1: %\documentstyle[]{mn}
2: \documentstyle[referee]{mn}
3: \def\baselinestretch{2}
4: \input{psfig.sty}
5: \def \ny{\~{n}}
6: \def \no{\noindent}
7: \def \be{\begin{equation}}
8: \def \ee{\end{equation}}
9: \def \bea{\begin{eqnarray}}
10: \def \eea{\end{eqnarray}}
11: \def \h{\hfil}
12:
13: \begin{document}
14:
15: \title[Hyperbolic character of the angular moment equations]
16: {Hyperbolic character of the angular moment equations of radiative
17: transfer and numerical methods.}
18:
19: \author[J.A. Pons, J. M$^{\underline{\mbox{a}}}$ Ib\'a\~nez
20: and J. A. Miralles]{J. A. Pons$^{1,2}$, J. M$^{\underline{\mbox{a}}}$
21: Ib\'a\~nez$^1$ and J. A. Miralles$^1$ \\
22: $^1$ Departament d'Astronomia i Astrof\'{\i}sica,
23: Universitat de Val\`encia,
24: 46100 Burjassot, (Val\`encia), Spain \\
25: $^2$Department of Physics \& Astronomy, SUNY at Stony Brook,
26: Stony Brook, New York 11794-3800 }
27:
28: \maketitle
29:
30: \begin{abstract}
31:
32: We study the mathematical character of the angular moment equations
33: of radiative transfer in
34: spherical symmetry and conclude that the system is hyperbolic
35: for general forms of the closure relation found in the literature.
36: Hyperbolicity and causality
37: preservation lead to mathematical conditions allowing to
38: establish a useful characterization of the closure relations.
39: We apply numerical methods specifically designed to solve hyperbolic
40: systems of conservation laws (the so-called Godunov-type methods),
41: to calculate numerical solutions of the radiation transport
42: equations in a static background. The feasibility of the method in any
43: kind of regime, from diffusion to free-streaming, is demonstrated by a
44: number of numerical tests and the effect of the choice of the closure
45: relation on the results is discussed.
46:
47: \end{abstract}
48:
49: \begin{keywords}
50: radiative transfer -- methods: numerical
51: \end{keywords}
52:
53: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
54: \section{Introduction}
55:
56: In standard problems of radiation\footnote{ We will use the term
57: {\it radiation} for both photons and neutrinos} hydrodynamics (RH)
58: where radiation contains a large fraction of the energy and momentum density,
59: the Boltzmann Equation (BE) must be coupled to the hydrodynamic
60: equations in order to obtain the evolution of the system as well as the
61: spectrum and angular distribution of the radiation field.
62: However, an algorithm built to solve the BE
63: numerically (a Boltzmann solver) in a non-stationary case is too
64: time-consuming, from a computational point of view, to allow for the
65: extension to more than one dimension of the existing numerical codes
66: \cite{MB93,Ya99}. In many cases, instead of the BE, its {\it angular
67: moments} are considered, obtaining then the multigroup (or
68: multi-frequency) equations or the even simpler energy averaged equations.
69:
70: Standard RH methods \cite{MM84} have found their
71: way into the literature, and one can find several RH codes with
72: different approaches for the radiation transport part,
73: from single-energy-group
74: (or two temperature) such as VISPHOT \cite{EB92} or TITAN \cite{GM94},
75: to multigroup radiative transfer, such as STELLA \cite{BB93}.
76: Other authors built codes devoted more specifically to the radiation transport,
77: but at the expense of detailed hydrodynamics. An example is the
78: code EDDINGTON \cite{EP93} in which free expansion is assumed.
79: The hydrodynamics part of most of the existing codes is a one dimensional
80: implicit finite difference scheme, including artificial viscosity terms
81: for problems requiring an accurate treatment of shock waves and discontinuities.
82: During the last decade, a new subclass of numerical methods, the so-called
83: Godunov-type methods, has been gradually substituting the schemes based
84: on numerical viscosity due to their easier extension to multidimensional
85: cases and their greater capabilities in the treatment of strong shocks.
86: The lack of a radiation transport method fully compatible with hydrodynamical
87: Godunov-type schemes is one of the motivations of these paper.
88:
89: Turning back to the radiation transport part, another important
90: issue related to the angular moment equations is the closure relation.
91: By taking angular moments of the distribution
92: function the complexity of BE is highly reduced, but the information about
93: the angular dependence of the radiation field is partially lost. Each
94: $n^{th}$ angular moment equation of the BE contains angular moments of
95: higher order, thus any truncated hierarchy of moment equations
96: contains more unknowns than equations and must be supplied with
97: additional equations or {\it closure relations} \cite{CB94,Gro96}.
98: Theoretically, if the
99: correct closure for a given problem is known, the solution obtained for
100: the first moments of the distribution function by solving the moment
101: equations should be the same as the solution obtained by solving the BE.
102:
103: Although the moment equations are much simpler to solve than the BE,
104: in many situations an additional simplification is made by neglecting
105: some other terms,
106: leading to the {\it diffusion approximation} (DA). In this approximation,
107: however, the resulting energy flux can be higher than the limit predicted
108: by causality, specially in the regions where the radiation mean free path
109: becomes comparable to the characteristic length. Different extensions of DA,
110: like {\it flux-limited diffusion}
111: \cite{BW82} or {\it artificial opacities} \cite{DJ92} have being used to
112: overcome this problem. These extensions
113: only partially solve the breakdown of the DA in the
114: semi-transparent and transparent layers, and all of them are based on the
115: same idea: to reintroduce some terms which had been dropped in the original
116: assumptions of the DA. The term which is usually kept out of the equations
117: is the time derivative of the fluxes. By neglecting this term, the character of
118: the system of equations is changed to parabolic, and causality is therefore
119: violated (disturbances propagate at infinite velocity). In order
120: to develop a method which is valid in all regimes
121: (from diffusion to free streaming) while preserving causality, one must solve
122: the full set of equations keeping its hyperbolic character.
123:
124: In this paper, we address the problem of establishing a well-defined
125: hyperbolic system of equations for the first angular moments of the BE in
126: the non-stationary spherically symmetric case. We will see that, besides
127: the important question of its hyperbolicity, some constraints on the
128: mathematical properties of the closure relations can be derived. They might
129: help to disregard among some of the closures previously proposed. Moreover,
130: the behaviour of the characteristic fields of the hyperbolic system,
131: which gives information on the velocity of perturbations, can also be used to
132: establish additional constraints to the closures.
133:
134: A direct consequence of considering the moment equations in
135: hyperbolic form is that it permits the application of
136: powerful numerical techniques that have been developed
137: in recent years for hyperbolic systems of conservation laws (the equations of
138: classical and relativistic fluid dynamics for perfect fluids, for example).
139: Among the different numerical techniques, the so-called
140: high resolution shock capturing (HRSC) methods
141: have a number of interesting features such that stability,
142: being conservative, convergence to physical solutions and
143: high accuracy in regions where the solution is smooth.
144: HRSC methods are based on the resolution of local Riemann problems
145: (an initial value problem with discontinuous data) at
146: the interfaces of numerical shells,
147: ensuring a consistent treatment of discontinuities and steep
148: gradients \cite{Go59}.
149: Their special relativistic extension \cite{MIM91}
150: has shown its potential in simulations of heavy ion
151: collisions and extra-galactic jets, and different attempts
152: to extend the method to General Relativity have been done
153: \cite{BFIMM97,PFIMM98}.
154: We refer the interested reader to the recent reviews by
155: Mart\'{\i} \& M\"uller \shortcite{MM99}
156: and Ib\'a\~nez \& Mart\'{\i} \shortcite{IM99}
157: for a detailed description of the current
158: status of HRSC techniques in numerical relativistic hydrodynamics.
159:
160: Although HRSC are specially designed for hyperbolic systems without
161: source terms which corresponds to the transparent regime, we will show in
162: numerical experiments that, by appropriate
163: treatment of the sources, accurate solutions can be obtained also in the
164: diffusion regime, where the source terms are dominant.
165:
166:
167: The structure of the paper is the following:
168: In \S 2 we summarise the deduction of the angular moment equations
169: of the BE in a spherically symmetric case, and for a static
170: background. We also describe the general form of the collision
171: terms when emission-absorption and iso-energetic scattering processes
172: are included. In \S 3 we discuss the most common techniques used
173: to solve the transport equations, such as diffusion or flux-limited
174: diffusion, remarking their main features and limitations.
175: In \S 4 the hyperbolic character of the equations and its implications
176: for the closure relations are discussed.
177: In \S5, the numerical techniques employed to solve the transport
178: equations as a hyperbolic system of conservation laws are discussed.
179: A number of numerical experiments solving the transport equations
180: in several test problems are displayed in \S 6,
181: and the feasibility of the
182: hyperbolic treatment in all kinds of regimes is demonstrated.
183: Finally, main conclusions
184: and advantages of our proposal are summarised in \S 7.
185:
186:
187: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
188:
189: \section{Transport equations in spherical coordinates}
190:
191: We shall start our discussion from the radiative transfer
192: equations in a static medium, deferring to a future work the
193: inclusion of fluid velocity
194: terms in the equations. Although the restriction to zero velocity
195: seems to be specific, there are some astrophysical scenarios where
196: the assumption that matter is at rest is a reasonable approach.
197: Moreover, the inclusion of velocity terms, in some cases of interest
198: does not change the essentials of our conclusions.
199:
200: In a static background, the Boltzmann Equation for massless particles in
201: spherical coordinates can be written as follows\footnote{We work in units
202: where $c=\hbar=1$.}
203:
204: \begin{equation}
205: \label{BE}
206: \omega \left[ \frac{\partial {\cal I}}{\partial t} +
207: \mu \frac{\partial {\cal I}}{\partial r} +
208: \frac{(1- {\mu}^2)}{r} \frac{\partial {\cal I}}{\partial \mu} \right]
209: = \left( \frac {d{\cal I}}{d \tau} \right)_{coll}
210: \end{equation}
211: \noindent where ${\cal I}={\cal I}(t,r,\omega,\mu)$ is the invariant
212: distribution function,
213: $\mu$ is the cosine of the angle of the particle momentum with respect
214: to the radial direction, $\omega$ the energy of the particle and the right
215: hand side term is the invariant source or collision term coming from the interaction
216: between the radiation field and matter.
217:
218: As stated in the introduction, a common method of solving the BE is the
219: method of moments
220: \cite{Tho81}, which involves
221: taking angular moments of the equation by applying the operator
222:
223: \begin{equation}
224: \frac{1}{4 \pi} \int_{-1}^{+1} \int_{0}^{2 \pi}
225: {\mu}^{i}~d{\mu}~d\Phi \,\,\,\,\,\,\, i=0,1,2,...
226: \end{equation}
227:
228: Denoting by $E$, $F$ and $P$ the angular moments of
229: the specific intensity $g \omega^3 {\cal I}$, $g$ being the
230: statistical weight ($g=2$ for photons and $g=1$ for neutrinos)
231: \begin{eqnarray}
232: \label{defp}
233: E=E(t,r,\omega)=
234: g \frac{\omega^3}{2}\int_{-1}^{+1} d\mu ~{\cal I},
235: \nonumber \\
236: F=F(t,r,\omega)=
237: g \frac{\omega^3}{2}\int_{-1}^{+1} d\mu~\mu ~{\cal I},
238: \nonumber \\
239: P=P(t,r,\omega)=
240: g \frac{\omega^3}{2}\int_{-1}^{+1} d\mu~\mu^2 ~{\cal I},
241: \end{eqnarray}
242: the equations
243: corresponding to the first two moments (0 and 1) of the BE are:
244:
245: \begin{eqnarray}
246: \label{mom0}
247: \partial_t E + \partial_r F + \frac{2 F}{r} = s^0
248: \end{eqnarray}
249: \begin{eqnarray}
250: \label{mom1}
251: \partial_t F + \partial_r P + \frac{ 3P-E}{r} = s^1
252: \end{eqnarray}
253: where the moments of the collision term are defined as
254: \begin{equation}
255: \label{q0}
256: s^0 = g \frac{{\omega}^2 }{2} \int_{-1}^{+1}
257: {\left(\frac {d{\cal I}}{d \tau}\right)}_{coll} d{\mu}
258: \end{equation}
259: \begin{equation}
260: \label{q1}
261: s^1 = g \frac{{\omega}^2 }{2} \int_{-1}^{+1}
262: {\left(\frac {d{\cal I}}{d \tau}\right)}_{coll} \mu d{\mu}
263: \end{equation}
264:
265: Notice that, in the absence of velocity fields or strong gravitational
266: fields, the different energy groups in a multifrequancy scheme are
267: only coupled through the source terms, rendering the mathematical
268: character of the left hand side of equations (\ref{mom0})-(\ref{mom1})
269: formally identical to the energy-averaged problem.
270:
271: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
272:
273: \subsection{Collision terms}
274:
275: The most general form for the collision term including emission, absorption
276: and isoenergetic scattering is the following:
277:
278: \begin{equation}
279: {\left(\frac {d{\cal I}}{d \tau}\right)}_{coll} =
280: \omega\left(j - \kappa_a {\cal I} + \kappa_s[{\cal I}]\right)
281: \end{equation}
282: where $j$ is the emissivity, $\kappa_a$ the absorption
283: opacity including final states blocking (for fermions) or
284: stimulated emission (for bosons) and $\kappa_s$ the scattering opacity related
285: to the scattering
286: reaction rate ($R^s$) through
287: \begin{equation}
288: \kappa_s(\omega) = { \left( \frac{\omega}{2 \pi} \right)}^3 \int_{-1}^{+1}
289: d \mu' \int_0^{2 \pi} d \varphi \, R^s(\omega,\cos \Theta) \,
290: ({\cal I}(\omega, \mu')- {\cal I}(\omega, \mu))
291: \end{equation}
292: where $\Theta$ is the angle between the in-going and outgoing particle
293: and $\varphi$ is the azimuthal angle. The opacities and the emissivity
294: are expressed in units of inverse length.
295:
296: For isotropic scattering the source terms appearing in equations
297: (\ref{mom0}) and (\ref{mom1}) can be written as
298: \begin{equation}
299: s^0 = \kappa_a \left( E^{eq}-E \right)
300: \label{S0}
301: \end{equation}
302: \begin{equation}
303: s^1 = - \kappa F
304: \label{S1}
305: \end{equation}
306: where $ \kappa = \kappa_a + \kappa_s $
307: and $E^{eq}$ is the value of $E$ in equilibrium with matter.
308: In the above, we have assumed matter in local thermodynamic equilibrium (LTE),
309: and only the radiation field (photons or neutrinos) is allowed to deviate
310: from equilibrium.
311: The conclusions of our study might be affected
312: by large velocity fields or for frequencies close to line
313: discontinuities, which are present in some astrophysical scenarios, and
314: deserve a more careful study \cite{MSKH76,Ku83,MK86}.
315:
316:
317: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
318:
319: \section{Flux-limited diffusion and artificial opacities}
320: One of the approaches most widely used to solve the transport
321: equations numerically is the {\it diffusion approximation}, in which
322: the invariant distribution function is assumed to be nearly isotropic
323: in the comoving frame.
324: Thus an expansion in terms of Legendre polynomials
325: to the order $O(\mu^2)$ is enough to maintain the main features of the
326: radiation field
327: \begin{equation}
328: {\cal I}(\mu)= {\cal I}_0 + 3~{\cal I}_1~\mu
329: \end{equation}
330: where ${\cal I}_1 \ll {\cal I}_0$. Consistent with this
331: assumption, the time derivative of the flux is neglected in equation
332: (\ref{mom1}), and together with equation (\ref{q1})
333: gives the following relation
334: for the flux in terms of the energy gradient
335: \begin{equation}
336: \label{f1}
337: F(\omega) = - \frac{1}{3 \kappa(\omega)}
338: \frac{\partial E(\omega)}{\partial r}
339: \end{equation}
340:
341: As stated before, the DA breaks down when the mean free path is large
342: compared to the typical scale of the problem, and the fluxes calculated
343: with the former formula may give non-causal values, in the sense that
344: the flux can be greater than the energy density.
345: This pathology comes from the fact that we have
346: neglected some terms in the momentum equation in order to obtain a simple
347: formula for the fluxes.
348:
349: Flux limiters have been introduced {\it ad hoc} in
350: the literature to avoid this non-causal behaviour; see, {\it e.g.}, Minerbo
351: \shortcite{Mi78}, Levermore \& Pomraining \shortcite{LP81} or Cernohorsky
352: \& Bludman \shortcite{CB94}. In order to illustrate where the problems stem
353: from, we begin by deriving the {\it flux-limited diffusion}
354: equations.
355: Let us define the {\it flux factor}, $f$, and the {\it Eddington factor}, $p$:
356: \begin{equation}
357: f = \frac{F}{E}
358: \end{equation}
359: \begin{equation}
360: p = \frac{P}{E}
361: \end{equation}
362: By subtracting equation (\ref{mom0}) multiplied by $f$ from (\ref{mom1}), and
363: after some algebra (now keeping all terms), one obtains:
364: \begin{equation}
365: F = - \frac{(p - f^2)}
366: {(\kappa_{ph}+\kappa_{J})} \partial_r E
367: \label{Fdif}
368: \end{equation}
369: where the quantities, including the flux and Eddington factors, are now energy
370: dependent. The explicit expression for the different {\it opacities} are
371: \begin{equation}
372: \kappa_{ph} = \kappa_s
373: + \kappa_a \frac{E^{eq}}{E}
374: \end{equation}
375: \begin{eqnarray}
376: f \kappa_{J} = {\partial_t f} + {\partial_r p} - f \partial_r f
377: + \frac{(3p-1-2f^2)}{r}
378: \end{eqnarray}
379:
380: The {\it physical opacity}, $\kappa_{ph}$, or effective albedo only depends
381: on the interaction of radiation with matter and includes an
382: out--of--equilibrium correction to absorption opacity. The second
383: term in the denominator, $\kappa_J$, is the multigroup
384: extension of the so-called {\it artificial opacity} \cite{DJ92} and takes into
385: account geometrical corrections and deviations from the diffusion limit
386: (it vanishes in the limit $f \ll 1$).
387:
388: The term in the numerator of equation (\ref{Fdif}), $(p - f^2)$, plays the role
389: of a flux--limiter. It has the value $\frac{1}{3}$ in the diffusion
390: limit and goes to zero as the free streaming limit ($f \rightarrow 1$
391: , $p \rightarrow 1$) is reached.
392: To close the set of equations, a closure relation for $p$
393: is needed, and it is directly related to flux limiter $(p-f^2)$.
394: %Giving a closure is the price
395: %one must pay to avoid solving the BE, which is more
396: %difficult technically and much more expensive computationally.
397: We will re-address the closure relation problem in next section.
398:
399: With this description of the transport equations, the simple structure
400: of the diffusion equation and its parabolic character is maintained when the
401: first term in $\kappa_J$ is neglected ($\partial_t f$ = 0).
402: The geometric and velocity field corrections can also be included
403: through artificial opacities as corrections to the physical opacity.
404: Although the flux-limiter helps to avoid non-causal behaviour in the
405: sense $F>E$, causality is also violated when the time derivative of
406: $f$ is neglected, because the character of the equations change to
407: parabolic, which results in an infinite speed of propagation
408: of the information. By keeping this term, the diffusion scheme is, in some
409: applications, numerically unstable and further numerical tools are needed
410: to handle these instabilities. This is the subject of the next section.
411:
412: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
413:
414: \section{Hyperbolic Formulation of Transport Equations}
415:
416: The system of equations (\ref{mom0}) and (\ref{mom1}) can be
417: written in a compact form as follows
418: \begin{equation}
419: \label{system}
420: \frac{\partial \vec{\cal U}}{\partial t}
421: + \frac{1}{r^2} \frac{\partial \left( r^2 \vec{\cal F} \right)}
422: {\partial r} = \vec{\cal S}
423: \end{equation}
424: where the vectors $ \vec{\cal U}$ and $\vec{\cal F}$ are
425: \begin{equation}
426: \vec{\cal U} =(E,F)
427: \end{equation}
428: \begin{equation}
429: \vec{\cal F} = (F, pE)
430: \end{equation}
431: and the sources $\vec{\cal S}$, according to equations (\ref{S0}-\ref{S1}),
432: are given by
433: \begin{equation}
434: \vec{\cal S} = \left( \kappa_a (E^{eq}-E), -\kappa F + \frac{E-P}{r} \right).
435: \end{equation}
436: Notice that the sources are free of partial derivative operators.
437:
438: The above system of equations (\ref{system}) is said to be
439: a hyperbolic system of
440: conservation laws (with source terms) when the Jacobian matrix
441: associated to the {\it fluxes} ,$\nabla_{\vec{\cal U}} \vec{\cal F}$, has
442: real eigenvalues
443: and there exists a complete set of corresponding eigenvectors.
444: To analyse the Jacobian matrix, we will assume that the Eddington factor
445: is only a function of the flux factor $p=p(f)$. This assumption is commonly
446: used by a number of authors who have proposed different closure relations or
447: flux-limiters in the context of flux-limited diffusion.
448: In this case, the Jacobian matrix is
449: \be
450: \nabla_{\vec{\cal U}}\vec{\cal F}
451: = \left ( \matrix{
452: & 0 & 1 \cr
453: & p - f p' & p' \cr } \right )
454: \ee
455: where $p' = \displaystyle{\frac{d p}{d f}}$ .
456:
457: The eigenvalues of the above matrix are
458: \begin{equation}
459: \label{lambda}
460: \lambda_{\pm} = \frac{p' \pm \sqrt{ {p'}^2 + 4(p - f p') } }{2}
461: \end{equation}
462: and the right eigenvectors are
463: \begin{equation}
464: \label{vector}
465: \vec{r}_{+} = (1,\lambda_{+}) ; ~~~ \vec{r}_{-} = (1,\lambda_{-})
466: \end{equation}
467: The mathematical character of the system of equations
468: (\ref{system}) depends on the value of the discriminant
469: in equation (\ref{lambda}), which can be written as follows
470: \be
471: \label{disc}
472: \Delta = {({p'}-2f)}^2 + 4(p -f^2).
473: \ee
474: Since $p$ and $f$ are normalised moments of a non--negative weight
475: function, on the unit sphere, there is an extra relation between $f$
476: and $p$, the {\it Schwarz inequality}
477: \be
478: \label{schw}
479: f^2 \le p \le 1 ~~~,
480: \ee
481: that ensures that for
482: any closure satisfying (\ref{schw}) the eigenvalues
483: are always real and hence the hyperbolic character is guaranteed. In addition,
484: causality imposes that the velocity of propagation of perturbations must be
485: smaller than the speed of light, which gives another constraint on the
486: eigenvalues
487: \be
488: \left| \lambda_{\pm} \right| \le 1 ~,
489: \ee
490: that can be expressed in terms of $f$, $p$ and $p'$ as follows
491: \be
492: -\frac{1-p}{1+f} \le p' \le \frac{1-p}{1-f}.
493: \label{ppb}
494: \ee
495: This restriction, together with equation (\ref{schw}) and the
496: condition $|f| \le 1$,
497: defines the region in the $[f,p,p']$ space allowed for causal closures.
498:
499: In many astrophysical problems, radiation diffuses out from a
500: central opaque region to an outer optically thiner region. In such a
501: scenario, the radiation angular distribution is nearly isotropic in
502: the central region ($|f| \ll 1$) and forward peaked in the transparent
503: region, far from the opaque centre ($f \rightarrow 1$). Thus, the closure
504: in that problem must satisfy
505: \be
506: p(f=0)=\frac{1}{3} \quad {\rm and} \quad p(f=1) = 1.
507: \ee
508: The first limit comes from the diffusion approximation while the
509: second limit comes from the Schwarz inequality (\ref{schw}), and both
510: are satisfied by all the closures found in the literature (see Table I).
511: Another general property is that, in the diffusion regime
512: the Eddington factor has the constant value
513: of $1/3$, thus another restriction on the derivative can be imposed
514: \be
515: \label{cons2}
516: p'(f=0) = 0 ~,
517: \ee
518: that leads to the characteristic speeds
519: $\lambda_{\pm} = \pm \frac{1}{\sqrt{3}}$.
520: In the limit $f=1$, equations (\ref{schw}) and (\ref{ppb}) gives the bounds
521: $ 0 \le p' \le 2 $ and
522: the eigenvalues in this limit are $\lambda_{+}=1$ and $\lambda_{-}=p'-1$.
523: According to the theory of hyperbolic systems, this result can be
524: interpreted as a first wave propagating outwards
525: at the speed of light and a second wave with a characteristic speed that
526: depends on the value of $p'$.
527: Although $\lambda_{-}$ depends on the form of the closure, for $f=1$
528: the amplitude of this second wave is zero and the value of $p'$ does not
529: affect the evolution. However, for $f$ close but not equal to one,
530: the choice of $p'$ can be important, as we will show in section \S 5.2.
531: Similar conclusions can be obtained for the case $f=-1$.
532:
533: In table 1 we summarise the main properties of some of the most widely
534: used closures: Minerbo \shortcite{Mi78}, LP \cite{LP81},
535: the linear formula by Auer \shortcite{Auer84},
536: Kershaw's parabolic approach \cite{Ke76}, the flux-limiters used by Bruenn
537: \shortcite{Bru85} and Bowers \& Wilson \shortcite{BW82}, Janka's fit
538: to Montecarlo simulations of neutrino transport in Supernovae \cite{Jan91}
539: and that corresponding to an isotropically emitting sphere \cite{CVB86},
540: named {\it vacuum}.
541:
542: There are several comments about Table I that we would like to address.
543: First, one closure relation (BW) leads to
544: non-causal values of one eigenvalue in the diffusion regime.
545: Secondly, four closures do not satisfy the condition $p'(f=0)=0$
546: (Auer, Bruenn, BW, Vacuum).
547: Auer's formula failure is due to an excessive simplification (it is
548: a linear approach) while 'Vacuum' has been obtained from the assumption
549: of an isotropically radiating sphere, which restricts their validity
550: to $f \ge 0.5$.
551: The closure relations proposed by Kershaw, Janka, Minerbo and LP have the
552: common features of preserving causality leading to
553: reasonable values of the characteristic velocities
554: ($\lambda_{\pm} = \pm \frac{1}{\sqrt{3}}$) in the diffusion regime.
555: Thirdly, it is worthy to point out that all the closures
556: except Auer's give $p'\ge 1$
557: when approaching the radially-streaming limit $(f=1)$, which makes
558: both eigenvalues be positive. It ensures that no information can be
559: propagated inwards for sufficiently high values of $f$.
560: It is clear that these closures cannot be used to treat problems
561: in which a small perturbation is propagating inwards from the transparent
562: region, as we will show in next section.
563:
564: %-------------------------------------------------------------------
565:
566: In a more general situation, the Eddington factor depends on both,
567: the flux factor and the occupation density corresponding to the zeroth
568: angular moment of the distribution function,
569: $e=E/\omega^3 $.
570: The Jacobian matrix associated to the fluxes is
571:
572: $$\frac{\partial \vec{\cal F}}{\partial \vec{U}}
573: = \left ( \matrix{
574: & 0 & & 1 \cr
575: & & & \cr
576: & p - f {\displaystyle \left. \frac{\partial p}{\partial f} \right|_e}+
577: e {\displaystyle \left. \frac{\partial p}{\partial e} \right|_f}
578: & & {\displaystyle \left. \frac{\partial p}{\partial f} \right|_e } \cr
579: } \right) $$
580: and now the eigenvalues are
581: \begin{equation}
582: \lambda_{\pm} = \frac{1}{2} \left(
583: \left. {\displaystyle \frac{\partial p}{\partial f}} \right|_e
584: \pm \sqrt{\Delta} \right)
585: \end{equation}
586: where the discriminant is
587: \begin{equation}
588: \Delta = {\left( \left. \frac{\partial p}{\partial f} \right|_e
589: - 2~f \right)}^2 + 4 \left( p - f^2 +
590: e \left. \frac{\partial p}{\partial e} \right|_f \right)
591: \end{equation}
592:
593: In contrast to (\ref{disc}), the sign of the discriminant might be
594: negative for negative values of
595: $\left. \frac{\partial p}{\partial e} \right|_f$, and the system
596: would become elliptic instead of hyperbolic.
597: As stated before, hyperbolic
598: equations are related to finite speeds of propagation of the
599: perturbations, while elliptic equations are related with boundary
600: problems and do not allow for solutions
601: with propagating signals. This fact indicates that, in order to
602: preserve hyperbolicity, only closures which
603: give positive values of the discriminant should be considered.
604:
605: A closure of this kind has been proposed by
606: Cernohorsky \& Bludman \shortcite{CB94} using arguments of maximum
607: entropy. Minerbo's, LP's and the Vacuum closures can be obtained as
608: limiting cases of CB closure
609: in the low occupation limit (classical statistics), large ocuppation limit
610: for Bose-Einstein statistics and maximal forward packing limit for
611: Fermi-Dirac statistics, respectively. We have checked that CB closure
612: gives real and causal eigenvalues for both, Bose-Einstein and Fermi-Dirac
613: statistics, in any regime.
614:
615: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
616: \section{The numerical approach.}
617:
618: The numerical code employed for the calculations presented in
619: \S 6 is a finite finite difference scheme in which
620: the left (L) and right (R) states for the Riemann problems set
621: up at each interface have been obtained
622: with a monotonic, piecewise linear reconstruction procedure
623: \cite{vL79}. The numerical fluxes are evaluated following the idea
624: of Harten, Lax \& van Leer \shortcite{HLL83}, that is based on approximate
625: solution of the original Riemann problem with a single intermediate state.
626: The resulting numerical flux is given by
627: \be
628: \hat{\cal F}^{HLL} = \frac
629: {\psi^+ {\cal F}_L - \psi^- {\cal F}_R +
630: \psi^+ \psi^- ({\cal U}_R - {\cal U}_L)}{\psi^+ - \psi^-}
631: \ee
632: where $\psi^+ = \mbox{max}(0,\lambda_+^R, \lambda_+^L)$ and
633: $\psi^- = \mbox{min}(0,\lambda_-^R, \lambda_-^L)$.
634:
635: The time-advance algorithm is a TVD-preserving, third order Runge-Kutta
636: that takes into account the influence of the source terms in intermediate
637: steps. The source term in this problem deserves a special mention.
638: It acts as a relaxation term which
639: leads to a long-time behaviour governed by the reduced parabolic
640: system described in section \S 3. Although high resolution methods
641: for hyperbolic conservation laws usually fail to capture this feature
642: (unless a very fine grid is used),
643: the use of a modified flux \cite{JL96}
644: and an implicit linearization of the source term in each Runge-Kutta
645: step allows us to succeed in the treatment of such stiff relaxation term
646: with coarse grids and
647: time-steps longer than the source time-scale ($\approx 1/\kappa$),
648: as we will show in subsection \S 6.1.
649: Although we have constrained the numerical experiments to situations
650: when only scattering processes are permitted, the inclusion of
651: absorption-emission processes can be treated in a similar way. However
652: it introduces energy and momentum exchange with the matter and makes
653: necessary to solve the equations of radiation transport coupled with
654: the equations of hydrodynamics or the equations of hydrostatic equilibrium
655: with number and heat transfer.
656: Such a problem is out of the scope of this paper, in which we focussed
657: on the radiation part, and will be addressed in forthcoming works.
658:
659: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
660: \section{Numerical experiments and discussion.}
661:
662: To check the applicability of HRSC techniques to radiation
663: transport problems, we have performed a number of numerical tests
664: considering several simplified models in which the
665: analytical solution is known. In the remainder of this section
666: we describe the different tests and display the results.
667: Our intention is to prove that our method can solve
668: radiation transport in all kinds of regimes, from diffusion (TEST 1)
669: to radially streaming (TEST 2), going through the semi-transparent
670: region (TEST 5) and in situations where strong gradients are
671: developed (TEST 4).
672:
673: Ideally, in a multi-group scheme we have as many systems of two equations
674: as energy bins
675: in which the radiation density is splitted. In a radiation hydrodynamics
676: problem all those pairs of equations are coupled to the matter equations
677: and must be solved together. Since our objective is to study the
678: applicability of HRSC techniques for the radiation part, in what follows
679: we assume to have only one energy group or, equivalently, we consider
680: the energy integrated equations.
681: Notice, though, that the conclusions
682: are not affected since, when velocity fields are not taken
683: into account, the equations for different energies are
684: only coupled through the source terms.
685: A different problem will arise when velocity fields are important or
686: gravitational and Doppler red-shift are considered. In those situations
687: either the anisotropy of the source term (in the observer frame) or
688: the advection and aberration terms (in the comoving frame) cause
689: the energy redistribution along the photon (or neutrino)
690: path and a more careful treatment needs to be done.
691:
692: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
693: \subsection{TEST 1: Diffusion limit}
694:
695: In the diffusion limit, the closure is obviously $p=1/3$
696: and the time derivative of the flux can be neglected, leading to
697: equation (\ref{f1}).
698: Setting the absorption opacity to zero and
699: the scattering opacity to a constant value, the analytical solution
700: of the system formed by (\ref{mom0}) and (\ref{f1}), corresponding to
701: an initial value problem consisting
702: in a Dirac delta located in $r=0$ at $t=0$, is the following
703: \be
704: E(r,t) = {\left( \frac{\kappa}{t} \right)}^{3/2}
705: \exp \left\{ - \frac{3 \kappa r^2}{4t} \right\}
706: \ee
707: \be
708: F(r,t) = \frac{r}{2t} E(r,t)
709: \ee
710: We imposed $F(r=0)=0$ as inner boundary conditions and outflow
711: boundary conditions in the outermost shell.
712:
713: We have performed two simulations. First, we set a value of $\kappa=100$
714: on a equally spaced grid of 100 shells in the domain $[0,1]$. It
715: corresponds to a Peclet number ($Pe=\kappa \Delta x$) of order unity.
716: The cell {\it Peclet number} is a measure of the spatial resolution relative
717: to the relaxation scale and a grid is coarse when $Pe \gg 1$.
718: This first simulation starts at $t=1$ to avoid
719: the numerical problems produced by an infinite value in $t=0$.
720: Results are shown in Figure \ref{fig1a}.
721: The energy density (left panel) and the flux (right panel) are displayed for
722: different times in dimension-less units, $t=1,2,3,5$ from top to bottom.
723: The solid line corresponds to the analytical solution and the crosses
724: correspond to the numerical solution.
725: Notice that, although we are finding numerical solutions of
726: system (\ref{system}), the analytical solution of the diffusion limit
727: is well reproduced by the numerical code, since
728: the opacity is large enough to be close to the diffusion limit.
729: The second simulation has been performed in a grid of 100 shells in
730: the domain $[0,0.5]$, and using a value of $\kappa=10^5$, which
731: corresponds to $Pe=5 \times 10^4$. The initial time in this case is 200
732: and the Courant time-step is of the same order ($10^{-2}$) as in the previous
733: simulation, three orders of magnitude larger than the relaxation scale.
734: Results are displayed in Fig. \ref{fig1b}. The agreement between the
735: analytical and the numerical solutions clearly proofs the capability
736: of the code to handle situations with very stiff source terms.
737:
738: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
739: \subsection{TEST 2: Radially Streaming limit}
740:
741: In the radially streaming limit ($|f|=1$, $p=1$), and assuming there is
742: no interaction between radiation and matter ($\kappa_a=0$, $\kappa=0$)
743: equations (\ref{system}) are reduced to
744: \begin{eqnarray}
745: \frac{\partial E(\omega)}{\partial t} +
746: \frac{1}{r^2} \frac{\partial (r^2 E(\omega))}{\partial r}
747: = 0
748: \end{eqnarray}
749: The solution to this equation is any linear combination of
750: ingoing and outgoing spherical waves, {\it i.e.}, any linear
751: combination of functions of the form
752: $E(t,r)=\frac{1}{r^2}
753: g(r \pm t)$. For the first numerical experiment we choose a Gaussian
754: spherical wave propagating outwards,
755: \be
756: E(r,t) = \frac{1}{r^2} \exp \left\{ - {(r-t)}^2 \right\}
757: \ee
758: \be
759: F(r,t) = E(r,t).
760: \ee
761: Our numerical grid consists on 200 equally spaced shells
762: going from $r_1=0.2$ to $r_2=10.2$ and the boundary conditions
763: consist in imposing the analytical solution in both boundaries.
764: We show the results of this test in Figure \ref{fig2a}. In the figure
765: we displayed the analytical solution (solid line) and the
766: numerical (crosses) for the energy density.
767: In order to analyse the effect of the particular choice for the closure,
768: we repeated the experiment using different values for $p'(f=1)$, which leads
769: to different speeds of propagation of the waves, and no difference
770: was observed. The final reason is that all the closures displayed in Table 1
771: give $\lambda_+ (f=1)=1$. This is a general property of any
772: closure, since $p-f^2$ is a positive defined function
773: that goes to zero as $f \rightarrow 1$ and, therefore, its derivative has to
774: be negative in a vicinity of $f=1$ and $p' \le 2f$. Thus taking
775: the limit $f \rightarrow 1$ in (\ref{lambda}), one obtains
776: \be
777: \lambda_+ = \frac{p'+\sqrt{(2-p')^2}}{2} = 1
778: \ee
779:
780: To understand why the value of $\lambda_-$ is not important in this
781: problem, we study the diagonalized system obtained by adding and
782: subtracting the two equations in (\ref{system}) after taking $p=1$,
783: \begin{eqnarray}
784: \frac{\partial (E+F)}{\partial t} +
785: \frac{1}{r^2} \frac{\partial (r^2 (E+F))}{\partial r} = 0
786: \\
787: \frac{\partial (E-F)}{\partial t} -
788: \frac{1}{r^2} \frac{\partial (r^2 (E-F))}{\partial r} = 0.
789: \end{eqnarray}
790: In this equations it is clearly seen that
791: the quantities propagated with velocities $\lambda_{\pm}$ (Riemann invariants)
792: are $E \pm F$, respectively. Since in our particular problem
793: we took $F=E$, the quantity propagated by $\lambda_-$ is
794: identically null and the value of the speed of propagation
795: does not affect the solution.
796:
797: A different situation is found when we add to the initial conditions
798: a second wave propagating in the opposite direction. Figure \ref{fig2b}
799: shows the results from such experiment, in which a second wave
800: with smaller amplitude and opposite direction is located initially at $r=7$.
801: The closure used was $p=1$, $p'=0$, which gives $\lambda_+=1$, $\lambda_-=-1$.
802: In the figure, the analytical solution is denoted by the solid lines and
803: the numerical solution by crosses and each panel corresponds to a different
804: time. It can be seen how the code is able to solve the interaction
805: of the two waves until they completely cross over.
806: The difference between the analytical and numerical solution near
807: the center for the last panel are due to the boundary conditions
808: (initially, it was an ingoing wave).
809:
810: To illustrate the effect of a wrong speed of propagation, the
811: experiment is repeated with Kershaw's closure, that gives a correct speed
812: of propagation when $f=\pm 1$ but a wrong value otherwise.
813: Results are shown in Figure \ref{fig2c} for the same time steps
814: as in the previous case. At the beginning both waves seem to propagate
815: correctly but, as they cross each other and the flux factor departs
816: from unity, the numerical result differs from the analytical solution.
817: A similar behaviour is obtained with other closures.
818: This simple example manifests the importance of the choice for
819: the closure, since it fixes the speeds of propagation of the waves.
820:
821: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
822: \subsection{TEST 3: Sphere radiating in the vacuum.}
823:
824: In the two previous tests we studied the limiting cases, diffusive
825: and free-streaming regimes.
826: Another academic problem in the limit of vanishing
827: opacities consists in a sphere of radius $R$ radiating isotropically into
828: vacuum (Figure \ref{vac}).
829: Given the explicit form of the distribution function at the
830: surface of the sphere, ${\cal I}_0(t)={\cal I}(t=0,r=R)$, the general solution
831: can be easily obtained
832:
833: \be
834: {\cal I}(t,r,\mu) = \left\{ {\matrix{
835: {\cal I}_0(t-s) & \mu \ge x(r,t) \cr
836: 0 & \mu < x(r,t) \cr}} \right.
837: \ee
838: where
839: \be
840: s=r \left( \mu - \sqrt{\mu^2-1+{(R/r)}^2} \right)
841: \ee
842: \be
843: x(r,t) = \left\{ {\matrix{
844: \sqrt{1-{(R/r)}^2} & t \ge \sqrt{r^2-R^2} \cr
845: \displaystyle{ \frac{t^2+r^2-R^2}{2tr}} & r-R<t<\sqrt{r^2-R^2} \cr
846: 1 & t < r-R \cr}} \right.
847: \ee
848: where $\mu=\cos \alpha$.
849:
850: We take a linear time dependence ${\cal I}_0(t)=C t$, being $C$
851: an arbitrary constant, so that the angular integrals can be done
852: analytically. The angular moments of the distribution
853: function are, then, given by
854: \be
855: E(t,r)=\frac{C}{4} \left[ 2t(1-x) -r (1-x^2) + r \sqrt{1-x^2}
856: + r x^2 \log \left( \frac{x}{1 + \sqrt{1-x^2}} \right) \right]
857: \ee
858:
859: \be
860: F(t,r)=\frac{C}{12} \left[ 3t(1-x^2)
861: -2r (1-x^3) + 2r (1-x^2)^{3/2} \right]
862: \ee
863:
864: \be
865: P(t,r)=\frac{C}{48} \left[ 8t(1-x^3) - 6r (1-x^4) + 3r (2-x^2) \sqrt{1-x^2}
866: + 3 r x^4 \log \left( \frac{x}{1 + \sqrt{ 1-x^2}} \right) \right].
867: \ee
868:
869: We set up a numerical grid of 200 points, uniformly distributed,
870: from $r=1$ to $r=200$. The initial conditions are $F=0$ and
871: $E=\alpha/r^2$ with $\alpha$ being a small number, different from
872: zero for numerical purposes. We have taken $\alpha=10^{-4}$.
873: The radiation field is
874: introduced through the inner boundary conditions,
875: $E(t,r=R)={2t}$ and $f(t,r=R)=0.5$. Outflow is assumed in
876: the outer boundary.
877:
878: In Figure \ref{fig3a} the evolution of the radiation energy density as
879: a function of the radial coordinate is shown for three different
880: evolution times. The test has been repeat for three different
881: closure relations from those described in table 1: Vacuum (dashed lines),
882: Kershaw (dotted lines) and Minerbo (dashed-dotted lines). As in the
883: previous figures the analytical solution is represented by the solid line.
884: Although the appropriate closure for this problem (vacuum) gives a better
885: approximation to the real solution, it is remarkable that other closures
886: also give similar results. The differences between different closures
887: stem from the slightly different propagation speeds of the simple waves
888: in the regime of interest. Unlike in the two-wave problem described in
889: previous section, for this specific problem any reasonable closure
890: might be used without major deviations from the real solution.
891:
892:
893: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
894: \subsection{TEST 4: Radiative cooling and heating}
895:
896: With this test we intend to simulate a more realistic scenario in which,
897: initially, radiation produced in the center of a star diffuses out with
898: constant luminosity. We begin by
899: setting up a sphere which is in radiative equilibrium at a given
900: luminosity $L$. For $p=\frac{1}{3}$, the stationary solution is given by
901: \be
902: E(r) = \frac{L}{4 \pi} \left[ \frac{1}{R^2} + 3 \kappa
903: \left( \frac{1}{r}-\frac{1}{R} \right) \right]
904: \ee
905: \be
906: F(r) = \frac{L}{4 \pi r^2}
907: \ee
908: where $R$ is the radius of the outer boundary, at which $f=1$.
909: Starting with the initial model corresponding to $L=4 \pi$, we
910: let it evolve in three different situations: 1) we take $L=0$ as boundary
911: conditions at the inner boundary (located at $r=1$), and the radiation field
912: should diffuse out continuously;
913: 2) we change the inner boundary condition to
914: $L=40 \pi$, ten times larger than the initial model. The behaviour
915: is now the opposite, the energy density must increase until the
916: new stationary solution for the new luminosity is reached;
917: 3) we again take $L=0$ at the inner boundary but we switch on a
918: central point--source in the middle of the star ($r=6$).
919: The radiation field must evolve towards the corresponding stationary
920: solution, which is different in the two parts of the star defined by the
921: source. Inside, the luminosity is zero and the energy density is constant.
922: In the region external to the energy source, the stationary
923: solution is defined by $L=\dot{s}$, where $\dot{s}$ is the
924: energy per unit time injected by the source.
925: The results for $\kappa=100$ are shown in Figures \ref{fig4a}-\ref{fig4c},
926: for cases 1 to 3, respectively.
927: All three simulations have been done on the same grid of 200 uniform zones
928: and outflow boundary conditions in the outer edge have been imposed.
929:
930: Left panels display the energy density as a function of radius.
931: The thick solid line in the figures corresponds to the initial model
932: and the thick dotted line in Figure \ref{fig4b} represent the stationary
933: solution with $L=40 \pi$. In Figure \ref{fig4b}, we can see how the
934: heat wave lasts about 1000 time-units (the diffusion time $R \kappa$)
935: to arrive to the surface and a
936: new stationary solution is successfully obtained after $\approx 15$ times
937: the diffusion time scale.
938:
939: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
940: \subsection{TEST 5: The semi-transparent regime.}
941:
942: Testing a method in the semi-transparent regime becomes a tough
943: issue due to a lack of analytical solutions to compare with.
944: To overcome this problem, we have performed Monte Carlo simulations,
945: which is the closest solution to an exact one, and results in an
946: excellent test of the performance of the method in the
947: semi-transparent regime.
948:
949: We have used a Monte Carlo code that simulates the transfer of a
950: radiation field in a spherically symmetric geometry. We consider a
951: region with inner boundary at $r = R_{in}$ and outer boundary at
952: $r = R_{out}$, divided in 200 spherical shells. Within
953: each shell a constant value of the scattering opacity (elastic and isotropic)
954: is assumed. We do not consider emission or absorption processes.
955: Our Monte Carlo procedure starts by injecting particles at the inner boundary
956: outwardly-directed. Then, we compute the trajectory of each particle as
957: it scatters off matter, according to the standard random walk laws
958: \cite{Lucy99}. The
959: trajectory ends when the particle escapes by crossing the outer boundary
960: or it re-enters the inner boundary. Once the path of a particle is computed
961: we proceed to calculate the contribution of this particle to the moments of
962: the radiation field. The contribution of each particle to the energy density
963: at a given numerical shell is calculated by adding the path length between
964: two successive scattering events, within the shell, divided by the volume
965: of the shell. Consequently, the total energy density is obtained just by
966: adding the contribution of each particle multiplied by an arbitrary factor
967: Analogously, to obtain the contribution of each particle to the flux,
968: the path length is weighted using the cosine of the angle with respect to the
969: radial direction. Finally, pressure is calculated by means of the same
970: procedure, but using the square of the cosine as weight of the path length.
971:
972: In our numerical experiment we have taken: $R_{in} = 10$ and
973: $R_{out} = 100$. The opacity law is an exponentially decreasing
974: function in radius,
975: taking the value $\kappa=2$ at the inner boundary and $\kappa=2\times 10^{-5}$
976: at the outer boundary.
977: The corresponding values of the optical depth are such that
978: an important part of the region is semi-transparent.
979: We have run the Monte Carlo code using $500\,000$ particles in order
980: to obtain the stationary solution.
981:
982: The set up for our transport code is the following:
983: i) the above opacity law,
984: ii) the inner boundary conditions given by the stationary solution (obtained
985: from Monte Carlo), and iii) the closure relation. We have performed
986: two different calculations using Kershaw's
987: closure relation and a cubic polynomial fit to Monte Carlo results, explicitly
988: \be
989: p = 0.3228+0.1902 f-0.0476 f^2 +0.5131 f^3
990: \ee
991: Starting with arbitrary initial profiles, we let
992: the radiation field evolve until the stationary solution is reached and
993: results can be compared to those obtained from the Monte Carlo simulations.
994:
995: In Figure \ref{fig6} we show the stationary solutions calculated from
996: Monte Carlo simulations (solid line) and from our code with the two
997: different closures, Kershaw's (dotted line) and the above fit (dashed line).
998: From the results displayed in the figure we can conclude that the
999: evolutionary code describes the semi-transparent regime accurately.
1000: The differences in the values of the energy density (left panel),
1001: normalized to its inner boundary value, amounts less than a few percentage in
1002: any case. The slight discrepancies in the flux factor are mainly due to the
1003: closure relation employed, being less than 6\% for Kershaw's closure and
1004: less than 3\% for the fit.
1005: The closure, therefore, sets the maximum accuracy that can be obtained,
1006: although qualitative results at a level of 10\% uncertainty can easily
1007: be obtained with any reasonable closure in the semi-transparent regime.
1008:
1009: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1010: \subsection{A toy model.}
1011: Our last experiment intends to illustrate how the numerical method
1012: we proposed can succeed in situations when standard flux-limited
1013: diffusion techniques fail.
1014: We set up a sphere of radius $r=5$ with a radiation energy density of $E=1$,
1015: embedded in a background where the opacity varies as follows:
1016: \be
1017: \kappa = \left\{ {\matrix{
1018: \kappa_0 & r \le 1 \cr
1019: \kappa_0/r^2 & 1 < r \le 10 \cr
1020: 0 & 10 < r \le 15 \cr
1021: \kappa_0 & 15 < r \le 17 \cr
1022: 0 & r \ge 17 \cr
1023: }} \right.
1024: \ee
1025: where we have taken $\kappa_0=10^3$. Outside the sphere we take $E=0$
1026: (in practice, $E=10^{-6}$ for numerical reasons),
1027: and initially the flux is zero everywhere.
1028: It tries to mimic a situation in which radiation scatters out in
1029: a central star, where the diffusion approximation remains valid, through
1030: an opacity-varying atmosphere up to the surface ($r=10$), where the
1031: opacity suddenly vanishes. Then, it crosses a transparent region to
1032: find a very opaque layer, where diffusion operates again.
1033: This sort of problem is obviously a tough challenge for diffusion-based
1034: codes, since they cannot work properly in the transparent region.
1035:
1036: In Figure \ref{fig6} we show results from the evolution of this problem
1037: on a grid of 200 uniform shells and using Kershaw's closure. In the left
1038: panel we show the energy density for different times.
1039: It can be seen how a gradient is developed in the diffusive regions
1040: $(r<10, 15<r<17)$ while radiative equilibrium $(dE/dr=0)$ in the transparent
1041: region between the two opaque layers is rapidly obtained.
1042: Notice also that only for the last snapshot ($t=1000$, dashed curve) there is
1043: a flux of energy in the outer transparent region, because it takes some
1044: time to the radiation to diffuse through the opaque layer and develop
1045: a gradient that keeps the outgoing flow.
1046:
1047: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1048:
1049: \section{Conclusions}
1050:
1051: We have studied the mathematical character of the angular moment equations of
1052: radiative transfer, considering the implications of
1053: using different closure relations.
1054: The role played by the closure and their derivatives is more apparent
1055: in the hyperbolic formulation.
1056: After this analysis, we conclude that the hyperbolic character is assured for
1057: a given variety of closures widely used in the literature: those given by
1058: $p=p(f)$. For general closures $p=p(e,f)$, hyperbolicity can not be guaranteed,
1059: although the closure proposed by Cernohorsky \& Bludman \shortcite{CB94},
1060: based on maximum entropy arguments, leads to real eigenvalues.
1061: Additional constraints on the closures are imposed on the base of the
1062: behaviour of the eigenvalues of the Jacobian matrix since they give the
1063: velocity of propagation of perturbations. Causality limitation is written in
1064: terms of the closure relation helping us to select, from this point of view,
1065: among the different closures reviewed. It turns
1066: out that only one of the closures analysed does not satisfy causality,
1067: allowing for velocities of the waves higher than the speed of light.
1068: For the other closures,
1069: although derived without taking into account
1070: the mathematical issues addressed in this paper,
1071: we found that they are consistent with hyperbolicity and causality.
1072:
1073: Writing the moment equations of the distribution
1074: function as a hyperbolic system of conservation laws
1075: allows one to apply numerical techniques specifically
1076: designed for such systems, the so-called HRSC methods.
1077: We have applied HRSC methods to solve the transport equations in a static
1078: background showing, for a number of test problems, the feasibility of the
1079: method. From the results, the main conclusion is that the numerical method
1080: can be used in any regime, from optically thick to transparent regions,
1081: obtaining numerical solutions with
1082: high accuracy. It turns out that the closure relation plays an important
1083: role, more evident than in traditional methods for radiation
1084: transfer, that can be useful to choose the more appropriate closure for
1085: a given problem.
1086:
1087: When the radiation transport equations are written as a system of conservation
1088: laws, the coupling with hydrodynamics is straightforward
1089: \cite{Morel99}. This feature make us
1090: to be confident in the possibility of applying the method to problems
1091: involving radiative flows. Moreover,
1092: we are optimistic in the applicability to multidimensional problems,
1093: since the extension of the method is straightforward.
1094:
1095: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1096: \section*{Acknowledgements}
1097: {This work has been supported by the Spanish
1098: DGICYT (grant PB97-1432). J.A.P. also acknowledges previous financial
1099: support from the Ministerio de Educaci\'on y Cultura.}
1100:
1101:
1102:
1103:
1104: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1105: \begin{thebibliography}{}
1106:
1107: \bibitem[\protect\citename{Auer }1984]{Auer84}
1108: Auer, L. H., 1984, in Methods in Radiative Transfer,
1109: ed. W. Kalkofen, Cambridge University Press, Cambridge.
1110:
1111: \bibitem[\protect\citename{Banyuls {\it et al.} }1997]{BFIMM97}
1112: Banyuls, F., Font, J. A., Ib\'a\~{n}ez J.M$^{\underline{\mbox{a}}}$.,
1113: Mart\'{\i}, J. M$^{\underline{\mbox{a}}}$., Miralles, J. A.,
1114: 1997, ApJ, 476, 221.
1115:
1116: \bibitem[\protect\citename{Blinnikov \& Bartunov }1993]{BB93} %JOSE
1117: Blinnikov, S. I., Bartunov, O. S., 1993, A\&A., 273, 106
1118:
1119: \bibitem[\protect\citename{Bowers \& Wilson }1982]{BW82}
1120: Bowers, R. L., Wilson, J. R., 1982, ApJ Suppl., 50, 115
1121:
1122: \bibitem[\protect\citename{Bruenn }1985]{Bru85}
1123: Bruenn S. W., 1985, ApJ Suppl., 58, 771
1124:
1125: \bibitem[\protect\citename{Cernohorsky \& Bludman }1994]{CB94}
1126: Cernohorsky, J., Bludman, S. A., 1994, ApJ, 433, 250
1127:
1128: \bibitem[\protect\citename{Cooperstein, van den Horn \& Baron }1986]{CVB86}
1129: Cooperstein, J., van den Horn, L.J., Baron, E.A.,
1130: 1986, ApJ, 309, 653.
1131:
1132: \bibitem[\protect\citename{Dgani \& Janka }1992]{DJ92}
1133: Dgani, R., Janka, H.-Th., 1992, A\&A, 256, 428
1134:
1135: \bibitem[\protect\citename{Eastman \& Pinto }1993]{EP93} %JOSE
1136: Eastman, R. G., Pinto, P. A., 1993, ApJ, 412, 731
1137:
1138: \bibitem[\protect\citename{Ensman \& Burrows }1992]{EB92} %JOSE
1139: Ensman, L., Burows, A., 1992, ApJ, 393, 742
1140:
1141: \bibitem[\protect\citename{Gehmyer \& Mihalas }1994]{GM94} %JOSE
1142: Gehmeyr, M., Mihalas, D., 1994, Physica D, 72, 320
1143:
1144: \bibitem[\protect\citename{Godunov }1959]{Go59}
1145: Godunov, S. K., 1959, Mat. Sb., 47, 271.
1146:
1147: \bibitem[\protect\citename{Groth {\it et al.} }1996]{Gro96}
1148: Groth, C. P. T., Gombosi, T. I., Roe, P., L., Brown, S. L., 1996,
1149: Proceedings of the {\it Fifth international conference on
1150: Hyperbolic Problems}, Eds. Glimm, J., Graham, M.J., Grove, J.W.,
1151: World Scientific.
1152:
1153: \bibitem[\protect\citename{Harten, Lax \& van Leer }1983]{HLL83}
1154: Harten, A., Lax, P.D., van Leer, B., 1983, SIAM Rev., 25, 35.
1155:
1156: \bibitem[\protect\citename{Ib\'a\~{n}ez \& Mart\'{\i} }1999]{IM99}
1157: Ib\'a\~{n}ez, J. M$^{\underline{\mbox{a}}}$.,
1158: Mart\'{\i}, J. M$^{\underline{\mbox{a}}}$., 1991,
1159: J. Comput. Appl. Math., in press.
1160:
1161: \bibitem[\protect\citename{Janka }1991]{Jan91}
1162: Janka H.-Th., 1991, PhD thesis, Technische Universit\"at M\"unchen.
1163:
1164: \bibitem[\protect\citename{Janka, Dgani \& van den Horn }1992]{JDH92}
1165: Janka H.-Th., Dgani, R., van den Horn, L. J., 1992, A\&A, 265, 345.
1166:
1167: \bibitem[\protect\citename{Jin \& Levermore }1996]{JL96}
1168: Jin, S., Levermore, C.D., 1996, J. Comput. Phys., 126, 449.
1169:
1170: \bibitem[\protect\citename{Kershaw }1976]{Ke76}
1171: Kershaw, D. S., 1976, report UCRL-78378, Lawrence Livermore
1172: National Laboratory, Livermore, CA.
1173:
1174: \bibitem[\protect\citename{Kunasz }1983]{Ku83}
1175: Kunasz, P. B., 1983, ApJ, 271, 321.
1176:
1177: \bibitem[\protect\citename{van Leer }1979]{vL79}
1178: van Leer, B., 1979, J. Comput. Phys., 32, 101.
1179:
1180: \bibitem[\protect\citename{LeVeque }1992]{Le92}
1181: LeVeque R. J., 1992,
1182: Numerical methods for conservation laws, Birkh\"auser, Basel.
1183:
1184: \bibitem[\protect\citename{Levermore \& Pomraining }1981]{LP81}
1185: Levermore, C.D., Pomraining, G. C., 1981, ApJ, 248, 321
1186:
1187: \bibitem[\protect\citename{Lindquist }1966]{Lin66}
1188: Lindquist, R. W., 1966, Ann. Phys., 37, 478
1189:
1190: \bibitem[\protect\citename{Lowrie, Morel \& Hittinger }1999]{Morel99}
1191: Lowrie, R.B., Morel, J.E., Hittinger, J.A., 1999, ApJ, 432, 521.
1192:
1193: \bibitem[\protect\citename{Lucy }1999]{Lucy99}
1194: Lucy, L.B., 1999, A\&A, 344, 282.
1195:
1196: \bibitem[\protect\citename{Mart\'{\i}, Ib\'a\~{n}ez \& Miralles }1991]{MIM91}
1197: Mart\'{\i}, J. M$^{\underline{\mbox{a}}}$., Ib\'a\~{n}ez,
1198: J. M$^{\underline{\mbox{a}}}$., Miralles, J. A., 1991, Phys. Rev., D43, 3794
1199:
1200: \bibitem[\protect\citename{Mart\'{\i} \& M\"uller }1999]{MM99}
1201: Mart\'{\i}, J. M$^{\underline{\mbox{a}}}$.,
1202: M\"uller, E., 1999, {\it to appear in} Living Reviews in Relativity,
1203: Vol. 2.
1204:
1205: \bibitem[\protect\citename{Mezzacappa \& Bruenn }1993]{MB93}
1206: Mezzacappa, A., Bruenn, S. W., 1993, ApJ, 405, 669
1207:
1208: \bibitem[\protect\citename{Mihalas \& Kunasz }1986]{MK86}
1209: Mihalas, D., Kunasz, P. B., 1986, JCP, 64, 1M.
1210:
1211: \bibitem[\protect\citename{Mihalas \& Mihalas }1984]{MM84}
1212: Mihalas, D., Mihalas, B. W., 1984, Foundations of Radiation
1213: Hydrodynamics, Oxford Univ. Press, New York.
1214:
1215: \bibitem[\protect\citename{Mihalas {\it et al.} }1976]{MSKH76}
1216: Mihalas, D., Shine, R.A., Kunasz, P. B., Hummer, D. G., 1976, ApJ, 205, 492.
1217:
1218: \bibitem[\protect\citename{Minerbo }1978]{Mi78}
1219: Minerbo, G. N., 1978, J. Quant. Spectrosc. Radiat. Transfer, 20, 541
1220:
1221: \bibitem[\protect\citename{Miralles, Van Riper \& Lattimer }1993]{MRL93}
1222: Miralles, J. A., Van Riper, K. A., Lattimer, J. M., 1993, ApJ, 407, 687.
1223:
1224: \bibitem[\protect\citename{Norman \& Winkler }1986]{NW86}
1225: Norman, M. L., Winkler, K-H. A., 1986,
1226: Astrophysical Radiation Hydrodynamics, Reidel.
1227:
1228: \bibitem[\protect\citename{Pons {\it et al.} }1998]{PFIMM98}
1229: Pons, J. A., Font, J. A., Ib\'a\~{n}ez J.M$^{\underline{\mbox{a}}}$.,
1230: Mart\'{\i}, J. M$^{\underline{\mbox{a}}}$., Miralles, J. A.,
1231: 1998, A\&A, 339, 638.
1232:
1233: \bibitem[\protect\citename{Schinder \& Bludman }1989]{ScB89}
1234: Schinder, P. J., Bludman S. A., 1989, ApJ, 346, 350
1235:
1236: \bibitem[\protect\citename{Thorne }1981]{Tho81}
1237: Thorne K. S., 1981, MNRAS, 194, 439
1238:
1239: \bibitem[\protect\citename{Turolla \& Nobili }1988]{TN88}
1240: Turolla, R., Nobili, L., 1988, MNRAS, 235, 1273
1241:
1242: \bibitem[\protect\citename{Turolla, Zampieri \& Nobili }1995]{TZN95}
1243: Turolla, R., Zampieri, L., Nobili, L., 1995, MNRAS, 272, 625
1244:
1245: \bibitem[\protect\citename{Yamada, Janka \& Suzuki }1999]{Ya99}
1246: Yamada, S., Janka, H.-Th., Suzuki, H., 1999, A\&A, 344, 533.
1247:
1248: \end{thebibliography}
1249:
1250: \newpage
1251:
1252: \begin{center}
1253: \begin{table}
1254: \caption[]{Closure and characteristic structure.}
1255: \begin{tabular}{|c|cccccc|}
1256: \hline \hline
1257: \, & $p(f=0)$ & $p(f=1)$ & $p'(f=0)$ & $p'(f=1)$ & $\lambda_{\pm} (f=0) $
1258: & $\lambda_{\pm} (f=1) $
1259: \\
1260: \hline
1261: Minerbo & ${1/3}$ & 1 & 0 & 2 & $(\frac{1}{\sqrt{3}}, -\frac{1}{\sqrt{3}})$
1262: & (1,1) \\
1263: LP & ${1/3}$ & 1 & 0 & 1 & $(\frac{1}{\sqrt{3}}, -\frac{1}{\sqrt{3}})$
1264: & (1, 0) \\
1265: Auer & ${1/3}$ & 1 & ${2/3}$ & ${2/3}$ & $(1, -\frac{1}{3})$
1266: & $(1, -\frac{1}{3})$ \\
1267: Kershaw & ${1/3}$ & 1 & 0 & ${4/3}$ &$(\frac{1}{\sqrt{3}}, -\frac{1}{\sqrt{3}})$
1268: & $(1, \frac{1}{3})$ \\
1269: Bruenn & ${1/3}$ & 1 & $-{1/3}$ & ${5/3}$ & $(-1\pm \sqrt{13})/6$
1270: & $(1, \frac{2}{3})$ \\
1271: BW & ${1/3}$ & 1 & $-{4/3}$ & ${5/3}$ & $(-2\pm \sqrt{7})/3$
1272: & $(1, \frac{2}{3})$ \\
1273: Janka & ${1/3}$ & 1 & 0 & 2 & $(\frac{1}{\sqrt{3}}, -\frac{1}{\sqrt{3}})$
1274: & (1,1) \\
1275: Vacuum & ${1/3}$ & 1 & $-{2/3}$ & 2 & $(-1, \frac{1}{3})$ & (1,1) \\
1276: \hline \hline
1277: \end{tabular}
1278: \end{table}
1279: \end{center}
1280:
1281: \newpage
1282:
1283: \begin{figure}
1284: \psfig{figure=f1a.eps,width=16cm}
1285: \caption{Moderate diffusion limit: Energy density (left panel) and
1286: flux (right panel) as a function
1287: of the radial coordinate for different times, ($t=1,2,3,5$ from top
1288: to bottom). The opacity has been set to $\kappa=100$, corresponding
1289: to a Peclet number of $Pe=1$.}
1290: \label{fig1a}
1291: \end{figure}
1292:
1293: \begin{figure}
1294: \psfig{figure=f1b.eps,width=16cm}
1295: \caption{Extreme diffusion limit: Energy density (left panel) and
1296: flux (right panel) as a function
1297: of the radial coordinate for different times, ($t=200,240,300,400$ from top
1298: to bottom). The opacity has been set to $\kappa=10^5$, corresponding
1299: to a Peclet number of $Pe=5 \times 10^4$.}
1300: \label{fig1b}
1301: \end{figure}
1302:
1303: \begin{figure}
1304: \psfig{figure=f2a.eps,width=16cm}
1305: \caption{Free streaming limit: A Gaussian spherical wave propagating outwards.
1306: The analytical solution is represented by the solid line and the numerical
1307: solution by crosses. Snapshots for four different times ($t=2.5,3.5,5.0,7.0$)
1308: are plotted.}
1309: \label{fig2a}
1310: \end{figure}
1311:
1312: \begin{figure}
1313: \psfig{figure=f2b.eps,width=16cm}
1314: \caption{Free streaming limit: Two Gaussian spherical waves crossing over.
1315: The solid line corresponds to the analytical solution and the crosses
1316: to the numerical solution. Each panel displays a snapshot for a
1317: different time ($t=2.5,3.5,4.5,6.0$). The closure employed is $p=1$,
1318: $p'=0$, which gives $\lambda_{\pm}=\pm 1$
1319: everywhere.}
1320: \label{fig2b}
1321: \end{figure}
1322:
1323: \begin{figure}
1324: \psfig{figure=f2c.eps,width=16cm}
1325: \caption{Same as Figure \ref{fig2b} but using Kershaw's closure.
1326: The effect of taking wrong characteristic speeds is clearly
1327: illustrated.}
1328: \label{fig2c}
1329: \end{figure}
1330:
1331: \begin{figure}
1332: \psfig{figure=sphere.eps,width=12cm}
1333: \caption{Schematic diagram of Test 3: Sphere radiating isotropically
1334: into vacuum. The analytical solution for the radiation distribution
1335: function can be obtained in terms of $r$ and $\mu= \cos \alpha$.}
1336: \label{vac}
1337: \end{figure}
1338:
1339: \begin{figure}
1340: \psfig{figure=f3a.eps,width=16cm}
1341: \caption{Sphere radiating in the vacuum: Energy density as a function
1342: of the radial coordinate for three different evolution times. The
1343: analytical solution (described in \S 5.3) is represented by solid lines
1344: while the numerical solutions obtained with three different closures
1345: are represented by dashes (vacuum), dots (Kershaw) and the dash-dotted
1346: line (Minerbo). }
1347: \label{fig3a}
1348: \end{figure}
1349:
1350: \begin{figure}
1351: \psfig{figure=f4a.eps,width=16cm}
1352: \caption{Radiation initially in radiative equilibrium ($L=4 \pi$)
1353: diffusing out after switching off the central energy source. The
1354: solid thick line represents the initial model. The energy density
1355: is plotted in the left panel and the luminosity in the right panel.}
1356: \label{fig4a}
1357: \end{figure}
1358:
1359: \begin{figure}
1360: \psfig{figure=f4b.eps,width=16cm}
1361: \caption{Radiative heating: starting from the same initial model
1362: (solid thick line) as in figure \ref{fig4a}, we increase the luminosity
1363: at the inner boundary to $L=40 \pi$. The energy density evolves towards
1364: the new stationary solution at higher luminosity.}
1365: \label{fig4b}
1366: \end{figure}
1367:
1368: \begin{figure}
1369: \psfig{figure=f4c.eps,width=16cm}
1370: \caption{Same as figure \ref{fig4a} but with a point--like energy source
1371: located at $r=6$. The region inner to the source reaches the stationary
1372: solution at $L=0$, while the outer region settles in radiative equilibrium
1373: at luminosity higher than the initial model. The sharp discontinuity in the
1374: luminosity is well resolved.}
1375: \label{fig4c}
1376: \end{figure}
1377:
1378: \begin{figure}
1379: \psfig{figure=f5.eps,width=16cm}
1380: \caption{
1381: Semi-transparent regime: Stationary solution of
1382: the radiation field in a background
1383: with a scattering opacity exponentially decreasing with radius.
1384: Solid lines correspond to the Monte Carlo simulation and dotted and
1385: dashed lines correspond to the results obtained with the
1386: evolutionary code with Kershaw's closure and a numerical fit to
1387: Monte Carlo results, respectively. Left panel shows
1388: the normalized (to its inner boundary value) energy
1389: density as a function of the optical depth.
1390: Right panel shows the flux factor.
1391: }
1392: \label{fig5}
1393: \label{fnueva}
1394: \end{figure}
1395:
1396:
1397: \begin{figure}
1398: \psfig{figure=f6.eps,width=16cm}
1399: \caption{Evolution of the radiation field in a background
1400: with strongly varying opacity. Snapshots of the energy density (left panel)
1401: and flux (right panel) are shown for different times
1402: $t=0.5, 100, 250, 500, 1000$. The thick curve corresponds to $t=0.5$
1403: and the dashed curve to $t=1000$. See the text for
1404: details on the opacity profile.}
1405: \label{fig6}
1406: \end{figure}
1407:
1408: \end{document}
1409:
1410:
1411:
1412:
1413: