maximum.
1095: 
1096: Defining $\tau_{\rm nuc}$ in this fashion as the induction time scale
1097: from the point where the energy generation reaches $1/e$ of its
1098: eventual maximum allows the computation of $\lambda = (\epsilon
1099: \tau_{\rm nuc}^3)^{1/2}$. For the case of carbon mass fraction in the
1100: fuel = 75\%, the induction time scale was calculated based on a
1101: starting carbon mass fraction of 37\%. The values of $\lambda$ so
1102: determined are given in Table 3. These values are about a factor of
1103: two less than given in \citet{Woo07} because they refer to a time
1104: scale derived from the width of the energy generating region. If one
1105: instead uses the larger width based upon temperature or carbon mass
1106: fraction, the derived values for $\lambda$ are about twice as large,
1107: consistent with \citet{Woo07}.
1108: 
1109: \subsection{Stirred Flames (SF Regime)}
1110: \lSect{stirred}
1111: 
1112: The stirred flame (SF) regime, which is characterized by Ka $\gg 1$
1113: and Da = $L/(U_L \tau) = (L/\lambda)^{2/3} > 1$, is the most complex
1114: of the three regimes of turbulent burning. In this case, there is no
1115: well-determined scale for the flame width. Structures of size $\sim
1116: \lambda$ persist to some extent, but since the flame experiences a new
1117: eddy of length $\lambda$, in approximately the same time it takes to
1118: burn a distance $\lambda$, it is subject to continual disruption. At
1119: times there may be almost no burning; at others, nearly simultaneous
1120: burning happens on scales much larger than $\lambda$.  The fact that
1121: the flame has no persistent steady state also reflects a poorly
1122: defined nuclear time scale. Each temperature has a different burning
1123: time scale and thus the characteristic widths of mixtures prepared by
1124: the turbulence is very temperature sensitive. Cooler mixtures have
1125: larger length scales.  In general though, the average time scales are
1126: longer and the flame structures larger than expected from our studies
1127: of the WSR.
1128: 
1129: Because of the large integral scale in the supernova, the SF regime is
1130: encountered as soon as the Karlovitz Number exceeds about 10. The
1131: Damk\"ohler number at that point is already much greater than unity and
1132: the WSR regime (Da $< 1$) is encountered much later, if ever.
1133: 
1134: \subsubsection{The transition from the flamelet to the SF regime}
1135: \lSect{transition}
1136: 
1137: Just before the laminar flame is disrupted at Ka $\sim$ 10, there are
1138: $\sim U_L/S_{\rm lam}$ flame surfaces folded in the flame brush. The
1139: average spacing between these flamelets is $d \sim (L/U_L) S_{\rm
1140:   lam}$ (Table 3) and the thickness of each is $\delta_{\rm lam}$. A
1141: short time later, as Ka continues to increase, these flamelets are
1142: smeared out to make a smaller number of broader, faster structures
1143: %%
1144: with characteristic average thickness $\lambda$.  Their number then is
1145: $\sim \sqrt{\rm Da} = (L/\lambda)^{1/3}$, and their spacing is $\sim
1146: L/\sqrt{{\rm Da}} = L^{2/3} \lambda^{1/3}$.  So long as Da $\gg 1$, as
1147: it is at the transition, the new broadened structures will, on the
1148: average, still be separated by distances much greater than their size
1149: and will not coalesce into one large mixture.
1150: 
1151: The quantity $\lambda$ (\eq{lambda}) that plays a such a critical role
1152: in this discussion can be thought of as a generalization of the Gibson
1153: length (\eq{Gibson}). For the flamelet regime, the Gibson scale is the
1154: size of the eddy that would be crossed by a laminar flame with speed
1155: $S_{\rm lam}$ on an eddy turnover time. It is smaller for larger
1156: turbulent energy. The quantity $\lambda$, on the other hand, is also
1157: the size of an eddy that turns over on a nuclear time scale, but its
1158: size depends only on turbulence properties, not the radiative
1159: diffusion that sets $S_{\rm lam}$. It is larger for larger turbulent
1160: energy. As the supernova expands, the Gibson scale shrinks as $S_{\rm
1161:   lam}$ declines. Eventually when Ka $= 1$, $l_G = \delta_{\rm lam}$,
1162: but except for differences due to a changing nuclear time scale
1163: (\Sect{scaling}), at Ka = 1, $\lambda$ is also approximately equal to
1164: $l_G$ and $\delta_{\rm lam}$. Thereafter, as the density decreases
1165: more, $l_G$ ceases to have much meaning since laminar flame speeds are
1166: no longer relevant. The new relevant scale, $\lambda = (\epsilon
1167: \tau^3)^{1/2}$, on the other hand, grows rapidly at lower density due
1168: to the increasing nuclear time scale.
1169: 
1170: What this means is that the transition from laminar burning to the SF
1171: regime is probably smooth and uneventful. Flamelets will grow
1172: gradually in speed and width as the density declines, not
1173: discontinuously. Because the nuclear time, and thus $\lambda$, lack
1174: precise definitions in the SF regime, and because of intermittency,
1175: one cannot completely rule out large scale transient mixing at the
1176: laminar-SF transition, especially if the turbulent energy is very
1177: large, but a detonation here seems unlikely. If it did happen at such
1178: high density, a very bright supernova would result due to the near
1179: complete incineration of the star.
1180: 
1181: \subsubsection{Complex structure}
1182: \lSect{complex}
1183: 
1184: As the density continues to decrease, $\lambda$ increases and there
1185: are fewer, thicker, faster turbulent flamelets in the flame brush
1186: (\Fig{andy3} and \Fig{andyvofl}). The limit of one flame of size
1187: $\lambda = L$ is reached when Da = 1. This is the condition suggested by
1188: \citet{Woo07} as likely for detonation.  The present study shows,
1189: however, that mixed regions larger than $\lambda$ can exist
1190: transiently even for Da $\gg 1$. Since the critical mass for detonation
1191: decreases rapidly with increasing density, this makes detonation
1192: easier.
1193: 
1194: Consider the case of 50\% carbon, a turbulent dissipation rate of
1195: 10$^{15}$ erg g$^{-1}$ s$^{-1}$, density 10$^7$ g cm$^{-3}$, and
1196: integral scale $l = 4.92$ km (Table 2). For these conditions,
1197: $\lambda$ = 64 m (Table 3) and Ka = $(4.92 \times 10^5/6.4 \times
1198: 10^3)^{2/3} = 18$. The average number of flame surfaces is $\sqrt{\rm
1199:   Ka} \sim 4$. \Fig{andy5} shows that mixed regions as big as L (and
1200: much bigger than $\lambda$) occasionally exist. The extent of mixing
1201: and hence the flame speed is highly variable (\Fig{andyvofl}). Most of
1202: the time, less mixing is seen than in \Fig{andy5} and only a few
1203: disjoint regions of burning are present, but structures like these can
1204: exist for an eddy turnover time or so. Characteristics of these
1205: regions include a carbon mass fraction much less than in the fuel,
1206: typically on the rapidly rising part of the energy generation curve in
1207: \Fig{taunuc}, and temperatures that are a substantial fraction of the
1208: ash temperature. The relatively high temperatures are a consequence of
1209: the temperature dependent nature of the heat capacity (\Sect{cp}).
1210: 
1211: 
1212: 
1213: % fig 13 - shows transient burning at large Da 
1214: \begin{figure*}
1215: \begin{center}
1216: \includegraphics[width=0.475\textwidth]{fig13a.eps} \hfill
1217: \includegraphics[width=0.475\textwidth]{fig13b.eps} 
1218: \caption{A single flame in the stirred flame regime for an integral
1219:   scale of 4.92 km. The turbulent speed on that scale is 78.9 km
1220:   s$^{-1}$ and the average flame speed and width are also close to
1221:   these values. Many complex and folded structures are seen like in
1222:   \Fig{andy3}, but the above two snapshots show that occasionally
1223:   quite regular, well-mixed regions of fuel and ash exist. Sometimes
1224:   several of these structures add together to create a mixed region
1225:   about as large, or even larger than the integral scale.
1226:   \lFig{andy5}}
1227: \end{center}
1228: \end{figure*}
1229: 
1230: \subsubsection{Ledges}
1231: \lSect{ledges}
1232: 
1233: The ledges seen in \Fig{andy5} play an important role in promoting
1234: detonation and it is thus worth spending a moment to discuss their
1235: credibility. They have not been seen previously in any combustion
1236: simulation or experiment.  For accessible terrestrial conditions, it
1237: is believed that chemical flames in this regime would extinguish
1238: \citep{Pet00}.  Numerical exploration of this regime, other than with
1239: LEM, has so far been impractical.
1240: 
1241: Nevertheless, it has long been known that property fields in
1242: turbulence exhibit intermittent behavior, including the occurrence of
1243: transient well-mixed regions separated by cliffs (sharp property
1244: changes).  This structure is seen, for instance, in 1D profile data
1245: from 3D numerical simulations \citep{Wat06}.  A quantitative signature
1246: of this structure is the saturation of high-order structure-function
1247: exponents, reflecting the dominant contribution of the cliff regions
1248: to high-order intermittency statistics \citep{Cel00}.  Measurements by
1249: \citet{Moi01} show clear evidence of the saturation of the scalar
1250: structure-function exponents.  LEM structure-function exponents
1251: exhibit a somewhat slower roll-off to saturation \citep{Ker91},
1252: suggesting that the prevalence and duration of well-mixed regions in
1253: LEM might be lower than actually occurs in turbulence.  This is
1254: plausible because maps in LEM are statistically independent events,
1255: but the eddies that they represent actually occur in bunches because
1256: each eddy breakdown is an energy source for subsequent eddy breakdown.
1257: This bunching contributes to the intermittency of turbulence.
1258: 
1259: There is some limited understanding of why turbulence exhibits these
1260: properties \citep{Cel00}.  Further clarification would provide
1261: insight into a mechanism that appears to play a key role in the timely
1262: occurrence of detonations in supernovae.
1263: 
1264: % fig 14 - development of detonation
1265: \begin{figure*}
1266: \begin{center}
1267: \includegraphics[angle=90,width=0.475\textwidth]{fig14a.ps}
1268: \hfill
1269: \includegraphics[angle=90,width=0.475\textwidth]{fig14b.ps}
1270: \includegraphics[angle=90,width=0.475\textwidth]{fig14c.ps}
1271: \hfill
1272: \includegraphics[angle=90,width=0.475\textwidth]{fig14d.ps}
1273: 
1274: \caption{The birth of a detonation. A sample mixture calculated using
1275:   LEM for a density $1.0 \times 10^7$ g cm$^{-3}$, v' = 500 km
1276:   s$^{-1}$, and L = 10 km was mapped into a compressible hydrodynamics
1277:   code and its subsequent evolution was followed. The time selected
1278:   was characterized by a very subsonic flame speed but temperature
1279:   gradients that looked ``interesting''. Following the remap, which
1280:   preserved distance scale, ash was on the left and fuel on the
1281:   right. A detonation developed that, barring large barriers of ash,
1282:   would explode the whole star (see text). Shown in the plot are
1283:   carbon mass fraction (gold), pressure (blue), temperature (black),
1284:   and velocity (red) at four different times - 0, 0.15, 0.30, and 0.43
1285:   ms after the mapping. The velocity has been divided by 500 km
1286:   s$^{-1}$ in frames 1 and 2, 2000 km s$^{-1}$ in frame 3 and 5000 km
1287:   s$^{-1}$ in frame 4. The pressure has been scaled by the background
1288:   value, $9.08 \times 10^{23}$ dyne, in frames 1, 2, and 3, and by an
1289:   additional factor of 5 in frame 4. The temperature has been divided
1290:   by $3.0 \times 10^9$ K.  \lFig{detonation}}
1291: \end{center}
1292: \end{figure*}
1293: 
1294: 
1295: \section{Conditions for Detonation}
1296: \lSect{ddt}
1297: 
1298: In order that a detonation occur, three conditions must be satisfied.
1299: First, some region must burn supersonically. The simplest example
1300: would be a hot, isothermal volume in which the nuclear induction time
1301: scale was a constant. As the temperature rises, the time scale
1302: decreases and for a sufficiently large volume there comes a
1303: temperature where the burning time is shorter than the sound crossing
1304: time. The pressure will increase in that region faster than expansion
1305: can damp its growth and, after a maximum temperature is reached (i.e.,
1306: the fuel is gone), the region expands at a speed that is somewhat
1307: higher than the sound speed in the surrounding medium. In reality, the
1308: temperature is never completely isothermal, but, as we shall see, it
1309: suffices to burn only some fraction of the fuel within the volume in a
1310: sound crossing time. A subset of fluid elements within the volume must
1311: have the same temperature to some level of tolerance that depends on
1312: their burning time scale. The smaller the fraction, the smaller the
1313: overpressure, but sonic expansion nevertheless occurs. The colder
1314: matter is an inert dilutant.
1315: 
1316: Second, the size of the ``detonator'' must exceed some critical value
1317: \citep{Nie97,Dur06}. That critical mass is larger if the mass fraction
1318: of combustible fuel is small or if the fraction of inert cold matter
1319: is large. Because some mixing of ash into the pre-detonation region is
1320: unavoidable and because an appreciable amount of carbon must burn to
1321: reach temperatures where the time scales start to approach sonic, the
1322: critical masses we compute here will be larger than those for mixtures
1323: of just carbon and oxygen with a smooth temperature structure.  For a
1324: given turbulent energy, density must fall to lower values to obtain
1325: these larger structures.
1326: 
1327: Third, and perhaps most subtly, there must exist, in a significant
1328: fraction of the mass, preferably at its edge, a nearly sonic phase
1329: velocity for the burning \citep{Zel85,Kho97}. This is the ``shock wave
1330: amplification by coherent energy release'' (SWACER) mechanism for
1331: initiating a detonation in an unconfined medium. One requires an
1332: appreciable boundary layer where
1333: \begin{equation}
1334: \frac{d \tau_{\rm nuc}(T)}{dx} \approx \ (c_{\rm sound})^{-1}.
1335: \end{equation}
1336: 
1337: Two properties of the flame in the SF regime assist in satisfying
1338: these three conditions. First is the unsteady nature of the burning
1339: (\Sect{complex}). As fuel and ash are mixed, burning may briefly
1340: almost go out, only to return with a vengeance after sufficient mixing
1341: and slow burning have occurred. This allows the creation of regions
1342: that, after some delay comparable to a turnover time, can consume fuel
1343: at a rate faster than a single flame moving at the turbulent {\sl rms}
1344: speed on the integral scale. Amplification factors of three are
1345: frequently observed, and larger values are presumably possible in rare
1346: instances.
1347: 
1348: Second, as discussed in \Sect{ledges}, transient well-mixed
1349: structures, ledges, are a frequent occurrence in the SF
1350: regime. Turbulence does not always lead to heterogeneity on
1351: macroscopic scales. These mixed regions have relatively constant
1352: induction time and they are at least occasionally bounded by regions
1353: in which the temperature decreases (and fuel concentration increases)
1354: fairly smoothly. The characteristic size of these regions is
1355: approximately $\lambda$, which increases with decreasing density.  As
1356: a result, the ratio of mixing time to sound crossing time decreases,
1357: since
1358: \begin{equation}
1359: \frac{\tau_t}{\tau_{\rm sonic}} \ \propto
1360: \ \frac{\lambda^{2/3}}{\lambda} \ = \lambda^{-1/3},
1361: \end{equation}
1362: and this helps detonation happen.  However, as the density decreases,
1363: the critical size required to initiate a detonation also increases
1364: dramatically \citep{Nie97,Woo07}, so mixing a smaller region at higher
1365: density may, in some cases, be better than a larger one at low
1366: density.  Detonation may occur before the characteristic size of the
1367: mixed region becomes equal to the integral scale.
1368: 
1369: Another important point favoring detonation is that once burning
1370: starts to happen on a time scale approaching sonic, it no longer
1371: occurs at constant pressure. The pressure rises with respect to the
1372: surrounding fuel and is not immediately relieved by expansion. At
1373: constant volume, a given amount of carbon burning raises the
1374: temperature more, thus appreciably shortening the time scale for
1375: additional burning.  That is, for the relevant conditions, the heat
1376: capacity at constant pressure is appreciably larger than the heat
1377: capacity at constant volume. The final temperature from burning all
1378: the fuel is also higher.
1379: 
1380: 
1381: \subsection{Detonation Observed for a Mixture Calculated Using LEM}
1382: \lSect{itworks}
1383: 
1384: In order to verify that some of the mixed regions calculated using LEM
1385: would actually detonate, the burning of select mixtures was followed
1386: using the compressible hydrodynamics code, Kepler
1387: \citep[][\Fig{detonation}]{Wea78,Woo02}. The composition and temperature
1388: structure were taken from an LEM simulation of flame-turbulence
1389: interaction at a density of 1.0 $\times 10^7$ g cm$^{-3}$ for a
1390: characteristic turbulent speed of 500 km s$^{-1}$ on an integral scale
1391: of 10 km (Table 2). Carbon had a mass fraction of 0.75 in the
1392: fuel. Use of 65536 zones gave a characteristic Reynolds number of
1393: 80,000 and a resolution not achievable in a multi-dimensional
1394: simulation on this length scale. The LEM calculation was run for 0.10
1395: s, or 5 eddy turnover times on the integral scale, during which 40
1396: dumps were created at equal time intervals. Visual inspection isolated
1397: several cases for further study. One of these was dump number 28, made
1398: 70 ms after the beginning of the run. The overall burning rate at this
1399: time was not particularly high, corresponding to an effective speed of
1400: only 250 km s$^{-1}$. However, most of the energy generation was
1401: occurring in regions where the carbon mass fraction was already low
1402: (\Fig{detonation}), so the local {\sl rate} of pressure increase was
1403: quite large, even though the integrated total change in pressure in
1404: the end was small.
1405: 
1406: Kepler, frequently used for studying stars and supernova explosions,
1407: is an implicit 1D hydrodynamics code with a spherical Lagrangian
1408: mesh. In order to simulate a problem with approximately plane parallel
1409: geometry, the 4 km of interest was mapped on top of a sphere of pure
1410: ash (mostly magnesium and silicon) with a radius 10 km so that LEM
1411: zones of constant thickness had approximately constant mass and
1412: thickness in Kepler. In the figure, that 10 km has been subtracted off
1413: of the x-coordinate. The overall thickness in km of the region of
1414: interest was the same in the Kepler study as in the LEM
1415: calculation. The temperature was by no means isothermal in this
1416: region. In fact, an isothermal runaway would not have initiated a
1417: detonation here, even if the burning were supersonic. However, several
1418: key elements favoring detonation were in place. First, there were some
1419: regions where the temperature was already high because of previous
1420: mixing and burning. The rate of carbon burning was quite high in these
1421: regions with a peak energy generation rate near 10$^{21}$ erg g
1422: s$^{-1}$. These were embedded in an extended region with X$_{12}$
1423: $\approx 0.4$ that was already quite warm due to mixing. Finally that
1424: mixture lay at the base of a region where the carbon mass fraction
1425: increased steadily, if somewhat noisily, in an outwardly direction. It
1426: is important to emphasize that this was the result of an LEM
1427: calculation, and not an artificial construct.
1428: 
1429: During the first phase, the ``initiator'' burned rapidly producing an
1430: overpressure in the surrounding zones of about 10 - 15\%. This
1431: increase was sufficient to cause the warm material to burn faster, on
1432: a roughly sonic time scale. Their expansion then compressed and
1433: ignited material at successively higher carbon mass fractions in the
1434: gradient. While weak at first, the detonation strengthened and by the
1435: last frame shown was strong enough that its permanent propagation was
1436: guaranteed. It was followed until it left the grid. Among the
1437: interesting implications here is that a region need not be highly
1438: isothermal, nor need all of it initially burn on a supersonic time
1439: scale to provoke a detonation, though near sonic speeds are needed. A
1440: pile up of strong acoustic pulses at around 2.3 km initiates the
1441: runaway.
1442: 
1443: % fig 15 - shows strong pulse but not detonation 
1444: \begin{figure*}
1445: \begin{center}
1446: \includegraphics[angle=90,width=0.475\textwidth]{fig15a.ps} \hfill
1447: \includegraphics[angle=90,width=0.475\textwidth]{fig15b.ps} 
1448: \caption{Development of a small scale explosion within a flame front
1449:   in the WSR regime. The highly variable gold line is the carbon mass
1450:   fraction. The dark line that is initially zero is the velocity. The
1451:   velocity scale is in units of 1000 km s$^{-1}$. Unlike in
1452:   \Fig{detonation}, the rapid burning of a mixed region does not
1453:   initiate a detonation here, but it does produce a strong subsonic
1454:   pulse. These may be very common events and would be a mechanism for
1455:   adding turbulent energy at the flame scale.  The two edits here are
1456:   from the same run that produced the detonation in \Fig{detonation}
1457:   sampled $3.9 \times 10^{-4}$ s apart.  \lFig{pulse}}
1458: \end{center}
1459: \end{figure*}
1460: 
1461: \subsection{Active Turbulent Combustion}
1462: \lSect{ATC}
1463: 
1464: Almost as interesting were the many other sample mixtures in this and
1465: other runs that, when mapped into Kepler, did {\sl not}
1466: detonate. Instead, the irregular burning that is characteristic of
1467: combustion in the SF regime produces strong pressure waves. A
1468: particularly strong pulse occurs roughly every turnover time on the
1469: integral scale because these big mixing events trigger a lot of
1470: burning. An example from the same run that produced the detonation in
1471: \Fig{detonation} is shown about one turnover time later (dump 39 at 95
1472: ms) in \Fig{pulse}. This time the burning did not produce a
1473: detonation, but substantial burning still occurred on less than
1474: a sonic time scale for the 4 km shown (the sound crossing time for
1475: this region is about 1 ms). This burning produced a strong pulse that
1476: sent matter moving inwards and outwards at $\sim600$ km s$^{-1}$, about the
1477: same value as assumed for turbulence on an integral scale of 10 km in
1478: the study.
1479: 
1480: In three dimensions, these pulses would be quasi-spherical and would
1481: occur all over the surface where burning and mixing are going on.
1482: Collisions between the fronts would pump additional energy into
1483: turbulence on a scale comparable to the width of the burning region,
1484: i.e., the integral scale, or about 10 km.  As noted, the velocities in
1485: these pulses are frequently larger than expected from the assumed
1486: turbulent energy on that length scale. As discussed by
1487: \citet{Ker96,Nie97}, this sets the stage for a potential runaway.
1488: More turbulence leads to increased mixing which leads to more
1489: violent irregular burning, which creates more turbulence. The
1490: culmination of this runaway could be a detonation.
1491: 
1492: % fig 16 - summary plot
1493: \begin{figure}
1494: \begin{center}
1495: \includegraphics[angle=90,width=0.475\textwidth]{fig16a.ps} \hfill
1496: \includegraphics[angle=90,width=0.475\textwidth]{fig16b.ps} 
1497: \caption{Speeds required on an integral scale of 10 km to establish
1498:   the necessary conditions for detonation. The first panel is for an
1499:   initial carbon mass fraction of 50\%; the second, for 75\%. The solid
1500:   line that increases rapidly from left to right is the condition Da =
1501:   10 and the two dashed line parallel to it are Da = 1 (upper) and 100
1502:   (lower). The nearly horizontal solid line is the sound speed divided
1503:   by 5 and the dash-dotted line beneath is the sound speed divided by
1504:   10. The lowest solid line is the condition Ka = 10. Scaling to other
1505:   values is given in the text. In order to detonate the speed on the
1506:   integral scale must be greater than that given by Ka = 10 and in the
1507:   vicinity of Da = 10. It is also necessary that the turbulent speed
1508:   be at least 1/10 the sound speed and, better still, 1/5 the sound
1509:   speed, but turbulent speeds above 20\% sonic may not be achieved in
1510:   the star. Thus detonation for the assumed conditions and a carbon
1511:   mass fraction of 0.5 probably occurs in the band between 0.8 and 1.6
1512:   $\times 10^7$ g cm$^{-3}$. For 75\% carbon, the likely density range
1513:   is 0.5 to 1.0 $\times 10^7$ g cm$^{-3}$. \lFig{darm}}
1514: \end{center}
1515: \end{figure}
1516: 
1517: \section{Conclusions}
1518: \lSect{concl}
1519: 
1520: Three regimes of turbulent flame propagation relevant to nuclear
1521: burning in a Type Ia supernova have been explored. At high density,
1522: for Karlovitz numbers less than about one, burning occurs in multiply
1523: folded laminar flames. The overall progress of burning is governed by
1524: the turbulent energy and has a speed that is independent of the
1525: laminar speed. A similar description of ``laminar flame brushes'' has been
1526: given many times in the literature, but this is the first time it has
1527: been simulated in a supernova (e.g., \Fig{flamelet}).
1528: 
1529: The average number of flamelets in the flame brush is $U_L/S_{\rm
1530:   lam}$, and initially is not large. Because of this, there will be
1531: considerable variation in the burning rate.  We estimate such
1532: variations to be as large as a factor of three. However, detonation
1533: does not happen so early because the turbulent speed then is very
1534: subsonic and the individual flamelets are thin.  Later, the number of
1535: flamelets becomes very large, reaching hundreds or even thousands, and
1536: variations in the overall burning rate are much smaller. Detonation
1537: remains impossible in the flamelet regime. No critical mass of hot
1538: fuel can be obtained.
1539: 
1540: As the Karlovitz number increases above about 10, hot ash and cold
1541: fuel can be mixed for the first time and detonation is, in principle,
1542: possible.  Two cases of turbulent burning were explored corresponding
1543: to the well-stirred reactor (WSR; Da $< 1$, \Sect{distrib}) and
1544: stirred flames (SF; Da $> 1$; \Sect{stirred}). In the WSR regime,
1545: turbulent diffusion substitutes for heat diffusion and turbulently
1546: broadened flames result. These flames can be much larger than the
1547: integral scale. Because the integral scale in a supernova is
1548: so large, this limiting case is probably never fully realized, but is
1549: amenable to numerical simulation \citep{Asp08}. We normalized an
1550: uncertain constant in the LEM calculation, $C$, (\Sect{LEM}) to those
1551: simulations. In the stirred flame regime, one again has something like
1552: a flame brush, but with turbulently broadened structures substituting
1553: for the individual flamelets in the flamelet regime. The structure of
1554: burning here is complex and highly variable. Because the burning time
1555: scale is very temperature sensitive, it is possible to mix large
1556: regions of ``warm'' fuel and ash in which are embedded smaller nuggets
1557: on the verge of explosion. Mixed regions of nearly constant temperature
1558: are sometimes observed (\Fig{andy5}, \Sect{ledges}). 
1559: 
1560: Spontaneous detonation can happen in the SF regime
1561: (\Fig{detonation}, \Sect{itworks}). It does not require complete mixing
1562: on the integral scale, and hence Da $\approx 1$ as discussed in
1563: \citet{Woo07}, but is favored by values of Da that are not large.
1564: The necessary conditions occur infrequently and require a favorable
1565: confluence of several mixed regions including: a) some region, perhaps
1566: not large and with a low mass fraction of carbon, where the burning
1567: occurs sufficiently rapidly to increase the pressure by a small
1568: fraction supersonically; and b) an extended region where the carbon
1569: mass fraction rises gradually and the temperature falls
1570: slowly. Neither region need be particularly homogeneous as long as
1571: some part of the composition burns supersonically for condition a) and
1572: there are neither large barriers of ash or abrupt, sustained
1573: increases in carbon mass fraction for condition b).
1574: 
1575: In general, the production of these situations require turbulent
1576: speeds that are already a considerable fraction of sonic, certainly
1577: within 10\% and probably 20\%.  Overall the progress of burning in the
1578: SF regime varies frequently by a factor of three up and down, and rare
1579: excursions to larger values probably also occur due to intermittency
1580: \citep{Pan08}. We also find that appreciable turbulence can be put
1581: into the burning region on a scale comparable to the integral scale by
1582: small subsonic explosions of mixtures that fail to detonate
1583: (\Sect{ATC}, \Fig{pulse}). Before the big bang in these stars, there is
1584: a lot of thunder. Though we have not demonstrated it here, we
1585: speculate that this energy input may exceed that put in at the large
1586: scale by the flame instabilities. If so, there is the possibility of a
1587: turbulent runaway in which mixing pumps energy into turbulence which
1588: in turn causes accelerated mixing \citep{Ker96}. The endpoint would be
1589: detonation.
1590: 
1591: These necessary conditions for detonation are summarized in
1592: \Fig{darm}. Detonation is estimated to occur for a reasonable range of
1593: turbulent energies in the nearly horizontal band between the lines Da
1594: = 10 and Da = 100.  It must be acknowledged that this figure is very
1595: approximate because the nuclear time scale (and hence Damk\"ohler
1596: number) are poorly determined. The effective nuclear time scale in Da
1597: is probably longer than in Table 3 and intermittency essentially
1598: raises the effective turbulent speed making detonation possible at a
1599: higher density. This is why the conditions Da = 10 - 100 in \Fig{darm}
1600: are probably more favored than say 1 to 10. Above Da = 100, the mixed
1601: regions may be too small to initiate a detonation.
1602: 
1603: Three-dimensional simulations by \citet{Roe07a} show that the necessary
1604: degree of turbulence for detonation, roughly $u'$ = 500 km s$^{-1}$, is
1605: realized in the full star models. 
1606: 
1607: We also find some dependence of the detonation conditions on initial
1608: carbon mass fraction \citep[see also][]{Woo07}. For lower carbon mass
1609: fractions, Damk\"ohler numbers of order 10 are reached at a
1610: higher density. If detonation does occur at a higher density, the
1611: explosion makes more $^{56}$Ni and a brighter supernova. However,
1612: detonation depends on achieving a significant overpressure on a sonic
1613: time scale. By the time that carbon-poor fuel carbon reaches a
1614: temperature where it burns rapidly (\Fig{taunuc}), the remaining
1615: burning produces too small an overpressure. Detonating carbon-rich fuel
1616: is, in this sense, easier \citep[see also][]{Ume99a,Ume99b}. It is
1617: important to note that the relevant location for detonation is
1618: probably in the outer layers where the density first declines below
1619: 10$^7$ g cm$^{-1}$ at the flame front, not near the center. The carbon
1620: abundance is higher in these outer layers.
1621: 
1622: All in all, our results are supportive of the hypothesis that
1623: some, perhaps even all Chandrasekhar mass white dwarfs explode by a
1624: delayed detonation that occurs shortly after the burning enters the SF
1625: regime. This would then make the location and number of detonation
1626: points, along with the ignition conditions \citep{Kuh06}, the principal
1627: determining factors in the intrinsic properties of a Type Ia
1628: supernova.
1629: 
1630: Several avenues for future investigation are opened up by this work.
1631: First, our results hinge on the calibration of the 1D LEM model to
1632: direct 3D simulation \citep{Asp08}. The 3D study used for
1633: normalization here was carried out in the WSR regime, but the results
1634: (e.g., the constant C) were assumed to be valid in the SF regime.
1635: Given the subgrid model developed here, it should be possible to carry
1636: out equivalent 3D studies for the SF regime ($Da \gg 1$).
1637: 
1638: The results for detonation (\Sect{itworks}) and active turbulent
1639: combustion (\Sect{ATC}) were obtained by mapping results from LEM
1640: using a linear grid into a 1D compressible hydro-code with spherical
1641: coordinates. It would be greatly preferable to see both the mixing and
1642: the strong pressure waves in the same, preferably 3D study. Because of
1643: the range in length scales, the need for a large effective Reynolds
1644: number, and the rare, transient nature of the phenomena, this will
1645: require a very major investment of computational resources, but should
1646: be practical in the near future.
1647: 
1648: Ultimately, of course, one would want to see these results applied to
1649: full scale models of the supernova and its light curve.
1650: 
1651: 
1652: 
1653: \acknowledgements
1654: 
1655: The authors gratefully acknowledges helpful conversations on the
1656: subject of the paper with Andy Aspden, John Bell, and Martin Lisewski.
1657: This research has been supported by the NASA Theory Program
1658: (NNG05GG08G) and the DOE SciDAC Program (DE-FC02-06ER41438). Work at
1659: Sandia was supported by the US Department of Energy, Office of Basic
1660: Energy Sciences, Division of Chemical Sciences, Geosciences and
1661: Biosciences. Sandia is a multiprogram laboratory operated by Sandia
1662: Corporation, a Lockheed Martin Company, for the US Department of
1663: Energy under contract DE-AC04-94AL85000.
1664: 
1665: \vskip 0.5 in
1666: 
1667: \begin{widetext}
1668: %\appendix
1669: \begin{appendices}
1670: \section{LEM flame speed in the flamelet regime}
1671: 
1672: 
1673: Properties of the triplet map imply a simple, exact expression for the
1674: turbulent burning velocity $v_{\rm turb}$ in terms of the LEM
1675: parameters $D_{\rm turb}$ (turbulent diffusivity) and $L$ (largest
1676: allowed map size) under the conditions $v_{\rm turb}\gg S_{\rm lam}$ (here
1677: assuming the flamelet regime) and $D_{\rm turb}\gg \nu /\rho$
1678: (implying Re $\gg 1$).  On the LEM domain (coordinate $x$), assume
1679: fuel on the right and ash on the left, so the flame advances rightward
1680: (direction of increasing $x$).  $v_{\rm turb}$ is evaluated by
1681: tracking the forward progress of the rightmost (forwardmost) ash
1682: location.  The only mechanism affecting this location, denoted $r$, is
1683: the triplet map because small $S_{\rm lam}$ implies that the contribution of
1684: laminar burning is negligible.
1685: 
1686: Any triplet map containing $r$ maps $r$ to three locations, the
1687: rightmost of which, denoted $r'$, exceeds $r$.  Specifically, if the
1688: map interval is $[x_1,x_2]$, then $r'=x_2-\frac{1}{3}(x_2-r)$, i.e.,
1689: the distance from $r'$ to $x_2$ is one third of the pre-map distance.
1690: The advancement of the rightmost ash location is therefore $r'-r
1691: =\frac{2}{3}(x_2-r)$.
1692: 
1693: Suppose that the interval $[x_1,x_2]$ is chosen arbitrarily.  If it
1694: contains $r$, then $r$ is equally likely to be anywhere within the
1695: interval, and hence is uniformly distributed in the interval.  The
1696: advancement is therefore averaged over $r$ values in the interval.
1697: Because the advancement is linear in $r$, this implies that the
1698: average value $(x_1+x_2)/2$ is substituted for $r$, giving the average
1699: advancement $\langle r'-r\rangle =\frac{1}{3}(x_2-x_1)=\frac{1}{3}l$,
1700: where $l$ is the map size.
1701: 
1702: Suppose that all maps were the same size $l$ and denote the frequency
1703: of maps containing a given location as $\phi$.  Then the average rate
1704: $v_{\rm turb}$ of advancement of $r$ is $\phi$ times the average
1705: advancement of $r$ per map, giving $v_{\rm turb}=\frac{1}{3}\phi l$.
1706: 
1707: $\phi$ can be expressed in terms of $D_{\rm turb}$.  From random walk
1708: theory, $D_{\rm turb}$ is one-half of the frequency of events that
1709: displace a point times the mean-square displacement per event.  The
1710: mean-square displacement induced by a size-$l$ triplet map is
1711: $\frac{4}{27}l^2$ cite{part5}, so $D_{\rm turb}=\frac{2}{27}\phi l^2$,
1712: giving $\phi = \frac{27}{2}D_{\rm turb}/l^2$ and thus $v_{\rm
1713:   turb}=\frac{9}{2}D_{\rm turb}/l$.  This illustrates the analysis but
1714: is not the case of interest because inertial-range turbulence is
1715: represented in LEM by a distribution of map sizes $l$.
1716: 
1717: As explained in \citet{Ker91}, in LEM the map size distribution is
1718: $f(l)=Al^{-8/3}$ for $l$ in the range $[\eta,L]$, where $A$ is a
1719: normalization factor.  The total frequency of maps of all sizes per
1720: unit domain length is denoted $\Lambda$.  The frequency of maps in the
1721: size range $[l,l+dl]$ that contain a given point is then $\Lambda l
1722: f(l) \, dl$, so $D_{\rm turb}=\frac{2}{27} \Lambda \int l^3 f(l)\, dl=
1723: \frac{1}{18} \Lambda A(L^{4/3}-\eta^{4/3})$.  For high Re, the $\eta$
1724: term is negligible and is dropped, giving $D_{\rm turb}= \frac{1}{18}
1725: \Lambda AL^{4/3}$.
1726: 
1727: To obtain $v_{\rm turb}$, the average advancement $\frac{1}{3}l$ for
1728: map size $l$ is multiplied by $\Lambda l f(l) \, dl$ and integrated
1729: over $l$ to obtain (ignoring the $\eta$ term) $v_{\rm turb} = \Lambda
1730: AL^{1/3}$.  In terms of $D_{\rm turb}$, the result $v_{\rm turb} = 18
1731: D_{\rm turb}/L$ is obtained.  The numerical factor is four times
1732: larger than if all maps were of size $L$.  A heuristic interpretation
1733: of this result is that the typical map size from a flame propagation
1734: viewpoint is $L/4$ when $f(l)$ is based on inertial-range scaling.
1735: 
1736: \end{appendices}
1737: \end{widetext}
1738: 
1739: \clearpage
1740: 
1741: \begin{thebibliography}{99}
1742: 
1743: \bibitem[Aspden et al.(2008)]{Asp08}
1744: Aspden, A. J., Bell, J. B., Day, M. S., Woosley, S. E., \& Zingale,
1745: M. 2008, submitted to \apj
1746: 
1747: \bibitem[Bell et al.(2004a)]{Bel04a} 
1748: Bell, J.~B., Day, M.~S., Rendleman, C.~A., Woosley, S.~E., \& Zingale,
1749: M.\ 2004, \apj, 606, 1029
1750: 
1751: \bibitem[Bell et al.(2004b)]{Bel04b} 
1752: Bell, J.~B., Day, M.~S., Rendleman, C.~A., Woosley, S.~E., \& Zingale,
1753: M.\ 2004, \apj, 608, 883
1754: 
1755: \bibitem[Bildsten \& Hall(2001)]{Bil01} 
1756: Bildsten, L., \& Hall, D.~M.\ 2001, \apjl, 549, L219
1757: 
1758: \bibitem[Celani et al.(2000)]{Cel00}
1759: Celani, A., Lanotte, A., Mazzino, A. \& Vergassola, M. 2000, \prl,
1760: 84, 2385
1761: 
1762: \bibitem[Damk\"ohler(1940)]{Dam40}
1763: Damk\"ohler, G. 1940, Z. Elektrchem, 46, 601
1764: 
1765: \bibitem[Dursi \& Timmes(2006)]{Dur06} 
1766: Dursi, L.~J., \& Timmes, F.~X.\ 2006, \apj, 641, 1071
1767: 
1768: \bibitem[Hansen et al.(1975)]{Han75} 
1769: Hansen, J.-P., McDonald, I.~R., \& Pollock, E.~L.\ 1975, \pra, 11,
1770: 1025
1771:  
1772: \bibitem[Hillebrandt \& Niemeyer(2000)]{Hil00} 
1773: Hillebrandt, W., \& Niemeyer J. 2000, ARAA, 38, 191b
1774:  
1775: \bibitem[Kerstein(1991)]{Ker91}
1776: Kerstein, A. 1991, J. Fluid Mech., 231, 361
1777: 
1778: \bibitem[Kerstein(1996)]{Ker96}
1779: Kerstein, A. 1996, Combust. Sci. Tech., 118, 189
1780: 
1781: \bibitem[Kerstein(2001)]{Ker01}
1782: Kerstein, A. 2001, Phys. Rev. E., 64, 066306
1783: 
1784: \bibitem[Khokhlov, Oran, \& Wheeler(1997)]{Kho97}
1785: Khokhlov, A., Oran, E. S., \& Wheeler, J. C. 1997, \apj, 478, 678
1786: 
1787: \bibitem[Kuhlen et al.(2006)]{Kuh06} 
1788: Kuhlen, M., Woosley, S.~E., \& Glatzmaier, G.~A.\ 2006, \apj, 640, 407
1789: 
1790: \bibitem[Landau \& Lifshitz(1959)]{Lan59} 
1791: Landau, L., \& Lifshitz, F. M. 1959, Course in Theoretical Physics,
1792: Vol. 6.  Fluid Mechanics (Oxford:Pergammon)
1793: 
1794: %\bibitem[Lisewski et al.(2000)]{Lis00a} 
1795: %Lisewski, A.~M., Hillebrandt, W., Woosley, S.~E., Niemeyer, J.~C., \&
1796: %Kerstein, A.~R.\ 2000, \apj, 537, 405
1797: 
1798: \bibitem[Lisewski et al.(2000)]{Lis00b} 
1799: Lisewski, A.~M., Hillebrandt, W., \& Woosley, S.~E.\ 2000, \apj, 538,
1800: 831
1801: 
1802: \bibitem[Moisey et al.(2001)]{Moi01}
1803: Moisy, F., Willaime, H., Andersen, J. S., \& Tabeling, P. 2001, \prl,
1804: 86, 4827
1805: 
1806: \bibitem[Nandkumar \& Pethick(1984)] {Nan84}
1807: Nandkumar, R., \& Pethick, C.~J.\ 1984, \mnras, 209, 511
1808: 
1809: \bibitem[Niemeyer(1995)]{Nie95a} 
1810: Niemeyer, J.~C.\ 1995, PhD Thesis, MPA, Munich, Germany: Technical University
1811: 
1812: \bibitem[Niemeyer \& Hillebrandt(1995)]{Nie95b} 
1813: Niemeyer, J.~C., \& Hillebrandt, W.\ 1995, \apj, 452, 779
1814: 
1815: \bibitem[Niemeyer \& Woosley(1997)]{Nie97} 
1816: Niemeyer, J. C., \& Woosley, S. E. 1997, \apj, 475, 740
1817: 
1818: \bibitem[Niemeyer \& Kerstein(1997)]{Nie97a} 
1819: Niemeyer, J.~C., \& Kerstein, A.~R.\ 1997, New Astronomy, 2, 239
1820: 
1821: %\bibitem[Niemeyer \& Kerstein(1997)]{Nie97b} 
1822: %Niemeyer, J.~C., \& Kerstein, A.~R.\ 1997, Comb. Sci.
1823: %Tech., 128, 343
1824: 
1825: \bibitem[Pan et al.(2008)]{Pan08} 
1826: Pan, L., Wheeler, J.~C., \& Scalo, J.\ 2008, \apj, 681, 470
1827: 
1828: \bibitem[Peters(1986)]{Pet86}
1829: Peters, N. 1986, Proc. Combust. Inst., 21, 1231
1830: 
1831: \bibitem[Peters(2000)]{Pet00} 
1832: Peters, N.\ 2000, Turbulent Combustion, by Norbert Peters,
1833: pp.~320.~ISBN 0521660823.~Cambridge, UK: Cambridge University Press,
1834: August, 2000
1835: 
1836: \bibitem[R{\"o}pke et al.(2004)]{Roe04} 
1837: R{\"o}pke, F.~K., Hillebrandt, W., \& Niemeyer, J.~C.\ 2004, \aap,
1838: 420, 411
1839: 
1840: \bibitem[R\"opke(2007)]{Roe07a}
1841: R\"opke. F. 2007, \apj, 668, 1103
1842: 
1843: \bibitem[Schmidt et al.(2006a)]{Sch06a} 
1844: Schmidt, W., Niemeyer, J.~C., \& Hillebrandt, W.\ 2006, \aap, 450, 265
1845: 
1846: \bibitem[Schmidt et al.(2006b)]{Sch06b} 
1847: Schmidt, W., Niemeyer, J.~C., Hillebrandt, W., R\"opke, F.~K.\
1848: 2006, \aap, 450, 283
1849: 
1850: \bibitem[Smith \& Menon(1997)]{Smi97} 
1851: Smith, T. M. \& Menon, S. 1997, Combust. Sci. Tech., 128, 99
1852: 
1853: \bibitem[Timmes \& Woosley(1992)]{Tim92} 
1854: Timmes, F.~X., \& Woosley, S.~E.\ 1992, \apj, 396, 649
1855: 
1856: \bibitem[Timmes(2000)]{Tim00a} 
1857: Timmes, F.~X.\ 2000, \apj, 528, 913 see also
1858: http://cococubed.asu.edu/code\_pages/eos.shtml
1859: 
1860: \bibitem[Timmes \& Swesty(2000)]{Tim00b} 
1861: Timmes, F.~X., \& Swesty, F.~D.\ 2000, \apjs, 126, 501, see also 
1862: http://cococubed.asu.edu/code\_pages/eos.shtml
1863: 
1864: \bibitem[Timmes et al.(2000)]{Tim00c} 
1865: Timmes, F.~X., Hoffman, R.~D., \& Woosley, S.~E.\ 2000, \apjs, 129,
1866: 377
1867: 
1868: \bibitem[Umeda et al.(1999a)]{Ume99a} 
1869: Umeda, H., Nomoto, K., Kobayashi, C., Hachisu, I., \& Kato, M.\ 1999a,
1870: \apjl, 522, L43
1871: 
1872: \bibitem[Umeda et al.(1999b)]{Ume99b} 
1873: Umeda, H., Nomoto, K., Yamaoka, H., \& Wanajo, S.\ 1999b, \apj, 513,
1874: 861
1875: 
1876: \bibitem[Watanabe \& Gotoh(2006)]{Wat06}
1877: Watanabe, T., \& Gotoh, T. 2006 Phys. Fluids, 18, 058105
1878: 
1879: \bibitem[Weaver, Zimmerman, \& Woosley(1978)]{Wea78} 
1880: Weaver, T. A., Zimmerman, G. B., \& Woosley, S. E. 1978, \apj, 225, 1021
1881: 
1882: \bibitem[Woosley et al.(2002)]{Woo02} 
1883: Woosley, S.~E., Heger, A., \& Weaver, T.~A.\ 2002, Reviews of Modern
1884: Physics, 74, 1015
1885: 
1886: \bibitem[Woosley et al.(2004)]{Woo04} 
1887: Woosley, S.~E., Wunsch, S., \& Kuhlen, M.\ 2004, \apj, 607, 921
1888: 
1889: \bibitem[Woosley(2007)]{Woo07} 
1890: Woosley, S.~E.\ 2007, \apj, 668, 1109
1891: 
1892: \bibitem[Zeldovich et al.(1985)]{Zel85}
1893: Zeldovich, Ya. B., Barenblatt, G. I., Librovich, V. B., \&
1894: Makhviladze, G. M. 1985, {\sl The Mathematical Theory of Combustion
1895:   and Explosions}, Consultants Burea, Plenum Press.
1896: 
1897: \end{thebibliography}
1898: 
1899: \clearpage
1900: 
1901: \begin{deluxetable}{cccccc} 
1902: \tablecaption{Properties of Laminar Flames}
1903: \tablehead{
1904: X$_{12}$ & $\rho$ & $S_{\rm lam}$ &$\delta_{\rm lam}(\epsilon)$ & $\delta_{\rm lam}(T)$ & $\delta_{\rm lam}(C)$ \\
1905:         & ($10^7$ g cm$^{-3}$)  & (cm/s) &  (cm)  &  (cm)       &   (cm) }
1906: \startdata
1907: 0.50 &  0.6  & 8.62(2) & 11.7   & 15.6  & 19.3   \\
1908: 0.50 &  0.8  & 1.77(3) &  4.35  & 6.01  & 7.03   \\
1909: 0.50 &  1.0  & 3.23(3) &  2.08  & 3.12  & 3.32   \\
1910: 0.50 &  1.2  & 4.99(3) &  1.15  & 1.78  & 1.82   \\
1911: 0.50 &  1.4  & 7.19(3) &  0.70  & 1.14  & 1.10   \\
1912: 0.50 &  1.6  & 9.66(3) &  0.45  & 0.76  & 0.72   \\
1913: 0.50 &  1.8  & 1.26(4) &  0.31  & 0.54  & 0.50   \\
1914: 0.50 &  2.0  & 1.58(4) &  0.22  & 0.40  & 0.36   \\        
1915: 0.50 &  2.5  & 2.47(4) &  0.12  & 0.23  & 0.19   \\
1916: 0.50 &  3.0  & 3.54(4) &  0.067 & 0.14  & 0.11   \\  
1917: 0.50 &  3.5  & 4.66(4) &  0.041 & 0.090 & 0.065  \\
1918:      &       &         &        &       &        \\
1919: 0.75 &  0.3  & 1.73(2) &   172  & 157   & 325    \\
1920: 0.75 &  0.4  & 8.65(2) &  21.4  & 22.4  & 40.0   \\
1921: 0.75 &  0.5  & 1.74(3) &  9.80  & 11.4  & 17.7   \\
1922: 0.75 &  0.6  & 2.80(3) &  5.02  & 6.12  & 8.94   \\
1923: 0.75 &  0.7  & 4.10(3) &  2.94  & 3.52  & 5.29   \\
1924: 0.75 &  0.8  & 5.86(3) &  1.80  & 2.27  & 3.21   \\
1925: 0.75 &  0.9  & 8.39(3) &  1.25  & 1.73  & 2.20   \\
1926: 0.75 &  1.0  & 1.08(4) &  0.85  & 1.21  & 1.49   \\
1927: 0.75 &  1.2  & 1.70(4) &  0.43  & 0.66  & 0.77   \\
1928: 0.75 &  1.4  & 2.37(4) &  0.25  & 0.40  & 0.45  
1929: \enddata
1930: \end{deluxetable}
1931: 
1932: \clearpage
1933: 
1934: 
1935: \begin{deluxetable}{ccccccccc} 
1936: \tablecaption{Flame Properties at $\rho = 1.0 \times 10^7$ g cm$^{-3}$}
1937: \tablehead{
1938: X$_{12}$ & $l$   & $U_L$   & Zones & $\Delta x$  & Transport & $v_f$ & 
1939: $\delta_f(\epsilon)$  & $\delta_f(T)$ \\
1940:         & (cm)  & (km/s) &       &  (cm)       &           &  (km/s) & cm & cm }
1941: \startdata
1942: 0.50 &  -      &   0   &  2048  & 0.0244 & rad   & 0.0323  &  2.1    & 3.1 \\
1943: 0.50 &  -      &   0   &   256  & 0.196  & rad   & 0.0314  &  2.1    & 3.1 \\
1944: 0.50 & 15      & 2.47  &  2048  & 0.244  & rad.  & 0.149   &  40     & 70  \\
1945: 0.50 & 15      & 2.47  &  2048  & 0.244  & SG    & 0.156   &  40     & 70  \\
1946: 0.50 & 120     & 4.93  &  2048  & 0.977  & SG    & 0.61    &  160    & 300 \\
1947: 0.50 & 960     & 9.86  &  2048  & 9.77   & SG    & 2.3     & $\sim$700 &
1948:  $\sim$1200 \\
1949: 0.50 & 960     & 9.86  & 16384  & 1.22   & SG    & 2.3     & $\sim$700 & 
1950:  $\sim$1200 \\
1951: 0.50 & 7680    & 19.7  &  8192  & 9.77   & SG    & 8.2     &  fluc. & fluc. \\
1952: 0.50 & 6.14(4) & 39.4  & 32768  & 9.16   & SG    & 26      &  fluc. & fluc.\\
1953: 0.50 & 4.92(5) & 78.9  & 65536  & 49.9   & SG    & 110     &  fluc. & fluc.\\
1954: 0.50 & 3.93(6) & 158   & 65536  & 360    & SG    & 280     &  fluc. & fluc.\\
1955: 0.50 & 3.93(6) & 340   & 65536  & 360    & SG    & 440 - 470? & fluc. & fluc.\\
1956:      &         &       &        &        &       &         &          &    \\
1957: 0.75 &  -      &  0    &  2048  & 0.0391 & rad   & 0.113   &  0.85 & 1.2 \\
1958: 0.75 & 10      & 2.16  &  2048  & 0.0977 & rad   & 0.415   &  10   & 20   \\
1959: 0.75 & 40      & 3.42  &  4096  & 0.0977 & rad   & 0.922   &  $\sim$25 & 
1960:  $\sim$60\\
1961: 0.75 & 80      & 4.31  &  8192  & 0.0977 & rad   & 1.46    &  $\sim$40 & 
1962:  $\sim$90 \\
1963: 0.75 & 80      & 4.31  &  8192  & 0.0977 & SG    & 1.54    &  $\sim$40 &
1964:  $\sim$90 \\
1965: 0.75 & 320     & 6.84  & 32768  & 0.0977 & SG    & 3.2     &  fluc.  & fluc.\\
1966: 0.75 & 2560    & 13.7  & 32768  & 0.391  & SG    & 12      &  fluc.  & fluc.\\
1967: 0.75 & 2.05(4) & 27.4  & 32768  & 3.05   & SG    & 40      &  fluc.  & fluc.\\
1968: 0.75 & 1.00(4) & 215   & 32768  & 2.50   & SG    & 96      &  fluc. & fluc. \\
1969: 0.75 & 1.00(6) & 500   & 32768  & 152    & SG    & 1020    &  fluc. & fluc. \\
1970: 0.75 & 1.00(6) & 500   & 65536  & 76.3   & SG    & 870     &  fluc. & fluc. \\
1971: \enddata
1972: \end{deluxetable}
1973: 
1974: \clearpage
1975: 
1976: \begin{deluxetable}{cccccc} 
1977: \tablecaption{Characteristic Scales in the WSR and ST Regimes for $U_L = 100$ km s$^{-1}$ and L = 10 km}
1978: \tablehead{
1979: X$_{12}$ & $\rho$ & $\tau_{\rm nuc}$ &$\lambda$ & $d$ & Ka \\
1980:         & ($10^7$ g cm$^{-3}$)  & (sec) &  (cm)  &  (cm)  &   }
1981: \startdata
1982: 0.50 & 0.6  & 6.1(-2) &  4.7(5) & 1.0(2) & 3.5(3)  \\
1983: 0.50 & 0.8  & 1.2(-2) &  4.0(4) & 1.8(2) & 8.9(2)  \\
1984: 0.50 & 1.0  & 3.5(-3) &  6.4(3) & 3.2(2) & 2.5(2)  \\
1985: 0.50 & 1.2  & 1.3(-3) &  1.5(3) & 5.0(2) & 9.6(1)  \\
1986: 0.50 & 1.4  & 5.8(-4) &  4.5(2) & 7.2(2) & 4.3(1)  \\
1987: 0.50 & 1.6  & 2.9(-4) &  1.6(2) & 9.7(2) & 2.2(1)  \\
1988: 0.50 & 1.8  & 1.6(-4) &  6.4(1) & 1.3(3) & 1.3(1)  \\
1989: 0.50 & 2.0  & 9.5(-5) &  2.9(1) & 1.6(3) & 7.5(0)  \\        
1990: 0.50 & 2.5  & 3.2(-5) &  5.7(0) & 2.5(3) & 2.8(0)  \\
1991: 0.50 & 3.0  & 1.3(-5) &  1.5(0) & 3.5(3) & 1.2(0)  \\ 
1992: 0.50 & 3.5  & 6.4(-6) &  5.1(-1)& 4.7(3) & 6.4(-1) \\              \\
1993: 0.75 & 0.3  & 3.6(-1) &  6.8(6) & 1.7(1) & 1.8(5)  \\  
1994: 0.75 & 0.4  & 6.0(-2) &  4.6(5) & 8.7(1) & 5.8(3)  \\
1995: 0.75 & 0.5  & 1.6(-2) &  6.4(4) & 1.7(2) & 1.4(3)  \\
1996: 0.75 & 0.6  & 5.3(-3) &  1.2(4) & 2.8(2) & 4.8(2)  \\
1997: 0.75 & 0.7  & 2.1(-3) &  3.0(3) & 4.1(2) & 2.1(2)  \\
1998: 0.75 & 0.8  & 1.0(-3) &  9.9(2) & 5.9(2) & 9.5(1)  \\
1999: 0.75 & 0.9  & 5.0(-4) &  3.5(2) & 8.4(2) & 4.6(1)  \\
2000: 0.75 & 1.0  & 2.8(-4) &  1.4(2) & 1.1(3) & 2.6(1)  \\
2001: 0.75 & 1.2  & 1.0(-4) &  3.3(1) & 1.7(3) & 9.4(0)  \\
2002: 0.75 & 1.4  & 4.5(-5) &  9.6(0) & 2.4(3) & 4.3(0)
2003: \enddata
2004: \end{deluxetable}
2005: 
2006: 
2007: \end{document}
2008: