1: %===========================================================================
2: \section{Introduction}
3: %
4: %
5: In 1915 Albert Einstein published a geometrical theory of gravitation:
6: {\it The General Theory of Relativity}. He presented a fundamentally new
7: description of gravity in the sense that the relative acceleration of
8: particles is not viewed as a
9: consequence of gravitational forces but results from the curvature of the
10: spacetime in which the particles are moving. As long as no non-gravitational
11: forces act on a particle, it is always moving on a ``straight line''.
12: If we consider curved manifolds there is still a concept of
13: straight lines which are called {\em geodesics}, but these will
14: not necessarily have the properties we intuitively associate with straight
15: lines from our experience in flat Euclidean geometry. It is, for example,
16: a well known fact that two distinct straight lines in 2-dimensional
17: flat geometry will
18: intersect each other exactly once unless they are parallel in which case
19: they do not intersect each other at all. These
20: ideas result from the fifth Euclidean
21: postulate of geometry which
22: plays a special role in the formulation of geometry.
23: It is a well known fact that one needs to impose it
24: separately from the first four Euclidean postulates in order to
25: obtain flat Euclidean geometry. It was not realised until
26: the work of Gauss, Lobachevsky, Bolyai and Riemann in the 19th century
27: that the omission of the fifth postulate leads to an entirely new
28: class of non-Euclidean geometries in curved manifolds. A fundamental
29: feature of non-Euclidean geometry is that straight lines in curved
30: manifolds can intersect each other more than once and correspondingly
31: diverge from and converge towards each other several times. In order
32: to illustrate how these properties give rise to effects we commonly
33: associate with forces such as gravitation, we consider two
34: observers on the earth's surface, say one in Southampton and one in
35: Hamburg. We assume that these two observers start moving due south
36: in ``straight lines'' as for example guided by an idealised compass
37: exactly pointing towards the south pole. If we follow their separate
38: paths we will discover exactly the ideas outlined above. As long as both
39: observers are in the northern hemisphere the proper distance between
40: them will increase and reach a maximum when they reach the equator. From
41: then on they will gradually approach each other and their paths
42: will inevitably cross at the south pole. In the framework of Newtonian
43: physics the observers will attribute the relative acceleration of
44: their positions to the action of a force. It is clear, however, that no
45: force is acting in the east-west direction on either observer
46: at any stage of their journey. In a geometric description the
47: relative movement of the observers finds a qualitatively new interpretation
48: in terms of the curvature of the manifold they are moving in, the curvature
49: of the earth's surface. With the development of general relativity
50: Einstein provided the exact mathematical foundation for applying these
51: ideas to the forces of gravitation in 4-dimensional spacetime. One may
52: ask why such a geometrical interpretation has only been
53: developed for gravitation. Or in other words which feature distinguishes
54: gravitation from the other three fundamental interactions? The answer
55: lies in the ``gravitational charge'', the mass. It is a common observation
56: that the gravitational mass $m_{\rm G}$ which determines the coupling of a
57: particle to the gravitational field is virtually identical to the inertial
58: mass $m_{\rm I}$ which describes the particle's
59: kinematic reaction to an external force. High precision experiments have
60: been undertaken to measure the difference between these two types of masses.
61: All these results are compatible with the assumption that the masses are
62: indeed equal. The mass will therefore drop out of the Newtonian equations
63: governing the dynamics of a particle subject exclusively to gravitational
64: forces $m a = GmM/r^2 $, where $a$ is the acceleration of the particle,
65: $G$ the gravitational constant, $M$ the mass of an external source and
66: $r$ the distance from this source. The particle mass $m$ can be factored
67: out so that the movement of the particle is described in purely
68: kinematic terms. The redundancy of the concept of a gravitational
69: force is naturally incorporated into a geometric theory of
70: gravity such as general relativity.
71: It is important to note that this behaviour distinguishes gravity from the
72: other fundamental interactions which are associated with different types
73: of charges, such as electric charge in the case of electromagnetic interaction.
74: It is not obvious how and whether it is possible to obtain similar
75: geometric formulations for the electromagnetic, weak and strong interaction.
76: The unification of these three fundamental forces with gravity in the
77: framework of quantum theory is one of the important areas of ongoing
78: research. \\
79: In order to formalize the ideas mentioned above, general relativity views
80: spacetime as a 4-dimensional manifold equipped with a metric
81: $\hbox{\vec g}_{\alpha \beta}$ of Lorentzian signature
82: where the Greek indices range from 0 to 3.
83: At any given point in the manifold the
84: signature enables one to distinguish between time-like, space-like
85: and null directions. The metric further induces a whole range of
86: higher level geometric concepts on the manifold. It defines a
87: scalar product between vectors which leads to the measurement
88: of length and the idea of orthogonality. From the metric and
89: its derivatives one can derive a {\em connection} on the manifold which
90: facilitates the definition of a {\em covariant derivative}. The notion
91: of a derivative is more complicated in a curved manifold than
92: in the common case of flat geometry and Cartesian coordinates because
93: the basis vectors will in general vary
94: from point to point in the manifold. It is therefore no longer possible
95: to identify the derivative of a tensor with the derivative of its
96: components. Instead one obtains extra terms involving the derivatives
97: of the basis vectors. In terms of a covariant derivative these
98: terms are represented by the connection. In general relativity one
99: uses a metric-compatible connection defined by
100: %
101: \begin{align*}
102: \Gamma^{\gamma}_{\alpha \beta} &= \frac{1}{2} \hbox{\vec g}^{\gamma \delta}
103: (\partial_{\alpha} \hbox{\vec g}_{\beta \delta} + \partial_{\beta}
104: \hbox{\vec g}_{\alpha \delta} - \partial_{\delta}
105: \hbox{\vec g}_{\alpha \beta}),
106: \end{align*}
107: %
108: where the {\em Einstein summation convention}, according to which
109: one sums over repeated upper and lower indices, has been used.
110: These connection coefficients
111: are also known as the {\em Christoffel symbols} and define a covariant
112: derivative of tensors of arbitrary rank by
113: %
114: \begin{align*}
115: \nabla_{\delta} \hbox{\vec T}^{\alpha \beta}{}_{\gamma}
116: &= \partial_{\delta} \hbox{\vec T}^{\alpha \beta}{}_{\gamma}
117: + \Gamma^{\alpha}_{\rho \delta} \hbox{\vec T}^{\rho \beta}{}_{\gamma}
118: + \Gamma^{\beta}_{\rho \delta} \hbox{\vec T}^{\alpha \rho}{}_{\gamma}
119: - \Gamma^{\rho}_{\gamma \delta} \hbox{\vec T}^{\alpha \beta}{}_{\rho},
120: \end{align*}
121: %
122: where $\partial_{\delta}$ represents the standard partial derivative with
123: respect to the coordinate $x^{\delta}$.
124: So for each upper index one adds a term containing the connection
125: coefficients and for each lower index a corresponding term is subtracted.
126: With the definition of a covariant derivative we can finally write
127: down the exact definition of a ``straight line'' in a curved manifold.
128: A geodesic is defined as the integral curve of a vector field
129: $\hbox{\vec v}$ which is parallel transported along itself
130: %
131: \begin{align*}
132: \hbox{\vec v}^{\alpha} \nabla_{\alpha} \hbox{\vec v}^{\beta} &= 0.
133: \end{align*}
134: %
135: Based on the covariant derivative we can also give a precise definition
136: of curvature. For this purpose the {\em Riemann tensor} is defined
137: by
138: %
139: \begin{align*}
140: \hbox{\vec R}^{\alpha}{}_{\beta \gamma \delta} &= \partial_{\gamma}
141: \Gamma^{\alpha}_{\delta \beta} - \partial_{\delta}
142: \Gamma^{\alpha}_{\gamma \beta} + \Gamma^{\alpha}_{\gamma \rho}
143: \Gamma^{\rho}_{\delta \beta} - \Gamma^{\alpha}_{\delta \rho}
144: \Gamma^{\rho}_{\gamma \beta}.
145: \end{align*}
146: %
147: If we use a coordinate basis, i.e. $\hbox{\vec e}_{\alpha}
148: = \partial/\partial x^{\alpha}$,
149: this definition can be shown to imply that for any vector field
150: $\hbox{\vec v}^{\alpha}$
151: %
152: \begin{align*}
153: \hbox{\vec R}^{\alpha}{}_{\beta \gamma \delta} \hbox{\vec v}^{\beta}
154: &= \nabla_{\gamma} \nabla_{\delta} \hbox{\vec v}^{\alpha}
155: -\nabla_{\delta} \nabla_{\gamma} \hbox{\vec v}^{\alpha},
156: \end{align*}
157: %
158: which is commonly interpreted by saying that a vector
159: $\hbox{\vec v}$ is changed by being
160: parallel transported around a closed loop unless the curvature vanishes
161: (see for example \citeNP{Misner1973}). In order to describe the effect
162: of the matter distribution on the geometry of spacetime one defines the
163: {\em Ricci tensor} as the {\em contraction} of the Riemann tensor
164: $\hbox{\vec R}_{\beta \delta} = \hbox{\vec R}^{\alpha}{}_{\beta \alpha
165: \delta}$, where again the Einstein summation convention for repeated
166: indices has been used. The geometry and the matter are then related by
167: the Einstein field equations
168: %
169: \begin{align*}
170: \hbox{\vec G}_{\alpha \beta}
171: := \hbox{\vec R}_{\alpha \beta} -1/2\,R\,\hbox{\vec g}_{\alpha \beta}\,
172: = 8\pi \hbox{\vec T}_{\alpha \beta},
173: \end{align*}
174: %
175: where $R=\hbox{\vec R}^{\alpha}{}_{\alpha}$ is the {\em Ricci scalar}
176: and $\hbox{\vec T}_{\alpha \beta}$ the {\em energy momentum tensor}.
177: The interaction between the matter distribution and the geometry of
178: spacetime can be summed up in the words of \citeANP{Misner1973}:
179: {\em ``Space acts on matter, telling it how to move. In turn, matter
180: reacts back on space, telling it how to curve''}. \\
181: Although the field equations look rather neat in the
182: compact notation we have given above, this should not hide the fact that
183: the {\em Einstein tensor}
184: $\hbox{\vec G}_{\alpha \beta}$ is in fact a complicated function of the
185: metric $\hbox{\vec g}_{\alpha \beta}$ and its first and second derivatives.
186: Due to the symmetry of the Einstein tensor and the energy momentum tensor
187: the field equations represent 10 coupled, non-linear partial differential
188: equations, which written explicitly
189: may contain of the order of 100,000 terms in the general case.
190: It therefore came as quite a surprise
191: when Karl Schwarzschild found a non-trivial, analytic
192: solution to these equations just some months after their
193: publication. Since then many analytic solutions have been found and
194: a whole branch of the studies of general relativity is concerned with
195: their classification. Enormous insight into the structure of general
196: relativity has been gained from these analytic solutions, but due to the
197: complexity of the field equations these solutions are normally
198: idealized and restricted by symmetry assumptions. In order
199: to obtain accurate descriptions of astrophysically relevant scenarios
200: one may therefore have to go beyond purely analytic studies. A
201: particularly important area of research connected with general relativity
202: that has emerged in recent years concerns the detection of
203: {\em gravitational waves}. In analogy to the prediction of electromagnetic
204: waves by the Maxwell equations of electrodynamics, the Einstein field
205: equations admit radiative solutions with a characteristic propagation
206: speed given by the speed of light. Due to the weak coupling constant
207: of the gravitational interaction, which is a factor of $10^{40}$ smaller
208: than the electromagnetic coupling constant, gravitational waves will
209: have an extremely small effect on the movement of matter and are
210: correspondingly difficult to detect. If one considers for example a metal
211: bar of a length of several kilometres, estimates have shown that the detection
212: of gravitational waves requires one to measure
213: changes in length orders of magnitude smaller than the diameter of an atomic
214: nucleus. Even though attempts to detect gravitational radiation go
215: back to the work of Joe Weber in the early sixties, it is only the
216: recent advance of computer and laser technology that provides
217: scientists with a realistic chance of success. The current generation of
218: gravitational wave detectors GEO-600, LIGO, TAMA and VIRGO that have
219: been constructed for this purpose are complex multi-national collaborations
220: and have recently gone online or are expected to go online in the near
221: future. Due to the extreme smallness of the signals, the accumulation of data
222: over several years is expected to improve the chances of a positive
223: identification of signals from extra-galactic sources. \\
224: Confidence in the
225: existence of gravitational waves has been significantly boosted by the
226: Nobel prize winning discovery of the
227: binary neutron star system PSR1913$+$16 (\citeNP{Hulse1975},
228: \citeNP{Taylor1989}). The spin-down of this system has
229: been found to agree remarkably well with the energy-loss predicted
230: by general relativity due to the emission of gravitational waves and
231: is generally accepted as indirect proof of the existence of gravitational
232: radiation. \\
233: In order to simplify the enormous task of detecting gravitational waves, it is
234: vital to obtain information about the structure of the signals one
235: is looking for. It is necessary for this purpose to accurately
236: model the astrophysical scenarios that are considered likely
237: sources of gravitational waves
238: and extract the corresponding signals from these models.
239: According to {\em Birkhoff's} \citeyear{Birkhoff1923} {\em theorem}
240: the Schwarzschild solution, which describes a static, spherically symmetric
241: vacuum spacetime, is the only spherically symmetric,
242: asymptotically flat solution to the Einstein vacuum field equations.
243: As a consequence a spherically symmetric spacetime,
244: even if it contains a radially
245: pulsating object, will necessarily have an exterior static region and
246: be non-radiating.
247: %According
248: %to a well known result of general relativity a spherically symmetric
249: %spacetime will necessarily be non-radiating, so that
250: It is necessary, therefore,
251: to use less restrictive symmetry assumptions in the modelling of
252: astrophysical sources of gravitational waves.
253: In fact the most promising
254: sources of gravitational waves currently under consideration are the
255: in-spiralling and merger of two compact bodies (neutron stars or black holes)
256: and complicated oscillation modes of neutron stars that increase in amplitude
257: due to the emission of gravitational waves
258: by extracting energy from the rotation of the star. Even though
259: a great deal of information about these scenarios has been gained
260: from approximative studies, such as the
261: {\em post-Newtonian} formalism or the use of
262: {\em perturbative techniques}, a detailed simulation will require the
263: solution of the Einstein equations in three dimensions.
264: The complicated
265: structure of the corresponding models in combination with the enormous
266: advance in computer technology has given rise to
267: {\em numerical relativity}, the computer based
268: generation of solutions to Einstein's field equations. \\
269: In order to numerically solve Einstein's field equations
270: it is necessary to cast the equations in a form suitable for a
271: computer based treatment. Among the formulations proposed for this
272: purpose by far the most frequently applied
273: is the canonical ``3+1'' decomposition of \lcite{Arnowitt1962},
274: commonly referred to as the ADM formalism. In this approach spacetime
275: is decomposed into a 1-parameter family of 3-dimensional space-like
276: hypersurfaces and the Einstein equations are put into the form
277: of an initial value problem. Initial data is provided on one
278: hypersurface in the form of the spatial 3-metric and its time derivative and
279: this data is evolved subject to certain constraints and the specification
280: of gauge choices.
281: It is a known problem, however, that the ADM formalism does not result
282: in a strictly hyperbolic formulation of the Einstein equations and in
283: combination with its complicated structure the stability properties
284: of the ensuing finite differencing schemes remain unclear. These difficulties
285: have given rise to the development of modified versions of the
286: ADM formulation in which the Einstein equations are written as a
287: hyperbolic system. These and similar modifications of the canonical ADM
288: scheme have been successfully tested, but
289: %hyperbolic ``3+1'' formulations (see for example \shortciteNP{Bona1995},
290: %\shortciteNP{Friedrich1996}, \shortciteNP{Anderson1997}) which
291: %in turn facilitate
292: %the application of powerful mathematical theorems.
293: %The conformal decomposition of \lcite{Shibata1995} and \lcite{Baumgarte1999}
294: %(``BSSN'' for short)
295: %aims in the same direction and has been successfully implemented
296: %in several cases (see for example \shortciteNP{Baumgarte1999},
297: %\shortciteNP{Alcubierre2001}).
298: %The final answer as to the optimal ``3+1''
299: %formalism has not yet been found, but attention has definitely shifted
300: %in recent years from the standard ADM-formulation to alternative schemes
301: %such as the ``BSSN''-scheme. The majority of these schemes, however,
302: %are based
303: %on the standard ADM decomposition of spacetime, so that a description of
304: %the canonical ADM-formalism is essential for their understanding.
305: an optimal ``3+1'' formulation has yet to be found
306: and it may well be possible
307: %that there is not {\em one} optimal formulation but instead
308: that an optimal ``3+1''-strategy depends sensitively on the problem that needs
309: to be solved. \\
310: An entirely different approach to the field equations
311: is based on the decomposition of spacetime into families of null-surfaces,
312: the characteristic surfaces of the propagation of gravitational
313: radiation. The Einstein field equations are again formulated
314: as an initial value problem and by virtue of a suitable choice of
315: characteristic coordinates one obtains
316: a natural classification of the equations into
317: evolution and hypersurface equations.
318: The characteristic initial value problem was first formulated
319: by \scite{Bondi1962} and \lcite{Sachs1962} in order to facilitate
320: a rigorous analysis of gravitational radiation which is properly described
321: at null infinity only. It is a generic drawback of ``3+1'' formulations
322: that null infinity cannot be included in the numerical grid by means
323: of compactifying spacetime and instead outgoing radiation boundary conditions
324: need to be used at finite radius. Aside from the non-rigorous analysis
325: of gravitational radiation at finite distances these artificial boundary
326: conditions give rise to spurious numerical reflections. A characteristic
327: formulation resolves these problems in a natural way but is itself
328: vulnerable to the formation of caustics in regions of strong curvature.
329: It is these properties of ``3+1'' formulations and the characteristic method
330: that resulted in the idea of {\em Cauchy characteristic matching} (CCM),
331: i.e. the combination of a ``3+1'' scheme applied in the interior and a
332: characteristic formalism in the outer vacuum region. This allows one to
333: make use of the advantages of both methods as we will illustrate in more
334: detail below.\\
335: This thesis consists of four parts. First we will investigate the Einstein
336: field equations from the numerical point of view. This includes a detailed
337: description of the ADM and the characteristic Bondi-Sachs formalism as
338: well as a general discussion of finite difference methods and numerical
339: concepts such as stability and convergence. Section \ref{ccm} is concerned with
340: Cauchy characteristic matching as a numerical tool to solve the field
341: equations. In particular we present
342: a long term stable CCM code for cylindrically symmetric
343: vacuum spacetimes containing both gravitational degrees of freedom.
344: In section \ref{cstr} we investigate the behaviour of static and
345: dynamic cosmic strings in cylindrical symmetry.
346: The numerical codes developed for the analysis are described together
347: with a detailed study of the oscillations of a cosmic string
348: excited by gravitational radiation.
349: Finally in section \ref{pert} we present a fully non-linear perturbative
350: approach to study non-linear radial oscillations of neutron stars.
351: The perturbative formulation enables us to study non-linear oscillations over
352: a large amplitude range with high precision. In an Eulerian formulation,
353: however, the surface of the star gives rise to numerical difficulties
354: which leads us to investigate a simplified neutron star model instead.
355: The section is concluded with the development of a Lagrangian formulation
356: of dynamic spherically symmetric stars
357: in which the surface problems are resolved in a natural way.
358: % The resulting
359: %code is tested for second order convergence and shown to accurately
360: %reproduce analytic solutions.
361: We use the exact treatment of the
362: surface for the analysis of shock formation near the surface for initial
363: data of low amplitude.
364:
365:
366:
367: