1: %=========================================================================
2: \section{Non-linear oscillations of spherically symmetric stars}
3: \label{pert}
4: %
5: %
6: %=========================================================================
7: \subsection{Introduction}
8: %
9: %
10: In this section we will turn our attention towards the study of compact
11: stars in the framework of general relativity. The discovery of stars
12: significantly more compact than the sun goes back to observations of the
13: binary star Sirius in the middle of the 19th century.
14: Sirius is the brightest star in
15: the sky as viewed from the earth. From an astrophysical point of view,
16: however, the faint companion of the bright main star, Sirius B has provoked
17: much more interest. The astronomer Bessel was the first to
18: infer the existence of an unseen companion of Sirius from a wobble in
19: the motion of the main star. It took another twenty years
20: before Alvin Clark managed to optically resolve Sirius B.
21: By the early twentieth century it became clear from the analysis of
22: its electromagnetic spectrum that Sirius B has a rather high surface
23: temperature of about 25,000$\,K$. In view of this result the extraordinarily
24: low luminosity of Sirius B lead to the conclusion that the star is very
25: small, about the size of the earth. This type of
26: high density star was consequently named a {\em white dwarf}.
27: It was understood at the time that white dwarfs mark the final stage in the
28: evolution of stars, but it remained a puzzle how such compact objects
29: were able to support themselves against gravitational contraction.
30: The answer was first provided by Eddington and
31: Fowler who suggested that the
32: star is supported by the degenerate electron pressure, a quantum
33: effect arising from the Pauli-exclusion principle. When Chandrasekhar
34: worked out the corresponding theory for a relativistic electron gas
35: he made the remarkable discovery that
36: the degenerate electron pressure will never be sufficient
37: to support white dwarfs above a mass of about $1.4\,M_{\odot}$.
38: In his words: {\em ``For a star of small mass the natural white dwarf
39: stage is an initial step towards complete extinction. A star of
40: large mass cannot pass into the white dwarf stage and one is left
41: speculating on other possibilities.''} It did not take long before
42: such speculations were directed towards the existence of neutron stars.\\
43: The first suggestion that stars made up of nucleons may exist
44: came from Landau in 1932 just two years after the discovery of the neutron.
45: Two years later Baade and Zwicky proposed the idea that neutron stars
46: may be the product of supernova explosions and thus mark the final stage in
47: the evolution of stars of large mass. The first theoretical models of
48: neutron stars were calculated in 1939 by Tolman, Oppenheimer and Volkoff.
49: It took another thirty years, however, before neutron stars were
50: actually discovered observationally. Furthermore this
51: discovery came in a completely unexpected way. In 1967 the then
52: Cambridge graduate student Jocelyne Bell and her supervisor Antony Hewish
53: were looking for scintillations of radio sources produced
54: by the interstellar medium. On the 28th of November 1967 they discovered
55: a source with an exceptionally regular pattern of radio pulses
56: which at the time even gave
57: rise to the speculation of an extra-terrestrial, intelligent origin.
58: These speculations were quickly abandoned, however, when
59: three more ``pulsars'' were found within the next few weeks. The extremely
60: short duration of the pulses and the high pulse frequencies
61: lead to the conclusion that these sources must be significantly smaller than
62: white dwarfs. An explanation for this phenomenon was finally found
63: when a pulsar was
64: detected at the centre of the crab nebula. From historical records it is
65: known that the crab nebula marks a supernova explosion that was observed
66: in the year 1054. Pulsars are therefore identified with neutron stars,
67: the remnants of supernova explosions. In the same way that the degeneracy
68: pressure of the electrons
69: supports white dwarfs against gravitational collapse,
70: the internal pressure in neutron stars
71: arises from the degenerate nucleons. A great deal
72: of work has gone into the observational and theoretical study of these
73: compact objects. From these studies it is known that neutron stars
74: have masses of about $1.4$ solar masses and radii of about
75: $10\,{\rm km}$. Neutron stars are believed to have a solid crust in which the
76: density increases from about $10^4\,{\rm g/cm}^3$ to a few times
77: $10^{11}\,{\rm g/cm}^3$. In this density range the matter is assumed to
78: consist of a degenerate electron gas and atomic nuclei that form a
79: crystal-like structure. At larger densities the atomic nuclei gradually
80: dissolve and at about $2\cdot10^{14}\,{\rm g/cm}^3$ the matter largely
81: consists of a highly incompressible neutron fluid with small amounts
82: of protons and electrons. An interesting property of this fluid
83: arises from the thermal temperature which is commonly believed to
84: be smaller than
85: $10^{8}\,{\rm K}$. Compared with the Fermi-temperature of the nucleons of about
86: $3\cdot 10^{11}\,{K}$ this implies that the matter behaves like
87: a fluid at zero temperature and becomes superfluid and, in the case of
88: the protons, super-conductive.
89: The structure of matter and the resulting equations
90: of state at higher densities become increasingly unclear and are
91: subject to ongoing research. Near the centre of a neutron star the density
92: is assumed to be of the order of $10^{15}\,{\rm g/cm}^3$ and the matter may
93: at least partly consist of hyperons, pions or quarks, so-called {\em strange
94: matter}. \\
95: The extreme compactness of neutron stars makes them particularly interesting
96: from a relativistic point of view. We have already mentioned the significance
97: of neutron stars in the context of the search for gravitational waves.
98: In this respect the importance of neutron star oscillations
99: arises from the
100: discovery of secularly unstable oscillation modes that increase
101: in amplitude due to the spin down of the neutron star
102: while energy is radiated away in the form of gravitational waves.
103: %This class of oscillations, called {\em r-modes},
104: %(\ref{Andersson1993})
105: %gains their energy from the rotational energy
106: %of the neutron star.
107: If the attempts to measure gravitational waves
108: are indeed successful, a whole new window for astrophysical observations
109: may be opened and facilitate a unique opportunity to directly observe the
110: interior of astrophysical objects such as neutron stars.
111: In this work, however, we will not directly study
112: neutron star oscillations in the context of gravitational radiation.
113: Instead we use the simpler case of spherically symmetric dynamic
114: neutron stars in order to probe a new numerical approach
115: which enables us to
116: numerically evolve non-linear oscillations of arbitrary amplitude with
117: high accuracy. While these evolutions will not lead to the generation
118: of gravitational waves because of the spherical symmetry, the numerical
119: results, the new techniques and the discussion of numerical
120: difficulties encountered
121: in the course of this work may still be relevant for numerical
122: simulations of more general types of neutron star oscillations. \\
123: The use of oscillations as a diagnostic tool to obtain information about
124: the interior structure of an object is an old idea and by no means restricted
125: to the realm of distant stars. For example the same technique has been
126: applied to the earth where the study of artificially induced oscillations
127: and, in particular, earthquakes has lead to invaluable insight into
128: the internal structure of our planet. In the same way
129: a great deal of knowledge has been obtained about the sun and more distant
130: stars by investigating their oscillations which reveal themselves in the
131: electromagnetic spectra of these objects.
132: Whereas Newtonian theory is perfectly adequate for studying ``normal''
133: stars, i.e. stars that gain their energy from continuous nuclear burning
134: of hydrogen and other light elements,
135: accurate modelling of compact objects like neutron stars
136: requires a general relativistic description. \\
137: %An accurate description of these objects can only be achieved in the
138: %framework of general relativity.
139: %Probably the most interesting aspect of studying oscillations of compact
140: %objects is the potential generation of gravitational waves sufficiently
141: %strong to be identified with gravitational detectors that are currently
142: %under construction or have recently gone online. \\
143: The first type of neutron star oscillations to be studied extensively
144: were linear radial oscillations
145: (see for example \shortciteNP{Chandrasekhar1964a},
146: \citeyearNP{Chandrasekhar1964b})
147: which today represent a well understood
148: problem that is described in the standard literature. The same is
149: not true, however, for nonlinear radial oscillations which lead to
150: qualitatively new problems. We have already mentioned that
151: % It is a well known result of general relativity
152: spherically symmetric spacetimes do not admit radiative solutions.
153: Instead the generation of gravitational waves
154: requires a time varying quadrupole or higher multi-pole ($l \ge 2$) moment
155: of the neutron star inertia.
156: %Consequently radial oscillations will not give rise to gravitational
157: %waves and from that point of view their study is not immediately
158: %interesting.
159: From that point of view, the study of radial oscillations is
160: not immediately interesting.
161: There are, however, several other important aspects
162: associated with radial oscillations. In the work mentioned above,
163: Chandrasekhar first revealed the existence of relativistic instability.
164: In the framework of radial oscillations this instability manifests itself
165: in the instability of the fundamental radial oscillation mode. If the frequency
166: of this mode becomes imaginary, an exponential growth of
167: physical quantities results and the star collapses or evaporates.
168: A fully non-linear evolution code
169: based on spectral methods has been developed by \lcite{Gourgoulhon1991}
170: and has been used to study various aspects of the stability of neutron stars
171: and their collapse into black holes (\citeNP{Gourgoulhon1993},
172: \shortciteNP{Gourgoulhon1994}).
173: Radial oscillations have also been considered from the point of view
174: of astrophysical observations.
175: The discovery of quasi-periodic radio sub-pulses in the spectra of pulsars
176: and periodicities in X-ray sources has lead to the suggestion that
177: radial oscillations of neutron stars may give rise to these features
178: (\citeNP{Boriakoff1976}, \citeNP{vanHorn1980}), which in turn
179: has stimulated further
180: research in this direction (see for example \shortciteNP{Marti1988},
181: \citeNP{Vaeth1992}). Furthermore the influence of radial oscillations
182: on the electromagnetic spectrum of neutron stars and their dependence
183: on the structure of matter at super-nuclear densities may provide
184: valuable information about the equation of state in the
185: high density range (\citeNP{Glass1983}).
186: The study of radial oscillations
187: is frequently carried out in the linear regime,
188: where all physical quantities have a harmonic time dependence
189: $f=f(r)e^{i\omega t}$ and the radial profiles $f(r)$ are determined
190: by an eigenvalue problem. In this work we will present explicit time
191: evolutions of the physical variables in the fully non-linear case. These
192: evolutions will serve two purposes. First we will be able to study deviations
193: from the known linearized behaviour, such as mode coupling and shock formation.
194: Secondly the spherically symmetric case can be used to investigate
195: numerical difficulties that are also expected in the more complicated
196: time evolutions in two or three spatial dimensions. A detailed analysis in
197: the computationally less expensive 1-dimensional case may lead to the
198: development of new advantageous numerical techniques or other types
199: of solutions to these problems.
200: % An explicit evolution
201: %of the time dependent quantities might, however, reveal unexpected numerical
202: %problems that are also present
203: %in more general scenarios with lower degrees of symmetry. This is
204: %particularly relevant for fully non-linear evolutions and the numerical
205: %investigation of non-linear radial oscillations will thus serve, among
206: %other purposes, also for probing numerical techniques with a wider range of
207: %applicability.
208: The work of \lcite{Gourgoulhon1991} for example has shown
209: among other results that the use of momentum densities as fundamental
210: variables may lead to
211: computation errors in passing from the momentum densities to the
212: velocity fields which can be avoided if velocity variables are used
213: in the first place. \\
214: In our discussion we will start with a
215: static spherically symmetric star which is governed by the
216: Tolman-Oppenheimer-Volkoff equations (\citeNP{Tolman1939},
217: \citeNP{Oppenheimer1939}). In section \ref{STATIC} we
218: will investigate these equations and describe the numerical methods
219: we use to calculate the resulting neutron star models. In section
220: \ref{DYNAMIC} we will use the static results in order to obtain a
221: fully non-linear perturbative formulation of dynamic spherically
222: symmetric stars. As a subclass we will discuss the linearized limit of
223: these equations in section \ref{PERT_LIN} and numerically
224: calculate the corresponding eigenmode solutions. It is interesting to
225: see that the surface of the star turns out to be a problematic
226: area even in this comparatively simple case. After a more detailed
227: discussion of the general problems one faces at the surface
228: in an Eulerian formulation we describe the numerical implementation of the code.
229: Even though the code is shown to perform well
230: in the linear regime for a large variety of neutron star models
231: in section \ref{EULER_LIN}, the surface problem is shown to give rise
232: to spurious results in some special cases.
233: In order to circumvent these problems
234: we use a simplified neutron star model in section
235: \ref{MODECOUPLING} to test the code in the non-linear regime and
236: to investigate the non-linear coupling of eigenmodes.
237: We conclude this work with the development of a fully non-linear
238: perturbative Lagrangian code in section \ref{LAGR}. We demonstrate how the
239: difficulties at the surface are resolved in such a formulation and
240: extensively test this code in the linear and non-linear regime. We
241: use this code to address the question whether non-linear effects
242: are present near the surface of the neutron star models in the
243: case of low amplitude oscillations.
244:
245:
246:
247: %=========================================================================
248: \subsection{Spherically symmetric static stars}
249: \label{STATIC}
250: %
251: %
252: In the fully non-linear perturbative approach to the study of
253: radial oscillations
254: we will decompose the time dependent physical quantities into static
255: background contributions and time dependent perturbations. The background
256: quantities will obey the corresponding static set of equations which will
257: then be used to remove terms of zero order from the fully non-linear
258: evolution equations in the time dependent case. In our studies we have
259: two principal choices for the static background: vacuum flat space in
260: which case we recover the standard non-perturbative formulation of the
261: problem and a static self-gravitating perfect fluid in spherical
262: symmetry which is described by the Tolman-Oppenheimer-Volkoff equations.
263: It is the second case which enables us to obtain highly accurate numerical
264: solutions for any given amplitude of the oscillations.
265: We will therefore first discuss in detail the Tolman-Oppenheimer-Volkoff (TOV)
266: equations as well as their numerical solution.\\
267:
268:
269: %=========================================================================
270: \subsubsection{The Tolman Oppenheimer Volkoff equations}
271: \label{TOV_EQ}
272: %
273: %
274: In the framework of the ``3+1'' formalism described in section
275: \ref{threep1}, we start by choosing coordinates $r$, $\theta$, $\phi$ on each
276: spatial hypersurface $\Sigma$. $\theta$ and $\phi$ are standard angular
277: coordinates and the radius $r$ is defined by the radial gauge condition,
278: so that the area of a surface $r={\rm const}$ is $4\pi r^2$.
279: The 3-dimensional line element is then given by
280: %
281: \begin{align}
282: ds^2 &= \mu^2 dr^2 + r^2(d\theta^2 + \sin^2\theta d\phi^2),
283: \end{align}
284: %
285: where in spherical symmetry $\mu$ is a function of $r$ only. If we label the
286: hypersurfaces $\Sigma$ by the coordinate $t$ we can apply the polar
287: slicing condition which combined with radial gauge can be shown to imply
288: a vanishing shift vector in spherical symmetry.
289: The 4-dimensional metric is then given by
290: %
291: \begin{align}
292: ds^2 &= -\lambda^2 dt^2 + \mu^2 dr^2 + r^2(d\theta^2 + \sin^2\theta d\phi^2).
293: \label{TOV_LINEELEMENT}
294: \end{align}
295: %
296: Here the lapse function $\lambda$ is also a function of $r$. Alternatively
297: this metric can be described by the variables $m$ and $\phi$ defined by
298: %
299: \begin{align}
300: \mu^2 &= \left( 1-\frac{2m}{r} \right) ^{-1}, \label{TOV_MUOFM} \\[10pt]
301: \lambda^2 &= e^{2\phi}.
302: \end{align}
303: %
304: In the Newtonian limit $\phi$ becomes the gravitational potential and $m$
305: the gravitating mass.\\
306: Our description of the matter is based on three simplifying assumptions, which
307: we will discuss in order.
308:
309: \vspace{0.5cm}
310: 1) \parbox[t]{15cm}
311: {
312: We will describe the matter as a single component perfect fluid.
313: This means that the fluid is seen as isotropic by a comoving observer.
314: %and it moves with a 4-velocity $\hbox{\vec u}^{\mu}$
315: %which may vary from point to point.
316: In particular no heat conduction, no shear stresses, anisotropic pressures
317: or viscosity must be present. The deviation from the
318: perfect fluid equilibrium due to anisotropic stresses resulting from the
319: solid crust are found to be $<10^{-5}$ even for rotating stars
320: (\citeNP{Friedman1992}). It is, however, not entirely clear to what
321: extent the treatment of the neutron star matter as a {\em single}
322: perfect fluid is too restrictive. It was suggested as early as 1959
323: by \citeANP{Migdal1959} that nucleons might be present in the form of
324: superfluids in the interior of neutron stars. In order to obtain
325: more realistic descriptions of neutron stars it might therefore be
326: necessary to describe the matter as a multicomponent fluid.
327: These issues are subject to ongoing research (see for example
328: \citeNP{Andersson2001}) and their investigation would exceed the scope
329: of this work.
330: We will therefore focus our discussion on single
331: component perfect fluids in which case we can write the energy-momentum
332: tensor in the form
333: %
334: \begin{align}
335: \hbox{\vec{T}}^{\mu \nu} &= (\rho + P) \hbox{{\vec u}}^{\mu}
336: \hbox{{\vec u}}^{\nu} + P \hbox{\vec{g}}^{\mu \nu}, \label{TOV_EMTENSOR}
337: \end{align}
338: %
339: }
340:
341: \hspace{0.5cm}\parbox[t]{15cm}
342: {
343: where $\rho$ is the energy density and $P$ the pressure measured by
344: a comoving observer. In the static spherically symmetric case $\rho$
345: and $P$ are functions of the radius $r$ and the 4-velocity
346: has a non-vanishing time component only. The normalisation condition
347: $\hbox{{\vec u}}^{\mu}\hbox{{\vec u}}_{\mu}=-1$ then implies
348: %
349: \begin{align}
350: \hbox{{\vec u}}^{\mu} &= \left[ \lambda^{-1},0,0,0 \right].
351: \label{TOV_4VELOCITY}
352: \end{align}
353: %
354: } \\
355:
356: \vspace{0.5cm}
357: 2) \parbox[t]{15cm}
358: {
359: The neutron star matter is assumed to be at zero temperature. This is
360: justified by comparing the thermal temperature of the stellar interior,
361: which is assumed to be smaller than $10^8\,{\rm K}$ in mature neutron stars,
362: with the relevant temperature scale given by the Fermi
363: temperature of the matter. Even though the thermal temperature is
364: large compared with terrestrial standards, it is orders of magnitude
365: below the Fermi temperature of matter at nuclear density ($\approx 3\cdot
366: 10^{11}$\,K), so that the thermal degrees of freedom are frozen out.
367: As a consequence the single component perfect fluid is described
368: by a 1-parameter equation of state which is
369: commonly chosen to be of the form $P=P(\rho)$.
370: } \\
371:
372: \vspace{0.5cm}
373: 3) \parbox[t]{15cm}
374: {
375: The equation of state (EOS) is assumed to be given by a polytropic law
376: %
377: \begin{align}
378: P &= K \rho^{\gamma}, \label{POLYTROPE}
379: \end{align}
380: %
381: where $K$ and $\gamma$ are constants. Instead of the polytropic exponent
382: $\gamma$ sometimes the polytropic index $n$ is used which is defined by
383: %
384: \begin{align}
385: \gamma = 1+\frac{1}{n}. \label{POLYTROPICINDEX}
386: \end{align}
387: %
388: The suitability of such an EOS
389: is certainly a debatable issue and the determination of realistic
390: equations of state of matter at super-nuclear densities represents
391: an entire branch of physical research. Conclusive answers have yet to be
392: obtained, however, and by using polytropes with different indices $n$
393: one is able to study the qualitative differences in the behaviour of
394: neutron stars with equations of state of varying stiffness. Furthermore
395: polytropes are given in analytic form
396: so that no additional numerical error arises from their use.
397: } \\
398:
399: \vspace{0.5cm}
400: We have got all ingredients now to derive the equations governing the static
401: spherically symmetric neutron star model. Starting with the metric
402: (\ref{TOV_LINEELEMENT}) and the energy-momentum tensor given by
403: Eq.\,(\ref{TOV_EMTENSOR}) with the 4-velocity (\ref{TOV_4VELOCITY})
404: the Einstein field equations $\hbox{\vec{G}}_{\mu \nu} = 8\pi
405: \hbox{\vec{T}}_{\mu \nu}$ result in two independent equations
406: %
407: \begin{align}
408: \frac{\lambda_{,r}}{\lambda} &=
409: \frac{\mu^2-1}{2r} + 4\pi r \mu^2 P, \label{TOV_LAMBDAR} \\[10pt]
410: \frac{\mu_{,r}}{\mu} &= -\frac{\mu^2-1}{2r} + 4\pi r \mu^2 \rho.
411: \label{TOV_MUR}
412: \end{align}
413: %
414: All other field equations are consequences of these two equations, their
415: derivatives and the matter equation (\ref{TOV_PR}).
416: In terms of the alternative variable $m(r)$ defined by
417: Eq.\,(\ref{TOV_MUOFM}), the equation for $\mu$ can be rewritten as
418: %
419: \begin{align}
420: m_{,r} &= 4\pi r^2 \rho. \label{TOV_MR}
421: \end{align}
422: %
423: From now on we will therefore refer to $m$ as the ``mass'' or
424: ``mass function'' of the neutron star.
425: Conservation of energy and momentum
426: $\nabla_{\mu}\hbox{\vec{T}}^{\mu \nu}=0$ results
427: in a single equation describing the hydrostatic equilibrium
428: %
429: \begin{align}
430: P_{,r} &= -\frac{\la_{,r}}{\la} (\rho + P). \label{TOV_PR}
431: \end{align}
432: %
433: The system of ODEs (\ref{TOV_LAMBDAR}), (\ref{TOV_MUR}), (\ref{TOV_PR}) was
434: first derived by \lcite{Tolman1939} and \lcite{Oppenheimer1939}
435: and is thus known as the Tolman-Oppenheimer-Volkoff or TOV equations.
436: Together with an
437: equation of state which we choose to be the polytropic law (\ref{POLYTROPE})
438: they describe a self-gravitating perfect fluid in spherical symmetry. \\
439: We finally need to specify appropriate boundary conditions for these
440: equations.
441: The condition for the radial component of the metric is $\mu=1$ at the origin
442: $r=0$ in order to avoid a conical singularity. This is also illustrated by
443: the requirement of a finite energy density $\rho$ at the centre which
444: according to Eq.\,(\ref{TOV_MR}) implies
445: that $m_{,r} = \mathscr{O}(r^2)$ near the centre. Consequently
446: $M= \mathscr{O}(r^3)$ and Eq.\,(\ref{TOV_MUOFM}) leads to $\mu=1$.
447: The lapse function $\lambda$ on the other hand appears in the equations
448: in the form $\lambda_{,r}/\lambda$ and is therefore only defined
449: up to a constant factor. Normally this factor is
450: chosen so that $\lambda$ takes on the value $\sqrt{1-2m/r}$ at the stellar
451: surface which matches the interior metric (\ref{TOV_LINEELEMENT}) to an
452: exterior Schwarzschild metric
453: %
454: \begin{align}
455: ds^2 &= -\left( 1-\frac{2M}{r} \right) dt^2 + \left( 1-\frac{2M}{r}
456: \right)^{-1} + r^2 d\theta^2 + r^2 \sin^2\theta d\phi^2,
457: \end{align}
458: %
459: where $M=m(R)$ and $R$ is the radius of the star.
460: Finally the surface of the star is defined by the vanishing of the
461: pressure $P$ which for the polytropic equation of state is
462: equivalent to $\rho=0$. We note that for some equations of state
463: the fluid extends to infinity and the energy density will vanish nowhere.
464: %Even though these equations of state do not describe neutron stars
465: %their study may still lead to valuable insights and
466: %an alternative boundary condition has to be used, for example by prescribing
467: %the energy density at some finite radius.
468: In this work, however, we will restrict ourselves to equations of state which
469: lead to stars of
470: finite size. We therefore summarise the boundary conditions as
471: %
472: \begin{align}
473: \mu &= 1 \label{TOV_MUBC}
474: \end{align}
475: %
476: at the origin $r=0$ and
477: %
478: \begin{align}
479: \lambda &= \sqrt{1-\frac{2m}{r}} = \frac{1}{\mu}, \label{TOV_LAMBDABC} \\[10pt]
480: \rho &= 0 \label{TOV_RHOBC}
481: \end{align}
482: %
483: at the surface $r=R$, i.e. three boundary conditions for the three
484: first order ODEs (\ref{TOV_LAMBDAR}), (\ref{TOV_MUR}), (\ref{TOV_PR}).
485: At first glance this seems to completely specify the physical scenario.
486: We have to note one subtlety however: the location of the stellar surface,
487: i.e. the extension of the numerical grid, is not determined at this
488: stage. For any given equation
489: of state we therefore expect a 1-parameter family of solutions parameterised
490: by the radius $R$. As we will see below we can alternatively parameterise the
491: family of solutions by the central density $\rho_{\rm c}$ of the star. Which
492: of these parameters we choose and therefore have to specify in addition to
493: the boundary conditions (\ref{TOV_MUBC})-(\ref{TOV_RHOBC}) depends on the
494: numerical approach we take towards solving the TOV-equations. There
495: are two main approaches to this problem.
496: %one strictly correct and one convenient.
497: % Before we turn our attention towards the numerical solution,
498: % however, we will briefly discuss the choice of physical units.
499:
500:
501: %=========================================================================
502: \subsubsection{The numerical treatment of the TOV-equations}
503: \label{TOV_NUM}
504: %
505: %
506: The problem we have to solve numerically is given by the TOV equations
507: (\ref{TOV_LAMBDAR}), (\ref{TOV_MUR}), (\ref{TOV_PR}), the boundary conditions
508: (\ref{TOV_MUBC})-(\ref{TOV_RHOBC}) and the prescription of the free parameter.
509: From a numerical point of view this is a two-point boundary value
510: problem and should be solved accordingly with shooting or relaxation methods.
511: This is the first of the two approaches we mentioned in the previous section.
512: Here we will discuss a relaxation algorithm.
513: In this case we set up a numerical grid, thus specifying the free parameter
514: in the form of the stellar radius, and finite difference the equations as
515: described in section \ref{relaxation}.
516: The three boundary conditions then provide
517: the remaining three algebraic equations and having specified an initial guess
518: the code relaxes to the solution of the TOV-equations. The main advantages
519: of this approach are:
520: %
521: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
522: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
523: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
524: \item all boundary conditions are exactly satisfied,
525: %\item if a neutron star model with a given radius is required the prescription
526: % of the radius is straightforward.
527: \item a neutron star model with a specified radius is obtained
528: straightforwardly by appropriately setting up the numerical grid.
529: \end{list}
530: %
531: This code suffers from some drawbacks, however, which can be summarized
532: as follows:
533: %
534: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
535: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
536: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
537: \item the specification of initial data is non-trivial and the convergence
538: of the code depends on a ``good'' initial guess,
539: \item obtaining high accuracy via a higher ($>2^{\rm nd}$) order
540: finite differencing scheme results in more complicated coefficient
541: matrices and inversion routines,
542: \item it is not clear how to obtain a neutron star model with a specified
543: central density,
544: \end{list}
545: %
546: It is quite remarkable that the second numerical approach has exactly the
547: opposite properties in that the advantages and drawbacks are reversed. In this
548: approach the outer boundary conditions are ignored initially
549: and instead one starts with three boundary conditions at the centre
550: %
551: \begin{align}
552: \mu &= 1, \\[10pt]
553: \lambda &= 1, \\[10pt]
554: \rho &= \rho_{\rm c}.
555: \end{align}
556: %
557: The TOV-equations can then be integrated outwards straightforwardly until the
558: energy density becomes negative and the out-most grid point will define the
559: surface of the star. Even though the energy density will not vanish exactly
560: at this point but take on a small positive value, the accuracy thus obtained
561: is good enough for most practical purposes. The remaining freedom to multiply
562: the lapse function $\lambda$ with an arbitrary constant
563: is used to enforce the
564: boundary condition (\ref{TOV_LAMBDABC}). Alternatively one can first integrate
565: Eqs.\,(\ref{TOV_MUR}), (\ref{TOV_PR}) for $\mu$ and $P$
566: which decouple from $\lambda$ and afterwards obtains $\lambda$ from
567: inward integration of Eq.\,(\ref{TOV_LAMBDAR}).\\
568: In a sense the two methods complement each other and for
569: example we use the quadrature approach to obtain an initial guess for the
570: relaxation scheme. Throughout this work we will use both numerical methods
571: and specify in each case how the TOV solutions were calculated. \\
572: Before we
573: investigate the solutions thus obtained, however, we have to discuss two
574: technical issues, the choice of physical units and a transformation to
575: a new radial coordinate which will provide higher resolution near
576: the surface of the star. Below we will see that sufficient resolution
577: in this region can be crucial for an accurate numerical
578: evolution in the time dependent case.
579:
580:
581:
582: %=========================================================================
583: \subsubsection{Physical units}
584: \label{TOV_UNITS}
585: %
586: %
587: Throughout this work we have worked with natural
588: units, i.e. $c=1=G$. This choice can be written in the form
589: %
590: \begin{align}
591: 1\,{\rm s} &= 2.9979 \cdot 10^{10}\, {\rm cm}, \label{C1} \\[5pt]
592: 1\,{\rm g} &= 7.4237 \cdot 10^{-29}\, {\rm cm}. \label{G1}
593: \end{align}
594: %
595: In astrophysics energy density is commonly measured in g/cm$^3$ and
596: pressure in dyne/cm$^2$, where 1\,dyne=1\,erg/cm. However, we prefer to
597: measure all quantities in km or corresponding powers thereof. Using
598: Eqs.\,(\ref{C1}) and (\ref{G1}) we can calculate that
599: %
600: \begin{align}
601: 1\, {\rm km}^{-2} &= 1.3477 \cdot 10^{18}\,\frac{{\rm g}}{{\rm cm^3}},\\[10pt]
602: 1\, {\rm km}^{-2} &= 1.2106 \cdot 10^{30}\,\frac{{\rm dyne}}{{\rm cm^2}}
603: = 1.2106 \cdot 10^{30}\, \frac{{\rm g}}{{\rm cm\, s^2}}.
604: \end{align}
605: %
606: The metric variables $\mu$ and $\la$ are dimensionless and it is obvious then
607: from Eqs.\,(\ref{TOV_MUOFM}) and (\ref{TOV_MUR}) that radius $r$ and
608: mass $m$ are
609: measured in km.
610: For example a typical central density for neutron stars is
611: $10^{15}$ g/cm$^3$ which in our units becomes 0.000742 km$^{-2}$.
612: We can also compare our results for radius and mass with the solar values
613: %
614: \begin{align}
615: M_{\odot} &= 1.4766\, {\rm km}, \\[10pt]
616: R_{\odot} &= 6.960 \cdot 10^5\, {\rm km}.
617: \end{align}
618: %
619: In contrast to these values typical radii and masses of neutron stars
620: are given by
621: %
622: \begin{align}
623: M_{\rm NS} &\approx 2\,{\rm km}, \\[10pt]
624: R_{\rm NS} &\approx 10\,{\rm km}.
625: \end{align}
626: %
627: It is a well known result that relativistic correction terms to
628: a Newtonian description of stars generally appear in terms of the ratio
629: $M/R$, so that this quotient describes the importance of relativistic effects.
630: In view of this result and the quotient $M_{\odot}/R_{\odot}=2.1\cdot 10^{-6}$
631: it is immediately obvious why a Newtonian description of the sun
632: and other ``normal'' stars is
633: perfectly adequate. In contrast we find $M/R \approx 0.2$ for neutron stars,
634: so that relativistic effects will play an important role in their behaviour
635: and accurate models need to be developed in the framework of
636: general relativity.
637:
638:
639: %=========================================================================
640: \subsubsection{Transformation to a new radial coordinate}
641: \label{TOV_RYTRAFO}
642: %
643: %
644: We have already mentioned that the surface of the star is defined by the
645: vanishing of the pressure which in the case of a polytropic
646: equation of state is equivalent to a zero energy density.
647: A dependent quantity frequently introduced in the study of neutron stars
648: is the speed of sound defined by
649: %
650: \begin{align}
651: C^2 &= \frac{\partial P}{\partial \rho}, \label{TOV_C2}
652: \end{align}
653: %
654: which in the polytropic case (\ref{POLYTROPE}) becomes
655: %
656: \begin{align}
657: C^2 &= K \gamma \rho^{\gamma-1}.
658: \end{align}
659: %
660: Consequently the speed of sound will also vanish at the surface if
661: $\gamma>1$ as will always be the case
662: for a star of finite mass. In particular we will show below that the
663: asymptotic behaviour of the speed of sound near the surface is given by
664: %
665: \begin{align}
666: C \sim \sqrt{R-r}. \label{TOV_ASSYMPTOTICC}
667: \end{align}
668: %
669: Taking into account the vanishing of the propagation speed
670: of sound waves at $r=R$ we now consider the qualitative behaviour
671: of a localized pulse travelling towards
672: the surface. As a result of the decreasing sound speed $C$
673: the front of the pulse will
674: in general travel more slowly than its tail and we would expect the pulse
675: to narrow. In particular the numerical resolution near the surface might
676: be inadequate to accurately evolve the pulse in this region and it might be
677: beneficial to work with a radial coordinate in terms of
678: which the propagation speed
679: is by and large independent of the position within the star.
680: In order to study the implications of a locally vanishing
681: propagation speed we consider the simpler scenario of the
682: 1-dimensional wave equation with variable propagation speed
683: %
684: \begin{align}
685: u_{,tt} &= c(r)^2 u_{,rr}, \label{WAVEEQ}
686: \end{align}
687: %
688: on a physical domain $0\le r\le R$.
689: Without loss of generality we will set $R=1$ for the rest
690: of this discussion. Eq.\,(\ref{TOV_ASSYMPTOTICC}) then suggests
691: to choose a propagation speed of the form
692: %
693: \begin{align}
694: c(r) &= \sqrt{1-r}. \label{WAVE_C}
695: \end{align}
696: %
697: For the numerical implementation we introduce the
698: auxiliary variables $F=u_{,t}$ and $G=u_{,r}$ and rewrite Eq.\,(\ref{WAVEEQ})
699: as a system of two first order PDEs
700: %
701: \begin{align}
702: F_{,t} &= c^2 G_{,r}, \label{WAVE1_FT} \\[10pt]
703: G_{,t} &= F_{,r}, \label{WAVE1_GT}
704: \end{align}
705: %
706: and impose the boundary conditions $u=0$, $F=0$ at both
707: boundaries. The system (\ref{WAVE1_FT}), (\ref{WAVE1_GT})
708: is linear and can be written in vectorial form
709: %
710: \begin{align}
711: \hbox{\vec{v}}_{,t} + A \hbox{\vec{v}}_{,r} &=0, \\[10pt]
712: \hbox{\vec{v}} &= \begin{pmatrix} F \\ G \end{pmatrix}, \\[10pt]
713: A &= \begin{pmatrix} 0 & -c^2 \\ -1 & 0 \end{pmatrix}.
714: \end{align}
715: %
716: The characteristics of the PDE are then given by
717: %
718: \begin{align}
719: \frac{dr}{dt} &= \Lambda_i,
720: \end{align}
721: %
722: where $\Lambda_1=c$, $\Lambda_2=-c$ are the eigenvalues of the
723: matrix $A$. At the outer boundary the slopes of the characteristics
724: collapse because of the vanishing of the wave speed $c$. \\
725: % It is
726: % this behaviour which gave rise to our suspicion in the first place.\\
727: This system has been evolved with the second order in space
728: and time McCormack finite differencing scheme described in section
729: \ref{McCormack} using a grid of 500 points.
730: In Fig.\,\ref{WAVES} we show the time evolution of $u$ obtained
731: for initial data in the form of a Gaussian pulse. Snapshots of $u$
732: are plotted at
733: $t_1 = 0.00$, $t_2=0.48$, $t_3=0.72$, $t_4=1.44$, $t_5=2.52$, $t_6=3.40$,
734: $t_7=4.44$, $t_8=4.60$, $t_9=5.60$, $t_{10}=6.56$, $t_{11}=7.20$
735: and $t_{12}=8.00$.
736: %
737: \begin{figure}[t]
738: \centering
739: \epsfig{file=u_s01_03.eps, height=200pt, width=150pt,angle=-90}
740: \epsfig{file=u_s04_06.eps, height=200pt, width=150pt,angle=-90}
741: \epsfig{file=u_s07_09.eps, height=200pt, width=150pt,angle=-90}
742: \epsfig{file=u_s10_12.eps, height=200pt, width=150pt,angle=-90}
743: \caption{The numerical evolution of an initial
744: Gaussian pulse according to the
745: wave equation in terms of the coordinate $r$ as obtained for
746: the varying propagation speed given by Eq.\,(\ref{WAVE_C})
747: which vanishes at $r=1$. The Snapshots are
748: shown for the times $t_1,\ldots,t_{12}$.}
749: \label{WAVES}
750: \end{figure}
751: %
752: %Nothing obviously wrong is seen in the plots although the broadening of
753: %the pulse after reflection at the outer boundary may be unexpected.
754: In order to shed light on the quality of the numerical evolution we
755: analyse the convergence properties of the code.
756: For this purpose we have performed the same runs
757: using 1000 and 2000 grid points and calculated the time dependent convergence
758: factor according to the method described in section \ref{CCM_CONVERGENCE}.
759: Again we use a high resolution reference solution obtained for 2000 grid points
760: in place of the analytic solution. The results shown
761: in Fig.\,\ref{CONV_WAVES} demonstrate that
762: the convergence of the code drops to
763: first order at about $t=2.5$ which coincides with the snapshot at $t_5$
764: when the pulse is reflected at the outer boundary for the first time. This
765: result is confirmed by high resolution runs in which no
766: broadening of the pulse similar to that shown in Fig.\,\ref{WAVES} is observed
767: after reflections at either boundary. We
768: conclude that a naive numerical evolution can lead to spurious results
769: in regions with a vanishing propagation speed and that this problem is
770: %
771: \begin{figure}[t]
772: \centering
773: \epsfig{file=wave_conv_s.eps, height=300pt, width=200pt, angle=-90}
774: \caption{The convergence factor obtained for 500 and 1000
775: grid points as a function of time. At $t\approx 2.5$ the
776: convergence drops to first order.}
777: \label{CONV_WAVES}
778: \end{figure}
779: %
780: due to an insufficient spatial resolution. \\
781: A solution to this problem is obtained by transforming to a new
782: spatial coordinate $y$ in terms of which the slopes of the characteristics
783: do not vary as drastically over the numerical domain and in particular do
784: not vanish at the boundary. A simple recipe is to define this new coordinate
785: by
786: %
787: \begin{align}
788: y = \int_0^r{\frac{1}{c(\tilde{r})}d\tilde{r}}, \label{WAVE_RY}
789: \end{align}
790: %
791: which implies
792: %
793: \begin{align}
794: \frac{\partial}{\partial r} &= \frac{1}{c} \frac{\partial}{\partial y},
795: \\[10pt]
796: dr &= c\,dy.
797: \end{align}
798: %
799: In the special case where the propagation speed is given by Eq.\,(\ref{WAVE_C})
800: the coordinates $r$ and $y$ are related by
801: %
802: \begin{align}
803: y &= 2-2\sqrt{1-r}, \\[10pt]
804: r &= y-\frac{y^2}{4},
805: \end{align}
806: %
807: so that the interval $r \in [0,1]$ is mapped to $y \in [0,2]$.
808: In terms of the new coordinate $y$ the system (\ref{WAVE1_FT}),
809: (\ref{WAVE1_GT}) can be rewritten as
810: %
811: \begin{align}
812: F_{,t} &= c G_{,y}, \label{WAVE2_FT} \\[10pt]
813: G_{,t} &= \frac{1}{c} F_{,y}, \label{WAVE2_GT}
814: \end{align}
815: %
816: and the characteristic curves are given by
817: %
818: \begin{align}
819: \frac{dy}{dt} &= \pm 1.
820: \end{align}
821: %
822: In order to compare the new scheme with the original approach, we evolve the
823: %
824: \begin{figure}[t]
825: \centering
826: \epsfig{file=u_x01_03.eps, height=200pt, width=150pt,angle=-90}
827: \epsfig{file=u_x04_06.eps, height=200pt, width=150pt,angle=-90}
828: \epsfig{file=u_x07_09.eps, height=200pt, width=150pt,angle=-90}
829: \epsfig{file=u_x10_12.eps, height=200pt, width=150pt,angle=-90}
830: \caption{The same evolution as in Fig.\,\ref{WAVES}, but obtained with
831: the new coordinate $y$ which results in a higher density
832: of grid points near the outer boundary $r=1$.}
833: \label{WAVEX}
834: \end{figure}
835: %
836: same initial data as above using the system (\ref{WAVE2_FT}), (\ref{WAVE2_GT})
837: on a $y$-grid again with 500 grid points and the same boundary conditions.
838: %
839: \begin{figure}[t]
840: \centering
841: \epsfig{file=wave_conv_x.eps, height=300pt, width=200pt, angle=-90}
842: \caption{The time dependent convergence factor obtained for the
843: numerical evolution of the wave equation on a $y$-grid with
844: 500 and 1000 grid points. Second order convergence is clearly
845: maintained throughout the evolution.}
846: \label{CONV_WAVEX}
847: \end{figure}
848: %
849: The result is shown in Fig.\,\ref{WAVEX} where we plot
850: the same snapshots as in Fig.\,\ref{WAVES}.
851: For comparison purposes the plots show $u$ as a function of the coordinate
852: $r$ but as a result of the computation on the $y$-grid,
853: the density of grid points is higher towards $r=1$ in Fig.\,\ref{WAVEX} whereas
854: the grid points are distributed homogeneously in Fig.\,\ref{WAVES}. In
855: contrast to the above evolution no broadening
856: of the pulse after reflection at the outer boundary is observed.
857: The time dependent convergence analysis shown
858: in Fig.\,\ref{CONV_WAVEX}
859: demonstrates second order convergence throughout the run even though
860: small variations in the convergence factor are visible when the pulse is
861: reflected at either boundary. We conclude that a transformation of the type
862: (\ref{WAVE_RY}) provides the necessary resolution in a region of
863: vanishing propagation speed and leads to satisfactory results at reasonable
864: grid resolutions. \\
865: We now have to apply this idea to the case of a static, spherically symmetric
866: neutron star. The role of the wave speed $c$ is now assumed by the speed
867: of sound $C$ defined in Eq.\,(\ref{TOV_C2}) and we introduce the new radial
868: coordinate
869: %
870: \begin{align}
871: y &= \int_0^r{\frac{1}{C(\tilde{r})}d\tilde{r}}. \label{TOV_YOFR}
872: \end{align}
873: %
874: This transformation has also been successfully used by \lcite{Ruoff2000} in the
875: linearized time evolution of radial oscillations for more realistic
876: equations of state.
877: The asymptotic behaviour of the sound speed
878: in the Tolman-Oppenheimer-Volkoff case given by Eq.\,(\ref{TOV_ASSYMPTOTICC})
879: is identical to that of the wave speed
880: in the toy problem. Consequently the radial
881: interval $r\in[0,R]$ of the
882: star will be mapped to a finite interval $y \in [0,Y]$.
883: % The numerical results
884: %shown below confirm this expectation.
885: In order to obtain a formulation which
886: includes both possible choices of the radial coordinate, we introduce
887: the variable $x$ in terms of which the TOV equations are written as
888: %
889: \begin{align}
890: r_{,x} &= \left\{ \parbox{7cm}
891: {
892: $1\hspace{1cm}{\rm if}\,\,\,x=r$\\
893: $C\hspace{0.9cm}{\rm if}\,\,\,x=y$,
894: } \right. \label{TOV_RY} \\[10pt]
895: \frac{\lambda_{,x}}{\lambda} &=
896: r_{,x} \left(\frac{\mu^2-1}{2r}
897: + 4\pi r \mu^2 P \right), \label{TOV_LAMBDAY}
898: \\[10pt]
899: \frac{\mu_{,x}}{\mu} &= r_{,x} \left( -\frac{\mu^2-1}{2r} + 4\pi r \mu^2 \rho
900: \right), \label{TOV_MUY} \\[10pt]
901: P_{,x} &= -\frac{\lambda_{,x}}{\lambda} (\rho + P) \label{TOV_PY}.
902: \end{align}
903: %
904: In the numerical code we are thus able to switch between the two alternative
905: modes of calculation by assigning the derivative $r_{,x}$ according to
906: either possibility of Eq.\,(\ref{TOV_RY}).
907: % No further modification is
908: %required.
909: In either case the boundary conditions are given by
910: Eqs.\,(\ref{TOV_MUBC})-(\ref{TOV_RHOBC}) supplemented with the requirement
911: that $r$ and $x$ vanish simultaneously at the origin
912: %
913: \begin{align}
914: r &= 0 \,\,\,{\rm at}\,\,\,x=0.
915: \end{align}
916: %
917: %Unless specified otherwise, from now on we will use the rescaled
918: %radial coordinate and thus set $r_{,x}=C$.
919: One subtlety concerning the relaxation method of calculating
920: TOV solutions has to be mentioned. In this case we need to specify the
921: radius of the star. If we use the rescaled radial coordinate, however, the
922: surface value $x_{\rm s}$ is not a priori known. In practice we
923: therefore specify the free parameter in the form of the
924: central density and solve the TOV equations via the quadrature
925: method first. This provides us with the outer boundary value of the
926: coordinate $x$ for the stellar model in question and we can solve the
927: TOV equations in a second step with the relaxation method.
928: % In the next section we will
929: %present numerical solutions for different polytropic models and also
930: %investigate the asymptotic behaviour of the solutions at both
931: %boundaries in detail.
932:
933:
934: %=========================================================================
935: \subsubsection{Asymptotic properties of the TOV equations}
936: \label{TOV_ASYMPTOTICS}
937: %
938: %
939: %For a given system of ordinary differential equations it is always worth
940: %studying the asymptotic behaviour of the solutions if the complexity
941: %of the equations allows such a study. In the case of the equations for
942: %a static cosmic string in section \ref{SECcsmink} for example the
943: %analysis of the
944: %asymptotic behaviour revealed the existence of exponentially diverging
945: %unphysical solutions. This result enabled us to choose a suitable numerical
946: %technique which automatically selected the non-diverging physical solutions.
947: %In the case of the TOV equations no such exponentially diverging solutions
948: %do exist which makes the quadrature method of solving the equations such
949: %a successful tool, but we will see below that the asymptotic behaviour
950: The asymptotic behaviour of the solutions of the TOV equations
951: (\ref{TOV_RY})-(\ref{TOV_PY}) at the surface of the star has
952: serious implications for the
953: simulation of dynamic neutron stars with certain equations of state in a
954: strictly Eulerian framework. We will therefore discuss the asymptotic
955: behaviour first and then compare the results with the numerically obtained
956: solutions. Since the introduction of the rescaled radial coordinate resulted
957: from numerical requirements only, we use $r_{,x}=1$ i.e. the original system
958: (\ref{TOV_LAMBDAR})-(\ref{TOV_PR}) for the asymptotic analysis.
959: We start with the behaviour at the origin, where we assume that
960: %
961: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
962: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
963: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
964: \item the energy density and thus the pressure are finite and positive,
965: \item the lapse function $\lambda$ is finite and positive.
966: \end{list}
967: %
968: We have already seen that the central value of the energy density is a
969: free parameter and the pressure follows from the equation of state.
970: The central value of the lapse function, on the other hand,
971: is determined by matching $\lambda$ to an exterior
972: Schwarzschild metric.
973: We also know from section \ref{TOV_EQ} that our assumptions imply
974: $\mu=1$ and $m=\mathscr{O}(r^3)$ at the origin. From Eq.\,(\ref{TOV_MUOFM})
975: we therefore conclude that $\mu=1+\mathscr{O}(r^2)$. Inserting this result into
976: Eq.\,(\ref{TOV_LAMBDAR}) and using the second assumption we find that
977: $\lambda_{,r}/\lambda \sim r$ and thus $\lambda =\lambda_{\rm c}
978: +\mathscr{O}(r^2)$. Using this result in Eq.\,(\ref{TOV_PR}) leads to
979: $P_{,r}\sim r$, i.e. $P=P_{\rm c} + \mathscr{O}(r^2)$ and the equation
980: of state then shows that the energy density has the same behaviour.
981: In summary the results near the origin are
982: %
983: \begin{align}
984: \lambda(r) &= \lambda_{\rm c} + \mathscr{O}(r^2), \\[10pt]
985: \mu(r) &= 1 + \mathscr{O}(r^2), \\[10pt]
986: \rho(r) &= \rho_{\rm c} + \mathscr{O}(r^2), \\[10pt]
987: P(r) &= K \rho_{\rm c}^{\gamma} + \mathscr{O}(r^2).
988: \end{align}
989: %
990: %We also note that $\lambda$, $\mu$, $\rho$ and $P$ are components of rank
991: %2 tensors and consequently their series expansion should contain
992: %only even powers of $r$ in spherical symmetry for regularity reasons. Our
993: %results are compatible with this requirement. \\
994: The corresponding analysis for the surface is more complicated and the results
995: will later prove to be of more significance. For this analysis it is convenient
996: to work with the radial variable
997: %
998: \begin{align}
999: z &:= R-r.
1000: \end{align}
1001: %
1002: We start with the following assumptions.
1003: %
1004: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
1005: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
1006: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
1007: \item The metric function $\mu$ is finite at the surface and also satisfies
1008: the inequality $\mu > 1$. This follows from Eq.\,(\ref{TOV_MUOFM})
1009: and the requirement that the mass satisfies
1010: the condition $0 < 2m(R) < R$. The first inequality follows
1011: from Eq.\,(\ref{TOV_MR}) for any non vacuum model and the
1012: second implies that
1013: the neutron star extends beyond its Schwarzschild radius.
1014: \item The lapse $\lambda$ is finite and positive at the surface.
1015: \item The energy density and the pressure vanish at the surface and their
1016: leading order terms are given by some positive powers of $z$.
1017: \end{list}
1018: %
1019: We write these assumptions as
1020: %
1021: \begin{align}
1022: \mu &= \mu_{\rm s} + \mathscr{O}(z^{\epsilon_1}),
1023: \label{TOV_ASSYMPTOTICANSATZMU} \\[10pt]
1024: \lambda &= \lambda_{\rm s} + \mathscr{O}(z^{\epsilon_2}), \\[10pt]
1025: \rho &= \rho_{\rm s} z^{\alpha} + \mathscr{O}(z^{\alpha+\epsilon_3}),
1026: \\[10pt]
1027: P &= P_{\rm s} z^{\beta} + \mathscr{O}(z^{\beta+\epsilon_4}),
1028: \label{TOV_ASSYMPTOTICANSATZP}
1029: \end{align}
1030: %
1031: where $\alpha$, $\beta$ and $\epsilon_1 , \ldots, \epsilon_4$ are positive
1032: constants we have yet to determine and $\mu_{\rm s}$, $\lambda_{\rm s}$,
1033: $\rho_{\rm s}$ and $P_{\rm s}$ are non vanishing constants subject to the
1034: restrictions mentioned above. We first insert the expressions for
1035: $\rho$ and $P$ into the equation of state (\ref{POLYTROPE}). Comparison of
1036: the leading order terms then leads to
1037: %
1038: \begin{align}
1039: \beta &= \alpha \gamma, \label{TOV_ALPHABETA1} \\[10pt]
1040: \epsilon_3 &= \epsilon_4,
1041: \end{align}
1042: %
1043: where $\gamma$ is the polytropic exponent. Similarly the leading order in
1044: Eq.\,(\ref{TOV_MUR}) results in
1045: %
1046: \begin{align}
1047: \epsilon_1 = 1.
1048: \end{align}
1049: %
1050: We then combine Eqs.\,(\ref{TOV_LAMBDAR}) and (\ref{TOV_PR}) to eliminate the
1051: lapse function and insert
1052: (\ref{TOV_ASSYMPTOTICANSATZMU})-(\ref{TOV_ASSYMPTOTICANSATZP}).
1053: The result of comparing the two leading orders is
1054: %
1055: \begin{align}
1056: \alpha+1 &= \beta, \\[10pt]
1057: \epsilon_4 &= 1.
1058: \end{align}
1059: %
1060: This provides a second condition for $\alpha$ and $\beta$ and
1061: with Eq.\,(\ref{TOV_ALPHABETA1}) we conclude that
1062: %
1063: \begin{align}
1064: \alpha &= \frac{1}{1-\gamma} = n, \\[10pt]
1065: \beta &= n+1,
1066: \end{align}
1067: %
1068: where $n$ is the polytropic index defined in (\ref{POLYTROPICINDEX}).
1069: Finally we use these results in Eq.\,(\ref{TOV_LAMBDAR}) for
1070: the lapse function and obtain
1071: %
1072: \begin{align}
1073: \epsilon_2 &= 1.
1074: \end{align}
1075: %
1076: We summarise the asymptotic behaviour at the surface:
1077: %
1078: \begin{align}
1079: \mu &= \mu_{\rm s} + \mathscr{O}(z), \label{TOV_ASSYMPTOTICMU}
1080: \\[10pt]
1081: \lambda &= \lambda_{\rm s} + \mathscr{O}(z), \\[10pt]
1082: \rho &= \rho_{\rm s} z^{n} + \mathscr{O}(z^{n+1}),
1083: \label{TOV_ASSYMPTOTICRHO} \\[10pt]
1084: P &= P_{\rm s} z^{n+1} + \mathscr{O}(z^{n+2}).
1085: \label{TOV_ASSYMPTOTICP}
1086: \end{align}
1087: %
1088: As a consequence we will not be able to Taylor expand $\rho$ and $P$
1089: about the surface $z=0$ unless a polytropic equation of state with
1090: integer index $n$ is chosen. Indeed a more extensive analysis carried out with
1091: the algebraic computing package GRTensor II
1092: shows that higher
1093: order terms containing the polytropic index $n$ also appear in the expansions
1094: of $\lambda$ and $\mu$ so that these functions are subject to the same
1095: limitations regarding Taylor expansion. \\
1096: The most important result of the asymptotic analysis concerns the
1097: behaviour of the energy density $\rho$ near the surface given by
1098: Eq.\,(\ref{TOV_ASSYMPTOTICRHO}). In particular we note that for a polytropic
1099: index $n < 1$ or exponent $\gamma>2$ the gradient of $\rho$
1100: with respect to the areal
1101: radius $r$ will be infinite at the surface. The case $n=1$, i.e. $\gamma=2$
1102: is the limiting case where $\rho$ has a finite gradient. This special case
1103: also implies
1104: that no fractional powers appear in the series expansions of $\lambda$,
1105: $\mu$, $\rho$ and $P$. $\gamma=2$ is considered to provide a
1106: qualitatively good description of the average stiffness of the equation
1107: of state of neutron stars and thus a popular choice for the polytropic
1108: exponent. For $n>1$ or $\gamma<2$ the energy density will have
1109: a vanishing gradient at the surface. \\
1110: It remains to check the asymptotic
1111: behaviour in terms of the rescaled radial coordinate $y$. From the definition
1112: of the speed of sound (\ref{TOV_C2}) and the results above we conclude that
1113: near the surface
1114: %
1115: \begin{align}
1116: C(z) &= \mathscr{O}(z^{1/2}), \label{TOV_ASSYMPTOTICC2}
1117: \end{align}
1118: %
1119: which implies that
1120: %
1121: \begin{align}
1122: \frac{\partial \rho}{\partial y} &= C \frac{\partial \rho}{\partial r}
1123: = \mathscr{O}(z^{n-1/2}).
1124: \end{align}
1125: %
1126: All other functions have vanishing gradients with respect to $y$ near the
1127: surface.
1128: Consequently the rescaled coordinate allows us to calculate neutron star models
1129: for polytropic exponents up to $\gamma=3$ without encountering infinite
1130: gradients and the corresponding numerical inaccuracies.
1131: % Since models with such
1132: %high polytropic exponents are not considered very realistic, we will restrict
1133: %ourselves to the range $\Gamma<3$. \\
1134:
1135:
1136: %=========================================================================
1137: \subsubsection{Solutions of the TOV equations}
1138: %
1139: %
1140: In view of the results of the asymptotic analysis we have numerically
1141: solved the TOV-equations for neutron star models with different polytropic
1142: exponents $\gamma<2$, $\gamma=2$ and $\gamma>2$.
1143: The corresponding models are listed
1144: in Table \ref{MODELS15} where we have included two further models
1145: with $\gamma=2$ but different polytropic factor $K$, which we will
1146: use to also study the variation of the solutions with $K$.
1147: %
1148: %\begin{table}[t]
1149: %\begin{center}
1150: %\caption{The parameters for five different neutron star models. We will
1151: % refer to these as models 1-5 in this work.}
1152: %\begin{tabular}{c|ccc}
1153: % \hline
1154: % \hline
1155: % model & $\gamma$ & $K$ & $\rho_{\rm c}$ \\
1156: % \hline
1157: % 1 & $1.75$ & $25\,\, {\rm km}^{1.5}$ &
1158: % $1.25\cdot 10^{-3}\, {\rm km}^{-2}$ \\
1159: % 2 & $2.00$ & $100\, {\rm km}^2$ &
1160: % $1.5\cdot 10^{-3}\, {\rm km}^{-2}$ \\
1161: % 3 & $2.00$ & $150\,\, {\rm km}^2$ &
1162: % $1.5\cdot 10^{-3}\, {\rm km}^{-2}$ \\
1163: % 4 & $2.00$ & $200\, {\rm km}^2$ &
1164: % $1.5\cdot 10^{-3}\, {\rm km}^{-2}$ \\
1165: % 5 & $2.30$ & $1800\, {\rm km}^{2.6}$ &
1166: % $1.0\cdot 10^{-3}\, {\rm km}^{-2}$ \\
1167: % \hline
1168: % \hline
1169: %\end{tabular}
1170: %\label{MODELS15}
1171: %\end{center}
1172: %\end{table}
1173: %
1174: %
1175: %
1176: \begin{table}[t]
1177: \begin{center}
1178: \caption{The parameters for five different neutron star models. We will
1179: refer to these as models 1-5 in this work.}
1180: \begin{tabular}{c|ccccc}
1181: \hline
1182: \hline
1183: model & $\gamma$ & $K$ & $\rho_{\rm c}$ [km$^{-2}$] & $M[M_{\odot}]$
1184: & $R$ [km] \\
1185: \hline
1186: 1 & $1.75$ & $25\,\, {\rm km}^{1.5}$ &
1187: $0.00125$ & 1.506 & 12.593 \\
1188: 2 & $2.00$ & $100\, {\rm km}^2$ &
1189: $0.0015$ & 1.130 & 9.653 \\
1190: 3 & $2.00$ & $150\,\, {\rm km}^2$ &
1191: $0.0015$ & 1.554 & 10.828 \\
1192: 4 & $2.00$ & $200\, {\rm km}^2$ &
1193: $0.0015$ & 1.878 & 11.646 \\
1194: 5 & $2.30$ & $1800\, {\rm km}^{2.6}$ &
1195: $0.0010$ & 1.756 & 11.710 \\
1196: \hline
1197: \hline
1198: \end{tabular}
1199: \label{MODELS15}
1200: \end{center}
1201: \end{table}
1202: %
1203: %
1204: In the remainder of this work we will refer to these stellar models as
1205: models 1-5.
1206: The code we have used for the calculation is based on the quadrature method
1207: described in section \ref{TOV_NUM} and uses a fourth order
1208: Runge-Kutta scheme for the integration (see for
1209: example \shortciteNP{Press1989}). We note, however,
1210: that the results of the relaxation
1211: method agree with those of the quadrature scheme with high precision and the
1212: corresponding plots are indistinguishable from those we show in this section.
1213: For the calculations in this section we use the rescaled coordinate $y$ and
1214: set $r_{,x}=C$ in Eq.\,(\ref{TOV_RY}).
1215: The code has been checked for convergence by calculating
1216: models 1-5 for different grid resolutions starting with 250 grid points.
1217: The resulting convergence factors $Q$ for the variables $\lambda$,
1218: $\mu$ and $\rho$ obtained for doubling the grid resolution is shown in
1219: Table \ref{CONV_TOV} for all 5 models. The high resolution
1220: reference solution has
1221: been calculated for 2000 grid points in all cases. For the fourth order
1222: Runge-Kutta scheme we would expect a convergence factor of 16. Even though
1223: the results show some variation around this value they are compatible
1224: with fourth order convergence.\\
1225: The numerical results obtained for the 5 stellar models we will now discuss
1226: %
1227: \begin{table}[t]
1228: \caption{The convergence factors obtained for doubling the grid resolution
1229: in a fourth order Runge-Kutta scheme for solving the TOV-equations
1230: via quadrature. The high resolution reference solution has been
1231: calculated for 2000 grid points.}
1232: \begin{center}
1233: \begin{tabular}{c|ccc}
1234: \hline \hline
1235: model & $Q_{\lambda}$ & $Q{\mu}$ & $Q_{\rho}$ \\
1236: \hline
1237: 1 & 14.23 & 15.55 & 9.69 \\
1238: 2 & 12.85 & 13.72 & 16.23 \\
1239: 3 & 17.98 & 18.40 & 18.76 \\
1240: 4 & 17.81 & 18.14 & 17.94 \\
1241: 5 & 11.64 & 16.51 & 21.13 \\
1242: \hline \hline
1243: \end{tabular}
1244: \label{CONV_TOV}
1245: \end{center}
1246: \end{table}
1247: %
1248: have all been calculated by
1249: using about 600 grid points.
1250: In Fig.\,\ref{TOV_GAMMA} we plot the metric functions $\lambda$, $\mu$,
1251: the energy density $\rho$, the pressure $P$,
1252: the mass $m$ and the sound speed $C$ as functions of the areal radius
1253: $r$ for models 1, 3 and 5. We note that the different central densities of
1254: these
1255: models have no impact on the qualitative behaviour of the solutions and have
1256: only been chosen to obtain neutron star models of similar size.
1257: %
1258: \begin{figure}[t]
1259: \centering
1260: \epsfig{file=r_lambdaGamma.eps, height=200pt, width=150pt, angle=-90}
1261: \vspace{0.5cm}
1262: \hspace{0.5cm}
1263: \epsfig{file=r_muGamma.eps, height=200pt, width=150pt, angle=-90}
1264: \vspace{0.5cm}
1265: \epsfig{file=r_rhoGamma.eps, height=200pt, width=150pt, angle=-90}
1266: \hspace{0.5cm}
1267: \epsfig{file=r_pGamma.eps, height=200pt, width=150pt, angle=-90}
1268: \epsfig{file=r_mGamma.eps, height=200pt, width=150pt, angle=-90}
1269: \hspace{0.5cm}
1270: \epsfig{file=r_cGamma.eps, height=200pt, width=150pt, angle=-90}
1271: \caption{The metric functions $\lambda$, $\mu$, the energy density $\rho$,
1272: the pressure $P$, the mass $m$ and the speed of
1273: sound $C$ are plotted
1274: as functions of radius for different polytropic
1275: indices $\gamma=1.75$ (model 1), $\gamma=2.00$ (model 3)
1276: and $\gamma=2.3$ (model 5).}
1277: \label{TOV_GAMMA}
1278: \end{figure}
1279: %
1280: The results demonstrate the dependence of the behaviour of the star near
1281: its surface on the polytropic
1282: exponent $\gamma$. According to the asymptotic analysis we expect the
1283: gradient of the energy density to be zero for $\gamma=1.75$ in model 1,
1284: finite for the critical case $\gamma=2$ in model 3 and infinite for
1285: model 5 where $\gamma=2.3$. This result is compatible with the plots
1286: of $\rho(r)$ in the middle left panel of Fig.\,\ref{TOV_GAMMA}.
1287: The pressure gradient
1288: on the other hand vanishes at the surfaces for any equation of state
1289: with positive $n$ according to Eq.\,(\ref{TOV_ASSYMPTOTICP})
1290: which agrees with the numerical results in the middle right panel. The
1291: speed of sound shows the opposite behaviour and has an infinite gradient
1292: independent of the polytropic index which is in
1293: agreement with the asymptotic result given by
1294: Eq.\,(\ref{TOV_ASSYMPTOTICC2}). With respect to the metric
1295: we note that the radial component $\mu$ has a local maximum, while the
1296: lapse $\lambda$ is monotonically increasing in the stellar interior.
1297: This behaviour becomes clear if we look at the corresponding equations
1298: for $\lambda$ and $\mu$. We already know that $\mu_{,r}$ vanishes
1299: at the centre. If we differentiate Eq.\,(\ref{TOV_MUR}) with respect to
1300: $r$ only one term on the right hand side is non zero at the centre,
1301: so that
1302: %
1303: \begin{align}
1304: \mu_{,rr}|_{r=0} &= 4\pi \rho_{\rm c},
1305: \end{align}
1306: %
1307: and $\mu_{,r}$ will
1308: %
1309: \begin{figure}[t]
1310: \centering
1311: \epsfig{file=r_lambdaK.eps, height=200pt, width=150pt, angle=-90}
1312: \vspace{0.5cm}
1313: \hspace{0.5cm}
1314: \epsfig{file=r_muK.eps, height=200pt, width=150pt, angle=-90}
1315: \vspace{0.5cm}
1316: \epsfig{file=r_rhoK.eps, height=200pt, width=150pt, angle=-90}
1317: \hspace{0.5cm}
1318: \epsfig{file=r_pK.eps, height=200pt, width=150pt, angle=-90}
1319: \epsfig{file=r_mK.eps, height=200pt, width=150pt, angle=-90}
1320: \hspace{0.5cm}
1321: \epsfig{file=r_cK.eps, height=200pt, width=150pt, angle=-90}
1322: \caption{The metric functions $\lambda$, $\mu$, the energy density
1323: $\rho$, the pressure $P$, the mass $m$ and the speed of sound $C$
1324: are plotted as functions of $r$ for different polytropic
1325: factors $K=100\,\,{\rm km}^2$ (model 2), $K=150\,\,{\rm km}^2$
1326: (model 3) and $K=200\,\,{\rm km}^2$ (model 4).}
1327: \label{TOV_K}
1328: \end{figure}
1329: %
1330: become positive as $r$ increases. At some point in the star,
1331: however, the negative first term on the right hand side of Eq.\,(\ref{TOV_MUR})
1332: will dominate the positive second term which goes to zero at the
1333: surface and $\mu_{,r}$ will become negative. Since Eq.\,(\ref{TOV_MUR})
1334: admits only one positive solution for $\mu$ if $\mu_{,r}=0$, $\mu$ will
1335: monotonically decrease beyond this point.
1336: We have already seen, however, that it cannot decrease to
1337: $1$ or below inside the star since this conflicts with the nonzero
1338: mass $m$ in
1339: Eq.\,(\ref{TOV_MUOFM}). Consequently $\mu>1$ inside the star and the
1340: right hand side of Eq.\,(\ref{TOV_LAMBDAR}) will be positive
1341: throughout the star which explains the monotonic behaviour of $\lambda$. \\
1342: In order to study the dependence of the solutions on the polytropic
1343: factor $K$ we compare the
1344: numerical results for models 2, 3 and 4 in Fig.\,\ref{TOV_K}. In contrast to
1345: the polytropic exponent, a variation of $K$ does not qualitatively change
1346: the results. A larger factor $K$ leads to a larger mass
1347: and radius of the neutron star model if all other parameters are kept fixed.
1348: This behaviour has been observed for various polytropic models
1349: and central densities
1350: and can be attributed to the larger pressure that follows from a larger $K$
1351: according to Eq.\,(\ref{POLYTROPE}). The star will thus be able to support more
1352: mass against self gravitation and extend to larger radii. \\
1353: We conclude the analysis of the TOV equations by studying the 1-parameter
1354: families of solutions corresponding to the five stellar models. For this purpose
1355: numerous solutions of the TOV equations with equations of state as given
1356: in Table \ref{MODELS15} have been
1357: calculated for various central densities. In Fig.\,\ref{TOV_FAMILIES} we
1358: plot the results in the form
1359: of relations between central density $\rho_{\rm c}$, total radius $R$,
1360: and total mass $M$ of the star.
1361: %
1362: \begin{figure}[t]
1363: \centering
1364: \epsfig{file=m_rhocGamma.eps, height=200pt, width=150pt, angle=-90}
1365: \epsfig{file=m_rGamma.eps, height=200pt, width=150pt, angle=-90}
1366: \epsfig{file=r_rhocGamma.eps, height=200pt, width=150pt, angle=-90}
1367: \caption{The 1-parameter families of static spherically symmetric neutron
1368: star models corresponding to models 1, 3 and 5
1369: are graphically illustrated by plotting the
1370: relations between the total mass, the central energy density
1371: and the radius of the star. The locations of neutron star
1372: models 1, 3 and 5 are indicated by crosses.}
1373: \label{TOV_FAMILIES}
1374: \end{figure}
1375: %
1376: One obvious result is the maximum of the mass curves $M(R)$ and
1377: $M(\rho_{\rm c})$ in the upper panels of the figure. It is a well known
1378: result that these maxima separate the stable and unstable branches of the
1379: neutron star families for a given equation of state
1380: (see for example \citeNP{Shapiro1983}).
1381: The stable branches consist of models with central densities below the
1382: critical value
1383: i.e. larger radii and the unstable branches correspond
1384: to larger central densities
1385: and smaller radii. In this context instability means that the frequency of
1386: the fundamental radial oscillation mode of the neutron star becomes imaginary
1387: and thus its amplitude will grow exponentially in time and the
1388: neutron star is unstable against arbitrarily small radial perturbations. The
1389: eigenmode spectrum of radial oscillations will be discussed in the next
1390: section when we look at dynamic spherically symmetric stars. \\
1391: Another interesting result is shown in the lower panel
1392: of Fig.\,\ref{TOV_FAMILIES} where
1393: we plot the radius as a function of the central density. The polytropic
1394: exponent $\gamma=2$ again appears as a critical value for which
1395: the radius converges to a finite value as the central density goes
1396: to zero. For smaller exponents the radius diverges in this
1397: limit whereas it goes to zero for exponents $\gamma>2$. We also discover
1398: this behaviour in the upper right panel where the mass is plotted as a
1399: function of radius. For $\gamma<2$ a unique value of $M$ can be assigned to
1400: any sufficiently large radius $R$.
1401: In the critical case $\gamma=2$ equilibrium models
1402: are only found for radii below a maximal value, but the relation $M(R)$ is
1403: still one to one near this maximum.
1404: For $\gamma>2$ this is no longer the case and
1405: for radii just below the maximal equilibrium radius we find two models
1406: with different mass. No such qualitatively different behaviour has been found
1407: when the polytropic factor $K$ is varied instead of $\gamma$.
1408: It is interesting to compare the mass-radius relation for
1409: $\gamma=2$ with the Newtonian case, where $\gamma=2$ is also a critical
1410: value and leads to the relation $R\sim M^0={\rm const}$
1411: (\citeNP{Shapiro1983}). The results in Fig.\,\ref{TOV_FAMILIES} indicate
1412: that relativistic effects break this kind of degeneracy. \\
1413: This completes our analysis of static spherically symmetric stars and in the
1414: next section we turn our attention to the dynamic case. The equations and
1415: results of this section will then be used to derive a fully
1416: non-linear perturbative formulation of radial oscillations on a static
1417: TOV background.
1418:
1419:
1420:
1421: %=========================================================================
1422: \subsection{Spherically symmetric dynamic stars in Eulerian coordinates}
1423: \label{DYNAMIC}
1424: %
1425: %
1426: %The study of radial oscillations of neutron stars is frequently carried
1427: %out using the linearized form of the equations. We will see in this section,
1428: %however, that a fully non-linear formulation of the problem leads to
1429: %a whole range of qualitatively new difficulties as well as
1430: % interesting features.
1431: %We have good reason to expect many of these difficulties and features to
1432: %also occur in more complicated oscillation modes in 2 or 3-dimensional
1433: %simulations. It is therefore well worth studying fully non-linear
1434: %radial oscillations with particular emphasis on non-linear effects and the
1435: %numerical problems resulting thereof.
1436: In this section we will develop
1437: an Eulerian formulation of a dynamic spherically symmetric neutron star.
1438: For code testing purposes
1439: it is interesting to also look at the corresponding scenario in the
1440: Cowling approximation, i.e. with the metric frozen at its equilibrium values.
1441: We will then use the results of the previous section to obtain a fully
1442: non-linear perturbative formulation of the problem. In this new
1443: approach to studying non-linear neutron star oscillations we eliminate
1444: terms of zero order in the perturbations but keep all higher order terms
1445: and thus obtain a formulation of the dynamic star which is equivalent to the
1446: original non-perturbative set of equations.
1447: % that
1448: %allows us to simulate oscillations of arbitrary amplitude with high accuracy.
1449: From the non-linear perturbative formulation it is easy to
1450: derive the linearized equations which we will use to investigate the eigenmode
1451: spectrum of radial neutron star oscillations. After describing the numerical
1452: methods used to evolve the dynamic neutron star in the non-linear case
1453: we have to discuss the ``surface problem'' which is intrinsic to any
1454: Eulerian formulation of non-linear oscillations that involve a radial
1455: displacement of the stellar surface. The numerical methods we have
1456: used to circumvent this problem will then be tested by comparing the
1457: numerical results obtained in the linear regime with the analytic solution of
1458: the linearized equations. By using vacuum flat space as the background,
1459: we can emulate a non-perturbative ``standard'' approach to the numerical
1460: evolution and compare the results with the perturbative scheme using
1461: the TOV background. Even though the perturbative
1462: scheme leads to highly accurate results for most stellar
1463: models, we have not been able to find a perfectly satisfactory solution
1464: to the surface problem.
1465: We have therefore decided to follow a more cautious approach and use a
1466: simplified neutron star model to investigate non-linear effects in the
1467: evolution of radial oscillations. This model has also been used to
1468: further test the performance of the code. The surface problem will be
1469: re-addressed with a Lagrangian approach in section \ref{LAGR}.
1470:
1471:
1472:
1473:
1474: %=========================================================================
1475: \subsubsection{The equations in the dynamic case}
1476: \label{NONP_EQ}
1477: %
1478: %
1479: We start the Eulerian formulation of the dynamic case with
1480: the line element in radial gauge and polar slicing
1481: %
1482: \begin{align}
1483: ds^2 &= -\hat{\lambda}^2 dt^2 + \hat{\mu}^2 dr^2
1484: + r^2 (d\theta^2 + \sin^2 \theta d\phi^2),
1485: \label{NONP_LINEELEMENT}
1486: \end{align}
1487: %
1488: where $\hat{\lambda}$ and $\hat{\mu}$ are now functions of $t$ and $r$ and
1489: the ``hat'' has been introduced to distinguish them from their static
1490: counterparts. As in the static case we describe the matter as a perfect fluid
1491: at zero temperature with a polytropic equation of state. As we have seen
1492: in section \ref{TOV_EQ} this enables us to write the energy
1493: momentum tensor in the form
1494: %
1495: \begin{align}
1496: \hbox{\vec{T}}^{\mu \nu} &= (\hat{\rho} +\hat{P}) \hbox{\vec{u}}^{\mu}
1497: \hbox{\vec{u}}^{\nu} + \hat{P} \hbox{\vec{g}}^{\mu \nu},
1498: \label{NONP_EMTENSOR}
1499: \end{align}
1500: %
1501: where again the ``hat'' on the functions $\hat{\rho}$, $\hat{P}$ means that
1502: they are functions of $t$ and $r$. The time dependent pressure and
1503: energy density are related by the polytropic law
1504: %
1505: \begin{align}
1506: \hat{P} &= K\hat{\rho}^{\gamma}, \label{DYNAMICPOLYTROPE}
1507: \end{align}
1508: %
1509: where the polytropic parameters $\gamma$ and $K$ are the same as in the
1510: static case. The time dependent speed of sound is defined in analogy to
1511: Eq.\,(\ref{TOV_C2}) by
1512: %
1513: \begin{align}
1514: \hat{C}^2 &= \frac{\partial \hat{P}}{\partial \hat{\rho}}. \label{PERT_C2}
1515: \end{align}
1516: %
1517: In contrast to the static case the
1518: 4-velocity will now have a non-vanishing radial component
1519: %
1520: \begin{align}
1521: \hbox{\vec{u}}^{\mu} &= (v,w,0,0),
1522: \end{align}
1523: %
1524: where $v=v(r,t)$ and $w=w(r,t)$. We have
1525: not denoted these quantities by a ``hat'' since we do not use
1526: static counterparts in their case. The normalisation condition
1527: $\hbox{\vec{u}}^{\mu}
1528: \hbox{\vec{u}}_{\mu}=-1$ relates these functions by
1529: %
1530: \begin{align}
1531: \hat{\lambda}^2 v^2 &= 1+\hat{\mu}^2 w^2. \label{PERT_NORMU}
1532: \end{align}
1533: %
1534: With the line element (\ref{NONP_LINEELEMENT}) and the energy momentum tensor
1535: (\ref{NONP_EMTENSOR}) the Einstein field equations
1536: $\hbox{\vec{G}}_{\mu \nu} = 8\pi \hbox{\vec{T}}_{\mu \nu}$
1537: result in two independent constraint equations
1538: %
1539: \begin{align}
1540: \frac{\hat{\lambda}_{,r}}{\hat{\lambda}} &= \frac{\hat{\mu}^2-1}{2r} + 4\pi r
1541: \hat{\mu}^2 \left[ \hat{P} + (\hat{\rho} + \hat{P})
1542: \,\hat{\mu}^2w^2 \right], \label{NONP_LAMBDAR} \\[10pt]
1543: \frac{\hat{\mu}_{,r}}{\hat{\mu}} &= -\frac{\hat{\mu}^2-1}{2r} + 4\pi r
1544: \hat{\mu}^2 \left[ \hat{\rho} + (\hat{\rho} + \hat{P})
1545: \hat{\mu}^2w^2 \right]. \label{NONP_MUR}
1546: \end{align}
1547: %
1548: It is a well known result that there are no gravitational degrees of freedom
1549: in spherical symmetry and we therefore expect to be able to determine
1550: the metric functions on each time slice without knowledge of their history.
1551: This is compatible with the result that the field equations can be given
1552: in the form of constraint equations only. The degrees of freedom of the
1553: scenario are thus entirely contained in the matter variables, whose evolution
1554: is determined by the equations of hydrodynamics
1555: $\nabla_{\mu} \hbox{\vec{T}}^{\mu \nu}=0$. In our case we can write these
1556: equations as a quasi linear system of PDEs
1557: %
1558: \begin{align}
1559: \hat{\rho}_{,t} + \tilde{\alpha}_{11} \hat{\rho}_{,r}
1560: + \tilde{\alpha}_{12} w_{,r} &= \tilde{b}_1,
1561: \label{NONP_RHOT} \\[10pt]
1562: w_{,t} + \tilde{\alpha}_{21} \hat{\rho}_{,r}
1563: + \tilde{\alpha}_{11} w_{,r} &= \tilde{b}_2,
1564: \label{NONP_WT}
1565: \end{align}
1566: %
1567: where the coefficients are given by
1568: %
1569: \begin{align}
1570: D =&\,\, v\left( 1- \hat{C}^2\frac{\hat{\mu}^2 w^2}{1+ \hat{\mu}^2 w^2}
1571: \right), \label{NONP_D} \\[10pt]
1572: \tilde{\alpha}_{11} =&\,\, \frac{w(1-\hat{C}^2)}{D}, \\[10pt]
1573: \tilde{\alpha}_{12} =&\,\, \frac{\hat{\rho} + \hat{P}}{(1+\hat{\mu}^2 w^2) D},
1574: \\[10pt]
1575: \tilde{\alpha}_{21} =&\,\, \frac{\hat{C}^2}{(\hat{\rho} + \hat{P})
1576: \hat{\mu}^2 D}, \label{NONP_ALPHA21} \\[10pt]
1577: \tilde{b}_1 =& \,\,-\frac{1}{D}(\hat{\rho} + \hat{P})
1578: \left(\frac{w\hat{\mu}_{,r}/\hat{\mu}
1579: +v\hat{\mu}_{,t}/\hat{\mu}}
1580: {1+\hat{\mu}^2 w^2} + 2\frac{w}{r} \right), \\[10pt]
1581: \tilde{b}_2 =& \,\, -\frac{1}{D}\left[
1582: w^2\left( \frac{\hat{\mu}_{,r}}{\hat{\mu}}
1583: +\frac{\hat{\lambda}_{,r}}{\hat{\lambda}}
1584: -\frac{2}{r}\hat{C}^2
1585: +2\frac{v}{w} \frac{\hat{\mu}_{,t}}{\hat{\mu}}
1586: \right)
1587: +\frac{\hat{\lambda}_{,r}/\hat{\lambda}}{\hat{\mu}^2}
1588: \right]. \label{NONP_B2}
1589: \end{align}
1590: %
1591: In practice we calculate the derivatives of the metric functions
1592: that appear in these coefficients from the constraint equations
1593: (\ref{NONP_LAMBDAR}), (\ref{NONP_MUR}) and a third field equation
1594: %
1595: \begin{align}
1596: \frac{\hat{\mu}_{,t}}{\hat{\mu}} &= -4\pi r \hat{\mu}^2 \hat{\lambda}^2
1597: v w (\hat{\rho} + \hat{P}), \label{NONP_MUT}
1598: \end{align}
1599: %
1600: which is an automatic consequence of the two constraints, their derivatives
1601: and the matter equations.
1602: We therefore calculate the coefficients $\tilde{\alpha}_{ij}$ and
1603: $b_{i}$ without approximating any derivatives with finite difference
1604: expressions. \\
1605: We have already mentioned in the discussion of the static case that a
1606: numerically superior performance is obtained if we transform to a new
1607: radial coordinate $y$ defined by Eq.\,(\ref{TOV_YOFR}). We note however that
1608: we need to calculate the corresponding static model first to obtain the
1609: static sound speed $C$. In the perturbative approach which we will discuss
1610: below that is done as a matter of course. There we will provide
1611: a formulation of the perturbative equations that includes both choices
1612: for the radial coordinate analogous to Eqs.\,(\ref{TOV_RY})-(\ref{TOV_PY}).
1613: In the Cowling approximation the set of equations corresponding
1614: to (\ref{NONP_RHOT})-(\ref{NONP_B2}) describes
1615: a dynamic, spherically symmetric perfect fluid in a fixed gravitational
1616: potential. We obtain these equations by the following modifications:
1617: %
1618: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
1619: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin0cm
1620: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
1621: \item the constraint equations for the dynamic metric functions
1622: (\ref{NONP_LAMBDAR}), (\ref{NONP_MUR}) are replaced by the
1623: corresponding TOV equations (\ref{TOV_LAMBDAR}), (\ref{TOV_MUR})
1624: which have to be solved only at the start of the evolution,
1625: \item in the coefficients $\tilde{\alpha}_{11}$, $\tilde{\alpha}_{12}$,
1626: $\tilde{\alpha}_{21}$ and $\tilde{b}_1$ all occurrences
1627: of $\hat{\mu}$, $\hat{\lambda}$,
1628: $\hat{\lambda}_{,r}$/$\hat{\lambda}$ and $\hat{\mu}_{,r}/\hat{\mu}$
1629: are replaced with their static
1630: analogues $\lambda$, $\mu$, $\lambda_{,r}/\lambda$ and $\mu_{,r}/\mu$
1631: respectively and $\hat{\mu}_{,t}/\hat{\mu}$ is set to zero,
1632: \item the coefficient function $\tilde{b}_2$ is replaced with the
1633: slightly modified version
1634: %
1635: \begin{align}
1636: \bar{b}_2 =& \,\, -\frac{1}{D}\left\{
1637: w^2\left[ \left( \frac{\mu_{,r}}{\mu}
1638: +\frac{\lambda_{,r}}{\lambda} \right)
1639: \left(1-\hat{C}^2 \right) -\frac{2}{r}\hat{C}^2 \right]
1640: +\frac{\lambda_{,r}/\lambda}{\mu^2}
1641: \right\}.
1642: \end{align}
1643: %
1644: \end{list}
1645: %
1646: These modifications are rather simple so that we incorporate both options, the
1647: evolution with time dependent metric and the Cowling
1648: approximation in one code. A user specified initial parameter determines
1649: which version is to be run. Before we describe the numerical
1650: implementation, we need to rewrite the equations of this subsection
1651: in a perturbative form.
1652:
1653:
1654: %=========================================================================
1655: \subsubsection{A fully non-linear perturbative formulation}
1656: \label{PERT_EQ}
1657: %
1658: %
1659: In this section we will decompose the time dependent quantities
1660: $\hat{\lambda}$, $\hat{\mu}$ and $\hat{\rho}$ into static
1661: background contributions and time dependent perturbations. We will see
1662: that the TOV equations are still present in the dynamic equations,
1663: for example in the terms
1664: $\alpha_{21} \hat{\rho}_{,r} - b_2$ in Eq.\,(\ref{NONP_WT}). It is the
1665: elimination of these zero order terms and the ensuing numerical
1666: inaccuracies which provides the motivation for our perturbative formulation.
1667: We start by decomposing the time dependent functions into a static
1668: background plus a time dependent perturbation
1669: %
1670: \begin{align}
1671: \hat{\lambda}(t,r) &= \lambda(r) + \delta \lambda(t,r), \label{PERT_LAMBDA}
1672: \\[10pt]
1673: \hat{\mu}(t,r) &= \mu(r) + \delta \mu(t,r), \label{PERT_MU} \\[10pt]
1674: \hat{\rho}(t,r) &= \rho(r) + \delta \rho(t,r), \label{PERT_RHO} \\[10pt]
1675: \hat{P}(t,r) &= P(r) + \delta P(t,r). \label{PERT_P}
1676: \end{align}
1677: %
1678: The radial velocity component $w$ vanishes in the static limit and therefore
1679: represents a perturbation in itself. The time dependent functions
1680: $\hat{P}$, $\hat{C}$ and $v$ are dependent variables and
1681: thus considered functions of the fundamental variables
1682: $\hat{\lambda}$, $\hat{\mu}$, $\hat{\rho}$ and $w$ according to
1683: Eqs.\,(\ref{DYNAMICPOLYTROPE}), (\ref{PERT_C2}) and (\ref{PERT_NORMU}).
1684: We stress that the perturbations are finite and that no assumption with
1685: regard to their size has been made. \\
1686: We start rewriting the dynamic
1687: equations with the constraint equation for $\hat{\lambda}$. If we
1688: insert Eqs.\,(\ref{PERT_LAMBDA})-(\ref{PERT_RHO}) into (\ref{NONP_LAMBDAR})
1689: and multiply with $\hat{\lambda}$ we obtain
1690: %
1691: \begin{align}
1692: \begin{split}
1693: \lambda_{,r} + \delta \lambda_{,r} =& \,\,\, \lambda
1694: \frac{\mu^2-1}{2r} + 4\pi r
1695: \lambda \mu^2 P + 4\pi r \lambda \mu^2
1696: \left[\delta P + (\rho + \delta \rho + \hat{P}) \hat{\mu}^2 w^2 \right]
1697: + \lambda \frac{2\mu \delta \mu + \delta \mu^2}{2r}
1698: \\[10pt]
1699: & + \delta \lambda \frac{\hat{\mu}^2-1}{2r}
1700: +4\pi r \left[\lambda(2\mu \delta \mu + \delta \mu^2) + \delta \lambda
1701: \hat{\mu}^2 \right] \left[ \hat{P}+(\rho + \delta \rho + \hat{P})
1702: \hat{\mu}^2 w^2 \right]. \label{COMPLETE_LAMBDAR}
1703: \end{split}
1704: \end{align}
1705: %
1706: The crucial terms are the first on the left and the first
1707: two terms on the right hand side. We know that these terms
1708: will cancel each other
1709: identically according to Eq.\,(\ref{TOV_LAMBDAR}) if a solution of the
1710: static equations is chosen as a background. Numerically, however,
1711: this will not be the case because of truncation errors.
1712: % This becomes immediately
1713: %obvious if we assume $\lambda$, $\mu$ and $P$ given as an exact solution
1714: %of the TOV equations.
1715: This residual error will inevitably contaminate
1716: the numerical evolution of the dynamic scenario. In other words the numerical
1717: accuracy we will obtain is limited by the numerical accuracy of the static
1718: background and not by that of the dynamic signal we are interested in.
1719: The severeness of this effect will depend on the relative size of the
1720: perturbations with respect to the background. For very large perturbations
1721: the numerical contamination will be less significant and for very small
1722: perturbations we may satisfy ourselves with a linearized code. For
1723: perturbations of intermediate strength, however, which are still
1724: smaller than the background but are large enough to give rise to non-linear
1725: effects, the numerical contamination will severely affect the
1726: evolution and may give rise to spurious phenomena. \\
1727: % The scenario many studies
1728: %of neutron star oscillations will ultimately target is an
1729: %oscillation mode which is unstable to the emission of gravitational waves.
1730: %Such a mode, starting with a small amplitude, will increase
1731: %in strength and gradually enter the non-linear regime. It is therefore
1732: %of particular interest to investigate the impact of non-linear effects
1733: %in the weakly and moderately non-linear regime. Such a study will necessarily
1734: %be a study in 2 or 3 spatial dimensions and is beyond the scope of this
1735: %work. The development of a fully non-linear numerical technique which
1736: %provides high accuracy for perturbations of any given amplitude will,
1737: %however, pave some of the way towards achieving this goal. A further benefit
1738: %of the high accuracy of this method is the possibility of testing the code
1739: %in the linear regime, the only regime where an exact solution is known
1740: %with (for all practical purposes) arbitrary accuracy. \\
1741: We return to Eq.\,(\ref{COMPLETE_LAMBDAR}) and continue the
1742: perturbative formulation of the dynamic case. Since we know that the
1743: zero order terms cancel each other, we can simply subtract them
1744: from the equation.
1745: % We thus keep higher order terms in the equations only
1746: %and ensure that the numerical accuracy we obtain will always be the
1747: %accuracy of the signal we are interested in.
1748: The perturbative equation for $\hat{\lambda}$ then becomes
1749: %
1750: \begin{align}
1751: \begin{split}
1752: \frac{\delta \lambda_{,x}}{r_{,x}}
1753: =& \,\,\, \lambda \frac{2\mu \delta \mu + \delta \mu^2}{2r}
1754: + \delta \lambda \frac{\hat{\mu}^2-1}{2r} + 4\pi r \lambda \mu^2
1755: \left[\delta P + (\rho + \delta \rho + \hat{P}) \hat{\mu}^2 w^2 \right]
1756: \\[10pt]
1757: & +4\pi r \left[\lambda(2\mu \delta \mu + \delta \mu^2) + \delta \lambda
1758: \hat{\mu}^2 \right] \left[ \hat{P}+(\rho + \delta \rho + \hat{P})
1759: \hat{\mu}^2 w^2 \right], \label{PERT_LAMBDAR}
1760: \end{split}
1761: \end{align}
1762: %
1763: where we have also implemented the transformation to the generalised
1764: radial coordinate $x$.
1765: Proceeding in the same way we rewrite the constraint equation for $\hat{\mu}$
1766: %
1767: \begin{align}
1768: \begin{split}
1769: \frac{\delta \mu_{,x}}{r_{,x}}
1770: =&\,\,\, -\mu \frac{2\mu \delta \mu + \delta \mu^2}{2r}
1771: -\delta \mu \frac{\hat{\mu}^2-1}{2r}
1772: + 4\pi r \mu^3 \left[ \delta \rho + (\rho + \delta \rho + \hat{P})
1773: \hat{\mu}^2 w^2 \right] \\[10pt]
1774: & +4\pi r (3 \mu^2 \delta \mu + 3 \mu \delta \mu^2 + \delta \mu^3)
1775: \left[\hat{\rho} + (\rho + \delta \rho + \hat{P})\hat{\mu}^2 w^2
1776: \right]. \label{PERT_MUR}
1777: \end{split}
1778: \end{align}
1779: %
1780: The reformulation of the matter equations (\ref{NONP_RHOT}) and (\ref{NONP_WT})
1781: is particularly simple due to their quasi linear nature. We obtain
1782: %
1783: \begin{align}
1784: \delta \rho_{,t} + \alpha_{11} \delta \rho_{,x} + \alpha_{12} w_{,x} &= b_1,
1785: \label{PERT_DRHOT} \\[10pt]
1786: w_{,t} + \alpha_{21} \delta \rho_{,x} + \alpha_{11} w_{,x} &= b_2,
1787: \label{PERT_WT}
1788: \end{align}
1789: %
1790: with the coefficient functions
1791: %
1792: \begin{align}
1793: D =&\,\, v\left( 1- \hat{C}^2\frac{\hat{\mu}^2 w^2}{1+ \hat{\mu}^2 w^2}
1794: \right), \label{PERT_D} \\[10pt]
1795: \alpha_{11} =&\,\, \frac{w(1-\hat{C}^2)}{r_{,x}D},
1796: \label{PERT_ALPHA11} \\[10pt]
1797: \alpha_{12} =&\,\, \frac{\rho + \delta \rho
1798: + \hat{P}}{(1+\hat{\mu}^2 w^2)r_{,x} D},
1799: \label{PERT_ALPHA12} \\[10pt]
1800: \alpha_{21} =&\,\, \frac{\hat{C}^2}{(\rho +\delta \rho + \hat{P})
1801: \hat{\mu}^2r_{,x} D}, \label{PERT_ALPHA21} \\[10pt]
1802: b_1 =& \,\,-\frac{1}{D}\left[(\rho + \delta \rho + \hat{P})
1803: \left(\frac{w\hat{\mu}_{,r}/\hat{\mu}
1804: +v\hat{\mu}_{,t}/\hat{\mu}}
1805: {1+\hat{\mu}^2 w^2} + 2\frac{w}{r} \right)
1806: + \rho_{,r} w (1-C^2)\right], \\[10pt]
1807: \begin{split}
1808: b_2 =& \,\, -\frac{1}{D}\left\{
1809: w^2\left( \frac{\hat{\mu}_{,r}}{\hat{\mu}}
1810: +\frac{\hat{\lambda}_{,r}}{\hat{\lambda}}
1811: -\frac{2}{r}\hat{C}^2
1812: +2\frac{v}{w} \frac{\hat{\mu}_{,t}}{\hat{\mu}}
1813: \right) +
1814: \frac{1}{\hat{\mu}^2 (\rho + \delta \rho + \hat{P})}
1815: \right. \\[10pt]
1816: & \left. \left[ (\hat{C}^2 - C^2) \rho_r
1817: + \frac{\delta \lambda}{\hat{\lambda}}
1818: C^2 \rho_r + \frac{\delta \lambda_r}{\hat{\lambda}} (\rho + P)
1819: + \frac{\hat{\lambda}_r}{\hat{\lambda}}(\delta \rho
1820: + \delta P) \right] \right\}. \label{PERT_B2}
1821: \end{split}
1822: \end{align}
1823: %
1824: Except for the coefficient $b_2$ where background terms have been eliminated
1825: by using the TOV-equations
1826: we note the similarity with the coefficients given
1827: in Eqs.\,(\ref{NONP_D})-(\ref{NONP_B2}) in the non-perturbative formulation. \\
1828: In order to derive the equations in the Cowling approximation we
1829: have to proceed in analogy to the previous section.
1830: %
1831: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
1832: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin0cm
1833: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
1834: \item The metric perturbations $\delta \mu$ and $\delta \lambda$ are
1835: set to zero.
1836: \item All occurrences of $\hat{\lambda}_{,r}/\hat{\lambda}$ and
1837: $\hat{\mu}_{,r}/\hat{\mu}$ are replaced with $\lambda_{,r}/\lambda$
1838: and $\mu_{,r}/\mu$ which are given by the TOV equations
1839: (\ref{TOV_LAMBDAR}), (\ref{TOV_MUR}).
1840: \item $\hat{\mu}_{,t}/\hat{\mu}$ is set to zero.
1841: \item The coefficient $b_2$ is replaced by
1842: %
1843: \begin{align}
1844: \begin{split}
1845: b_2 =& \,\, -\frac{1}{D}\left\{
1846: w^2\left[ \left( \frac{\hat{\mu}_{,r}}{\hat{\mu}}
1847: +\frac{\hat{\lambda}_{,r}}{\hat{\lambda}} \right)
1848: (1-\hat{C}^2)-\frac{2}{r}\hat{C}^2
1849: +2\frac{v}{w} \frac{\hat{\mu}_{,t}}{\hat{\mu}}
1850: \right] +
1851: \frac{1}{\hat{\mu}^2 (\rho + \delta \rho + \hat{P})}
1852: \right. \\[10pt]
1853: & \left. \left[ (\hat{C}^2 - C^2) \rho_r
1854: + \frac{\delta \lambda}{\hat{\lambda}} C^2 \rho_r
1855: +\frac{\delta \lambda_r}{\hat{\lambda}} (\rho + P)
1856: + \frac{\hat{\lambda}_r}{\hat{\lambda}}(\delta \rho
1857: + \delta P) \right] \right\}.
1858: \end{split}
1859: \end{align}
1860: %
1861: \end{list}
1862: %
1863: This completes our derivation of the equations for a dynamical spherically
1864: symmetric neutron star. In later sections we will numerically investigate
1865: the system of partial differential equations
1866: (\ref{PERT_LAMBDAR})-(\ref{PERT_WT}) with the
1867: coefficient functions (\ref{PERT_D})-(\ref{PERT_B2}) and
1868: the corresponding system in the Cowling approximation. Before that, we will
1869: turn our attention towards the linearized equations and the resulting
1870: eigenmode spectrum. These results will not only be used as initial data,
1871: but also provide one of the fundamental test beds for the code.
1872:
1873:
1874: %=========================================================================
1875: \subsubsection{The linearized equations and the eigenmode spectrum}
1876: \label{PERT_LIN}
1877: %
1878: %
1879: {\em (a) The equations} \\[5pt]
1880: In this section we will discuss the linearized equations for a dynamic
1881: spherically symmetric neutron star. For this purpose we will
1882: explicitly assume that the background is given by a non-vacuum solution
1883: of the TOV equations. If we further assume that all perturbations are small
1884: compared with their background
1885: values and the radial velocity $w$ is small compared with the speed of light,
1886: i.e. $w \ll 1$, the higher order terms in
1887: Eqs.\,(\ref{PERT_LAMBDAR})-(\ref{PERT_B2}) become negligible and can
1888: be omitted from the equations. It is convenient to follow e.g.
1889: \lcite{Misner1973} and introduce the variable $\xi$ which measures
1890: the displacement of the fluid elements. An observer who is comoving with the
1891: fluid and is located at $r_0$ in the equilibrium case will find herself at
1892: position $r_0+\xi(t,r_0)$ during the evolution. The displacement vector $\xi$
1893: is therefore related to our variables by
1894: %
1895: \begin{align}
1896: \xi_{,t} &= \lambda w. \label{LIN_XI}
1897: \end{align}
1898: %
1899: We note that the background value of the lapse function is used in this
1900: equation because higher order terms have been neglected.
1901: Another variable which facilitates a particularly simple formulation of the
1902: resulting equations is the rescaled displacement $\zeta$ defined by
1903: %
1904: \begin{align}
1905: \zeta &= \frac{r^2}{\lambda} \xi \label{LIN_ZETA}.
1906: \end{align}
1907: %
1908: If we insert this definition into the linearized form of
1909: equation (\ref{PERT_WT}) and use the linearized versions of
1910: Eqs.\,(\ref{PERT_LAMBDAR})-(\ref{PERT_DRHOT}) to eliminate the perturbations
1911: $\delta \lambda$, $\delta \mu$ and $\delta \rho$ we obtain the second order
1912: in time and space differential equation
1913: %
1914: \begin{align}
1915: W \zeta_{,tt} &= \frac{1}{r_{,x}}\left( \frac{\Pi}{r_{,x}}
1916: \zeta_{,x} \right)_{,x}
1917: + Q \zeta, \label{LIN_ZETATT}
1918: \end{align}
1919: %
1920: where the auxiliary functions $W$, $\Pi$ and $Q$ are defined by
1921: %
1922: \begin{align}
1923: \Pi &= C^2(\rho + P) \frac{\mu \lambda^3}{r^2}, \label{LIN_PI} \\[10pt]
1924: W &= (\rho +P) \frac{\mu^3 \lambda}{r^2}, \label{LIN_W} \\[10pt]
1925: Q &= \frac{\mu \lambda^3}{r^2}(\rho + P) \left[ \left(\frac{\lambda_{,r}}
1926: {\lambda} \right)^2 + 4\frac{\lambda_{,r}}{r\lambda} -8\pi \mu^2P
1927: \right].
1928: \label{LIN_Q}
1929: \end{align}
1930: %
1931: These equations describe the dynamics of a
1932: spherically symmetric neutron star in the linearized limit.
1933: If we insert the ansatz $\zeta(t,x)={\zeta(x)} f(t)$ into
1934: Eq.\,(\ref{LIN_ZETATT}) we find that
1935: the solution has harmonic time dependence
1936: %
1937: \begin{align}
1938: \zeta(t,x) &= {\zeta}(x) e^{i \omega t}.
1939: \end{align}
1940: %
1941: and the spatial profile is determined by the
1942: ordinary differential equation
1943: %
1944: \begin{align}
1945: \frac{1}{r_{,x}} \left( \frac{\Pi}{r_{,x}} \zeta_x\right)_x
1946: + (\omega^2 W + Q) \zeta=0. \label{LIN_ZETARR}
1947: \end{align}
1948: %
1949: %where we have omitted the ``tilde'' from the spatial function $\zeta(r)$.
1950: For the ensuing discussion it is convenient
1951: to work with the areal radius
1952: $r$ and therefore set $r_{,x}=1$. The ordinary differential equation
1953: (\ref{LIN_ZETARR}) can then be written in the form
1954: %
1955: \begin{align}
1956: \mathcal{L} \zeta &= -\omega^2 \zeta,
1957: \end{align}
1958: %
1959: where the differential operator $\mathcal{L}$ is defined by
1960: %
1961: \begin{align}
1962: \mathcal{L} &= \frac{1}{W}\left[ \frac{d}{dr} \left( \Pi \frac{d}{dr} \right)
1963: - Q\right].
1964: \end{align}
1965: %
1966: This type of ODE is called an {\em eigenvalue problem} and the
1967: particular structure of the differential operator $\mathcal{L}$
1968: classifies it as a {\em Sturm-Liouville problem} if the function $\zeta$
1969: satisfies so-called homogeneous boundary conditions
1970: (see for example \citeNP{Simmons1991}).
1971: Due to the asymptotic behaviour of the background
1972: solutions the functions $\Pi$, $W$ and $Q$ will either diverge or vanish
1973: at the boundaries, however, and the problem we are facing is
1974: a {\em singular Sturm-Liouville problem}. An important subclass of this type
1975: of problems is the {\em self-adjoint eigenvalue problem}
1976: which is defined by the requirement that
1977: %
1978: \begin{align}
1979: \langle \mathcal{L}u,v \rangle &= \langle \mathcal{L}v,u \rangle,
1980: \label{SELFADJOINED}
1981: \end{align}
1982: %
1983: for all solutions $u$, $v$. Here the inner product is defined by the
1984: {\em weighting function} $W(r)$
1985: %
1986: \begin{align}
1987: \langle f,g \rangle &= \int_a^b{W(r)\, f(r)\, g(r)\,dr},
1988: \label{LIN_INNERPRODUCT}
1989: \end{align}
1990: %
1991: where $a$ and $b$ are the boundaries, i.e. the centre and surface of the
1992: star in our case. A short calculation shows that condition
1993: (\ref{SELFADJOINED}) is ensured if the solutions satisfy the
1994: self-adjoint boundary condition
1995: %
1996: \begin{align}
1997: \left[ \vphantom{\frac{1}{x}} \Pi (v u_{,r} - u v_{,r}) \right]_a^b &= 0.
1998: \label{BCSELFADJOINED}
1999: \end{align}
2000: %
2001: Below we shall see that any solution $\zeta$ of the eigenvalue problem
2002: (\ref{LIN_ZETARR}) will be $\mathscr{O}(r^3)$ at the origin and
2003: be finite at the surface. In combination with the asymptotic behaviour of the
2004: TOV solutions determined in section \ref{TOV_ASYMPTOTICS} we can see that
2005: Eq.\,(\ref{BCSELFADJOINED}) is satisfied so that
2006: the differential equation (\ref{LIN_ZETARR}) represents a self-adjoint
2007: eigenvalue problem. For this type of equations one can show the
2008: following properties (see for example \citeNP{Coddington1955})
2009: %
2010: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
2011: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin0cm
2012: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
2013: \item There exist an infinite number of solutions $\zeta_1(r)$,
2014: $\zeta_2(r)$, $\zeta_3(r),\ldots$ which are called eigenfunctions
2015: and the corresponding eigenvalues are real and can be ordered
2016: %
2017: \begin{align}
2018: (\omega^2)_1 < (\omega^2)_2 < (\omega^2)_3 <\ldots.
2019: \end{align}
2020: %
2021: We note that in our case the real eigenvalues are $\omega^2$ and
2022: the corresponding frequencies will be imaginary if $\omega^2<0$.
2023: % Below we will see that this will indeed be the case for the fundamental
2024: % eigenmodes of gravitationally unstable neutron stars.
2025: \item After appropriate normalisation the eigenfunctions form an orthonormal
2026: set, i.e.
2027: %
2028: \begin{align}
2029: \langle \zeta_i, \zeta_j \rangle &= \delta_{i,j}.
2030: \label{LIN_ORTHONORMALITY}
2031: \end{align}
2032: %
2033: \item The eigenfunctions $\zeta_i$ form a complete set, i.e. any function
2034: $f(r)$ which satisfies the self-adjoint boundary conditions
2035: (\ref{SELFADJOINED}) can be expanded in a series of eigenmodes
2036: %
2037: \begin{align}
2038: f(r) &= \sum_i{A_i \zeta_i(r)},
2039: \end{align}
2040: %
2041: where the eigenmode coefficients of the function $f$ are given by
2042: %
2043: \begin{align}
2044: A_i &= \langle f, \zeta_i \rangle.
2045: \end{align}
2046: %
2047: \end{list}
2048: %
2049: Before we investigate Eq.\,(\ref{LIN_ZETARR}) numerically, we
2050: consider the asymptotic behaviour of the solutions. At the
2051: origin the displacement vectors $\xi$ and $\zeta$ have to vanish because
2052: of the spherical symmetry. If we therefore assume $\zeta(r) \sim r^{\alpha}$
2053: near the origin where $\alpha>0$, insert this ansatz into
2054: Eq.\,(\ref{LIN_ZETARR}) and use the asymptotic behaviour of the TOV solution,
2055: we obtain the leading order
2056: %
2057: \begin{align}
2058: \zeta(r) \sim \mathscr{O}(r^3). \label{LIN_ASSYMPTOTICZETA}
2059: \end{align}
2060: %
2061: At the surface we only require $\xi$ and $\zeta$ to be finite but allow
2062: for non-zero displacements
2063: %
2064: \begin{align}
2065: \zeta(z) \sim \mathscr{O}(z^0).
2066: \end{align}
2067: %
2068: It is of particular interest to consider the impact of these results
2069: on the asymptotic behaviour of the energy density perturbation
2070: $\delta \rho$ which is related to the displacement by the linearized
2071: version of Eq.\,(\ref{PERT_DRHOT})
2072: %
2073: \begin{align}
2074: \delta \rho &= -\frac{\lambda}{r^2} \left[(\rho + P) \zeta_{,r}
2075: + \rho_{,r} \zeta \right]. \label{LIN_DRHOOFZETA}
2076: \end{align}
2077: %
2078: At the centre the $r^3$ behaviour of the displacement $\zeta$ results in
2079: %
2080: \begin{align}
2081: \delta \rho \sim \mathscr{O}(r^0),
2082: \end{align}
2083: %
2084: so that the condition we imposed on $\zeta$ also guarantees a finite
2085: energy density perturbation at the origin. At the surface, however, the
2086: leading term on the right hand side of Eq.\,(\ref{LIN_DRHOOFZETA}) is the
2087: term involving the derivative of the background energy density. This
2088: term is responsible for the asymptotic behaviour of $\delta \rho$ at
2089: the surface
2090: %
2091: \begin{align}
2092: \delta \rho &\sim \mathscr{O}(z^{n-1}). \label{LIN_ASSYMPTOTICDRHO}
2093: \end{align}
2094: %
2095: Consequently the energy density perturbation is zero at the surface
2096: for $n > 1$, finite
2097: for $n=1$ and it diverges for $n<1$ i.e. $\gamma>2$. Even worse we
2098: also obtain the result
2099: %
2100: \begin{align}
2101: \frac{\delta \rho}{\rho} \sim \mathscr{O}(z^{-1})\label{DRHOOVERRHO}
2102: \end{align}
2103: %
2104: independent of the polytropic index.
2105: The energy density perturbation will therefore necessarily be larger than
2106: the background $\rho$ in a finite interval around the surface.
2107: This is in obvious conflict with the initial assumption
2108: $\delta \rho \ll \rho$ we used in the linearisation process and raises
2109: doubts about the validity of the results. Below we will see, however,
2110: that the linearized equations can be derived without any implicit
2111: contradiction from the fully non-linear Lagrangian formulation of
2112: the problem. This is already illustrated by a closer investigation of
2113: Eq.\,(\ref{LIN_DRHOOFZETA}) which can be rewritten as
2114: %
2115: \begin{align}
2116: \delta \rho &= \Delta \rho - \xi \rho_{,r}. \label{LIN_DRHOEULERLAGR}
2117: \end{align}
2118: %
2119: Here $\Delta \rho$ is the Lagrangian energy density perturbation measured
2120: by an observer moving with the fluid and is given by
2121: %
2122: \begin{align}
2123: \Delta \rho &= -\frac{\lambda}{r^2}(\rho + P) \zeta_{,r}.
2124: \label{LIN_LAGRDRHO}
2125: \end{align}
2126: %
2127: [cf. Eq.\,(\ref{LLIN_DRHO})].
2128: The asymptotic behaviour of $\Delta \rho$ is perfectly regular
2129: $\Delta \rho \sim x^{n}$ and the difficulties purely originate from the term
2130: $\xi \rho_{,r}$ on the right hand side of Eq.\,(\ref{LIN_DRHOEULERLAGR}).
2131: This correction term which facilitates the transformation
2132: between the Eulerian and
2133: Lagrangian perturbations is based on a Taylor expansion of $\rho$ which, as
2134: we have already seen above, is not generally permissible. For polytropic
2135: indices $n<1$ the derivative of $\rho$ does indeed diverge
2136: and Eq.\,(\ref{LIN_DRHOEULERLAGR}) is not a valid relation between the Eulerian
2137: and Lagrangian quantities. This is the first indication that a Lagrangian
2138: formulation is a somewhat more natural way of describing radial oscillations
2139: of neutron stars. From this point of view it is a remarkable fact
2140: that the linearisation of the Eulerian case leads to the ``correct''
2141: equations in spite of the internal inconsistency of the derivation.
2142: Finally it is worth pointing out that the irregular behaviour of
2143: $\delta \rho$ is not merely down to a poor choice of dependent variables.
2144: It is certainly possible to formulate the problem in Eulerian coordinates
2145: in terms of regular variables such as $\zeta$
2146: or $\xi$. We have seen, however, that such a regular formulation of the
2147: problem still leads to the unphysical result of a diverging
2148: total energy density $\rho+ \delta \rho$
2149: if the equations of state has an asymptotic power law behaviour $P\sim
2150: \rho^n$ with $n<1$. In view of these difficulties one may ask the question
2151: why we have decided to use an Eulerian rather than a Lagrangian
2152: formulation in the first place. Our main motivation for studying
2153: Eulerian schemes is to probe a method in spherical symmetry
2154: which enables one to accurately
2155: model a wide range of different types of non-linear neutron star oscillations.
2156: Below we shall see that the Lagrangian approach is a very powerful tool
2157: for the study of dynamic stars in spherical symmetry. However, it is a
2158: generic problem of Lagrangian methods that it is not clear how to generalise
2159: them to two or three spatial dimensions, where the paths of fluid elements
2160: may intersect and give rise to caustics. The vast majority of neutron star
2161: oscillations on the other hand will only be present if one drops
2162: the assumption of spherical symmetry, so that their numerical simulation
2163: requires the use of two or three spatial dimensions. In default
2164: of higher dimensional generalisations of Lagrangian techniques these
2165: simulations are generally performed in an Eulerian framework. \\
2166: We will now turn our attention towards the numerical solution of the
2167: linearized equations. From the asymptotic behaviour, we expect, however, that
2168: the results we obtain for $n<1$ will diverge at the surface
2169: and thus not represent a physical solution.
2170: From a numerical point of view it turns out to be beneficial to reformulate
2171: Eq.\,(\ref{LIN_ZETARR}) in terms of the displacement vector $\xi$. This is due
2172: to the asymptotic behaviour of $\zeta$ at the origin given
2173: by Eq.\,(\ref{LIN_ASSYMPTOTICZETA}).
2174: Below we will use the numerically calculated eigenmodes as initial data for
2175: the fully non-linear evolutions and for that purpose the solution for
2176: $\zeta$ would have to be converted into data for $w$ or in the Lagrangian
2177: code discussed in section \ref{LAGR} for $\xi$. The corresponding division by
2178: $r^2$ combined with the second order accuracy of the numerical eigenmode
2179: solutions results in poor accuracy of these initial data near the origin.
2180: We therefore rewrite Eq.\,(\ref{LIN_ZETARR}) in terms of $\xi$
2181: and introduce the auxiliary variable $A$
2182: to write the result as a first order system
2183: %
2184: \begin{align}
2185: &\Pi \xi_{,x} - A = 0, \label{LIN_XIY} \\[10pt]
2186: & A_{,x} + (r_{,x})^2\frac{\lambda^2}{r^4} \left( \frac{r^2}{\lambda r_{,x}}
2187: \right)_{,r} A + (r_{,x})^2\left\{ \frac{\lambda}{r^2}
2188: \left[ \Pi \left(\frac{r^2}{\lambda}\right)_{,r}
2189: \right]_{,r} + \omega^2 W + Q \right\} \xi =0.
2190: \label{LIN_AY}
2191: \end{align}
2192: %
2193: We note that the occurrence of $r$-derivatives in equation (\ref{LIN_XIY})
2194: is purely a convenient notation. In practice all these
2195: derivatives are eliminated
2196: via the TOV equations. If we use the rescaled radial coordinate, we have
2197: $r_{,x}=C$ and the $r$-derivative of $r_{,x}$ can
2198: be calculated from the relation
2199: %
2200: \begin{align}
2201: C^2_{,r} &= (\gamma-1) \frac{P_{,r}}{\rho},
2202: \end{align}
2203: %
2204: which is a consequence of the equation of state and the definition of
2205: the sound speed. The only derivatives
2206: in Eqs.\,(\ref{LIN_XIY}), (\ref{LIN_AY}) that have to be represented
2207: by finite differencing are the $x$-derivatives of $\xi$ and
2208: $A$. \\
2209: In the Cowling approximation all these results remain unchanged except for
2210: the function $Q$ which has to be replaced by
2211: %
2212: \begin{align}
2213: \tilde{Q} &= \lambda^2 (\rho + P) \left[ \left( \frac{\lambda_{,r}
2214: C^2 \mu}{r^2} \right)_r - \mu \left( \frac{\lambda_{,r}}{r^2}
2215: \right)_r + \lambda \mu
2216: \left( \frac{C^2 \mu_r}{r^2 \mu} \right)_r \right].
2217: \end{align}
2218: %
2219: and the relation between displacement and energy density perturbation which
2220: becomes
2221: %
2222: \begin{align}
2223: \delta \rho &= -(\rho + P) \frac{\lambda}{r^2} \left[
2224: \left(\frac{\lambda_{,r}}{\lambda} + \frac{\mu_{,r}}{\mu} \right)
2225: \zeta +\zeta_{,r} \right] - \frac{\lambda}{r^2} \rho_{,r} \zeta.
2226: \end{align}
2227: %
2228: It is an interesting fact that in both cases the results are simpler
2229: due to the cancellation of terms if gravity is included. \\
2230:
2231: {\em (b) The numerical implementation} \\[5pt]
2232: We have numerically calculated solutions of the eigenvalue problem
2233: (\ref{LIN_XIY}), (\ref{LIN_AY}) using a relaxation method. % as described
2234: %in section\,\ref{relaxation}.
2235: For this purpose
2236: we introduce an additional differential equation for the eigenvalues
2237: %
2238: \begin{align}
2239: (\omega^2)_{,x} &= 0, \label{LIN_OMEGAY}
2240: \end{align}
2241: %
2242: which states that the eigenmode frequency is constant throughout the star.
2243: The value of $\omega$ is not known at this stage but will result from the
2244: relaxation algorithm.
2245: In order to solve the system (\ref{LIN_XIY}), (\ref{LIN_AY}),
2246: (\ref{LIN_OMEGAY}) we need to supply three boundary conditions. At the centre
2247: we require that
2248: %
2249: \begin{align}
2250: \xi(0) &= 0, \\[10pt]
2251: A(0) &= \mathrm{const} \ne 0.
2252: \end{align}
2253: %
2254: The vanishing of the displacement $\xi$ at the origin is a necessary
2255: condition in spherical symmetry. The value of $A$ at the origin is allowed
2256: to take on any non-zero value because an eigenfunction is only defined
2257: up to a constant factor. At the outer boundary we have the condition
2258: %
2259: \begin{align}
2260: A &= 0, \label{LIN_BCOUTA}
2261: \end{align}
2262: %
2263: which follows from the definition of $A$ and the vanishing of the energy
2264: density at the surface of the star.
2265: An initial guess for $\omega$ enables us to calculate
2266: the initial functions $\xi$ and $A$ by integrating Eqs.\,(\ref{LIN_XIY}),
2267: (\ref{LIN_AY}) outwards. The solution including the eigenvalue $\omega^2$
2268: is then obtained by relaxation as described in section\,\ref{relaxation}. \\
2269:
2270: {\em (c) Testing the code} \\[5pt]
2271: For sufficiently low eigenmodes both alternative choices of the radial
2272: coordinate lead to good agreement between the predicted frequencies
2273: up to the fourth significant digit. As we will see below high order
2274: %
2275: \begin{table}[t]
2276: \caption{The convergence factors obtained for doubling the grid resolution
2277: in the relaxation code for calculating the eigenmodes of
2278: the neutron star models 1 - 5. Grid resolutions
2279: of 500, 1000 and 2000 points have been used.}
2280: \begin{center}
2281: \begin{tabular}{c|cc}
2282: \hline \hline
2283: model & fundamental mode & $10^{\rm th}$ eigenmode \\
2284: \hline
2285: 1 & 4.75 & 5.05 \\
2286: 2 & 4.76 & 4.85 \\
2287: 3 & 4.80 & 3.97 \\
2288: 4 & 4.75 & 4.82 \\
2289: 5 & 4.75 & 4.82 \\
2290: \hline \hline
2291: \end{tabular}
2292: \label{CONV_EIG}
2293: \end{center}
2294: \end{table}
2295: %
2296: eigenmode profiles show rapid oscillations near the surface of the star
2297: %
2298: \begin{table}[t]
2299: \caption{Radius, mass and frequencies of the lowest three eigenmodes
2300: for three randomly chosen models of Kokkotas and Ruoff have
2301: been recalculated with our codes and agree well with their values.}
2302: \begin{center}
2303: \begin{tabular}{l|ccccccccc}
2304: \hline \hline
2305: & $\gamma$ & $K$ & $\rho_{\rm c}$ & $R$ & $M$
2306: & $\nu_1$ & $\nu_2$ & $\nu_3$ \\
2307: & & & $[10^{15}\,\,{\rm g/cm}^3]$
2308: & $[{\rm km}]$ & $[M_{\odot}]$
2309: & [kHz] & [kHz] & [kHz] \\
2310: \hline \\[-10pt]
2311: Kokkotas \& Ruoff & 2.00 & $100\,\,{\rm km}^2$ & 5.000 & 7.787
2312: & 1.348 & 1.129 & 7.475 & 11.365 \\
2313: this work & 2.00 & $100\,\,{\rm km}^2$ & 5.000 & 7.788
2314: & 1.348 & 1.128 & 7.470 & 11.355 \\
2315: \hline \\[-10pt]
2316: Kokkotas \& Ruoff & 2.25 & $700\,\,{\rm km}^{2.5}$ & 4.000 & 8.199
2317: & 1.600 & 1.455 & 7.610 & 11.573 \\
2318: this work & 2.25 & $700\,\,{\rm km}^{2.5}$ & 4.000 & 8.200
2319: & 1.600 & 1.443 & 7.594 & 11.544 \\
2320: \hline \\[-10pt]
2321: Kokkotas \& Ruoff & 3.00 & $2\cdot 10^5\,\,{\rm km}^4$ & 2.200 & 9.419
2322: & 1.988 & 2.716 & 8.305 & 12.516 \\
2323: this work & 3.00 & $2\cdot 10^5\,\,{\rm km}^4$ & 2.200 & 9.419
2324: & 1.988 & 2.637 & 8.215 & 12.389 \\
2325: \hline \hline
2326: \end{tabular}
2327: \label{COMPEIGS}
2328: \end{center}
2329: \end{table}
2330: %
2331: which
2332: % depending on the number of grid points
2333: may not be well resolved if we
2334: work with the areal radius $r$. The frequencies deviate more significantly
2335: in these cases. In the rest of this section we will therefore work with
2336: the rescaled coordinate and set $r_{,x}=C$.
2337: The resulting code has been checked in four independent ways. First
2338: we have computed the eigenfunctions of the fundamental and the tenth
2339: mode for the neutron star models listed in Table \ref{MODELS15}
2340: and checked for convergence using 500, 1000 and 2000 grid points. The results
2341: shown in Table \ref{CONV_EIG} clearly demonstrate second order convergence
2342: as expected for the second order finite differencing scheme applied
2343: in the relaxation algorithm. \\
2344: Next we have randomly chosen three of the stellar models
2345: listed in \lcite{Kokkotas2001} and recalculated radius, mass of the neutron
2346: stars as well as the frequencies of the lowest three eigenmodes. The
2347: results are compared in Table \ref{COMPEIGS} and show good agreement. \\
2348: For the third test we recall the 1-parameter families of neutron stars
2349: shown in Fig.\,\ref{TOV_FAMILIES}. We have already
2350: mentioned that the maxima in the mass vs. central density plots separate the
2351: stable and unstable branches of neutron star models and that the frequency
2352: of the fundamental eigenmode becomes zero at the critical point
2353: and imaginary on the unstable branch. We have therefore
2354: determined the critical central densities for the five neutron star
2355: models of Table \ref{MODELS15}
2356: and calculated the frequency of the fundamental modes
2357: just below and above the critical densities. The numerical results
2358: are shown in Table \ref{CRITFREQS} and confirm this picture. The frequencies
2359: of the fundamental mode are very small but real for central densities
2360: %
2361: \begin{table}[t]
2362: \caption{The critical central densities corresponding to
2363: the neutron star models 1-5
2364: are given to four significant digits together with the
2365: frequency of the fundamental mode just below and above the
2366: critical point. Above the critical density the frequencies
2367: become imaginary as expected.}
2368: \begin{center}
2369: \begin{tabular}{l|ccccccccc}
2370: \hline \hline
2371: model & $\rho_{\rm c,crit}$
2372: & $\nu (\rho_{\rm c,crit}-10^{-6})\,{\rm km^{-2}}$
2373: & $\nu (\rho_{\rm c,crit}+10^{-6})\,{\rm km^{-2}}$ \\
2374: & [km$^{-2}$] & [kHz] & [kHz] \\
2375: \hline \\[-10pt]
2376: 1 & 0.002179 & 0.0294 & 0.0477$\,i$ \\
2377: 2 & 0.004205 & 0.0578 & 0.0429$\,i$ \\
2378: 3 & 0.002804 & 0.0629 & 0.0350$\,i$ \\
2379: 4 & 0.002103 & 0.0409 & 0.0592$\,i$ \\
2380: 5 & 0.002233 & 0.0591 & 0.0627$\,i$ \\
2381: \hline \hline
2382: \end{tabular}
2383: \label{CRITFREQS}
2384: \end{center}
2385: \end{table}
2386: %
2387: just below the critical value and become imaginary for larger densities.\\
2388: A further test for the eigenmode frequencies arises from a relation
2389: between the period of the fundamental mode $T_1$ of a neutron star model
2390: and the deviation of the radius $R$ from the critical radius $R_{\rm c}$
2391: that has been suggested by \scite{Harrison1965} [see their Eq.\,(155)]
2392: %
2393: \begin{align}
2394: (R-R_{\rm c})\cdot T_1^{\,2} &= \mathrm{const}. \label{LIN_HARRISON}
2395: \end{align}
2396: %
2397: In Table \ref{LIN_RCT0} we show the results obtained for neutron star
2398: models identical to model 1 and 3 with central densities as indicated.
2399: %
2400: \begin{table}[t]
2401: \caption{Equation\,(\ref{LIN_HARRISON}) is checked for neutron star models
2402: 1 and 3 for various central densities.}
2403: \begin{center}
2404: \begin{tabular}{ccc|ccc}
2405: \hline \hline
2406: \multicolumn{3}{c|}{model 1} & \multicolumn{3}{c}{model 3} \\
2407: $\rho_{\rm c}$ & $\omega_1$ & $(R-R_{\rm crit})/\omega_1^2$
2408: & $\rho_{\rm c}$ & $\omega_1$ & $(R-R_{\rm crit})/\omega_1^2$ \\
2409: $[{\rm km}^{-2}]$ & [km$^{-1}$] & [km$^3$] & [km$^{-2}$] & [km$^{-1}$]
2410: & [km$^3$] \\
2411: \hline
2412: 0.0021785 & \hspace{0.3cm} 0.000187 & 187.96 &
2413: 0.0028035 & \hspace{0.3cm} 0.000172 & 97.55 \\
2414: 0.0021780 & \hspace{0.3cm} 0.000616 & 192.72 &
2415: 0.0028030 & \hspace{0.3cm} 0.000773 & 85.21 \\
2416: 0.0021775 & \hspace{0.3cm} 0.000850 & 192.34 &
2417: 0.0028025 & \hspace{0.3cm} 0.001080 & 84.89 \\
2418: 0.0021750 & \hspace{0.3cm} 0.001563 & 192.59 &
2419: 0.0028020 & \hspace{0.3cm} 0.001317 & 84.77 \\
2420: 0.0021700 & \hspace{0.3cm} 0.002424 & 193.00 &
2421: 0.0028000 & \hspace{0.3cm} 0.002002 & 84.87 \\
2422: 0.0020000 & \hspace{0.3cm} 0.010939 & 207.93 &
2423: 0.0027000 & \hspace{0.3cm} 0.010838 & 86.97 \\
2424: 0.0015000 & \hspace{0.3cm} 0.020261 & 236.69 &
2425: 0.0020000 & \hspace{0.3cm} 0.029622 & 106.64 \\
2426: 0.0011775 & \hspace{0.3cm} 0.023606 & 235.61 &
2427: 0.0015000 & \hspace{0.3cm} 0.036427 & 128.95 \\
2428: \hline \hline
2429: \end{tabular}
2430: \label{LIN_RCT0}
2431: \end{center}
2432: \end{table}
2433: %
2434: Even though a deviation from Eq.\,(\ref{LIN_HARRISON}) up to $20\,\%$
2435: is observed for both models,
2436: this is rather small if one considers the variation of the frequency
2437: $\omega_1$ over several orders of magnitude. \\
2438:
2439: {\em (d) The eigenmode solutions} \\[5pt]
2440: We will now turn our attention to the eigenmode profiles of the
2441: physical variables.
2442: We have already noted that the eigenvalue problem has an enumerable infinite
2443: set of solutions which can be ordered with respect to their eigenvalues.
2444: This order is also reflected in the spatial profiles of the corresponding
2445: eigenfunctions. We have numerically calculated the first four
2446: eigenmodes in terms of the displacement vector $\xi$ for model 3
2447: with polytropic exponent $\gamma=2$. The velocity $w$,
2448: the rescaled displacement $\zeta$ and the energy density perturbation
2449: $\delta \rho$ then follow from Eqs.\,(\ref{LIN_XI}) where we use harmonic time
2450: dependence, (\ref{LIN_ZETA}) and (\ref{LIN_DRHOOFZETA}).
2451: The results are shown in Fig.\,\ref{MODES1_4},
2452: where we have also included the solution
2453: for $\xi$ corresponding to the tenth eigenmode.
2454: Since the eigenmode solutions are determined
2455: up to a constant factor only, we have rescaled them to about unit amplitude.
2456: For all variables we see that the number of nodes is given by the order
2457: of the mode and the number of local maxima or minima is given by
2458: the order minus one.
2459: This behaviour remains valid for higher modes and is characteristic of the
2460: eigenmode solutions.
2461: In order to illustrate the significance of the transformation to the
2462: rescaled radius $y$ we have plotted $\xi$ as a function of $r$ as well.
2463: In the upper panels of Fig.\,\ref{MODES1_4} we can see that the
2464: oscillations in the spatial profile of the eigenmodes become more
2465: concentrated towards larger radii $r$ the higher the order of the mode.
2466: In terms of the rescaled radius $y$, however, the oscillations are evenly
2467: distributed over the entire interval. This behaviour is reminiscent of the
2468: narrowing of the wave pulse we observed in section \ref{TOV_RYTRAFO}
2469: and illustrates why
2470: a superior numerical performance is obtained when using the coordinate $y$,
2471: especially when higher order modes are present in the evolution. \\
2472: %
2473: \begin{figure}[t]
2474: \centering
2475: \epsfig{file=m3_eigrxi.eps, height=200pt, width=150pt, angle=-90}
2476: \epsfig{file=m3_eigxi.eps, height=200pt, width=150pt, angle=-90}
2477: \epsfig{file=m3_eigw.eps, height=200pt, width=150pt, angle=-90}
2478: \epsfig{file=m3_eigdrho.eps, height=200pt, width=150pt, angle=-90}
2479: \caption{The displacement $\xi$ as a function of the areal radius $r$
2480: and the rescaled radius $y$ as well as the velocity $w$ and
2481: the energy density $\delta \rho$ as a function of $y$ are
2482: shown for the first four eigenmodes of model 3.
2483: For $\xi$ we have also
2484: plotted mode 10 to illustrate the concentration of oscillations
2485: towards larger $r$.}
2486: \label{MODES1_4}
2487: \end{figure}
2488: %
2489: %
2490: \begin{figure}[ht]
2491: \centering
2492: \epsfig{file=m1_eigdrho.eps, height=200pt, width=150pt, angle=-90}
2493: \epsfig{file=m5_eigdrho.eps, height=200pt, width=150pt, angle=-90}
2494: \caption{The energy density perturbation $\delta \rho$ obtained for the
2495: first four eigenmodes of stellar model 1 (left panel) and 5
2496: (right panel) is plotted as a function of $y$.}
2497: \label{EIG_DRHO}
2498: \end{figure}
2499: %
2500: The corresponding eigenmodes obtained for the other stellar models look
2501: qualitatively similar in all variables except for the energy density
2502: perturbation $\delta \rho$. We have already noted that the asymptotic
2503: behaviour of $\delta \rho$ depends on the polytropic exponent $\gamma$.
2504: This is confirmed by the numerical solutions shown in Fig.\,\ref{EIG_DRHO}
2505: where we plot the profiles of the energy density perturbation obtained
2506: for the stellar models 1 and 5 with polytropic exponents $\gamma=1.75$
2507: and $2.3$ respectively. For model 1 the energy density perturbation
2508: goes to zero at the surface, although with a non-zero gradient. In
2509: comparison the gradient of the background density of the same model vanishes
2510: in Fig.\,\ref{TOV_GAMMA} and the
2511: quotient $\delta \rho/\rho$ can indeed be shown to diverge
2512: in agreement with Eq.\,(\ref{DRHOOVERRHO}).
2513: For the larger polytropic index 2.3 the
2514: perturbation $\delta \rho$ itself diverges at the surface as expected
2515: from Eq.\,(\ref{LIN_ASSYMPTOTICDRHO}). \\
2516: The corresponding results obtained in the Cowling approximation are
2517: very similar to those shown above. The only notable difference is the
2518: frequency of the fundamental mode which does not decrease towards
2519: zero as the central density approaches the critical value but instead
2520: remains real and positive. This result is to be expected since a fluid
2521: will not become gravitationally unstable if the gravitational field is
2522: kept fixed. \\
2523: The eigenmode solutions obtained in this section will be used extensively
2524: as initial data in the non-linear evolutions. We have seen, however,
2525: that the stellar surface represents a problematic area even in the linearized
2526: case. The difficulties are more pronounced in the non-linear case
2527: and need to be investigated in more detail before we can study the
2528: fully non-linear numerical evolutions.
2529:
2530:
2531: %=========================================================================
2532: \subsubsection{The surface problem}
2533: \label{SURFACE}
2534: %
2535: %
2536: When we formulated the description of non-linear radial oscillations
2537: of neutron stars in section \ref{NONP_EQ} we consciously omitted the issue
2538: %
2539: \begin{figure}[t]
2540: \centering
2541: \epsfig{file=surface.eps, height=300pt, width=200pt, angle=-90}
2542: \caption{The energy density profile of an oscillating neutron
2543: star is schematically plotted at three different
2544: stages of one oscillation.
2545: Initially the stellar radius is at its equilibrium value,
2546: at the later time $t_1$ the star has expanded and at $t_2$
2547: it has shrunk below its initial radius. The vertical
2548: line indicates the extension of the numerical grid.}
2549: \label{NONL_SURFACE}
2550: \end{figure}
2551: %
2552: of boundary conditions. The difficulties involved in specifying outer boundary
2553: conditions in an Eulerian code are so complex that we dedicate a whole
2554: subsection to this topic.
2555: We have already mentioned that the surface is defined by the condition
2556: $\hat{P}=0$ which is equivalent to $\hat{\rho}=0$ for a polytropic
2557: equation of state. With respect to the fixed numerical grid,
2558: however, the surface of the star is moving and we cannot apply this condition
2559: at the outer grid boundary. This is a further indication that one may
2560: have to emulate a Lagrangian treatment of the surface in order to
2561: accurately model neutron star oscillations involving radial displacements
2562: of the surface.
2563: The situation is graphically illustrated in
2564: Fig.\,\ref{NONL_SURFACE} where the total energy density profile is
2565: schematically plotted as a function of radius. At time
2566: $t_0$ an equilibrium star (solid curve)
2567: is perturbed with a velocity field that causes the star to
2568: expand. The initial
2569: configuration also determines the extension of the numerical grid indicated
2570: by the vertical line. At a later time $t_1$ the star has expanded
2571: (long dashed curve). The outer part of the star has therefore moved out of
2572: the numerical grid (dotted part of the curve) and the corresponding
2573: information would be lost in a non-linear numerical evolution.
2574: At time $t_2$ the star
2575: has shrunk and is completely contained inside the numerical grid.
2576: Outside of the star the energy density will be zero. In general,
2577: therefore, the energy density profile or its derivatives will have a
2578: discontinuity at the stellar surface. Worse from a numerical point of view
2579: is the region between the stellar surface and the outer
2580: grid boundary.
2581: Even though the energy density will be zero at these points theoretically,
2582: numerically this will not exactly be the case. At some of these points the
2583: total energy density will have small negative values due to numerical
2584: noise, unless the
2585: values are manipulated in some form. A negative energy density, however,
2586: means that the pressure can no longer be calculated from the equation of
2587: state which normally terminates the evolution. There are several
2588: possibilities for dealing with these difficulties. We will discuss
2589: four methods and implement two of them in the course of this work. \\
2590: %
2591: \begin{list}{\rm{\arabic{count}.)}}{\usecounter{count}
2592: \labelwidth1cm \leftmargin1.0cm \labelsep0.4cm \rightmargin0cm
2593: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
2594: \item
2595: The first method consists in embedding the star in an atmosphere
2596: of low density. In this method the numerical grid extends
2597: well beyond the size of the neutron star and no information is lost at
2598: any stage of the evolution. The boundary conditions are then applied
2599: to the atmosphere whereas the star will always be confined to the interior
2600: numerical grid and the surface of the star is entirely described by
2601: the interior numerical evolution, for example by shock capturing methods.
2602: It is a non-trivial question, however, to what extent the atmosphere
2603: and the numerical treatment of the surface discontinuities
2604: will affect the evolution of the neutron star.
2605: % especially at reasonably
2606: %small amplitudes.
2607: For this reason it seems plausible to use an atmosphere
2608: of low density. A low density, however, will in general be accompanied by a
2609: small speed of sound and we have already seen in the discussion of the
2610: wave equation in section \ref{TOV_RYTRAFO} that such regions require
2611: a careful numerical treatment. An insufficient resolution
2612: may result in spurious phenomena. In terms of a rescaled radius such as
2613: the coordinate $y$ defined in Eq.\,(\ref{TOV_YOFR}) we have been able to obtain
2614: a sufficient resolution, but a large number of grid points would be required
2615: to simulate an atmosphere of significant spatial extension. \\
2616: An interesting variation of this method consists in viewing the surface
2617: of the star as an interface to an exterior vacuum region and explicitly
2618: tracking the movement of the interface. Sophisticated techniques such as
2619: {\em level set methods} and {\em fast marching methods}
2620: have been developed for these
2621: purposes (see for example \citeNP{Sethian1999}) and may provide an answer to
2622: the surface problem in Eulerian formulations. One may even go a step further
2623: and recall the strikingly similar concept of Cauchy-characteristic
2624: matching and, thus, consider a combination of these ideas. It is, however,
2625: well beyond the scope of this work to investigate these methods in more
2626: detail and we will therefore focus on simpler techniques.
2627: %
2628: \item
2629: The second method is a modified version of the atmosphere approach discussed
2630: above. Instead of using an external atmosphere, we
2631: modify the equation of state of the neutron star at low densities
2632: and thus view the outer layers of the neutron star itself as an atmosphere.
2633: For that purpose we use an equation of state given by
2634: %
2635: \begin{align}
2636: P &= K\,\rho^{\gamma}\hspace{4.17cm} \mathrm{if }\,\,
2637: \rho > \rho_{\rm t}, \\[10pt]
2638: P &= a_1 \rho + a_2\rho^2 + a_3 \rho^3 + a_4 \rho^4 \hspace{0.8cm}
2639: \mathrm{if }\,\, \rho \le \rho_{\rm t},
2640: \end{align}
2641: %
2642: where $a_2$, $a_3$ and $a_4$ are coefficients determined by the continuity
2643: of $P$ and its first two derivatives with respect to $\rho$. The
2644: coefficient $a_1$ and the transition density $\rho_{\rm t}$ are free
2645: parameters that are specified by the user. A consequence of this definition
2646: is that $P\sim \rho$ at low densities and the behaviour will
2647: be similar to that of a $\gamma=1$ polytrope in this region, i.e.
2648: extend beyond the surface of the original purely polytropic model. The
2649: low density part of the neutron star can thus be viewed as an atmosphere
2650: smoothly attached to a polytropic neutron star truncated at $\rho_{\rm t}$.
2651: Whenever the energy density falls below a threshold value $\rho_{\rm min}$
2652: during the evolution, it is set to this threshold value. The
2653: parameter $\rho_{\rm min}$ also needs to be specified by the user.
2654: This requirement avoids the occurrence of negative total energy densities,
2655: but introduces ad hoc discontinuities in the $\delta \rho$ profile. We take
2656: care of these discontinuities by introducing artificial viscosity of the
2657: %``standard'' modification of the
2658: modified von Neumann-Richtmyer form (see for example \citeNP{Fox1962})
2659: %
2660: \begin{align}
2661: q &= \left\{ \parbox{7cm}
2662: {
2663: $\displaystyle{b \Delta y^2 \hR w_{,y}^2}
2664: \hspace{1.5cm} {\rm if}\,\,\, w_{,y} < 0 \\[10pt]
2665: \displaystyle{0 \hspace{2.9cm} {\rm if} \,\,\, w_{,y} \ge 0},$
2666: } \right. \label{qrv2}
2667: \end{align}
2668: %
2669: where $b$ is the viscosity parameter. In many
2670: cases $b=2$ leads to satisfactory results.
2671: This viscosity term is added to the pressure perturbation $\delta P$
2672: wherever it occurs in the equations. With careful choices of the
2673: free parameters $a_1$, $\rho_{\rm t}$, $\rho_{\rm min}$ and $b$ we have
2674: obtained long term stable evolutions of localised wave pulses. The particular
2675: values we have to choose for a stable evolution,
2676: especially the density values $\rho_{\rm t}$
2677: and $\rho_{\rm min}$, do however depend sensitively on the initial data.
2678: Furthermore the manipulation of the energy density perturbation $\delta \rho$
2679: in cases of a negative total energy density leads to a contamination
2680: of the evolution of eigenmodes in the low density range. The resulting
2681: disturbances then travel into the stellar interior within a few
2682: oscillation periods. In view of these difficulties we have decided to
2683: use a different treatment of the stellar surface. \\
2684: %
2685: \item
2686: A fully satisfactory solution to the surface problem in one spatial
2687: dimension can be obtained with
2688: a Lagrangian formulation either of the surface or the whole star.
2689: In the first case this can be implemented by rescaling to a new
2690: radial coordinate
2691: %
2692: \begin{align}
2693: s&:= \frac{r}{R(t)},
2694: \end{align}
2695: %
2696: where $R$ is the time dependent total radius of the star. This transformation
2697: leads to a few extra terms in the equations in the radial gauge,
2698: but is more complicated
2699: to implement in terms of the rescaled coordinate $y$. For this reason
2700: and because of the wider range of applications we have chosen instead
2701: to reformulate the non-linear radial oscillations entirely within
2702: a Lagrangian framework. Combined with the singularity avoiding properties
2703: of the polar slicing condition the resulting code can not only be used for
2704: the simulation of radial oscillations but also allows high resolution
2705: studies of spherically symmetric gravitational collapse. This code and the
2706: corresponding testing will be discussed in detail in section \ref{LAGR}. \\
2707: Even though Lagrangian methods represent a formidable
2708: tool for 1-dimensional problems,
2709: we have already mentioned that there is no
2710: straightforward generalisation to two or three spatial
2711: dimensions, where the paths of fluid elements may intersect and give
2712: rise to caustics. \\
2713: %
2714: \item
2715: The method we will be using in the remainder of this section can be considered
2716: the inverse of the atmospheric treatments discussed above. Instead
2717: of adding matter in the form of an atmosphere the outer layers of the
2718: star are removed. In this context it is worth remembering that
2719: the solution of the TOV equations via quadrature does not go all the way
2720: out to $\rho=0$ and a fully non-linear perturbative code working with
2721: such a background intrinsically describes a truncated neutron star.
2722: The percentage of mass that we will remove from the star will be very
2723: small in most cases ($\ll 1 \%$). We will see below that the
2724: resulting code behaves well in the linearized limit in most cases.
2725: % The truncation
2726: % density below which the outer layers are removed from the background star,
2727: %has to be chosen larger for higher amplitude perturbations, however, which
2728: %leads to spurious numerical effects in the evolution of marginally
2729: %stable neutron stars, i.e. stars with a central density just below
2730: %the critical value. Consequently we will follow a more conservative approach
2731: %and study non-linear effects in a simplified neutron star model. Such
2732: %a model, albeit less realistic, is numerically under control in the sense
2733: %that no modifications of the model are required throughout the
2734: %amplitude range of interest and thus spurious numerical effects are
2735: %avoided. This model will enable us to study non-linear effects
2736: %in radial oscillations of a neutron star like dynamic system. \\
2737: \end{list}
2738:
2739:
2740: %=========================================================================
2741: \subsubsection{The numerical implementation in Eulerian coordinates}
2742: \label{PERT_NUMERICS}
2743: %
2744: %
2745: In section \ref{PERT_EQ} we have derived the equations for a fully non-linear
2746: perturbative formulation of a dynamic spherically
2747: symmetric star in terms of the generalised coordinate $x$. In the
2748: remainder of the Eulerian discussion we will restrict ourselves to the
2749: rescaled version and set $r_{,x} = C$ and $x=y$.
2750: In order to numerically solve these equations, we
2751: also have to specify appropriate boundary conditions.
2752: We start with the origin and recall that the displacement $\xi$ of a
2753: fluid element at the centre of a spherically symmetric star
2754: vanishes. As a consequence the radial velocity will also vanish at
2755: the origin.
2756: As far as the energy energy density is concerned, we note that
2757: $\hat{\rho}$ is a component of a rank 2 tensor
2758: and therefore the spatial derivative $\hat{\rho}_{,y}$ will vanish
2759: in spherical symmetry. The same is true for the background density
2760: $\rho$ and therefore we obtain the inner boundary condition
2761: $\delta \rho_{,y}=0$. Finally we require the vanishing of
2762: $\delta \mu$ to avoid a conical singularity. \\
2763: At the outer boundary we match
2764: the lapse function to an exterior Schwarzschild metric as in the static
2765: case which results in the condition $\hat{\lambda}\cdot \hat{\mu}=1$. As
2766: far as the matter variables are concerned, the situation is a bit
2767: more complicated. For the velocity we use the regularity condition $w_{,y}=0$.
2768: In view of the definition of the radial coordinate $y$ this is equivalent
2769: to demanding that the velocity
2770: has a finite gradient with respect to $r$ at the surface.
2771: This condition is satisfied by the eigenmode solutions obtained
2772: in section \ref{PERT_LIN}.
2773: In Fig.\,\ref{MODES1_4} we can see that the gradient $w_{,y}$ vanishes
2774: for all three polytropic exponents $\gamma=1.75$, $2.00$
2775: and $2.3$.
2776: In contrast to the velocity gradient the derivative of the energy density
2777: perturbation $\delta \rho_{,y}$ will in general not vanish at the surface.
2778: If we consider the stellar models listed in Table \ref{MODELS15}
2779: it can be shown that $\delta \rho_{,y}$ will only vanish in the
2780: case $\gamma=2$ which is also illustrated in Figs.\,\ref{MODES1_4}
2781: and \ref{EIG_DRHO}. In summary the boundary conditions are
2782: %
2783: \begin{align}
2784: \delta \rho_{,y} &= 0, \label{PERT_BCINDRHO} \\[10pt]
2785: w &= 0, \label{PERT_BCINW} \\[10pt]
2786: \delta \mu &= 0. \label{PERT_BCINDMU}
2787: \end{align}
2788: %
2789: at the origin and
2790: %
2791: \begin{align}
2792: w_{,y} &= 0, % \hspace{0.5cm} \mathrm{or} \hspace{0.5cm} w =0,
2793: \label{PERT_BCOUTW} \\[10pt]
2794: \hat{\lambda} \cdot \hat{\mu} &= 1
2795: \end{align}
2796: %
2797: at the surface. \\
2798: In this context it is worth mentioning a subtlety concerning
2799: second order finite differencing schemes used for evolution
2800: equations such as (\ref{PERT_DRHOT}), (\ref{PERT_WT}).
2801: In general this system of equations
2802: has one ingoing and one outgoing characteristic at each boundary
2803: and physical information has to be specified in the form of one
2804: condition for either $w$ or $\delta \rho$ at either boundary. The centred
2805: finite differencing scheme (or variation thereof) used in second order
2806: techniques, however, cannot be applied at the grid boundaries and the
2807: variables must be evolved in an alternative way. The physical
2808: boundary conditions do not necessarily provide enough information
2809: for this. In our case, for example, we have two variables $\delta \rho$, $w$
2810: that need to be updated at
2811: two grid points respectively which requires four conditions, but only
2812: two conditions are required to provide information for the
2813: characteristics entering the numerical grid.
2814: The remaining boundary values not determined by these two conditions
2815: have to be obtained in alternative ways,
2816: for example by extrapolation or the use of one sided derivatives in the
2817: evolution equations. We have obtained optimal performance in the evolution
2818: of $\delta \rho$ and $w$ by using conditions (\ref{PERT_BCINDRHO})
2819: and (\ref{PERT_BCINW}) at the centre and (\ref{PERT_BCOUTW}) at the
2820: surface. The outer boundary value of $\delta \rho$ is then obtained
2821: by extrapolation on each new time slice.
2822: It is worth pointing out that this problem is not apparent in the implicit
2823: finite difference methods applied to the cosmic string in section \ref{cstr}
2824: or the Lagrangian code in section \ref{LAGR}. \\
2825: Before we schematically outline the computational steps involved in
2826: the time evolution we need to discuss one final numerical issue,
2827: the CFL stability condition. We have mentioned in section \ref{stability}
2828: that the stability criterion of Courant, Friedrichs and Lewy requires
2829: the physical domain of dependence to be included in the numerical
2830: domain of dependence. A standard method to ensure that this criterion
2831: is met in a hydrodynamical evolution is based on calculating the
2832: slopes of the characteristics at each point on the numerical grid.
2833: In our case we consider the system of evolution equations
2834: (\ref{PERT_DRHOT}), (\ref{PERT_WT}). The quasi-linear nature of this
2835: system enables us to calculate the characteristics from
2836: %
2837: \begin{align}
2838: \frac{dy_i}{dt} &= \Lambda_i,
2839: \end{align}
2840: %
2841: where $\Lambda_i$ are the eigenvalues of the coefficient matrix
2842: and are defined by the equation
2843: %
2844: \begin{align}
2845: \left[ \begin{pmatrix} \alpha_{11} & \alpha_{12} \\
2846: \alpha_{21} & \alpha_{11} \\
2847: \end{pmatrix} -\Lambda \cdot \hbox{\oneone{1}} \right]
2848: \begin{pmatrix} \delta \rho \\ w
2849: \end{pmatrix} &= 0.
2850: \end{align}
2851: %
2852: The solution for the coefficient functions
2853: (\ref{PERT_ALPHA11})-(\ref{PERT_ALPHA21}) is given by
2854: %
2855: \begin{align}
2856: \Lambda &= \frac{1}{r_{,x}D} \left( w(1-\hat{C}^2) \pm \frac{\hat{C}}
2857: {\hat{\mu}\hat{\lambda}v} \right).
2858: \end{align}
2859: %
2860: If the characteristics are straight lines, the Courant-Friedrichs-Lewy
2861: condition is satisfied if the time step $dt$ obeys the inequality
2862: %
2863: \begin{align}
2864: dt &\le \frac{dy}{\max{|\Lambda_i|}}. \label{PERT_DT}
2865: \end{align}
2866: %
2867: We therefore calculate the eigenvalue fields $\Lambda_1$, $\Lambda_2$ on each
2868: time slice and determine the value of $\max{|\Lambda_i|}$.
2869: Even though the characteristics will in general not be straight lines,
2870: the deviation is small on time scales of $dt$ and we allow for
2871: this effect by multiplying the resulting time step by a factor of 0.9. With
2872: that choice and about 500 grid points
2873: we have obtained stable evolutions over several 100000
2874: time steps which corresponds to more than $0.1$ s of proper time
2875: as measured by an observer at infinity.\\
2876: We have got all ingredients now to summarise the individual steps involved in
2877: the fully non-linear numerical evolution.
2878: %
2879: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
2880: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
2881: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
2882: \item A static background model is calculated according to the
2883: TOV equations (\ref{TOV_RY})-(\ref{TOV_PY}), where
2884: we set $r_{,x}=C$. For this
2885: purpose the polytropic exponent $\gamma$, the polytropic factor
2886: $K$, the central density $\rho_{\rm c}$ and the surface density
2887: $\rho_{\rm s}$ need to be specified by the user. A non-zero
2888: surface density will result in a truncated neutron star model.
2889: The results are given in the form of data
2890: files containing the background variables $\lambda$, $\mu$,
2891: $\rho$ and $r$ as functions of $y$.
2892: \item If initial data is required in the form of eigenmode profiles,
2893: the eigenmode solutions can be calculated according to the
2894: method described in section \ref{PERT_LIN}. The order of the eigenmode is
2895: determined by the initial guess for the frequency which needs
2896: to be specified. The amplitude of the eigenmode is a free parameter
2897: in the evolution code.
2898: \item There are several alternative choices for the initial data. Among
2899: these are localised perturbations of Gaussian shape and
2900: linear combinations of different eigenmodes.
2901: \item With the initial velocity $w$ and energy
2902: density $\delta \rho$ specified, the metric
2903: perturbations follow from the constraint
2904: equations (\ref{PERT_LAMBDAR}) and (\ref{PERT_MUR}).
2905: These equations are numerically integrated with a fourth
2906: order Runge-Kutta scheme.
2907: \item The initial data is evolved according to the second order in
2908: space and time McCormack scheme described in section \ref{McCormack}.
2909: One evolution cycle
2910: consists of the following steps. \\[4pt]
2911: a) Calculation of the Courant factor, \\
2912: b) predictor step for $\delta \rho$ and $w$,\\
2913: c) application of the inner boundary conditions for
2914: $\delta \rho$ and $w$,\\
2915: d) \parbox[t]{12.7cm}{integration of the constraint equations
2916: to obtain preliminary values for $\delta \lambda$ and $\delta \mu$
2917: on the new time slice,}\\[10pt]
2918: e) corrector step for $\delta \rho$ and $w$,\\
2919: f) application of boundary conditions for $\delta \rho$ and $w$,\\
2920: g) \parbox[t]{12.7cm}{integration of the constraint equation on the
2921: new slice to obtain the final values of $\delta \lambda$
2922: and $\delta \mu$.}
2923: \end{list}
2924:
2925:
2926: %=========================================================================
2927: \subsubsection{The performance of the code in the linear regime}
2928: \label{EULER_LIN}
2929: %
2930: %
2931: We will now investigate the performance of the code in the linear regime,
2932: where we know the exact solution with high accuracy.
2933: If initial data is provided in the form of an eigenmode profile $w_i(y)$
2934: and zero $\delta \rho$, we know that
2935: the time dependent solution in the linear regime is given by
2936: %
2937: \begin{align}
2938: \delta \rho(t,y) &= -\delta \rho_i(y) \sin{\omega_i t}, \label{LIN_DRHOOFT}
2939: \\[10pt]
2940: w(t,y) &= w_i(y) \cos{\omega_i t}. \label{LIN_WOFT}
2941: \end{align}
2942: %
2943: For finite amplitudes this solution is not exact, but for
2944: sufficiently small amplitudes
2945: the deviation of the exact solution from (\ref{LIN_DRHOOFT}),
2946: (\ref{LIN_WOFT}) is negligible
2947: compared with the truncation error of the numerical scheme.
2948: We have therefore calculated the fundamental mode for stellar model
2949: 3 of Table \ref{MODELS15} using 1600 grid points and a truncation density
2950: $\rho_{\rm s}=1.0\cdot 10^{-7}\,\,{\rm km}^{-2}$. This
2951: density corresponds to the removal of about $3\cdot 10^{-8}$ of the
2952: neutron star mass which is one order of magnitude smaller than the accuracy
2953: of the numerically calculated total mass.
2954: The amplitude of the eigenmode corresponds to an oscillation
2955: of the stellar radius of about $10\,\,{\rm cm}$, i.e. a relative
2956: displacement of about $10^{-5}$. In Fig.\,\ref{PERT_LINEVOL}
2957: we show the time evolution
2958: of $\delta \rho$ and $w$ together with the deviation from
2959: the analytic solution (\ref{LIN_DRHOOFT}), (\ref{LIN_WOFT}).
2960: %
2961: \begin{figure}[t]
2962: \centering
2963: \epsfig{file=pert_test_drho.ps, height=400pt, width=175pt, angle=-90}
2964: \epsfig{file=pert_test_w.ps, height=400pt, width=175pt, angle=-90}
2965: \caption{The left panels show the time evolution of $\delta \rho$ and $w$
2966: obtained for neutron star model 3 with $\gamma=2.00$.
2967: The initial perturbation
2968: is given in the form of the fundamental mode in the velocity field
2969: $w$. The right panels show the deviation from the exact
2970: solution of the linearized equations.}
2971: \label{PERT_LINEVOL}
2972: \end{figure}
2973: %
2974: The numerical evolution reproduces the expected harmonic time dependence
2975: with high accuracy. Because of its low frequency the fundamental mode
2976: is particularly suitable for this graphical illustration. The code
2977: reproduces the sinusoidal evolution of higher modes with comparable
2978: accuracy, but the large number of oscillations is not well resolved in plots
2979: similar to Fig.\,\ref{PERT_LINEVOL}.
2980: For the same reason we have shown the earlier stages
2981: of the evolution up to $t=600\,\,{\rm km}$ only in the figure. The whole run
2982: lasts more than ten times longer and shows no significant loss
2983: of accuracy. It is worth
2984: mentioning that the accuracies obtained here are limited not only by
2985: the evolution code but also by the results for the static background,
2986: the eigenmode profiles and, most importantly, the eigenmode
2987: frequencies used in the calculation of the analytic solution.
2988: The same long term stability and high accuracy has been observed in
2989: similar evolutions for a variety of different neutron star models with
2990: polytropic indices $\gamma\le 2$. Below we will see, however, that the code
2991: does not perform equally satisfactorily if we use a larger truncation density
2992: in combination with a marginally stable
2993: neutron star model with a central density just below the
2994: critical value. \\
2995: For neutron star models sufficiently far away from the stability limit,
2996: we can also check the performance of the code in the linear regime by
2997: comparing the frequency spectrum of the time evolution with the
2998: values predicted by the eigenmode calculations of section \ref{PERT_LIN}.
2999: For this purpose initial velocity fields have been calculated
3000: for models 1 and 3 by
3001: adding the first ten eigenmode profiles whereas the initial density
3002: perturbation is set to zero. The combined amplitude of the perturbations is
3003: similar to that used above for determining the
3004: deviation from harmonic time dependence.
3005: In Fig.\,\ref{PERT_LINFOUR} we show the Fourier spectra for the
3006: corresponding time evolutions of the
3007: %
3008: \begin{figure}[t]
3009: \centering
3010: \epsfig{file=four_gamma175.eps, height=200pt, width=150pt, angle=-90}
3011: \epsfig{file=four_gamma200.eps, height=200pt, width=150pt, angle=-90}
3012: \caption{The power spectra of the time evolution of the central
3013: density for neutron star models 1 and 3. The vertical bars
3014: indicate the frequencies predicted by the linear analysis.}
3015: \label{PERT_LINFOUR}
3016: \end{figure}
3017: %
3018: central density perturbation $\delta \rho(t,0)$. The frequencies
3019: predicted by the eigenmode analysis are indicated by vertical bars
3020: and show good agreement with the peaks in the Fourier spectra. \\
3021: Next we compare the performance of the perturbative
3022: approach with that of a ``standard'' non-perturbative method.
3023: We have already mentioned that we can simulate a non-perturbative approach
3024: by using vacuum flat space for the background variables. In this case we
3025: only use the TOV-model to determine the numerical grid as well as the areal
3026: radius $r$ and the sound speed $C$ as functions of $y$.
3027: The background variables, however, are
3028: specified as $\lambda=1$, $\mu=1$ and $\rho=0$. If we insert these
3029: values into the perturbative equations
3030: (\ref{PERT_LAMBDAR})-(\ref{PERT_B2}) they will become identical to the
3031: non-perturbative system (\ref{NONP_LAMBDAR})-(\ref{NONP_B2})
3032: (after transformation to the radial coordinate $y$) with
3033: $\hat{\lambda}$, $\hat{\mu}$, $\hat{\rho}$ and $\hat{P}$ replaced by
3034: $1+\delta \lambda$, $1+\delta \mu$, $\delta \rho$ and $\delta P$. The
3035: occurrence of the constant 1 in the metric variables has no implications
3036: on the numerical performance. We have thus evolved
3037: initial data in the form of a fundamental eigenmode profile in the
3038: velocity field $w$ for neutron star model 3.
3039: First we have used the TOV-background and a resolution of 600
3040: grid points. We have then repeated the evolution with a flat space
3041: background using 600 and 1200 grid points in order to check the dependence
3042: of the non-perturbative results on the spatial resolution.
3043: It is important
3044: to note that the same code and the same evolution algorithm has been
3045: used in both cases.
3046: %We have only altered the outer boundary
3047: %conditions in the non-perturbative case where
3048: %we use the first order evolution of the predictor
3049: %step. With this choice we have obtained the best performance of
3050: %our non-perturbative code.
3051: %If we use the original boundary conditions we obtain
3052: %virtually the same result but on a shorter time scale.
3053: The amplitude of the perturbation corresponds
3054: to an oscillation of the surface of several metres. For this
3055: amplitude we still expect the evolution to be dominated by the
3056: harmonic time dependence, although the results of section \ref{MODECOUPLING}
3057: below indicate the presence of weak non-linear effects.
3058: The numerical results are shown in Fig.\,\ref{PERT_NONP}, where the
3059: central density perturbation is plotted as a function of time.
3060: %
3061: \begin{figure}[t]
3062: \centering
3063: \epsfig{file=pert_nonp.eps, height=320pt, width=220pt,
3064: angle=-90}
3065: \caption{The central density perturbation corresponding to the
3066: fundamental oscillation mode of model 3 as obtained
3067: with a perturbative method for 600 grid points
3068: (dotted curve) and a non-perturbative method for 600
3069: (solid)
3070: and 1200 grid points (dashed curve). See text for details.}
3071: \label{PERT_NONP}
3072: \end{figure}
3073: %
3074: We clearly see that the perturbative evolution results in the expected
3075: sinusoidal time dependence. In the non-perturbative case the
3076: central density shows similar oscillations but simultaneously
3077: the mean value decreases significantly.
3078: In longer runs this decrease is revealed
3079: to be exponential and thus indicates a starting evaporation of the star.
3080: Neutron star model 3, however, is located on the
3081: stable branch as we can clearly see in Fig.\,\ref{TOV_FAMILIES}
3082: and no collapse or evaporation is expected.
3083: Indeed the higher resolution run indicates that the non-perturbative scheme
3084: converges to the harmonic solution. In order
3085: to understand this behaviour of the non-perturbative scheme we recall
3086: the presence of background terms in the evolution equation
3087: for $w_{,t}$. If we consider the coefficients $\tilde{b}_2$ and
3088: $\tilde{\alpha}_{21}$ given by Eqs.\,(\ref{NONP_ALPHA21}), (\ref{NONP_B2}),
3089: we can see that the evolution equation (\ref{NONP_WT}) contains the
3090: background in the form
3091: %
3092: \begin{align}
3093: e(y) &:= \frac{1}{\mu^2 D} \left[ \frac{\lambda_{,r}}{\lambda}
3094: + \frac{C^2 \rho_{,r}}{(\rho + P)} \right]. \label{NONP_ERROR}
3095: \end{align}
3096: %
3097: We know that this term vanishes by virtue of the TOV equation
3098: %
3099: \begin{figure}[t]
3100: \centering
3101: \epsfig{file=sourcew.eps, height=300pt, width=220pt,
3102: angle=-90}
3103: \caption{The source terms of the evolution equation for $w$
3104: (\ref{PERT_WT}) at the
3105: first computational step are shown for the perturbative (solid curve)
3106: and the non-perturbative scheme (dashed curve). The dotted
3107: curve shows the numerical error of the background terms
3108: and demonstrates the significance of the spurious source terms.}
3109: \label{SOURCEW}
3110: \end{figure}
3111: %
3112: (\ref{TOV_PR}) and it has been removed from the equations in the perturbative
3113: formulation. In the non-perturbative case, however, it will manifest itself
3114: in the form of a residual numerical error.
3115: This error is shown in Fig.\,\ref{SOURCEW} for the first
3116: step in the evolution with
3117: 600 grid points together with the entire source terms of $w_{,t}$ as
3118: given by Eq.\,(\ref{NONP_WT}) in the non-perturbative and
3119: Eq.\,(\ref{PERT_WT}) in the perturbative case. Because of the cosine time
3120: dependence of the velocity the source terms should nearly vanish at $t=0$.
3121: It can be seen, however, that
3122: the source terms are dominated by the residual numerical
3123: error in the non-perturbative scheme which is particularly large
3124: at the centre and the surface. On the time scale of one oscillation period,
3125: about $150\,\,{\rm km}$, the spurious acceleration of up to
3126: $10^{-4}\,\,{\rm km}^{-1}$ will have a significant impact on the
3127: oscillation of several metres of the star. A closer investigation
3128: of the velocity field reveals that the integral effect of the residual
3129: error is an increase in the velocity field near the surface. We
3130: attribute the gradual evaporation of the star to this disturbance
3131: in the velocity field which gradually radiates matter off the numerical
3132: grid. \\
3133: Considering that the same code has been used for the comparison just described,
3134: it is necessary to check the perturbative scheme for similar
3135: spurious effects. After all the main advantage of the perturbative
3136: scheme lies in higher accuracy which may postpone the onset
3137: of a spurious collapse or evaporation but not necessarily avoid it.
3138: We have already mentioned, however, that no significant deviation
3139: from the harmonic time dependence has been observed in the case of model 3
3140: and initial data in the form of eigenmodes over very long times.
3141: In order to avoid even longer integration times and the associated
3142: computational costs, we have
3143: chosen an alternative way of testing the code for this behaviour. We use a
3144: stellar model identical to model 3 but with a central density of
3145: $\rho_{\rm c}=0.002802\,\,{\rm km}^{-2}$ which is just below the
3146: critical value given in Table \ref{CRITFREQS}.
3147: The initial data consist of the
3148: fundamental velocity mode with an amplitude corresponding to a
3149: surface displacement of about $10\,\,{\rm cm}$
3150: and we use a numerical grid with 600
3151: grid points. In the first calculation we have imposed a truncation density
3152: of $\delta \rho_{\rm s}=5\cdot 10^{-6}\,\,{\rm km}^{-2}$ and in a second run
3153: the intrinsic value of the TOV
3154: %
3155: \begin{figure}[t]
3156: \centering
3157: \epsfig{file=marg_stable.eps, height=300pt, width=220pt,
3158: angle=-90}
3159: \caption{The central density resulting from the evolution of the
3160: fundamental eigenmode of
3161: a neutron star corresponding to model 3 with a central density just
3162: below the critical value is plotted for a truncation
3163: density of $5\cdot 10^{-6}\,\,{\rm km}^{-2}$ (dashed curve)
3164: and $2.5\cdot 10^{-8}\,\,{\rm km}^{-2}$ (solid curve).
3165: }
3166: \label{PERT_MARGSTABLE}
3167: \end{figure}
3168: %
3169: code $\delta \rho_{\rm s}=2.5\cdot 10^{-8}\,\,{\rm km}^{-2}$ is used.
3170: In Fig.\,\ref{PERT_MARGSTABLE}
3171: we show the resulting central density perturbation as a function of time.
3172: For the small truncation density we obtain the expected sinusoidal
3173: time dependence whereas the larger value significantly affects the evolution,
3174: even though only a fraction of $10^{-5}$ of the stellar mass has been neglected
3175: in this case. This result demonstrates the limitations of the code in its
3176: current form. For larger truncation densities it does not necessarily
3177: guarantee mass conservation which we attribute to the boundary condition
3178: (\ref{PERT_BCOUTW}) which is strictly valid only if the numerical grid
3179: extends to $\rho=0$. For sufficiently small truncation densities
3180: the resulting numerical error is
3181: negligible and has no significant effect on the evolution. For larger
3182: truncation densities, however, it can result in spurious phenomena similar
3183: to those observed in the non-perturbative case. This is particularly
3184: problematic
3185: since the investigation of non-linear effects will require perturbations
3186: of larger amplitudes and consequently larger truncation densities are
3187: necessary in order to avoid total negative energy densities. From here
3188: on we will therefore proceed in two different ways. In the remainder
3189: of section \ref{DYNAMIC} we will investigate a simplified
3190: neutron star model for
3191: which the code ensures mass conservation for arbitrary amplitudes
3192: and negative energy densities are still avoided. This model will
3193: necessarily provide a less realistic description of a neutron star, but
3194: the general structure of the eigenmodes remains the same and it
3195: is not unrealistic to expect that non-linear effects such as mode coupling
3196: will be qualitatively similar in
3197: more realistic models. Considering the sensitivity of the numerical
3198: evolutions to the treatment of the surface, it is, however, desirable to
3199: develop a formulation of the dynamic neutron star which unambiguously
3200: provides a correct treatment of the surface. This will be done in
3201: section \ref{LAGR} where we develop a fully non-linear perturbative Lagrangian
3202: code.
3203:
3204:
3205: %=========================================================================
3206: \subsubsection{A simplified neutron star model}
3207: \label{NEWMODEL}
3208: %
3209: %
3210: In the previous section we have seen that a sufficiently large
3211: truncation density in combination with the boundary condition
3212: (\ref{PERT_BCOUTW}) may result in a continuous loss or gain of mass.
3213: In order to avoid total negative energy densities, however, we have to use
3214: sufficiently large truncation densities when we study
3215: non-linear effects in the time evolution of large amplitude
3216: perturbations. We have therefore decided to ensure mass conservation
3217: by using the alternative boundary condition
3218: %
3219: \begin{align}
3220: w &= 0 \label{PERT_BCOUTWMOD}
3221: \end{align}
3222: %
3223: at the surface instead of Eq.\,(\ref{PERT_BCOUTW}). This means that
3224: the surface of the star remains at a fixed position in space and
3225: only fluid elements in the interior of the star are displaced during the
3226: evolution. It is the fixed location of the surface
3227: which avoids the main problems
3228: we have encountered with the Eulerian formulation so far.
3229: The model we use for the following analysis
3230: has the same equation of state as model 3 of Table \ref{MODELS15}
3231: and a central density $\rho_{\rm c}=1.224\cdot 10^{-3}
3232: \,\,{\rm km}^{-2}$ which implies a radius $R=11.34\,\,{\rm km}$
3233: and a total mass $M=2.18\,\,{\rm km}$.
3234: The truncation density is fixed at $\rho_{\rm s}=2.0\cdot 10^{-4}
3235: \,\,{\rm km}^{-2}$ which means that the simplified model contains
3236: $90\,\%$ of the mass of the original star and extends to $84\,\%$ of the
3237: original radius. Apart from changing the
3238: truncation density in the calculation of the TOV-background
3239: and implementing the new boundary condition in the evolution code
3240: only one further modification in the numerical setup described in
3241: section \ref{PERT_NUMERICS} is required. The outer
3242: boundary condition (\ref{LIN_BCOUTA})
3243: in the calculation of the eigenmodes is replaced by
3244: %
3245: \begin{align}
3246: \xi(R) &= 0.
3247: \end{align}
3248: %
3249: The resulting eigenmodes can be ordered in the same way as described in
3250: section \ref{PERT_LIN} and the evolution of eigenmodes
3251: in the linear regime again results
3252: in harmonic time dependence as in the original case with
3253: the frequencies predicted by the eigenmode calculation. The first
3254: four eigenmode profiles of $\delta \rho$ and $w$ for the model mentioned
3255: above are shown in Fig.\,\ref{MOD_MODES}.
3256: %
3257: \begin{figure}[t]
3258: \centering
3259: \epsfig{file=mod_eigdrho.eps, height=200pt, width=150pt, angle=-90}
3260: \hspace{0.5cm}
3261: \epsfig{file=mod_eigw.eps, height=200pt, width=150pt, angle=-90}
3262: \caption{The profiles of the lowest four eigenmodes in $\delta \rho$
3263: and $w$ for the simplified neutron star model.}
3264: \label{MOD_MODES}
3265: \end{figure}
3266: %
3267: The plots show that the number of local maxima and minima of the
3268: %
3269: \begin{figure}[b]
3270: \centering
3271: \epsfig{file=mod_eigdrho2.eps, height=220pt, width=150pt, angle=-90}
3272: \hspace{0.5cm}
3273: \epsfig{file=mod_eigddrho2.eps, height=200pt, width=150pt, angle=-90}
3274: \caption{The evolution of the central density for initial data in the form
3275: a fundamental eigenmode in the velocity field for model 3
3276: with a central density $\rho_{\rm c}=0.002802\,\,{\rm km}^{-2}$.}
3277: \label{MOD_DRHOC}
3278: \end{figure}
3279: %
3280: profiles still corresponds to the order of the mode.
3281:
3282: %=======================================================================
3283: \subsubsection{Testing the code with the new model}
3284: %
3285: %
3286: The only modification of the code that needed to be implemented for the
3287: new model is the outer boundary condition (\ref{PERT_BCOUTWMOD}).
3288: The performance of the code in the linear regime is thus well established
3289: by the results of section \ref{EULER_LIN} and we merely have to demonstrate
3290: that no spurious results are obtained for larger truncation densities.
3291: This is the only case where we will depart from the model parameters listed
3292: in the previous section and use a central density
3293: $\rho_{\rm c}=0.002802\,\,{\rm km}^{-2}$ instead. We thus recover the
3294: parameters of the model which lead to a spurious evaporation of the star in
3295: Fig.\,\ref{PERT_MARGSTABLE}.
3296: For this model we have again evolved initial data in the form of the
3297: fundamental mode of the velocity with an amplitude of $10\,{\rm cm}$
3298: using 600 grid points. In Fig.\,\ref{MOD_DRHOC} we show the resulting
3299: central density $\delta \rho_{\rm c}$
3300: together with the deviation from the harmonic
3301: solution of the linearized case. For presentation purposes
3302: we only show the evolution up to $t=6000\,\,{\rm km}$. The harmonic
3303: time dependence is reproduced with reasonable accuracy as the deviation
3304: increases linearly up to about $1\,\%$. In general we have found the
3305: eigenmode frequency the quantity most vulnerable to numerical error as
3306: can be seen for example by varying the resolution. Because of this
3307: observation and the oscillatory character of the deviation in
3308: the figure
3309: we attribute the error mainly to the limited accuracy of the frequency
3310: rather than the numerical error of the time evolution itself.
3311: The increasing phase shift between the numerical and the analytic solution
3312: arising from the limited accuracy of the frequency
3313: %
3314: \begin{figure}[t]
3315: \centering
3316: \epsfig{file=qdrho.eps, height=200pt, width=175pt, angle=-90}
3317: \epsfig{file=qw.eps, height=200pt, width=175pt, angle=-90}
3318: \epsfig{file=qdlam.eps, height=200pt, width=175pt, angle=-90}
3319: \epsfig{file=qdmu.eps, height=200pt, width=175pt, angle=-90}
3320: \caption{The convergence factor obtained for evolving the second
3321: eigenmode with an amplitude of $70\,\,{\rm m}$ is shown
3322: for the variables $\delta \rho$, $w$, $\delta \lambda$
3323: and $\delta \mu$.}
3324: \label{MOD_CONV}
3325: \end{figure}
3326: %
3327: will result in a linear increase
3328: of the deviation as observed in Fig.\,\ref{MOD_DRHOC}.
3329: In spite of the small deviation this calculation
3330: is in sharp contrast with that shown in Fig.\,\ref{PERT_MARGSTABLE}, where
3331: a much smaller truncation density resulted in an exponential decay of the
3332: central density. We conclude that using a large truncation density
3333: in combination with the boundary condition (\ref{PERT_BCOUTWMOD})
3334: the code performs well in the linearized regime. \\
3335: We now return to the model parameters of the previous section and use
3336: $\rho_{\rm c}=0.001224\,\,{\rm km}^{-2}$.
3337: In order to test the code for convergence in the non-linear regime
3338: we have evolved the second eigenmode with an amplitude
3339: corresponding to a maximal displacement of fluid elements of $70\,\,{\rm m}$.
3340: % For this amplitude strong non-linear effects are present as we will
3341: %see below.
3342: The calculation has been carried out with 400, 800 and
3343: 1600 grid points and the resulting convergence factors
3344: are shown in Fig.\,\ref{MOD_CONV}. In spite of variations
3345: around the expected value 4, the results for all variables are compatible
3346: with second order convergence. \\
3347: For the next test we will use the code in the Cowling approximation, since
3348: the static metric provides a straightforward recipe to calculate
3349: conserved quantities. We have seen in section \ref{PERT_EQ} that
3350: only minor modifications are required
3351: to switch between the Cowling approximation and a dynamic metric.
3352: The conservation properties with a fixed metric will therefore represent
3353: a good test for the matter evolution in the general case.
3354: The first step in the derivation of a conserved quantity is to find a time-like
3355: Killing field. The existence of such a vector
3356: field follows from the static nature of the metric in the Cowling
3357: approximation. The Killing vector can be found by looking at the
3358: Killing equation
3359: %
3360: \begin{align}
3361: \nabla_{\mu} \hbox{\vec X}_{\nu} + \nabla_{\nu} \hbox{\vec X}_{\mu} &= 0.
3362: \end{align}
3363: %
3364: The resulting 10 differential equations can be solved rather easily
3365: and define the solution up to a constant
3366: factor. We choose this factor so that the Killing field can be written as
3367: %
3368: \begin{align}
3369: \hbox{\vec X}_{\mu} &= \left[ \lambda^2,0,0,0 \right].
3370: \end{align}
3371: %
3372: The conserved quantity then follows from contraction of the Killing field
3373: with the energy momentum tensor
3374: %
3375: \begin{align}
3376: \hbox{\vec J}^{\mu} &= \hbox{\vec{T}}^{\mu \nu} \hbox{\vec X}_{\nu}.
3377: \end{align}
3378: %
3379: By virtue of conservation of energy momentum this vector satisfies the
3380: condition
3381: %
3382: \begin{align}
3383: \nabla_{\mu} \hbox{\vec J}^{\mu} &= 0.
3384: \end{align}
3385: %
3386: With the metric (\ref{TOV_LINEELEMENT}) and the energy momentum tensor
3387: %
3388: \begin{figure}[t]
3389: \centering
3390: \epsfig{file=time_E.eps, height=300pt, width=250pt, angle=-90}
3391: \caption{The numerical evolution of the function $E$ obtained in the
3392: Cowling approximation. The quantity is conserved with an accuracy
3393: better than $10^{-4}$.}
3394: \label{CONSERVED}
3395: \end{figure}
3396: %
3397: (\ref{NONP_EMTENSOR}) this equation can be written in conservative form
3398: %
3399: \begin{align}
3400: \partial_t \left( \lambda \mu r^2 \hbox{\vec J}^t \right) + \partial_r \left(
3401: \lambda \mu r^2 \hbox{\vec J}^r \right) &= 0,
3402: \end{align}
3403: %
3404: where the $t$ and $r$ components of $\hbox{\vec{J}}$ are given by
3405: %
3406: \begin{align}
3407: \hbox{\vec J}^t &= (1+\mu^2 w^2) \hat{\rho} + \mu^2 w^2 \hat{P}, \\[10pt]
3408: \hbox{\vec J}^r &= \frac{w}{v} (1+\mu^2 w^2) (\hat{\rho} +\hat{P}).
3409: \end{align}
3410: %
3411: If we consider a general conservation law in one dimension
3412: %
3413: \begin{align}
3414: u_{,t} + F(u)_{,r} &= 0,
3415: \end{align}
3416: %
3417: we obtain after integration over $t$ and $r$
3418: %
3419: \begin{align}
3420: \int_0^R{[u(T,r)-u(0,r)]}dr + \int_0^T{[F(t,R)-F(t,0)]}dr &= 0.
3421: \end{align}
3422: %
3423: In our case the flux function is given by $F = \lambda \mu r^2 \hbox{\vec J}^r$
3424: and vanishes at $r=0$ and $r=R$ because the velocity $w$ vanishes
3425: at both boundaries. Consequently
3426: %
3427: \begin{align}
3428: E &= \int_0^R{\lambda \mu r^2 \hbox{\vec J}^t}dr
3429: \end{align}
3430: %
3431: is a conserved quantity. \\
3432: In order to test the conservation properties of the code
3433: we have evolved the same initial data as in the convergence
3434: analysis with the metric fixed at the background values.
3435: In Fig.\,\ref{CONSERVED}
3436: we show $E$ as a function of time as calculated with 800 grid points.
3437: The quantity
3438: is conserved with a relative accuracy better than
3439: $10^{-4}$. Even higher accuracy is
3440: obtained for smaller amplitudes of the initial data.
3441: We have thus demonstrated that the code performs well in the linear
3442: as well as the non-linear regime. The applicability of the
3443: code to a wide range of amplitudes will be crucial when we study
3444: non-linear effects in the evolution of eigenmodes in the next subsection.
3445:
3446:
3447: %=========================================================================
3448: \subsubsection{Non-linear mode coupling}
3449: \label{MODECOUPLING}
3450: %
3451: %
3452: {\em (a) Measuring the eigenmode coefficients} \\[5pt]
3453: We will now use the simplified neutron star model described in
3454: section\,\ref{NEWMODEL} to study the coupling of eigenmodes
3455: in non-linear evolutions of
3456: radial oscillations.
3457: % It is necessary therefore to find a way of measuring
3458: %the presence of the individual eigenmodes in the evolution. For this
3459: %purpose we recall the {\em Sturm-Liouville problem} (\ref{LIN_ZETARR})
3460: %which determines the eigenmode solutions in terms of the displacement
3461: %vector $\zeta$.
3462: In order to measure the presence of the individual eigenmodes
3463: in the evolution we recall the {\em Sturm-Liouville problem}
3464: (\ref{LIN_ZETARR}) which determines the eigenmode solutions
3465: in terms of the rescaled displacement vector $\zeta$.
3466: In section \ref{PERT_LIN} we have seen that the solutions
3467: $\zeta_i$ form a complete orthonormal system with respect to
3468: the inner product defined in Eq.\,(\ref{LIN_INNERPRODUCT}). This
3469: property enables us to quantify the contributions of the different
3470: eigenmodes in the evolution at any given time.
3471: We need to calculate the displacement $\zeta(t,r)$ of a
3472: fully non-linear evolution from the fundamental variables $\delta \rho$
3473: and $w$. For this purpose we eliminate $\xi$ from Eqs.\,(\ref{LIN_XI})
3474: and (\ref{LIN_ZETA}) and obtain
3475: %
3476: \begin{align}
3477: \zeta_{,t} &= r^2 w. \label{MC_ZETAT}
3478: \end{align}
3479: %
3480: The initial values of $\zeta$ follow from the initial data
3481: which we provide in the
3482: form of an eigenmode in the velocity field $w$ and
3483: zero energy density perturbation $\delta \rho$. We can see from
3484: Eq.\,(\ref{LIN_DRHOOFZETA}) that the initial displacement $\zeta$
3485: vanishes as a consequence.
3486: At any time $t$ we can then expand the non-linear
3487: displacement $\zeta(t,r)$ in terms of the eigenmodes
3488: %
3489: \begin{align}
3490: \zeta(t,r) &= \sum_i{A_i(t) \zeta_i(r)}, \label{ZETAEXPANSION}
3491: \end{align}
3492: %
3493: where the time dependent coefficients are given by the inner product
3494: %
3495: \begin{align}
3496: A_i(t) &= \langle \zeta(t,r),\zeta_i(r) \rangle .
3497: \end{align}
3498: %
3499: In practice we prefer to calculate the eigenmode coefficients from
3500: the time derivative of this equation
3501: %
3502: \begin{align}
3503: \frac{\partial A_i}{\partial t} &= \langle \zeta_{,t}, \zeta_i \rangle,
3504: \end{align}
3505: %
3506: where we have dropped the $t$ and $r$ dependence for convenience.
3507: If we substitute Eq.\,(\ref{MC_ZETAT}) for $\zeta_{,t}$ we obtain the
3508: final result
3509: %
3510: \begin{align}
3511: \frac{\partial A_i}{\partial t} &= \langle r^2 w, \zeta_i \rangle.
3512: \end{align}
3513: %
3514: We can thus calculate the time derivative of the coefficients
3515: and use the initial values to obtain the coefficients at
3516: any given time $t$. In our case all coefficients are zero initially
3517: because of the vanishing of $\zeta$. The integral appearing
3518: in the definition of the inner product is calculated with the
3519: fourth order Simpson method (see for example \shortciteNP{Press1989}). \\
3520: It is also interesting to consider the relative coefficients defined by
3521: %
3522: \begin{align}
3523: R_i(t) &= \frac{\langle \zeta , \zeta_i \rangle}
3524: {\langle \zeta, \zeta \rangle},
3525: \end{align}
3526: %
3527: whenever $\zeta$ is a non-zero function. If we multiply this equation
3528: %
3529: \begin{figure}[t]
3530: \centering
3531: \epsfig{file=Rsum.eps, height=300pt, width=250pt, angle=-90}
3532: \caption{The sum of the first ten $R_i$ has been calculated
3533: for evolving the second eigenmode with an amplitude of
3534: $70\,\,{\rm m}$.}
3535: \label{MOD_SUMR}
3536: \end{figure}
3537: %
3538: by $A_i$ and sum over $i$, we can use Eq.\,(\ref{ZETAEXPANSION}) to
3539: obtain the relation
3540: %
3541: \begin{align}
3542: \sum_i{R_i} &= 1, \label{MC_RSUM}
3543: \end{align}
3544: %
3545: which can be used to check the completeness of the numerically
3546: calculated eigenmodes. For this purpose we have
3547: evolved the second eigenmode with a large amplitude corresponding to
3548: a maximum displacement of
3549: $70\,\,{\rm m}$ and calculated the sum of the first ten weighted coefficients
3550: $R_i$ using 600 grid points. The result is shown in Fig.\,\ref{MOD_SUMR}
3551: and demonstrates
3552: that Eq.\,(\ref{MC_RSUM}) is satisfied to within less than one per cent.
3553: This does not only confirm the completeness of the system of eigenmodes,
3554: but also indicates that the energy essentially remains within the lowest
3555: ten eigenmodes.
3556: In order to check the orthonormality we have calculated the inner products
3557: of the eigenmodes. The results for
3558: the lowest five eigenmodes are shown in Table \ref{MOD_ORTHONORMALITY}
3559: and demonstrate that
3560: %
3561: \begin{table}[t]
3562: \caption{The inner product $\langle \zeta_i, \zeta_j \rangle$ between
3563: the five lowest eigenmodes.}
3564: \begin{center}
3565: \begin{tabular}{c|ccccc}
3566: \hline \hline
3567: & $\zeta_1$ & $\zeta_2$ & $\zeta_3$ & $\zeta_4$ & $\zeta_5$\\
3568: \hline
3569: $\zeta_1$ & 1.0 & $-2.1\cdot 10^{-6}$ & $-6.3\cdot 10^{-6}$
3570: & $-1.2\cdot 10^{-6}$ & $-2.6\cdot 10^{-6}$ \\
3571: $\zeta_2$ & $-2.1\cdot 10^{-6}$ & 1.0 & $-8.0\cdot 10^{-6}$
3572: & $-1.5\cdot 10^{-5}$ & $-6.3\cdot 10^{-6}$ \\
3573: $\zeta_3$ & $-6.3\cdot 10^{-6}$ & $-8.0\cdot 10^{-6}$ & 1.0
3574: & $-1.8\cdot 10^{-5}$ & $-2.7\cdot 10^{-5}$ \\
3575: $\zeta_4$ & $-1.2\cdot 10^{-6}$ & $-1.5\cdot 10^{-5}$
3576: & $-1.8\cdot 10^{-5}$ & 1.0 & $-3.2\cdot 10^{-5}$ \\
3577: $\zeta_5$ & $-2.6\cdot 10^{-6}$ & $-6.3\cdot 10^{-6}$
3578: & $-2.7\cdot 10^{-5}$ & $-3.2\cdot 10^{-5}$ & 1.0 \\
3579: \hline \hline
3580: \end{tabular}
3581: \label{MOD_ORTHONORMALITY}
3582: \end{center}
3583: \end{table}
3584: %
3585: the orthonormality condition (\ref{LIN_ORTHONORMALITY}) is satisfied with high
3586: accuracy. \\
3587:
3588: {\em (b) Non-linear coupling between eigenmodes} \\[5pt]
3589: In order to study the coupling of modes due to non-linear effects we have
3590: provided initial data in the form of one velocity eigenmode. The order
3591: of the eigenmode $j$ and the amplitude of the initial data $K_j$ are
3592: free parameters that determine the physical setup. We will specify the
3593: amplitude of the initial perturbation by the maximum value of the
3594: eigenmode profile of the displacement vector $\xi$ corresponding to the
3595: initial velocity perturbation. This is a measure
3596: for the maximum displacement a fluid element of the interior of the
3597: star will undergo. During the evolution
3598: we calculate the eigenmode coefficients
3599: $A_i(t)$ with $1\le i \le 10\,\,{\rm or}\,\,15$
3600: according to the method described above. Due to the oscillatory character
3601: of the modes, the coefficients will also oscillate during the evolution.
3602: This is shown in Fig.\,\ref{MOD_EVOLA2A4}
3603: where we plot the coefficients $A_2(t)$ and
3604: $A_4(t)$ for evolving the second eigenmode. A large amplitude
3605: corresponding to a maximum displacement of $70\,\,{\rm m}$ has been used
3606: %
3607: \begin{figure}[t]
3608: \centering
3609: \epsfig{file=time_A2.eps, height=200pt, width=150pt, angle=-90}
3610: \epsfig{file=time_A4.eps, height=200pt, width=150pt, angle=-90}
3611: \caption{The coefficients $A_2(t)$ and $A_4(t)$ obtained for initial
3612: data in the form of the second eigenmode with an amplitude
3613: of $70\,{\rm m}$.}
3614: \label{MOD_EVOLA2A4}
3615: \end{figure}
3616: %
3617: for this calculation and we can clearly see the transfer of energy
3618: between the second and the fourth mode. It is interesting to see
3619: that the energy transferred to the fourth mode does not remain there
3620: but instead is periodically passed back and forth between the two modes.
3621: We observe
3622: a qualitatively similar behaviour for the other eigenmodes, although these are
3623: excited less efficiently.
3624: If we want to investigate this coupling between eigenmodes more systematically,
3625: we need to quantify the degree to which a particular mode has been excited
3626: in an evolution. For this purpose we will use
3627: the maximum value of the corresponding coefficient
3628: obtained during that evolution. We will refer to these maxima by
3629: $A_i$ as opposed to $A_i(t)$ used for the time dependent coefficients.
3630: We have thus evolved the eigenmodes $i=1$, 2 and 3, referred to
3631: %
3632: \begin{figure}[t]
3633: \centering
3634: \epsfig{file=A0ia.eps, height=375pt, width=250pt, angle=-90}
3635: \epsfig{file=A0ib.eps, height=375pt, width=250pt, angle=-90}
3636: \caption{The eigenmode coefficients for the first ten eigenmodes are
3637: shown as a function of the amplitude $K_1$
3638: for initial data in the form of the fundamental velocity
3639: mode.}
3640: \label{MOD_COUPLING1A}
3641: \end{figure}
3642: %
3643: as case 1, 2 and 3 from now on, with amplitudes
3644: ranging between $1\,\,{\rm cm}$ and $100\,\,{\rm m}$.
3645: At some stage in the range between about $50\,{\rm m}$
3646: and $100\,{\rm m}$ we observed the onset of shock formation.
3647: The accuracy of the eigenmode coefficients resulting from these
3648: evolutions is not clear. In this discussion we have therefore only
3649: used amplitudes for which no discontinuities are observed.
3650: %
3651: \begin{figure}[h]
3652: \centering
3653: \epsfig{file=square01a.eps, height=375pt, width=250pt, angle=-90}
3654: \epsfig{file=square01b.eps, height=375pt, width=250pt, angle=-90}
3655: \caption{The excitation of eigenmodes has
3656: been fitted with quadratic power
3657: laws in the range between
3658: $K_1=1\,\,{\rm m}$ and $10\,\,{\rm m}$.}
3659: \label{MOD_COUPLING1B}
3660: \end{figure}
3661: %
3662: For the numerical runs we have used 3200 grid points and an integration time
3663: of $1500\,\,{\rm km}$. Test runs over significantly longer times
3664: did not lead to significantly different results for the $A_i$
3665: which is compatible with the periodic exchange of energy shown in
3666: Fig.\,\ref{MOD_EVOLA2A4}.
3667: The high grid resolution on the other hand enables us to measure small
3668: eigenmode coefficients with good accuracy. \\
3669:
3670: {\sl Case 1:} \\[5pt]
3671: We start our analysis with case 1, where the fundamental mode is excited
3672: initially. In Fig.\,\ref{MOD_COUPLING1A} we plot the coefficients
3673: $A_i$ as a function of the initial amplitude $K_1$ for the first ten
3674: %
3675: \begin{figure}[h]
3676: \centering
3677: \epsfig{file=combined01a.eps, height=375pt, width=250pt, angle=-90}
3678: \caption{The eigenmode coefficients $A_2$, $A_3$
3679: and $A_4$ are fitted with linear combinations of power laws
3680: according to Eqs.\,(\ref{FIT_01A2})-(\ref{FIT_01A4}).}
3681: \label{MOD_COUPLING1C}
3682: \end{figure}
3683: %
3684: eigenmodes.
3685: We find that the coefficient $A_1$ increases linearly with the amplitude
3686: $K_1$ as expected. A closer investigation of the higher eigenmode
3687: coefficients, however, reveals the presence of two distinct regimes.
3688: %
3689: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
3690: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
3691: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
3692: \item In a weakly non-linear regime
3693: for amplitudes up to about $10\,\,{\rm m}$ all coefficients
3694: $A_2,\ldots A_{10}$ increase quadratically with the amplitude $K_1$.
3695: Deviations from this quadratic power law at very small amplitudes
3696: are due to the limited numerical accuracy in calculating
3697: the coefficients.
3698: \item At larger amplitudes all eigenmode coefficients except for $A_2$
3699: show a transition to power laws with larger exponent which marks
3700: a moderately non-linear regime.
3701: \end{list}
3702: %
3703: We have illustrated this behaviour in
3704: Fig.\,\ref{MOD_COUPLING1B}
3705: where the eigenmode coefficients have been approximated with
3706: quadratic power laws
3707: %
3708: \begin{align}
3709: A_i &= c_i\cdot K_1^{\,\,2}. \label{MOD_SQUARE01}
3710: \end{align}
3711: %
3712: The coupling coefficients $c_i$ which represent the coupling strength in the
3713: weakly non-linear regime have been obtained from least square fits of
3714: quadratic power laws to the eigenmode coefficients in amplitude ranges
3715: between $0.1\,\,{\rm m}$ and $10\,\,{\rm m}$. It is interesting to investigate
3716: the dependence of the coupling coefficients
3717: on the order of the eigenmodes. This is shown in the upper left panel
3718: of Fig.\,\ref{MOD_CI}, where we plot $c_i$ over the order $i-1$.
3719: %
3720: \begin{figure}[t]
3721: \centering
3722: \epsfig{file=ci.eps, height=215pt, width=175pt, angle=-90}
3723: \epsfig{file=ci02.eps, height=215pt, width=175pt, angle=-90}
3724: \epsfig{file=ci03.eps, height=215pt, width=175pt, angle=-90}
3725: \caption{The coupling coefficients $c_i$ defined in Eq.\,\ref{MOD_SQUARE01}
3726: are plotted as a function of the mode number $i-1$ for case 1
3727: in the upper left panel. In the upper right and lower panel
3728: we plot the corresponding
3729: coefficients for case 2 and 3
3730: as a function of the mode number $i-3$ and $i-5$ respectively.
3731: In all cases the coefficients can be approximated with
3732: inverse cubic power laws as indicated by the solid lines.}
3733: \label{MOD_CI}
3734: \end{figure}
3735: %
3736: The solid line in this figure shows a power law fit for these coupling
3737: coefficients given by
3738: %
3739: \begin{align}
3740: c_i &= 3.2\cdot 10^{-7}\cdot (i-1)^{-3}. \label{CI}
3741: \end{align}
3742: %
3743: This result is compatible with the expectation that $c_i \rightarrow 0$
3744: as $i\rightarrow \infty$. Otherwise an infinite number of modes would each be
3745: excited with a finite amount of energy.
3746: In the moderately non-linear regime the eigenmode coefficients
3747: $A_3,\ldots,A_{10}$ show a higher order growth with the amplitude
3748: $K_1$. For the most efficiently excited modes 2, 3 and 4 we have been able
3749: to approximate the eigenmode coefficients with the following combinations
3750: of power laws
3751: %
3752: \begin{align}
3753: A_2 &= 3.6\cdot10^{-7}\cdot K_1^{\,2}, \label{FIT_01A2} \\[10pt]
3754: A_3 &= 3.4\cdot 10^{-8}\cdot K_1^2+9.7\cdot 10^{-10}\cdot K_1^3,
3755: \label{FIT_01A3} \\[10pt]
3756: A_4 &= 1.0\cdot 10^{-8}\cdot K_1^2+1.2\cdot 10^{-11}\cdot K_1^4.
3757: \label{FIT_01A4}
3758: \end{align}
3759: %
3760: Here the higher order power laws have been obtained from fitting
3761: the eigenmode coefficients after subtracting the quadratic contributions.
3762: The resulting fits are shown in Fig.\,\ref{MOD_COUPLING1C}.
3763: The higher order contributions for the higher eigenmodes is rather
3764: weak so that it is difficult to obtain accurate measurements of the
3765: %
3766: \begin{figure}[t]
3767: \centering
3768: \epsfig{file=even_A0i.eps, height=375pt, width=250pt, angle=-90}
3769: \epsfig{file=odd_A0i.eps, height=375pt, width=250pt, angle=-90}
3770: \caption{The eigenmode coefficients for the first ten eigenmodes are
3771: shown for initial data in the form of the second velocity mode.}
3772: \label{MOD_COUPLING2A}
3773: \end{figure}
3774: %
3775: corresponding power law exponents. It is thus not clear whether the regular
3776: pattern suggested by Eqs.\,(\ref{FIT_01A2})-(\ref{FIT_01A4}) remains valid
3777: for higher modes. The steepening of the curves in the
3778: %
3779: \begin{figure}[t]
3780: \centering
3781: \epsfig{file=square02a.eps, height=375pt, width=250pt, angle=-90}
3782: \epsfig{file=square02b.eps, height=375pt, width=250pt, angle=-90}
3783: \caption{The excitation of eigenmodes in case 2 has been fitted with
3784: quadratic power laws in the range between $K_1=0.1\,\,{\rm m}$
3785: and $10\,\,{\rm m}$.}
3786: \label{MOD_COUPLING2B}
3787: \end{figure}
3788: %
3789: moderately non-linear regime, however, can be clearly seen in
3790: Fig.\,\ref{MOD_COUPLING1B}. \\
3791:
3792: {\sl Case 2:} \\[5pt]
3793: We will now address the question to what extent these results remain valid
3794: if we initially excite higher modes. For this purpose we have repeated the
3795: %
3796: \begin{figure}[t]
3797: \centering
3798: \epsfig{file=higher02.eps, height=375pt, width=250pt, angle=-90}
3799: \epsfig{file=combined02.eps, height=375pt, width=250pt, angle=-90}
3800: \caption{In the upper panel we show the higher order power law contributions
3801: of Eqs.\,(\ref{FIT02_04})-(\ref{FIT02_10}) which fit the
3802: even eigenmode coefficients rather well in the moderately
3803: non-linear regime. The lower panel shows the
3804: resulting fits obtained from the sum of the quadratic and the higher
3805: order power laws according to the same equations.}
3806: \label{MOD_COUPLING2C}
3807: \end{figure}
3808: %
3809: numerical analysis by providing initial data in the form of the second
3810: velocity mode. The resulting eigenmode coefficients are shown
3811: as a function of the amplitude $K_2$ in Fig.\,\ref{MOD_COUPLING2A}.
3812: The presence of the two distinct regimes is again clearly demonstrated by the
3813: figures and a closer investigation confirms the quadratic growth of the
3814: eigenmode coefficients in the weakly non-linear regime. This is demonstrated
3815: in Fig.\,\ref{MOD_COUPLING2B} where the corresponding quadratic power law
3816: fits are shown for the eigenmodes. We also observe a similar dependence
3817: of the quadratic coupling coefficients $c_i$ on the mode number. In
3818: case 1 we observed a power law relation given by Eq.\,(\ref{CI}) between the
3819: coefficients $c_i$ and the mode number $i-1$. In case 2 we can also
3820: approximate the coefficients $c_i$ reasonably well with
3821: an inverse cubic power law if we use the number $i-3$ instead which is
3822: demonstrated in the right panel of Fig.\,\ref{MOD_CI}. The lower order
3823: modes 1 and 3 do not fit into this pattern and we shall
3824: %In Fig.\,\ref{MOD_COUPLING2B} we furthermore observe the
3825: %super-quadratic dependence of the eigenmode
3826: %coefficients $A_i$ on the amplitude $K_2$ in the moderately non-linear
3827: %region. \\
3828: readdress their behaviour in the quadratic regime below when we discuss
3829: case 3. \\
3830: Apart from these similarities there are some interesting
3831: differences between case 1 and case 2:
3832: %
3833: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
3834: \labelwidth1cm \leftmargin1.0cm \labelsep0.4cm \rightmargin0cm
3835: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
3836: \item The transition from the weakly to the moderately non-linear
3837: regime occurs at smaller amplitudes than in case 1. This
3838: is particularly pronounced in the case of
3839: mode 6 (see Fig.\,\ref{MOD_COUPLING2B}).
3840: \item The regular pattern observed in case 1 in the moderately non-linear
3841: regime for the strongly excited modes 2, 3
3842: and 4, which is expressed in Eqs.\,(\ref{FIT_01A2})-(\ref{FIT_01A4}),
3843: is now being observed for the eigenmodes of even order
3844: $2n$. We obtain excellent fits for the data if we model the even
3845: eigenmode coefficients with the following linear combinations of
3846: power laws.
3847: %
3848: \begin{align}
3849: A_4&=1.9\cdot10^{-6}\cdot K_1^2, \label{FIT02_04} \\[10pt]
3850: A_6&=2.5\cdot10^{-8}\cdot K_1^2+8.7\cdot10^{-9}\cdot K_1^3,
3851: \end{align}
3852: \begin{align}
3853: A_8&=4.9\cdot10^{-9}\cdot K_1^2+6.2\cdot10^{-11}\cdot K_1^4, \\[10pt]
3854: A_{10}&=1.8\cdot10^{-9}\cdot K_1^2+4.9\cdot10^{-13}\cdot K_1^5.
3855: \label{FIT02_10}
3856: \end{align}
3857: %
3858: In Fig.\,\ref{MOD_COUPLING2C} we show the curves resulting from the
3859: higher order power laws as well as those corresponding to the
3860: linear combinations.
3861: For the odd modes the higher order contributions are rather
3862: small so that we cannot accurately measure the corresponding power
3863: law indices. The steepening of the curves and thus the onset
3864: of the moderately non-linear regime, however, is clearly visible.
3865: \item Whereas the quadratic coupling coefficients $c_i$ shown in the right
3866: panel of Fig.\,\ref{MOD_CI} show a continuous decrease with the
3867: order of the mode starting with mode 4,
3868: a clear preference of the second mode to
3869: couple to modes
3870: of even order $2n$ is observed in the moderately non-linear regime.
3871: This is indicated by the rather efficient coupling to mode 4
3872: and the significantly steeper increase of the
3873: eigenmode coefficients $A_6$, $A_8$ and $A_{10}$ for
3874: larger amplitudes $K_2$ in Fig.\,\ref{MOD_COUPLING2B}.
3875: \item A small flattening of the even eigenmode coefficients at
3876: large amplitudes in Fig.\,\ref{MOD_COUPLING2C} may indicate the onset
3877: of saturation effects. A possible mechanism for saturation is
3878: the formation of discontinuities. As we have already mentioned
3879: we have chosen an amplitude range in which no shock formation is
3880: observed. At the high end of our amplitude range, it may be
3881: possible, however, that
3882: similar dissipative effects due to the strong non-linearity
3883: start having an effect on the coupling of eigenmodes.
3884: \end{list}
3885: %
3886: {\sl Case 3:} \\[5pt]
3887: Next we consider case 3 where we perturb the star with the third velocity
3888: mode. The fundamental observations we have made in the previous two cases
3889: %
3890: \begin{table}[t]
3891: \begin{center}
3892: \caption{The quadratic coupling coefficients $c_i$ for the lower modes
3893: in case 3.}
3894: \vspace{0.5cm}
3895: \begin{tabular}{c|c}
3896: \hline \hline
3897: $i$ & $c_i$ \\
3898: \hline
3899: 1 & $2.0\cdot10^{-7}$ \\
3900: 2 & $1.2\cdot10^{-7}$ \\
3901: 4 & $6.7\cdot10^{-8}$ \\
3902: 5 & $3.0\cdot10^{-8}$ \\
3903: \hline \hline
3904: \end{tabular}
3905: \label{MOD_CI03}
3906: \end{center}
3907: \end{table}
3908: %
3909: are confirmed by the results in this case. In the weakly
3910: non-linear regime all eigenmode coefficients (except for $A_3$) grow
3911: quadratically with the amplitude $K_3$. The corresponding quadratic
3912: coupling coefficients can once more be approximated with a power law
3913: with exponent $-3$. We find, however, that the relevant mode number is
3914: now $i-5$. This behaviour is graphically illustrated in the lower panel
3915: of Fig.\,\ref{MOD_CI} where the coupling coefficients are shown together
3916: with the power law approximation.
3917: The results of this figure suggest the following
3918: regular pattern: For initial data in the form of eigenmode $j$ the quadratic
3919: coupling coefficients starting with mode $2j$
3920: are well approximated by an inverse cubic power law of a
3921: relative mode number $i+1-2j$ which is $1$ for mode $2j$, $2$ for
3922: mode $2j+1$ and so on. \\
3923: We still have to analyse the quadratic coupling
3924: coefficients of the modes below $2j$. In case 1 and 2 we did not have
3925: enough data to derive any results for these modes. For case 3 we have listed
3926: the corresponding coefficients $c_i$ in Table \ref{MOD_CI03}.
3927: The coefficients $c_i$ are approximately reduced by a factor of 2 each time the
3928: mode number is increased which may indicate an exponential decrease of the
3929: quadratic coupling coefficients for the low order modes.
3930: %
3931: \begin{figure}[t]
3932: \centering
3933: \epsfig{file=combined03a.eps, height=375pt, width=250pt, angle=-90}
3934: \caption{The eigenmode coefficient $A_3$, $A_6$,$\ldots$,$A_{15}$ are
3935: shown for case 3 together with the resulting fits
3936: according to Eqs.\,(\ref{FIT03_06})-(\ref{FIT03_15}).}
3937: \label{MOD_COUPLING3A}
3938: \end{figure}
3939: %
3940: This is only a vague conclusion from a small data set, however, and needs to
3941: be confirmed by studies of higher eigenmodes. \\
3942: In the moderately non-linear regime we have seen for case 2
3943: a preferred coupling to modes with an even order
3944: $2n$. In analogy we find that the
3945: third eigenmode couples more efficiently to modes of order $3n$ for larger
3946: amplitudes. Again
3947: we can approximate the results with good accuracy with combinations of
3948: two power laws analogous to Eqs.\,(\ref{FIT02_04})-(\ref{FIT02_10})
3949: %
3950: \begin{align}
3951: A_6 &= 1.2\cdot 10^{-6}\cdot K_3^2, \label{FIT03_06} \\[10pt]
3952: A_9 &= 0.9\cdot 10^{-8}\cdot K_3^2+6.5\cdot 10^{-9}\cdot K_3^3, \\[10pt]
3953: A_{12} &= 2.2\cdot 10^{-9}\cdot K_3^2+4.6\cdot 10^{-11}\cdot K_3^4, \\[10pt]
3954: A_{15} &= 7.8\cdot 10^{-10}\cdot K_3^2+4.5\cdot 10^{-13}\cdot K_3^5.
3955: \label{FIT03_15}
3956: \end{align}
3957: %
3958: We recognise the same pattern of increasing integer power law indices in the
3959: higher order terms that we have already found in case 1 and 2.
3960: These results are graphically illustrated in
3961: Fig.\,\ref{MOD_COUPLING3A}. Again the higher order contributions in the
3962: other eigenmodes is clearly present but too weak to facilitate an accurate
3963: measurement of the exponents. \\
3964: We conclude the study of non-linear mode coupling with a summary of the
3965: key results.
3966: %
3967: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
3968: \labelwidth1cm \leftmargin1.0cm \labelsep0.4cm \rightmargin0cm
3969: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
3970: \item We clearly observe two distinct regimes in the non-linear
3971: coupling of eigenmodes. In the weekly non-linear regime,
3972: normally up to amplitudes of several metres, all eigenmode
3973: coefficients grow
3974: quadratically with the amplitude $K_j$. In the moderately
3975: non-linear regime we observe a steeper increase of
3976: the coefficients $A_i$.
3977: % which in many cases can be reasonably well modelled
3978: % with higher order power laws.
3979: \item In the quadratic regime the coupling coefficients $c_i$ generally
3980: decrease with increasing order of the eigenmodes. If the initial
3981: perturbation is given in the form of mode $j$, we can model the
3982: behaviour of the quadratic coupling coefficients with an inverse
3983: cubic power law of the mode number
3984: starting with mode $2j$. The coupling to lower
3985: modes does not obey the same pattern, but we also observe a decrease
3986: of the $c_i$ with increasing mode number for these modes.
3987: This decrease may have exponential character.
3988: \item In the moderately non-linear regime an initially present mode $j$
3989: shows a preference to couple to modes of order $n\cdot j$ where
3990: $n\ge 2$ is an integer number. In these
3991: cases we can accurately model the dependence of the eigenmode
3992: coefficients on
3993: the amplitude $K_j$ with the sum of a quadratic and a higher order power
3994: law with exponent $n$: $A_i=c_i\cdot K_j^{\,2} + d_i \cdot K_j^{\,n}$
3995: for $i=n\cdot j$.
3996: \item In some cases we observe a flattening of the eigenmode coefficients
3997: at amplitudes of about $50\,{\rm m}$ which may indicate the onset
3998: of saturation.
3999: \end{list}
4000: %
4001:
4002:
4003: %=========================================================================
4004: \subsubsection{Discussion of the non-linear mode-coupling}
4005: %
4006: %
4007: In the previous section we have studied the coupling of eigenmodes due
4008: to non-linear effects by evolving a single eigenmode with varying
4009: amplitude. Concerning the transfer of energy to other modes
4010: we have found two distinct regimes, a weakly non-linear regime where the
4011: excitation of modes grows quadratically with the initial amplitude and
4012: a moderately non-linear regime, where this increase can be reasonably
4013: well described by power laws of higher order. \\
4014: % Given an initial mode
4015: %$j$ we have also seen that the coupling to modes $n\cdot j$ is particularly
4016: %efficient in the moderately non-linear regime. \\
4017: In the analytic study of non-linear mode coupling one normally views the
4018: eigenmode coefficients as harmonic oscillators and the non-linear interaction
4019: between eigenmodes is represented in the form of driving terms which are
4020: quadratic or of higher order in the amplitudes (see for example
4021: \citeNP{VanHoolst1996})
4022: %
4023: \begin{align}
4024: \frac{d^2 A_i}{dt^2} + \omega_i^2 A_i &= c_i^{jk} A_jA_k + d_i^{jkl}
4025: A_j A_k A_l + \ldots,
4026: \label{FORCEDOSCILLATOR}
4027: \end{align}
4028: %
4029: where the $c_i^{jk}$, $d_i^{jkl},\ldots$ are the quadratic, cubic
4030: and higher order coupling coefficients and summation over $j,k,l$
4031: is assumed. In our analysis the initial data
4032: consists in one isolated eigenmode $j$, so that the right hand side
4033: can be approximated by $c_i A_j^2 + d_i A_j^3 + \ldots$
4034: In analytic studies this series expansion is normally truncated at second
4035: or third order. In view of our results
4036: the omission of higher order terms
4037: seems to be justified in the weakly non-linear regime, where
4038: our fully non-linear simulations confirm that quadratic terms in the
4039: initial amplitude dominate the coupling between eigenmodes.
4040: This is no longer true,
4041: however, in the moderately non-linear regime, where higher order terms
4042: are more important. In particular the regular pattern suggested
4043: for example by Eqs.\,(\ref{FIT02_04})-(\ref{FIT02_10}) indicates that
4044: the excitation of higher order modes is dominated by increasingly
4045: higher order powers of the initial amplitude. It is not clear how
4046: this behaviour can be modelled in the framework of a finite
4047: series expansion of the type (\ref{FORCEDOSCILLATOR}). It rather seems
4048: that the use of fully non-linear methods such as the numerical
4049: technique described in this work is necessary in order to obtain
4050: a comprehensive description of the coupling between eigenmodes
4051: in the moderately non-linear regime. In terms of the maximum displacement
4052: of fluid elements in the star this corresponds to initial amplitudes
4053: as low as a couple of metres. \\
4054: We have also observed that given an initial mode
4055: $j$ the coupling to modes $n\cdot j$ is particularly
4056: efficient in the moderately non-linear regime. We interprete this as a
4057: resonance effect, which we illustrate in the simple case of a forced oscillator
4058: %
4059: \begin{align}
4060: \frac{d^2 A_i}{dt^2} + \omega_i^2 A_i = F\sin{\Omega t},
4061: \end{align}
4062: %
4063: where $\Omega$ is the frequency and $F$ the amplitude of the external force.
4064: The particular
4065: integral of this ordinary differential equation is
4066: %
4067: \begin{align}
4068: A_i(t) &= \frac{F}{\omega_i^2 - \Omega^2} \sin{\Omega t},
4069: \end{align}
4070: %
4071: which implies resonance if $\omega_i=\Omega$.
4072: If we assume that resonance occurs for any integer multiple
4073: of the frequency $\Omega$ in the general non-linear case,
4074: we can schematically write the eigenmode coefficients in the form
4075: %
4076: \begin{align}
4077: A_i(t) &= \sum_n \frac{F_n}{\omega_i^2 - (n\Omega)^2}, \label{RESONANCE}
4078: \end{align}
4079: %
4080: where the $F_n$ may depend on the frequencies.
4081: The analytic study of non-linear mode coupling up to
4082: cubic order leads to eigenmode coefficients which resemble this pattern
4083: [see for example Eqs.\,(18), (19) of \citeNP{VanHoolst1996}].
4084: In our case the
4085: external force is provided by the non-linear coupling to the initial mode $j$,
4086: so that $\Omega=\omega_j$. We therefore obtain
4087: resonance in Eq.\,(\ref{RESONANCE}) if $\omega_i = n \omega_j$. As can
4088: be seen for example in Fig.\,\ref{PERT_LINFOUR}, the eigenfrequencies
4089: of radial neutron star oscillations
4090: are fairly equally spaced in the frequency domain with the exception of the
4091: fundamental mode and we can reasonably well approximate
4092: $\omega_i\approx (i \omega_j)/j$ for $i,j\ge 2$. The condition for resonance
4093: then becomes
4094: %
4095: \begin{align}
4096: i &= n\cdot j,
4097: \end{align}
4098: %
4099: which is exactly the relation we have observed in section \ref{MODECOUPLING}. \\
4100: From the relativistic point of view
4101: the non-linear coupling of eigenmodes in the weak and moderately
4102: non-linear regime is of particular interest in the discussion of
4103: unstable modes of rotating neutron stars. The underlying principle of
4104: these unstable oscillation modes is the increase in amplitude of the
4105: oscillation due to the emission of gravitational waves. The
4106: increased amplitude in turn gives rise to stronger gravitational radiation
4107: and so on. The
4108: conservation of energy is ensured in this case by the spin-down of the
4109: neutron star and the resulting decrease of rotational energy which sets
4110: a natural upper limit on this run-away effect.
4111: The physical mechanism which facilitates this remarkable instability
4112: is known as the
4113: CFS-instability (\citeNP{Chandrasekhar1970}, \citeNP{Friedman1978}).
4114: In order for a neutron star oscillation mode to be subject to the
4115: CFS-instability two conditions must be satisfied: (1) the mode must
4116: be retrograde with respect to the star but prograde with respect to a
4117: distant inertial observer and (2) the energy loss in the rotating frame
4118: due to dissipative effects must be smaller than the amount of energy
4119: gained from the gravitationally driven instability.
4120: The particular importance of
4121: the so-called $r$-modes in this respect arises from the fact that the
4122: dominating $l=m=2$ $r$-mode satisfies the first
4123: CFS-condition for arbitrarily small
4124: values of the angular frequency of the neutron star
4125: (\citeNP{Andersson1998}). One of the most important questions
4126: raised in connection
4127: with the $r$-modes concerns the efficiency with which energy is dissipated for
4128: example due to viscosity or non-linear effects. \\
4129: Considering the gradual increase in the oscillation amplitude, it is
4130: important to understand how the instability of the mode is affected
4131: in the weakly non-linear regime. To our knowledge the
4132: numerical studies presented
4133: in this work provide the first fully non-linear time evolutions
4134: of neutron star oscillations with high accuracy for amplitudes going
4135: all the way down to the weakly non-linear regime. Our results may
4136: therefore pave some of the way towards understanding non-linear effects
4137: in a wider class of neutron star oscillations.
4138: In particular we have managed to quantify the transfer of energy
4139: from low into higher eigenmodes. The picture that emerges from these
4140: evolutions is that only a rather small fraction of energy is shifted away
4141: from the low eigenmodes. In particular the results shown in
4142: Fig.\,\ref{MOD_EVOLA2A4} indicate that the energy shifted towards higher
4143: eigenmodes does not accumulate in time but is rather transferred back and
4144: forth between the initially present and the higher mode.
4145: Correspondingly we do not observe an efficient cascade of
4146: energy into higher modes.
4147: It is not clear,
4148: however, to what extent this picture will change if the energy residing
4149: in the higher order modes is gradually dissipated.
4150: In the context of $r$-modes it is expected that
4151: the energy in higher order modes is dissipated on a much
4152: shorter timescale than that of the dominating $l=m=2$ mode.
4153: The numerical techniques and the code developed in this work may
4154: facilitate a corresponding study in the framework of radial oscillations
4155: by introducing an artificial damping of higher order modes
4156: and an external force which drives the fundamental mode.
4157: One may then look for steady state situations arising from
4158: this model, where the amount of energy transfered to higher modes and thus
4159: dissipated equals that gained from the external driving mechanism.\\
4160: From a numerical point of view we emphasise the new perturbative
4161: approach which enabled us to obtain highly accurate fully non-linear evolutions
4162: over a large range of amplitudes. This technique can be applied for
4163: any physical problem where there exists a non-trivial static limit.
4164: The dynamic evolution can always be considered a finite perturbation
4165: of the static case and a corresponding perturbative formulation will
4166: provide a numerical accuracy that is determined by the amplitude of the
4167: perturbation rather than the static background. We expect this method to
4168: be particularly effective in higher dimensional evolutions where
4169: the grid resolution is rather limited by computational costs and
4170: the ensuing residual error arising from background terms in a non-perturbative
4171: formulation will be more significant.
4172:
4173:
4174: \newpage
4175: %=========================================================================
4176: \subsection{Radial oscillations in a Lagrangian formulation}
4177: \label{LAGR}
4178: %
4179:
4180: In the previous section we have seen that an Eulerian description of
4181: radial oscillations encounters difficulties at the stellar surface for several
4182: reasons. For certain equations of state the eigenmode profiles
4183: predicted by the linearized theory result in a diverging energy density
4184: perturbation. A purely numerical problem arises from the movement of the
4185: stellar surface with respect to the numerical grid. Highly sophisticated
4186: techniques may be required to adequately describe the surface of a neutron star
4187: in Eulerian coordinates
4188: and it is not clear to what extent these will lead to a fully satisfactory
4189: performance in the linear regime where the exact solution is
4190: known to high accuracy and facilitates a quantitative test for the code.
4191: It is interesting to see that these problems
4192: vanish immediately once the problem is described in a formalism where
4193: the coordinates follow the movement of the fluid elements. Even though
4194: it is not obvious how to generalise a Lagrangian approach to
4195: scenarios in two or three spatial dimensions, it still seems to
4196: be the natural choice for the 1-dimensional case. Lagrangian codes
4197: have often been based on the formulation of \lcite{May1966}
4198: and \citeyear{May1967}
4199: who following \lcite{Misner1964} use a vanishing shift vector and define
4200: the radial coordinate in terms of the interior rest mass.
4201: In order to facilitate a simple comparison with the Eulerian code
4202: discussed in section \ref{DYNAMIC}, however, it will be convenient
4203: for us to use as similar a gauge choice to the Eulerian case as possible.
4204: For this purpose we will follow \scite{Schinder1988} and use a Lagrangian
4205: gauge in combination with the polar slicing condition which is also
4206: implemented in the Eulerian code (cf. section \ref{NONP_EQ}).
4207: As a particularly useful consequence the singularity avoiding
4208: properties of this condition in combination with the Lagrangian gauge make
4209: this code highly suitable for studying spherically symmetric
4210: gravitational collapse. We will
4211: not exhaustively study this type of scenarios in this work, but will
4212: use the analytic solution by \lcite{Oppenheimer1939b} which
4213: describes the collapse of a homogeneous dust sphere
4214: for testing the code. \\
4215:
4216:
4217: %=========================================================================
4218: \subsubsection{The equations in the Lagrangian formulation}
4219: %
4220: %
4221: The derivation of the Lagrangian equations for a dynamic spherically symmetric
4222: neutron star was largely inspired by the work of \scite{Schinder1988}.
4223: We will, however, slightly deviate from their approach
4224: and work with a different set of variables and equations. \\
4225: We start by considering the line element of a spherically
4226: symmetric space time in polar slicing and Lagrangian gauge.
4227: As a result of the polar slicing condition, we are able to choose the same
4228: time coordinate $t$ as in the Eulerian case. The radial coordinate $x$ will
4229: label the fluid elements and generally differ from the
4230: areal radius $r$ which is intrinsically not comoving with the matter.
4231: Finally we choose standard angular coordinates
4232: $\theta$ and $\phi$ as above. Below we will see that the polar slicing
4233: condition implies a non-vanishing shift vector so that the line
4234: element
4235: % in polar slicing and Lagrangian gauge
4236: becomes
4237: %
4238: \begin{align}
4239: ds^2 &= \left(-\hat{\lambda}^2+\frac{\hat{\Gamma}\beta^2}{\hat{r}_{,x}^2}
4240: \right) dt^2
4241: +2\beta dt\, dx + \frac{\hat{r}_{,x}^2}{\hat{\Gamma}} dx^2
4242: + \hat{r}^2 (d\theta^2 +\sin^2{\theta} d\phi^2).
4243: \label{LAGR_LINEELEMENT}
4244: \end{align}
4245: %
4246: It turns out to be convenient for our discussion if we introduce the variables
4247: %
4248: \begin{align}
4249: w &= \frac{\hat{r}_{,t}}{\hat{\lambda}}, \label{LAGR_U} \\[10pt]
4250: \Omega^2 &= \hat{\Gamma} - w^2, \label{LAGR_OMEGA2} \\[10pt]
4251: \hat{m} &= \frac{\hat{r}}{2}(1-\hat{\Gamma}), \label{LAGR_M}
4252: \end{align}
4253: %
4254: where the velocity is identical to that used in the Eulerian case.
4255: As before we use the ``hat'' to distinguish
4256: between the time dependent variables and their counterparts in the static
4257: case. We note that we need to distinguish between the time dependent
4258: areal radius $\hat{r}$ and the static value $r$, since the areal
4259: radius corresponding to the position of a fluid element
4260: is a dependent variable and will generally vary with time. In
4261: the Eulerian case the areal radius was a coordinate and therefore
4262: intrinsically independent of time.
4263: If we compare the Lagrangian line element
4264: (\ref{LAGR_LINEELEMENT}) with the Eulerian one given by
4265: Eq.\,(\ref{NONP_LINEELEMENT}) we therefore have to use the time dependent
4266: $\hat{r}$ in the latter line element instead of $r$. The coordinate
4267: transformation relating the two line elements is described by
4268: %
4269: \begin{align}
4270: \hat{r} &= \hat{r}(t,x).
4271: \end{align}
4272: %
4273: The transformation law for the metric components corresponding to the
4274: transformation from coordinates $\hbox{\vec $x$}^{\mu}=(t,x,\theta, \phi)$ to
4275: $\hbox{\vec $x$}^{'\mu}=(t,r,\theta,\phi)$ is given by
4276: %
4277: \begin{align}
4278: \hbox{\vec g}'_{\mu\nu} &= \frac{\partial \hbox{\vec $x$}^{\rho}}
4279: {\partial \hbox{\vec $x$}^{'\mu}}
4280: \frac{\partial \hbox{\vec $x$}^{\sigma}}
4281: {\partial \hbox{\vec $x$}^{'\nu}} \hbox{\vec g}_{\rho \sigma},
4282: \end{align}
4283: %
4284: and leads to the two non-trivial equations
4285: %
4286: \begin{align}
4287: \beta &= \hat{r}_{,x} \hat{r}_{,t} \hat{\mu}^2, \\[10pt]
4288: \hat{\Gamma}&= \frac{1}{\hat{\mu}^2}. \label{LAGR_GAMMAOFMU}
4289: \end{align}
4290: %
4291: As a consequence the shift vector $\beta$ is related to the
4292: components of the Lagrangian metric by
4293: %
4294: \begin{align}
4295: \beta = \hat{\lambda} \frac{w\,\hat{r}_{,x}}{\hat{\Gamma}}. \label{LAGR_BETA}
4296: \end{align}
4297: %
4298: In terms of the extrinsic curvature defined in Eq.\,(\ref{BACKGR_EXTRCURV})
4299: this relation can be
4300: written as $\hbox{\vec{K}}^{\theta}_{\,\,\,\theta}=
4301: \hbox{\vec{K}}^{\phi}_{\,\,\,\phi}=0$ and we
4302: have recovered the polar slicing condition. The non-vanishing shift vector
4303: (\ref{LAGR_BETA}) is the price we have to pay for keeping the
4304: polar slicing condition in the Lagrangian gauge. \\
4305: As far as the matter is concerned, we use again a single component
4306: perfect fluid
4307: and thus the energy momentum tensor given by Eq.\,(\ref{NONP_EMTENSOR}).
4308: Since the fluid elements do not move with respect to the radial coordinate
4309: $x$, the 4-velocity has zero spatial components and is determined by the
4310: normalisation $\hbox{\vec u}^{\mu} \hbox{\vec u}_{\mu}=-1$
4311: %
4312: \begin{align}
4313: \hbox{\vec u}^{\mu} &= \left( \frac{\sqrt{\hat{\Gamma}}}{\hat{\lambda}
4314: \Omega}, 0,0,0 \right).
4315: \end{align}
4316: %
4317: The resulting field equations
4318: $\hbox{\vec{G}}_{\mu \nu} = 8\pi \hbox{\vec{T}}_{\mu \nu}$ can
4319: be written as
4320: %
4321: \begin{align}
4322: \frac{\hat{\lambda}_{,x}}{\hat{\lambda}} &= \frac{\hat{r}_{,x}}
4323: {\hat{r} \hat{\Gamma}}
4324: \left( \frac{\hat{m}}{\hat{r}} + 4\pi \hat{r}^2 \frac{w^2 \hat{\rho}
4325: + \hat{\Gamma} \hat{P}} {\Omega^2} \right), \label{LAGR_LAMBDAX} \\[10pt]
4326: \hat{m}_{,x} &= 4\pi \hat{r}^2 \hat{r}_{,x} \frac{\hat{\Gamma} \hat{\rho}
4327: +w^2 \hat{P}}{\Omega^2} \label{LAGR_MX}, \\[10pt]
4328: \hat{m}_{,t} &= -4\pi \hat{r}^2 \hat{\lambda} w \hat{P} \label{LAGR_MT}.
4329: \end{align}
4330: %
4331: Similarly the conservation of energy and
4332: momentum $\nabla_{\mu} \hbox{\vec{T}}^{\mu}_{\,\,\,\,{\nu}}=0$
4333: leads to two evolution equations for the matter variables
4334: %
4335: \begin{align}
4336: &\hat{\rho}_{,t} +(\hat{\rho}+\hat{P}) \frac{\Omega}{\hat{r}^2\hat{r}_{,x}}
4337: \left( \frac{\hat{r}^2 \hat{r}_{,x}}{\Omega} \right)_{,t}=0,
4338: \label{LAGR_RHOT} \\[10pt]
4339: &\Omega^4 \frac{\hat{P}_{,x}}{\hat{r}_{,x}} + \hat{P}_{,t}
4340: \frac{w}{\hat{\lambda}} \Omega^2
4341: + (\hat{\rho} + \hat{P})\left[ \hat{\gamma}\frac{w_{,t}}{\hat{\lambda}}
4342: +(\hat{\gamma}-2w^2) \left( \frac{\hat{m}}{\hat{r}^2} + 4\pi \hat{r}\hat{P}
4343: \right) \right] = 0, \label{LAGR_UT}
4344: \end{align}
4345: %
4346: and the system is closed by the polytropic equation
4347: of state (\ref{DYNAMICPOLYTROPE}).
4348: It is worth pointing out that the appearance of the time derivative in the
4349: field equation (\ref{LAGR_MT})
4350: does not contradict the absence of gravitational degrees of
4351: freedom in spherical symmetry. This equation can be shown to be a consequence
4352: of the constraints (\ref{LAGR_LAMBDAX}), (\ref{LAGR_MX}) and
4353: the matter equations (\ref{LAGR_RHOT}), (\ref{LAGR_UT}).
4354: In this sense the degrees of freedom still reside in
4355: the matter variables and the metric is determined at each time irrespective
4356: of its history. In practice, however,
4357: we will use the rather simple equation (\ref{LAGR_MT}) to evolve the variable
4358: $\hat{m}$ instead of evolving $\hat{\rho}$ via the matter equation
4359: (\ref{LAGR_RHOT}). \\
4360: If we consider the static limit of the system of equations
4361: (\ref{LAGR_LAMBDAX})-(\ref{LAGR_UT}) we expect to recover the
4362: Tolman-Oppenheimer-Volkoff equations (\ref{TOV_RY})-(\ref{TOV_PY}).
4363: That this is indeed the case can be seen if we set all time derivatives
4364: including the velocity $w$ to zero and assume that $x$ is identical to the
4365: coordinate $x$ we used in the static case.
4366: The second condition can always be satisfied since the fluid elements are
4367: not moving and can be labelled by the areal radius of their position or
4368: the rescaled coordinate $y$ defined in Eq.\,(\ref{TOV_YOFR}).
4369: Then Eq.\,(\ref{LAGR_MX}) directly reduces to Eq.\,(\ref{TOV_MR}) or the
4370: transformed version thereof expressed in terms of the coordinate $y$.
4371: From Eq.\,(\ref{LAGR_GAMMAOFMU}) we conclude that $\Gamma=1/\mu^2$ and the
4372: constraint (\ref{LAGR_LAMBDAX}) becomes identical to (\ref{TOV_LAMBDAY}).
4373: Finally the matter equation (\ref{LAGR_UT}) reduces to Eq.\,(\ref{TOV_PY}) and
4374: the evolution equations (\ref{LAGR_MT})
4375: and (\ref{LAGR_RHOT}) vanish identically.
4376:
4377:
4378: %=========================================================================
4379: \subsubsection{The linearized evolution equations}
4380: %
4381: %
4382: We have seen that the static limit of the evolution equations
4383: (\ref{LAGR_LAMBDAX})-(\ref{LAGR_RHOT}) is given by the TOV equations.
4384: We can therefore linearise the dynamic equations around this background and
4385: compare the results with the Eulerian case described in section \ref{PERT_LIN}.
4386: In order to distinguish between Eulerian and Lagrangian perturbations
4387: we will use a capital $\Delta$ in the Lagrangian case.
4388: The only exception is the radial displacement which is identical
4389: in both formulations so that we keep the variable name $\xi$. \\
4390: We start the linearisation with the definition of the radial velocity
4391: $w$ (\ref{LAGR_U}). In terms of the radial displacement this equation
4392: becomes
4393: %
4394: \begin{align}
4395: w &= \frac{\xi_{,t}}{\lambda}. \label{LLIN_U}
4396: \end{align}
4397: %
4398: We note that the background value of the lapse $\lambda$ appears
4399: in the denominator instead of the time dependent $\hat{\lambda}$.
4400: In the same way we will neglect higher order terms in the other equations.
4401: If we substitute this expression for $w$ in the evolution
4402: equation (\ref{LAGR_MT})
4403: for $m$ and integrate over time, we obtain
4404: %
4405: \begin{align}
4406: \Delta m &= -4\pi r^2 P \xi. \label{LLIN_DM}
4407: \end{align}
4408: %
4409: The constant of integration vanishes because a zero displacement $\xi$
4410: of the fluid elements implies $\Delta m=0$.
4411: We can use this expression for $\Delta m$ in the definition (\ref{LAGR_M})
4412: to obtain the result for the auxiliary variable $\hat{\Gamma}$
4413: %
4414: \begin{align}
4415: \Delta \Gamma &= 8\pi r P \xi + \frac{\xi}{r} (1-\Gamma). \label{LLIN_DGAMMA}
4416: \end{align}
4417: %
4418: The energy density perturbation then follows from substituting
4419: Eqs.\,(\ref{LLIN_U})-(\ref{LLIN_DGAMMA}) in the evolution equation
4420: (\ref{LAGR_RHOT})
4421: and integrating over time. With the constant of integration vanishing as
4422: before the result is
4423: %
4424: \begin{align}
4425: \Delta \rho &= \frac{(\rho +P)}{r_{,x}}
4426: \left( \frac{\xi r^2}{\lambda} \right)_{,x}.
4427: \label{LLIN_DRHO}
4428: \end{align}
4429: %
4430: From the definition of the speed of sound we can calculate the pressure
4431: perturbation
4432: %
4433: \begin{align}
4434: \Delta P &= C^2 \Delta \rho. \label{LLIN_DP}
4435: \end{align}
4436: %
4437: If we substitute the results (\ref{LLIN_U})-(\ref{LLIN_DP}) in the
4438: evolution equation (\ref{LAGR_UT}) we get exactly the second order
4439: differential equation (\ref{LIN_ZETATT}) of the Eulerian case with
4440: the coefficient functions (\ref{LIN_PI})-(\ref{LIN_Q}). No substitution
4441: for $\Delta \lambda$ is necessary here, because
4442: all terms containing $\Delta \lambda$ drop out by virtue of the
4443: TOV background equations. Writing the displacement as a product
4444: $\xi(x) f(t)$ we obtain again harmonic time dependence
4445: and finally arrive at the ordinary differential
4446: equation (\ref{LIN_ZETARR}) so that we can use the whole machinery
4447: developed in section \ref{PERT_LIN} to calculate the eigenmodes.
4448: It is interesting,
4449: however, to contrast Eq.\,(\ref{LLIN_DRHO}) for the Lagrangian $\Delta \rho$
4450: with the Eulerian analogue Eq.\,(\ref{LIN_DRHOOFZETA}). We have seen in section
4451: \ref{PERT_LIN} that the extra term in the Eulerian relation leads to the
4452: problematic asymptotic behaviour of $\delta \rho$ at the surface. No
4453: such problem occurs in the Lagrangian case which thus provides a
4454: self-consistent way of deriving the linearized equations.
4455:
4456:
4457: %=========================================================================
4458: \subsubsection{The equations for the numerical implementation}
4459: %
4460: %
4461: The Lagrangian evolution of a dynamic neutron star in spherical
4462: symmetry is described by the system of equations (\ref{LAGR_U}),
4463: (\ref{LAGR_LAMBDAX})-(\ref{LAGR_MT}), (\ref{LAGR_UT}), where the auxiliary
4464: variables $\hat{\Gamma}$ and $\Omega$ are defined by Eqs.\,(\ref{LAGR_OMEGA2})
4465: and (\ref{LAGR_M}). This choice of variables and equations, however, did
4466: not lead to an entirely satisfactory performance of the code. This
4467: became most obvious in the simulation of the Oppenheimer-Snyder dust
4468: collapse where the energy density
4469: showed an increasing deviation from the analytic solution near the centre
4470: of the star. When the dust sphere had collapsed close to its
4471: Schwarzschild radius,
4472: the deviation was larger than $10\,\%$. In order to understand this
4473: inaccuracy, we consider Eq.\,(\ref{LAGR_MX}) which relates
4474: the energy density to the mass. If we solve this equation for
4475: $\hat{\rho}$ we see that the mass appears in the form
4476: $\hat{m}_{,x}/\hat{r}^2$, which will be of the order $\mathscr{O}(1)$
4477: near the origin. The second order accuracy of the finite differencing
4478: scheme we have used, however, implies that the variable
4479: $\hat{m}$ is known with a local error
4480: $\mathscr{O}(\Delta x^3)$ only and consequently
4481: the numerical derivative $\hat{m}_{,x}$ has an error
4482: $\mathscr{O}(\Delta x^2)$. Near the origin
4483: the radius $\hat{r}$ is of the same order of magnitude as $\Delta x$ and
4484: the error of $\hat{m}_{,x}/\hat{r}^2$ and thus the energy density $\hat{\rho}$
4485: is large. This problem is a consequence of the $\hat{r}^3$ behaviour of the
4486: mass $\hat{m}$ near the origin combined with the strong variation
4487: of the variables in the dust collapse and persists in a perturbative
4488: formulation. In the numerical evolution we therefore use the
4489: variable
4490: %
4491: \begin{align}
4492: \hat{N} &= \frac{\hat{m}}{\hat{r}^2}, \label{NLAGR_NOFM}
4493: \end{align}
4494: %
4495: instead of the mass $\hat{m}$.
4496: The Lagrangian equations
4497: (\ref{LAGR_U}), (\ref{LAGR_LAMBDAX})-(\ref{LAGR_MT}), (\ref{LAGR_UT})
4498: then become
4499: %
4500: \begin{align}
4501: & \hat{\Gamma} \Omega^2 \hat{\lambda}_{,x} - \hat{r}_{,x} \hat{\lambda} \left[
4502: \Omega^2 \hat{N} + 4\pi \hat{r} (w^2 \hat{\rho} + \hat{\Gamma} \hat{P})
4503: \right] = 0, \label{NLAGR_LAMBDAX} \\[10pt]
4504: & \hat{r}\Omega^2 \hat{N}_{,x} + 2\Omega^2 \hat{r}_{,x} \hat{N}
4505: -4\pi \hat{r} \hat{r}_{,x}
4506: (\hat{\Gamma} \hat{\rho} + w^2 \hat{P}) = 0, \label{NLAGR_NX} \\[10pt]
4507: & \hat{r}\hat{N}_{,t} + 2\hat{\lambda} w \left( \hat{N} + 2\pi \hat{r}
4508: \hat{P} \right) =0, \\[10pt]
4509: & \hat{r}_{,t} - \hat{\lambda} w = 0, \\[10pt]
4510: & \hat{\lambda} \Omega^4 \hat{P}_{,x} + \hat{r}_{,x} w \Omega^2 \hat{P}_{,t}
4511: + \hat{r}_{,x} (\hat{\rho}+\hat{P}) \left[ \hat{\Gamma} w_{,t}
4512: + \hat{\lambda} (\hat{\Gamma}-2w^2)(\hat{N} + 4\pi \hat{r} \hat{P}) \right],
4513: \label{NLAGR_UT}
4514: \end{align}
4515: %
4516: where $\hat{\Gamma}$ is now defined by
4517: %
4518: \begin{align}
4519: \hat{\Gamma} &= 1-2\hat{N}\hat{r}.
4520: \end{align}
4521: %
4522: In the static limit these equations reduce to the TOV equations
4523: %
4524: \begin{align}
4525: & \Gamma \lambda_{,x} - r_{,x}\lambda (N + 4\pi r P ), \\[10pt]
4526: & r N_{,x} + r_{,x} \left( 2 N - 4\pi r \rho \right) = 0, \\[10pt]
4527: & \Gamma P_{,x} + r_{,x} (\rho+P) (N+4\pi r P)=0, \\[10pt]
4528: & \Gamma = 1-2Nr.\end{align}
4529: %
4530: In order to derive a fully non-linear perturbative formulation, we decompose
4531: the time dependent quantities into static background contributions
4532: and time dependent perturbations
4533: %
4534: \begin{align}
4535: \hat{r}(t,x) &= r(x) + \xi(t,x), \\[10pt]
4536: \hat{\lambda}(t,x) &= \lambda(x) + \Delta \lambda(t,x), \\[10pt]
4537: \hat{N}(t,x) &= N(x) + \Delta N (t,x), \\[10pt]
4538: \hat{\Gamma}(t,x) &= \Gamma(x) + \Delta \Gamma(t,x), \\[10pt]
4539: \hat{\rho}(t,x) &= \rho(x) + \Delta \rho(t,x).
4540: \end{align}
4541: %
4542: With these definitions the fully non-linear perturbative version
4543: of Eqs.\,(\ref{NLAGR_LAMBDAX})-(\ref{NLAGR_UT}) becomes
4544: %
4545: \begin{align}
4546: \begin{split}
4547: & \hat{\Gamma}^2 \Delta \lambda_{,x} + \Delta \Gamma (2\Gamma
4548: + \Delta \Gamma) \lambda_{,x} - (\xi_{,x} \lambda \Gamma
4549: + \hat{r}_{,x} \Delta \lambda \Gamma + \hat{r}_{,x}
4550: \hat{\lambda}\Delta \Gamma ) (N+4\pi r P)\\[10pt]
4551: & + w^2 \left[ -\hat{\Gamma} \hat{\lambda}_{,x} + \hat{r}_{,x}\hat{\lambda}
4552: (\hat{N}-4\pi \hat{r} \hat{\rho}) \right] - \hat{r}_{,x} \hat{\lambda}
4553: \hat{\Gamma} \left[ \Delta N + 4\pi (\xi P + \hat{r} \Delta P)
4554: \right] = 0,
4555: \end{split} \label{NLAGR_DLAMBDAX} \\[20pt]
4556: \begin{split}
4557: & -w^2(\hat{r}\hat{N}_{,x} +2\hat{r}_{,x} \hat{N} + 4\pi \hat{r}
4558: \hat{r}_{,x} \hat{P}) + \Delta \Gamma (\hat{r}\hat{N}_{,x}
4559: + 2\hat{r}_{,x} \hat{N} -4\pi \hat{r}\hat{r}_{,x} \hat{\rho} )
4560: \\[10pt]
4561: & + \Gamma \left[ \xi N_{,x} + \hat{r} \Delta N_{,x} + 2\xi_{,x} N
4562: + 2\hat{r}_{,x} \Delta N - 4\pi( \xi \rho r_{,x} + \hat{r}\xi_{,x}
4563: \rho + \hat{r} \hat{r}_{,x} \Delta \rho ) \right] = 0,
4564: \end{split} \label{NLAGR_DNX} \\[20pt]
4565: & \hat{r}\hat{N}_{,t} + 2\hat{\lambda} w (\hat{N} + 2\pi \hat{r} \hat{P}) = 0,
4566: \label{NLAGR_DNT} \\[20pt]
4567: & \xi_{,t} - \hat{\lambda} w = 0, \label{NLAGR_XIT} \\[20pt]
4568: \begin{split}
4569: & \hat{\lambda}(-2\hat{\Gamma} w^2 + w^4) \hat{P}_{,x} + \hat{r}_{,x}
4570: w \Omega^2 \hat{P}_{,t} + (\hat{\rho}+\hat{P}) \hat{r}_{,x} \left[
4571: \hat{\Gamma} w_{,t} -2\hat{\lambda}w^2(\hat{N} + 4\pi \hat{r} \hat{P})
4572: \right] \\[10pt]
4573: & + (\Delta \lambda \Gamma + \hat{\lambda} \Delta \Gamma)
4574: \left[ \hat{\Gamma} \hat{P}_{,x} + (\hat{\rho} + \hat{P}) \hat{r}_{,x}
4575: (\hat{N} + 4\pi \hat{r} \hat{P}) \right] + \lambda \Gamma \left\{
4576: \Delta \Gamma P_{,x} + \hat{\Gamma} \Delta P_{,x} \right. \\[10pt]
4577: & \left. + \left[ (\Delta \rho
4578: + \Delta P)r_{,x} + (\hat{\rho} + \hat{P}) \xi_{,x} \right] (N+4\pi r P)
4579: + (\hat{\rho} + \hat{P}) \hat{r}_{,x} (\Delta N + 4\pi \xi P
4580: + 4\pi \hat{r}\Delta P) \right\} = 0.
4581: \end{split} \label{NLAGR_DPT}
4582: \end{align}
4583: %
4584: This is the final system of equations used in the numerical implementation.
4585:
4586:
4587: %=========================================================================
4588: \subsubsection{Initial data and boundary conditions}
4589: %
4590: %
4591: In order to numerically evolve the system of partial differential
4592: equations (\ref{NLAGR_DLAMBDAX})-(\ref{NLAGR_DPT}) we have to specify
4593: initial data and boundary conditions. We will start with the initial data. \\
4594: In the Eulerian case we have determined the physical setup by providing
4595: initial data for the matter variables $\hat{\rho}$ and $w$. This
4596: gave us energy density and velocity at each radial position $\hat{r}$.
4597: In order to provide the same information in the Lagrangian case
4598: it is not sufficient to give initial data in the form of
4599: $\hat{\rho}(x)$ and $w(x)$ because the meaning of the spatial
4600: coordinate $x$ is not determined at this stage. Indeed it can easily be
4601: seen that the system of equations
4602: (\ref{NLAGR_LAMBDAX})-(\ref{NLAGR_UT}) is invariant under
4603: any transformation $x\rightarrow \bar{x}(x)$ which corresponds to a relabelling
4604: of the fluid elements. Consequently we also need to establish a relation
4605: between the Lagrangian coordinate $x$ and the areal radius $\hat{r}$ on
4606: the initial slice. The initial data for $\hat{r}(x)$ serve this purpose.
4607: Alternatively this additional requirement becomes obvious if we
4608: consider the structure of the system (\ref{NLAGR_LAMBDAX})-(\ref{NLAGR_UT}).
4609: These equations contain the time derivatives of $\hat{r}$, $w$,
4610: $\hat{N}$ and $\hat{P}$. In addition to the lapse function $\hat{\lambda}$
4611: only one of these quantities is determined by the constraint equations
4612: (\ref{NLAGR_LAMBDAX}), (\ref{NLAGR_NX}). The remaining three variables follow
4613: from the time evolution and thus require the specification of initial
4614: data. In the perturbative formulation the background functions
4615: $r(x)$, $\rho(x)$, $N(x)$ and $\lambda(x)$ follow from the
4616: solution of the TOV equations and we prescribe initial data for the
4617: perturbations $\xi$, $w$ and $\Delta \rho$. The values of $\Delta N$
4618: and $\Delta \lambda$ are then calculated from the constraint equations
4619: (\ref{NLAGR_DLAMBDAX}) and (\ref{NLAGR_DNX}). For this purpose we use
4620: an implicit second order scheme based on the finite differencing
4621: given for these equations in appendix \ref{LAGR_FDE}. \\
4622: The specification of boundary conditions, in particular at the
4623: stellar surface, turned out to be the most problematic part in the
4624: Eulerian formulation of the dynamic star. In contrast the boundary
4625: conditions are well defined in the Lagrangian case. At the centre
4626: we demand
4627: %
4628: \begin{align}
4629: \xi &= 0, \label{NLAGR_BCINXI} \\[10pt]
4630: w &= 0, \\[10pt]
4631: \Delta N &= 0.
4632: \end{align}
4633: %
4634: The first two conditions guarantee that the centre of the star does not
4635: move which immediately follows from the spherical symmetry and the third
4636: condition
4637: avoids the appearance of a conical singularity. At the surface we require
4638: %
4639: \begin{align}
4640: \Delta \rho &= 0, \label{NLAGR_BCOUTDRHO} \\[10pt]
4641: \hat{\lambda}^2 &= 1-2\hat{N}\hat{r}, \label{NLAGR_BCOUTLAMBDA}
4642: \end{align}
4643: %
4644: which follows from the definition of the surface and the matching
4645: to an exterior Schwarzschild metric.
4646: % For the numerical evolution
4647: %we will use an implicit second order in space and time finite
4648: %differencing scheme.
4649: If $K$ is the number of grid points used, the
4650: finite differencing of the evolution equations
4651: (\ref{NLAGR_DLAMBDAX})-(\ref{NLAGR_DPT}) results in $5K-5$ algebraic
4652: relations between the $5K$ function values. The boundary
4653: conditions (\ref{NLAGR_BCINXI})-(\ref{NLAGR_BCOUTLAMBDA}) provide
4654: the remaining 5 relations to determine the evolution and no additional
4655: treatment of boundary values is required.
4656:
4657:
4658: %=========================================================================
4659: \subsubsection{The finite differencing of the equations}
4660: %
4661: %
4662: We numerically solve the system of partial differential equations
4663: (\ref{NLAGR_DLAMBDAX})-(\ref{NLAGR_DPT}) by using an implicit
4664: second order in space and time finite differencing scheme. The
4665: particular choice of stencils has been guided by the presence of
4666: derivatives in the individual differential equations. This is illustrated
4667: in Fig.\,\ref{LAGR_STENCILS} where the grid points $k$ and $k+1$
4668: are shown for the time levels
4669: %
4670: \begin{figure}[t]
4671: \centering
4672: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4673: \begin{picture}(0,0)%
4674: \epsfig{file=lagrgrid.pstex}%
4675: \end{picture}%
4676: \setlength{\unitlength}{3947sp}%
4677: %
4678: \begingroup\makeatletter\ifx\SetFigFont\undefined%
4679: \gdef\SetFigFont#1#2#3#4#5{%
4680: \reset@font\fontsize{#1}{#2pt}%
4681: \fontfamily{#3}\fontseries{#4}\fontshape{#5}%
4682: \selectfont}%
4683: \fi\endgroup%
4684: \begin{picture}(5717,1620)(301,-811)
4685: \put(1170,299){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k$}}}
4686: \put(2861,-800){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k$}}}
4687: \put(5230,-800){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k$}}}
4688: \put(2566,-589){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$n$}}}
4689: \put(4934,-589){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$n$}}}
4690: \put(831,-800){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k$}}}
4691: \put(536,-589){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$n$}}}
4692: \put(3200,299){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k$}}}
4693: \put(5568,299){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k$}}}
4694: \put(3376,-811){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k+1$}}}
4695: \put(5776,-811){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k+1$}}}
4696: \put(1351,-811){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k+1$}}}
4697: \put(3751,314){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$k+1$}}}
4698: \put(301, 89){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$n+1$}}}
4699: \put(2326, 89){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$n+1$}}}
4700: \put(4651, 89){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\itdefault}$n+1$}}}
4701: \put(301,689){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}Eqs.\,(\ref{NLAGR_DLAMBDAX}), (\ref{NLAGR_DNX})}}}
4702: \put(2551,689){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}Eqs.\,(\ref{NLAGR_DNT}), (\ref{NLAGR_XIT})}}}
4703: \put(5101,689){\makebox(0,0)[lb]{\smash{\SetFigFont{11}{13.2}{\rmdefault}{\mddefault}{\updefault}Eq.\,(\ref{NLAGR_DPT})}}}
4704: \end{picture}
4705: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4706: \caption{The stencils used for the finite differencing of
4707: Eqs.\,(\ref{NLAGR_DLAMBDAX})-(\ref{NLAGR_DPT}).}
4708: \label{LAGR_STENCILS}
4709: \end{figure}
4710: %
4711: $n$ and $n+1$. The filled circles indicate grid points that have been used
4712: for the finite differencing, the crosses those points which have not
4713: been used. The constraint equations (\ref{NLAGR_DLAMBDAX}) and
4714: (\ref{NLAGR_DNX}) contain spatial derivatives only. It is therefore
4715: suitable to use two neighbouring grid points on the new time slice $n+1$.
4716: In contrast Eqs.\,(\ref{NLAGR_DNT}) and (\ref{NLAGR_XIT}) contain time
4717: derivatives only and we use two grid points at spatial position $k+1$
4718: on neighbouring time slices for the finite differencing. Both
4719: kinds of derivatives are present in Eq.\,(\ref{NLAGR_DPT}) and we need to use
4720: all four grid points as a consequence. Fig.\,{\ref{LAGR_STENCILS}
4721: also illustrates an
4722: extra option that has been included in the finite differencing.
4723: In the case of the Oppenheimer-Snyder dust collapse it turns out
4724: to be necessary to interpret the values of the energy density
4725: $\rho^n_k$, $\Delta \rho^n_k$ as cell averages and correspondingly
4726: use a staggered grid for these variables. This is indicated by the
4727: empty circles in Fig.\,\ref{LAGR_STENCILS}.
4728: In the finite differencing equations we will
4729: therefore introduce a parameter $\sigma$ which allows us to switch between
4730: a staggered and the ``normal'' grid for $\rho$ and $\Delta \rho$.
4731: The staggering, however, is only needed for the dust collapse and will not
4732: be used when we simulate neutron stars. \\
4733: The resulting finite difference equations are listed in appendix
4734: \ref{LAGR_FDE}
4735: together with the additional relations we use to calculate auxiliary
4736: functions and derivatives of the background variables. The parameter
4737: $\sigma$ will be zero in all cases except for the simulation
4738: of the Oppenheimer-Snyder dust collapse, where we will use the staggered grid
4739: for the energy density and set $\sigma=1$. Before we turn our attention
4740: towards solving this system of algebraic equations, we need to comment on some
4741: of its properties.
4742: %
4743: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
4744: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
4745: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
4746: \item If we use the staggered grid to calculate the energy density,
4747: the outer boundary condition (\ref{FDE_BCOUTDRHO}) is only a
4748: formal condition because $\Delta \rho_K$ decouples from the
4749: remaining $5K-1$ variables. In the analysis of the dust collapse
4750: we will therefore use the interior values
4751: $\Delta \rho_k$ for $k=1,\ldots ,K-1$ only.
4752: \item We also note that the finite difference expression (\ref{FDE_EQ5DRHOX})
4753: for $\Delta \rho_{,x}$ is only a first order accurate approximation
4754: if the staggered grid is used for the energy density. This does
4755: not affect the accuracy of the numerical scheme, however, since
4756: this derivative appears in the form of the pressure gradient
4757: $\Delta P_{,x}$ only in Eq.\,(\ref{NLAGR_DPT}). The only scenario where
4758: we use the staggering is the dust collapse, where
4759: the pressure and thus its gradient vanish identically.
4760: \item Finally we note that the finite differencing scheme used here
4761: slightly differs from that used for the evolution of
4762: cosmic strings in section \ref{dynstring}. The scheme used here
4763: was partly inspired by the work of \scite{Schinder1988}
4764: and partly resulted
4765: from attempts to eliminate numerical noise that we encountered
4766: during the development of the code. It turned out, however, that
4767: this noise originated from the numerical inaccuracy associated with
4768: the $\hat{r}^3$ behaviour of the variable $\hat{m}$ we discussed above.
4769: We have no reason therefore to question the applicability
4770: of the Crank-Nicholson scheme described in section \ref{FDE_CN}.
4771: \end{list}
4772: %
4773: In order to solve the system of $5K$ non-linear algebraic relations we
4774: use the Newton-Raphson method described in section \ref{relaxation}.
4775: The initial
4776: guess is given by the values on the previous time slice and convergence
4777: is typically achieved after three iterations.
4778:
4779:
4780: %=========================================================================
4781: \subsubsection{Testing the code}
4782: %
4783: %
4784: In order to check the performance of the code we subject it to
4785: three independent tests. As in the Eulerian case, we will compare
4786: the numerical results with the
4787: approximative analytic solution obtained from the linearized equations of a
4788: dynamic spherically symmetric neutron star. Secondly we will test the
4789: convergence properties of the code in the non-linear regime. Finally
4790: we calculate the deviation of the numerical results from the analytic
4791: solution by \lcite{Oppenheimer1939b} which describes the collapse of
4792: a homogeneous dust sphere. \\
4793: We start by testing the performance of the code in the linear regime.
4794: In the Eulerian analysis we have seen that the eigenmodes for
4795: stellar models with polytropic indices $\gamma > 2$
4796: lead to a diverging energy density perturbation at
4797: the surface and thus could not be used for a time evolution. We have seen,
4798: however, that this divergence results from a coordinate singularity
4799: at the stellar surface and the Lagrangian energy density perturbation is
4800: well behaved for any polytropic index. It is tempting therefore to
4801: use a stellar model with a large polytropic index to test the performance
4802: of the Lagrangian code in the linear regime. We choose a model with
4803: polytropic exponent $\gamma=3.0$, polytropic factor $K=2\cdot10^5\,\,
4804: {\rm km}^{-2}$ and central density $\rho_{\rm c}= 2.2\cdot
4805: 10^{15}\,\,{\rm g/cm}^3$. This is the third model of Table \ref{COMPEIGS}
4806: where we compared our results of the eigenmode frequencies with
4807: those of \lcite{Kokkotas2001}. \\
4808: In general we have achieved better performance with
4809: the Lagrangian code if the outer boundary condition $\rho=0$ is satisfied
4810: %
4811: \begin{figure}[t]
4812: \centering
4813: \epsfig{file=llincomp_Drho.ps, height=400pt, width=175pt, angle=-90}
4814: \epsfig{file=llincomp_xi.ps, height=400pt, width=175pt, angle=-90}
4815: \epsfig{file=llincomp_u.ps, height=400pt, width=175pt, angle=-90}
4816: \caption{The left panels show the time evolution of $\Delta \rho$,
4817: $\xi$ and $w$ obtained for 1600 grid points. The initial
4818: perturbation is given as the fundamental mode in the
4819: displacement vector $\xi$. The right panels show the
4820: deviation from the exact solution of the linearized equations.}
4821: \label{LAGR_LINTEST}
4822: \end{figure}
4823: %
4824: exactly. In the remainder of the Lagrangian discussion we will therefore
4825: use the relaxation method described in section \ref{TOV_NUM}
4826: to calculate the TOV background. Unless specified otherwise we will
4827: use the rescaled coordinate $y$ for this calculation and the time evolution
4828: and thus set $r_{,x}=C$.\\
4829: The next step
4830: consists in calculating the eigenmode profiles for the variables
4831: $\xi$, $w$ and $\Delta \rho$. These results enable us to specify initial
4832: data and calculate the analytic solutions. In this case the initial
4833: perturbation of the star consists in a displacement $\xi$ of the fluid
4834: elements corresponding to the fundamental mode with a surface amplitude
4835: of about $5\,\,{\rm cm}$. The initial velocity is set to zero and
4836: the energy density corresponding to this eigenmode follows from
4837: Eq.\,(\ref{LLIN_DRHO}). The remaining initial variables are calculated
4838: from the constraint equations (\ref{NLAGR_DLAMBDAX}), (\ref{NLAGR_DNX}).
4839: The resulting
4840: data on the initial slice are then evolved in time according to the method
4841: described in the previous section. The analytic solution for the
4842: fundamental variables $\xi$, $w$, $\Delta \rho$ is given by
4843: %
4844: \begin{align}
4845: \xi(t,x) &= \xi_1(x) \cos{\omega t}, \\[10pt]
4846: w(t,x) &= -w_1(x) \sin{\omega t}, \\[10pt]
4847: \Delta \rho &= \Delta \rho_1(x) \cos{\omega t},
4848: \end{align}
4849: %
4850: where $\omega$ is the frequency derived from the eigenmode calculation.
4851: In Fig.\,\ref{LAGR_LINTEST} we show the numerical
4852: results obtained for 1600 grid points together with their
4853: deviation from the harmonic solutions.
4854: These results show that the code reproduces the analytic solution
4855: %
4856: \begin{figure}[t]
4857: \centering
4858: \epsfig{file=lfour175.eps, height=200pt, width=150pt, angle=-90}
4859: \epsfig{file=lfour300.eps, height=200pt, width=150pt, angle=-90}
4860: \caption{Frequency spectra obtained for stellar models with polytropic
4861: indices $\gamma = 1.75$ (left) and 3.0 (right). The initial
4862: data consists of a displacement $\xi$ given by the sum of the
4863: first 10 eigenmodes. The vertical bars indicate the frequencies
4864: predicted by the eigenmode calculations.}
4865: \label{LAGR_FOUR}
4866: \end{figure}
4867: %
4868: with a relative accuracy of about $10^{-4}$. For presentation purposes
4869: the time evolution is shown up to $t=500\,\,{\rm km}$ only. No
4870: significant loss of accuracy has been observed for longer evolutions. \\
4871: We have also compared the frequency spectrum resulting from time evolutions
4872: with the corresponding predictions by the eigenmode calculation. For
4873: this purpose we have used the same stellar model as in the previous test
4874: as well as model 1 of Table \ref{MODELS15} which has a polytropic index
4875: $\gamma = 1.75$. In both cases the initial
4876: perturbation is given by the sum of the first ten eigenmodes in
4877: the displacement $\xi$. The combined amplitude is about $10\,\,{\rm cm}$
4878: in both cases, so that the deviation from the linear approximation should
4879: again be very small. In Fig.\,\ref{LAGR_FOUR} we show
4880: the Fourier spectra of the central energy density perturbation
4881: $\Delta \rho(t,0)$ obtained
4882: for time evolutions over $1500\,\,{\rm km}$ using 600 grid points.
4883: The vertical bars indicate the frequencies predicted for the first 10
4884: eigenmodes and coincide well with the peaks in the power spectra. \\
4885: In order to test the performance of the code in the non-linear regime
4886: %
4887: \begin{figure}[t]
4888: \centering
4889: \epsfig{file=lqxi.eps, height=180pt, width=130pt, angle=-90}
4890: \hspace{1cm}
4891: \epsfig{file=lqu.eps, height=180pt, width=130pt, angle=-90}
4892: \epsfig{file=lqDN.eps, height=180pt, width=130pt, angle=-90}
4893: \hspace{1cm}
4894: \epsfig{file=lqDrho.eps, height=180pt, width=130pt, angle=-90}
4895: \epsfig{file=lqDlam.eps, height=180pt, width=130pt, angle=-90}
4896: \caption{The convergence factor for $\xi$, $w$, $\Delta N$, $\Delta \rho$
4897: and $\Delta \lambda$ obtained for 400 and 800 grid points.
4898: The reference solution has been calculated for 1600
4899: grid points.}
4900: \label{LAGR_CONV}
4901: \end{figure}
4902: %
4903: we have performed a convergence
4904: analysis for an initial displacement with the profile of the second
4905: eigenmode and an amplitude of about $50\,\,{\rm m}$ for the stellar model
4906: with $\gamma=3$ and $K=2.0\cdot 10^5\,\,{\rm km}^{-2}$. In this amplitude range
4907: non-linear effects are present, but shock formation is not yet expected
4908: for initial data with sufficiently weak spatial variation. We have
4909: evolved these initial data using 400, 800 and 1600 grid points and
4910: have calculated the time dependent convergence factor according to the
4911: method described in section \ref{CCM_CONVERGENCE}. Since the exact
4912: solution is not known, we use the reference solution for 1600 grid points
4913: in its place. The result obtained for the
4914: variables $\xi$, $w$, $\Delta N$, $\Delta \rho$ and $\Delta \lambda$
4915: is shown in Fig.\,\ref{LAGR_CONV} and
4916: demonstrates second order convergence throughout the evolution. \\
4917: Finally we have tested the code with the analytic solution
4918: by \lcite{Oppenheimer1939b}
4919: which describes the collapse of a homogeneous spherically symmetric
4920: dust cloud. \scite{Petrich1986} have expressed this analytic solution in
4921: polar slicing combined with radial or isotropic gauge. Even though
4922: we are using a Lagrangian gauge condition here, we can use their
4923: results for a comparison with our numerical simulation. \\
4924: %In their calculation Petrich et al. consider the dust collapse from the
4925: %point of view of the closed Friedmann metric
4926: %%
4927: %\begin{align}
4928: % ds^2 &= -d\tau^2 + a(\tau)^2 (d\chi^2 + \sin^2{\chi} d\Omega^2).
4929: %\end{align}
4930: %%
4931: %Furthermore they use the conformal time $\eta$
4932: %which is related to $\tau$ via
4933: %%
4934: %\begin{align}
4935: % \tau &= (\sin{\chi_{\rm s}})^{-3} (\eta - \sin{\eta}).
4936: %\end{align}
4937: %%
4938: %Here $\chi_{\rm s}$ is the value of the coordinate $\chi$ at the surface
4939: %of the dust cloud. The time parameter $\eta$ varies between $-\pi$ and $0$
4940: %as the dust sphere collapses from initial radius to $r=0$. On given
4941: %time slice $t=\mathrm{const}$, where $t$ is the time coordinate defined
4942: %by polar slicing $\eta$ is given as a function of $\chi$ by
4943: %
4944: %
4945: In their calculation of the analytic solution \shortciteANP{Petrich1986} use
4946: a Lagrangian coordinate $\chi$ and a time parameter $\eta$ which
4947: varies from $-\pi$ to $0$ as the dust sphere collapses from initial radius to
4948: $\hat{r}=0$. On a given time slice $t=\mathrm{const}$, where $t$ is the time
4949: coordinate defined by the polar slicing condition, $\eta$ is given as a
4950: function of $\chi$ by
4951: %
4952: \begin{align}
4953: \cos{\frac{\eta}{2}} &= \cos{\frac{\eta_{\rm s}}{2}}
4954: \sqrt{\frac{\cos{\chi_{\rm s}}}{\cos{\chi}}},
4955: \label{OS_ETAOFCHI}
4956: \end{align}
4957: %
4958: where $\eta_{\rm s}$ and $\chi_{\rm s}$ are the values of
4959: $\eta$ and $\chi$ at the surface of the dust cloud. If we label
4960: the initial slice by $\eta_{\rm s}=-\pi$, this equation implies that
4961: $\eta=-\pi$ everywhere on the initial slice.
4962: At any given time $t$ the areal radius is then shown to be related to the
4963: coordinate $\chi$ by
4964: %
4965: \begin{align}
4966: \hat{r} &= 2M \frac{\sin{\chi}}{\sin^3{\chi_{\rm s}}} \left( 1- \cos^2
4967: \frac{\eta_{\rm s}}{2} \cdot \frac{\cos{\chi_{\rm s}}}{\cos{\chi}}
4968: \right), \label{OS_ROFCHI}
4969: \end{align}
4970: %
4971: where $M$ is the Schwarzschild mass of the dust cloud.
4972: If we consider the special case of this equation at the surface and on
4973: the initial slice we can calculate $\chi_{\rm s}$ from
4974: %
4975: \begin{align}
4976: \sin^2{\chi_{\rm s}} &= \frac{2M}{R},
4977: \end{align}
4978: %
4979: where $R$ is the initial radius of the dust sphere. For reasons that will
4980: be given below we will identify the radial coordinate $x$ with the
4981: areal radius of the initial location of the fluid elements. We
4982: can therefore set $\eta_{\rm s}=-\pi$ and $\hat{r}=x$ in Eq.\,(\ref{OS_ROFCHI})
4983: and use the result to calculate $\chi(x)$ on the initial slice.
4984: Since
4985: both coordinates are comoving with the fluid elements, this relation
4986: between $\chi$ and $x$
4987: remains valid at any time $t$. In order to calculate $\eta(x)$ at a given time
4988: $t$ we still need to find the value $\eta_{\rm s}$. This is done by
4989: inverting the relation
4990: %
4991: \begin{align}
4992: t &= M \frac{\cos{\chi_{\rm s}}}{\sin^3{\chi_{\rm s}}} \left\{
4993: (\eta_{\rm s} - \sin{\eta_{\rm s}}) + 2 \sin^2{\chi_{\rm s}}
4994: \left[ \eta_{\rm s} - 2\tan{\chi_{\rm s}} \artanh{\left(
4995: \tan{\chi_{\rm s}} \cot{\frac{\eta_{\rm s}}{2}}\right)} \right]
4996: \right\},
4997: \end{align}
4998: %
4999: for which we use a Newton-Raphson method. Once $\eta_{\rm s}$ has been
5000: calculated, we can use Eq.\,(\ref{OS_ETAOFCHI}) to calculate $\eta(x)$
5001: on that time slice. The physical variables $\hat{r}$, $\hat{\rho}$,
5002: $\hat{\Gamma}$ and $\hat{\lambda}$ then follow from Eq.\,(\ref{OS_ROFCHI})
5003: and further relations by \shortciteANP{Petrich1986} which we write in the form
5004: %
5005: \begin{align}
5006: \hat{\rho} &= 6 \frac{a_0}{a^3} \frac{1}{8\pi M^2}, \\[10pt]
5007: \hat{\Gamma} &= \frac{\cos^3{\chi}-\cos^2{\frac{\eta_{\rm c}}{2}}}
5008: {\cos{\chi}-\cos^2{\frac{\eta_{\rm c}}{2}}}, \\[10pt]
5009: \hat{\lambda} &= -\frac{\hat{\lambda}_{\rm c}}{\sin{\frac{\eta_{\rm c}}{2}}}
5010: \frac{\cos{\chi}-\cos^2{\frac{\eta_{\rm c}}{2}}}
5011: {\sqrt{\cos^3{\chi}-\cos^2{\frac{\eta_{\rm c}}{2}}}},
5012: \end{align}
5013: %
5014: where $\lambda_{\rm c}$ is the central value of the lapse function
5015: and $a_0$ and $a$ are given by
5016: %
5017: \begin{align}
5018: a_0 &= \frac{1}{\sin^3{\chi_{\rm s}}}, \\[10pt]
5019: a &= a_0 (1-\cos{\eta}).
5020: \end{align}
5021: %
5022: In practice we specify the initial energy density and radius
5023: of the dust sphere and set the velocity to zero.
5024: The functions $\hat{N}$ and $\hat{\lambda}$ are
5025: then calculated from the constraint equations and the total mass
5026: of the sphere follows from the definition (\ref{NLAGR_NOFM}). \\
5027: From the numerical point of view the dust collapse is a special
5028: %
5029: \begin{figure}[t]
5030: \centering
5031: \epsfig{file=OSr.ps, height=400pt, width=175pt, angle=-90}
5032: \epsfig{file=OSrho.ps, height=400pt, width=175pt, angle=-90}
5033: \caption{The numerical simulation of the Oppenheimer-Snyder
5034: dust collapse for a dust sphere of $10\,\,{\rm km}$ radius
5035: and initial density $2\cdot 10^{-4}\,\,{\rm km}^{-2}$.
5036: The left panels show the numerical results for the
5037: radius $\hat{r}$ and the energy density $\hat{\rho}$, the right
5038: panels the deviation from the analytic solution.}
5039: \label{OS_RRHO}
5040: \end{figure}
5041: %
5042: %
5043: \begin{figure}[t]
5044: \centering
5045: \epsfig{file=OSGam.ps, height=400pt, width=175pt, angle=-90}
5046: \epsfig{file=OSlam.ps, height=400pt, width=175pt, angle=-90}
5047: \caption{Same as Fig.\,\ref{OS_RRHO} for the metric variables
5048: $\hat{\Gamma}$ and $\hat{\lambda}$.}
5049: \label{OS_GAMLAM}
5050: \end{figure}
5051: %
5052: case in several aspects which restricts our choice of the
5053: available options of the code.
5054: \begin{list}{\rm{(\arabic{count})}}{\usecounter{count}
5055: \labelwidth1cm \leftmargin1.5cm \labelsep0.4cm \rightmargin1cm
5056: \parsep0.5ex plus0.2ex minus0.1ex \itemsep0ex plus0.2ex}
5057: \item By definition the pressure vanishes in the dust sphere. As
5058: a result there is no static configuration analogous to the
5059: static neutron star governed by the TOV equations. We therefore
5060: need to use vacuum flat space as the background and run the
5061: code in the non-perturbative mode.
5062: \item The vanishing of the pressure also implies that the speed of sound
5063: is zero throughout the dust sphere so that it cannot be used
5064: to rescale the radial coordinate according to Eq.\,(\ref{TOV_YOFR}).
5065: The radial coordinate $x$ is therefore
5066: defined by the areal radius of the initial positions of the
5067: fluid elements and we use the condition $r_{,x}=1$ in the code.
5068: \item The surface of a neutron star with a polytropic equation of
5069: state is defined by the vanishing of the energy density $\hat{\rho}$
5070: which provided the outer boundary condition in the numerical evolution.
5071: For the dust sphere this relation is not valid any more and the
5072: energy density is finite at the outer boundary. The exact value,
5073: however, is not known, so that we cannot use it to derive an
5074: alternative boundary condition. The boundary condition $\hat{P}=0$
5075: is trivially satisfied in the case of a dust sphere and does not
5076: provide any extra information either. If we consider the
5077: structure of equations (\ref{NLAGR_LAMBDAX})-(\ref{NLAGR_UT}),
5078: however, we can see that all spatial derivatives of the energy density
5079: appear in the form of pressure gradients. These terms are
5080: identically zero in this case and disappear from the equations.
5081: We can therefore use the staggered grid for the energy density
5082: and thus eliminate the need of a boundary condition for $\hat{\rho}$.
5083: For this purpose we set the parameter $\sigma$ to 1 in the evolution
5084: of the dust sphere.
5085: \end{list}
5086: %
5087: In Figs.\,\ref{OS_RRHO} and \ref{OS_GAMLAM}
5088: we show the results obtained for a dust sphere with initial
5089: density $\hat{\rho}_0 = 2\cdot10^{-4}\,\,{\rm km}^{-2}$ and radius
5090: $R_0=10\,\,{\rm km}$ which corresponds to a total mass of
5091: $M=0.838\,\,{\rm km}$.
5092: A grid resolution of 800 points has been used for this calculation.
5093: The results demonstrate the good accuracy with which the code
5094: reproduces the analytic solution. Near the surface of the dust sphere,
5095: however, the numerical error increases significantly as the
5096: sphere approaches its Schwarzschild radius. We attribute this behaviour
5097: to the steep gradient of the lapse function near the surface that
5098: arises in the late stages of the evolution. \\
5099: This simulation also illustrates the singularity avoiding properties of the
5100: polar slicing condition. As the dust sphere collapses towards its Schwarzschild
5101: radius, the lapse function decreases towards zero and the evolution is
5102: practically frozen. This effect, the so called {\em collapse of the lapse},
5103: is responsible for the apparent slow down of the collapse of the
5104: radial function $r$ that can be seen in the upper left panel of
5105: Fig.\,\ref{OS_RRHO}. It is this property that makes polar
5106: slicing a popular choice for
5107: the numerical analysis of 1-dimensional gravitational collapse.
5108:
5109:
5110: %=======================================================================
5111: \subsection{Do shocks form at the surface for low amplitude oscillations?}
5112: %
5113: %
5114: We will now address a question that implicitly arose in the discussion
5115: of the linearized equations in the Eulerian formulation. We have seen
5116: in Eq.\,(\ref{DRHOOVERRHO})
5117: that the linearized equations predict a diverging ratio $\delta \rho /\rho$
5118: at the surface. For polytropic indices $\gamma>2$ we know that the
5119: divergence of $\delta \rho$ is a result of the Taylor expansion used
5120: to relate the Eulerian energy density perturbation to the Lagrangian one
5121: in Eq.\,(\ref{LIN_DRHOOFZETA}) and thus a non-physical result. For
5122: polytropic exponents $\gamma\le 2$, however, Eq.\,(\ref{LIN_DRHOOFZETA})
5123: represents a valid relation to first order in the perturbations,
5124: so that the Eulerian density perturbation
5125: will indeed be large compared with the background value near the surface.
5126: This behaviour raises the question whether non-linear effects will
5127: affect the evolution near the surface and give rise to
5128: the formation of discontinuities. From a different point of view one may
5129: consider the speed of sound which vanishes at the surface for a
5130: polytropic exponent $\gamma > 1$ and the particle speed $w$ which
5131: is finite because of the movement of the stellar surface. Consequently the
5132: velocity of the fluid elements will exceed the speed of sound and
5133: one may again ask whether this leads to shock formation.
5134: We will investigate this by using the exact treatment of the
5135: surface provided by the Lagrangian code. \\
5136: For this purpose we consider the neutron star model 3 of Table \ref{MODELS15}
5137: and provide initial data in the form of a displacement $\xi$ corresponding
5138: to a single eigenmode. For reasonably low order eigenmodes and amplitudes up to
5139: several metres we have not observed any significant deviation from
5140: the expected harmonic time dependence. For eigenmodes of very high order,
5141: however, this picture changes. We illustrate this in the case of an
5142: initial displacement of the fluid elements corresponding to
5143: a high order eigenmode (about 50) and an amplitude of about $1\,{\rm m}$
5144: at the surface. The high resolution of 3200 grid points has been used for
5145: this calculation to adequately resolve the high order mode. We stress that
5146: this evolution is only possible because of the high resolution near the
5147: surface provided by the rescaled variable $y$.
5148: %
5149: \begin{figure}[t]
5150: \centering
5151: \epsfig{file=sh0010.eps, height=200pt, width=150pt, angle=-90}
5152: \epsfig{file=sh2030.eps, height=200pt, width=150pt, angle=-90}
5153: \epsfig{file=sh4050.eps, height=200pt, width=150pt, angle=-90}
5154: \epsfig{file=sh6063.eps, height=200pt, width=150pt, angle=-90}
5155: \caption{The numerical evolution of the energy density
5156: perturbation $\Delta \rho$ as a function of $y$
5157: obtained for an initial displacement corresponding to about the
5158: 50th eigenmode with amplitude $1\,{\rm m}$.
5159: Snapshots are shown
5160: at times $t_1\ldots,t_8$.}
5161: \label{LAGR_SHOCKS}
5162: \end{figure}
5163: %
5164: In Fig.\,\ref{LAGR_SHOCKS} we show snapshots of the time evolution of
5165: the energy density perturbation at times $t_1=0.0$, $t_2=0.5$,
5166: $t_3=1.0$, $t_4=1.5$, $t_5=2.0$, $t_6=2.5$, $t_7=3.0$ and
5167: $t_8=3.1\,{\rm km}$. We note that only the small radial
5168: range $28\,{\rm km} \le y \le 31.7\,{\rm km}$ is shown in the figure.
5169: In terms of the areal radius this corresponds to a range of about
5170: $120\,{\rm m}$ below the surface. We can see that for this small
5171: amplitude a steep gradient forms near the surface
5172: after about $t=3.1\,{\rm km}$ which
5173: corresponds to less than two oscillation periods of the eigenmode.
5174: This indicates the formation of a discontinuity. At later times than shown
5175: here the code fails to converge which we attribute to the numerical
5176: noise caused by the shock formation and the extreme sensitivity of the
5177: code near the surface of the star.
5178: In order to demonstrate that this result is not merely due to
5179: numerical inaccuracies, we have evolved the same initial data with
5180: the smaller amplitude of $1\,{\rm cm}$. In Fig.\,\ref{LAGR_NOSHOCKS}
5181: we show the same snap shots for this evolution
5182: as in Fig.\,\ref{LAGR_SHOCKS}. In this case we
5183: obtain harmonic time dependence as expected in the linear limit.
5184: %
5185: \begin{figure}[t]
5186: \centering
5187: \epsfig{file=sm0010.eps, height=200pt, width=150pt, angle=-90}
5188: \epsfig{file=sm2030.eps, height=200pt, width=150pt, angle=-90}
5189: \epsfig{file=sm4050.eps, height=200pt, width=150pt, angle=-90}
5190: \epsfig{file=sm6063.eps, height=200pt, width=150pt, angle=-90}
5191: \caption{The same as Fig.\,\ref{LAGR_SHOCKS} but for an amplitude of
5192: $1\,{\rm cm}$.}
5193: \label{LAGR_NOSHOCKS}
5194: \end{figure}
5195: %
5196: By using eigenmodes with even higher order we have observed shock formation
5197: at the surface for smaller amplitudes.
5198: % In the case of initial
5199: %displacements in the form of lower order modes
5200: %we can go up to amplitudes significantly larger than those used here
5201: %without any indication of discontinuities.
5202: In view of the results for low order modes where no significant non-linear
5203: effects are observed for similar amplitudes,
5204: we conclude that the magnitude
5205: of non-linear effects is not only determined by the size of the perturbations
5206: relative to the background variables, but also by the length scale on which
5207: the perturbations vary significantly. We finally note that the surface of
5208: a neutron star is too complicated to be accurately described by
5209: the polytropic equation of state used for these evolutions. It is not clear
5210: whether discontinuities will form in the same way for more realistic
5211: descriptions of neutron stars. Nevertheless our results demonstrate that
5212: the surface requires a careful numerical treatment.
5213:
5214:
5215:
5216:
5217: