gr-qc0410040/ch1.tex
1: \resetcounters
2: 
3: \chapter{Introduction} \lab{intro}
4: 
5: 
6: 
7: 
8: This thesis is concerned with the numerical simulation of boson stars
9: within the framework of Einstein's theory of general relativity.
10: Boson stars are self-gravitating compact objects~\footnote
11: {
12: 	By compact, we mean {\em gravitationally compact}, so that the size, $R$,
13: 	of the star is comparable to its Schwarzschild radius, $R_S$.  The
14: 	Schwarzschild radius associated with a mass $M$ is $R_S = (2G/c^2) M$, where
15: 	$G$ is Newton's gravitational constant, and $c$ is the speed of light.
16: }
17: composed of scalar particles~\cite{kaup:1968},
18: and are the bosonic
19: counterparts of the more well-known compact fermionic stars, which include 
20: white dwarfs and especially neutron stars.  
21: In contrast to fermionic stars, there
22: is no observational evidence that boson stars exist in nature,
23: nor has the existence of any {\em fundamental} scalar particle yet been verified by experiments.
24: It has been proposed, however, that 
25: boson stars could be a candidate for, or at least make up a considerable fraction
26: of, the dark matter in our universe~\cite{jetzer}.  Were this true, it is quite probable
27: that studies of their properties would lead to a better understanding of 
28: astrophysical phenomena.  However, at the current time boson stars remain purely
29: theoretical entities.
30: 
31: Their current hypothetical nature notwithstanding, boson stars might indeed exist
32: in our universe, and, more importantly for this thesis, they  
33: are excellent matter models for the numerical study of compact objects in strong 
34: gravitational fields, 
35: a subject that continues to be a key concern of numerical
36: relativity~\cite{luis:2001} ({\em numerical relativity:} the numerical simulation of 
37: the Einstein field equations, as well as the field equations for
38: any matter fields being modeled).
39: The boson stars we consider can be viewed as zero-temperature, ground-state, Bose-Einstein
40: condensates, with enormous occupation numbers, so that the stellar material
41: is described by a single complex scalar field, $\phi(t,{\bf x})$, that satisfies 
42: a simple, classical Klein-Gordon equation.
43: Many of the nice modeling properties associated with boson stars 
44: derives from the fact that the dynamics is governed by a partial differential
45: equation (PDE) that does {\em not} tend to develop discontinuities from
46: smooth initial data.  This is not the case for 
47: fermionic stars, which are usually treated as perfect
48: fluids, and which thus satisfy hydrodynamical equations with phenomenologically-determined
49: equations of state.  Relativistic fluid evolution will generically produce shock waves and
50: other discontinuities, even from smooth initial data.
51: Therefore, today's state-of-the-art relativistic hydrodynamics codes use sophisticated 
52: {\em high resolution shock capturing} (HRSC) schemes in order to more accurately and efficiently simulate the 
53: fluid physics.  The fact that the solutions of the Klein-Gordon equation stay smooth
54: is a tremendous boon when it comes to discretizing the equation in a stable and 
55: accurate manner.
56: 
57: In addition, the effective integration of HRSC schemes with other advanced
58: numerical techniques such as adaptive mesh refinement~\cite{BO}  (AMR) is still 
59: in its infancy, at least for the case of general relativistic applications.  
60: For bosonic matter, on the other hand, it has proven relatively easy to 
61: use rather generic AMR algorithms (such as that of Berger and Oliger~\cite{BO}) in
62: conjunction with straightforward second-order finite-difference discretization to 
63: achieve essentially unbounded dynamical range in an efficient
64: fashion~\cite{choptuik,fransp:phd,graxi:2003}.
65: 
66: Of course, we do not primarily study boson stars because they are 
67: easy to simulate.  As mentioned already, a good fraction of the work 
68: in numerical relativity is concerned with the strong-field 
69: dynamics of gravitationally compact
70: objects~\cite{jetzer,schunck:2003,baumgarte:2003b}.
71: Despite the large amount of 
72: effort that has been devoted to this subject, it is fair to say that much 
73: remains to be learned, and much of what remains to be discovered is 
74: likely to be found through simulation.  Although the strong-field
75: gravitational physics of boson stars may not compare in detail with that of 
76: fermionic stars in all respects,
77: there are clearly some key features of the usual
78: stars that are shared by their bosonic counterparts.  For example,
79: in analogy with relativistic fermionic stars, spherically symmetric 
80: boson stars typically come in one-parameter families, where the 
81: parameter can be viewed as the central density of the star (or 
82: an analogue thereof).  Moreover, in relativistic cases
83: the boson
84: star families share with the fermionic sequences the surprising 
85: property that there comes a point when increasing the density of 
86: the star at the center actually {\em decreases} the total gravitating 
87: mass of the star.  In both cases this leads to maximal masses for any given
88: family of stars.  In addition, in both instances, stars near that limit are 
89: naturally very strongly self-gravitating.
90: 
91: Suffice it to say, then, that a careful study of general relativistic boson 
92: stars is likely to lead to insights into strong-field gravitational
93: physics, even if there are no immediate astrophysical applications, and that
94: this is the primary motivation for the calculations described below. 
95: In particular, and again as with the fermionic case, black 
96: hole formation is a crucial process that can, and will, generically 
97: occur in the strong-field dynamics of one or more boson stars.
98: For the most part, this is simply because (1) as gravitationally compact
99: objects, boson stars are, by definition, close to the point of collapse, and 
100: (2) the process of gravitational collapse involving matter with positive 
101: energy tends to be very unstable in many senses, including the fact 
102: that black hole areas can only increase. 
103: 
104: Over the past decade or so, the careful study of gravitational collapse 
105: and black hole formation has lead to the discovery of black hole 
106: critical phenomena~\cite{choptuik}, wherein the process of black hole formation,
107: studied in solution space, takes on many of the features of 
108: a phase transition in a statistical mechanical system.
109: A primary goal of the work presented in this thesis is to study
110: so-called Type I critical phenomena of boson stars in both spherical
111: symmetry and axisymmetry, and this will be explained in more detail
112: below.  The simulations are carried out via finite difference solution
113: of the governing PDEs. 
114: We also develop and apply a new algorithm for constructing solutions 
115: of the time-independent form of the PDEs in axisymmetry; these solutions 
116: represent relativistic {\em rotating} boson stars.
117: 
118: \section{An Overview of Boson Stars} 
119: 
120: 
121: The study of boson stars can be traced back to the work of Wheeler.
122: Wheeler studied self-gravitating objects whose constituent element is the 
123: electromagnetic field and named the resulting ``photonic'' configurations geons~\cite{wheeler:1955}.
124: Wheeler's original intent was to construct a self-consistent, classical and field-theoretical notion of body, 
125: thus providing a divergence-free model for the Newtonian concept of body.  In the late 1960s,
126: Kaup \cite{kaup:1968} adopted the geon idea, but coupled a massive complex scalar field, rather than
127: the electromagnetic field, to general-relativistic gravity. 
128: Assuming time-independence and spherical symmetry he found solutions of the coupled equations which he called Klein-Gordon 
129: geons.  Subsequently, Ruffini \& Bonazzola \cite{ruffini:1969} 
130: studied field quantization of a real scalar field and considered the ground state
131: configurations of a system of such particles.  The expectation value of the field
132: operators gives the same energy-momentum tensor as those given by Kaup, and hence the
133: different approaches followed in the two studies give essentially the same macroscopic results.
134: 
135: Later, these Klein-Gordon geons were given the name boson stars, and the 
136: nomenclature {\em boson star} now
137: generally refers to compact self-gravitating objects that are
138: regular everywhere and that are made up of scalar fields.  Variations of the 
139: original model studied by Kaup and Ruffini \& Bonazzola include
140: self-interacting boson stars (described by a Klein-Gordon field with one or more self-interaction terms), charged boson stars
141: (boson stars coupled to the electromagnetic field) and rotating boson stars (boson stars possessing
142: angular momentum), to name a few.
143: When stable, all of these objects are held together by the balance between the
144: attractive gravitational force and a pressure that can be viewed as
145: arising from 
146: Heisenberg's uncertainty principle, as well as any explicit repulsive self-interaction
147: between the bosons that is incorporated in the model.
148: Depending on the mass of the constituent particles, and on
149: the value of the self-interaction coupling constant(s), the size of the stars can in 
150: principle vary from the atomic scale to an astrophysically-relevant scale~\cite{schunck:2003}. 
151: As mentioned previously, any given
152: boson star is typically only one member of a continuous family of equilibrium solutions. 
153: In spherical symmetry, the family can be conveniently parametrized using 
154: the central value of the modulus of the scalar field, in analogy
155: to the central pressure of perfect fluid stars.  Moreover, the equilibrium configurations
156: generally have an exponentially decaying tail at large distances from the stellar core, in contrast
157: to fluid-models stars which tend to have sharp, well-defined edges.
158: 
159: \subsection{Stability of Boson Stars} \lab{stability_BS}
160: 
161: The dynamical stability of equilibrium, compact fermionic (fluid) stars
162: against gravitational collapse can be studied
163: using the linear perturbation analysis of infinitesimal radial oscillations
164: that conserve the total particle number, $N$,  and mass/energy, $M$. One important theorem in this 
165: regard
166: concerns the transition between stable and unstable equilibrium
167: \cite{HTW:grav_theory}\cite[pp.305]{weinberg:grav_cos}.  
168: The theorem 
169: states that a perfect fluid star with constant chemical composition and constant entropy per nucleon 
170: becomes unstable with respect to some radial mode only at central densities $\rho(0)$ 
171: such that 
172: \bea
173:   \fr{\pa M(\rh(0),s,\cdots)}{\pa \rh(0)} &=& 0\,, \\
174:   \fr{\pa N(\rh(0),s,\cdots)}{\pa \rh(0)} &=& 0\,.
175: \eea
176: \noi
177: In other words, a change in stability can only occur at those points in a curve 
178: of total mass $M(\rho(0))$ {\em vs} $\rho(0)$, where an extrema is attained.  
179: This results directly
180: from the fact that the eigenvalues, $\si^2$, of the associated pulsation equation change
181: sign at those points.  
182: 
183: Similar results hold
184: for various boson star models.  
185: As just mentioned, in the case of boson stars we use the
186: central value of the scalar field modulus, denoted $\ph_0(0)$, as the parameter for the family
187: of solutions, and it has been shown that the pulsation equation has a zero mode
188: at the stationary points in the $M(\ph_0(0))$ plot~~\cite{lee_pang:1989,glesier_watkins:1989}.  
189: Numerical verification of the instability of configurations past the 
190: mass maximum using the full dynamical equations in spherical symmetry was studied in
191: \cite{seidel_suen:1990,Balakrishna:1998} and \cite{scott_matt:2000}.
192: 
193: 
194: \subsection{Maximum Mass of Boson Stars}
195: As already mentioned, 
196: another important feature of stellar structure that is largely 
197: due to relativistic effects is the existence of
198: a maximum allowable mass for a particular species of stars.  Depending on
199: the mechanism of stellar pressure (degenerate electron pressure for white
200: dwarfs, degenerate neutron pressure for neutron stars, Heisenberg
201: uncertainty principle for boson stars) these values are different.  
202: White dwarfs and neutron stars share the same dependence on the mass of the
203: constituent particles ($\sim M_{\rm pl}^3/m^2$), while the dependence for 
204: boson stars is quite
205: different ($\sim M_{\rm pl}^2/m$).  The origin of
206: the difference can be understood via the following 
207: heuristic argument.
208: The ground state stationary boson stars are macroscopic quantum states of
209: cold, degenerate bosons, whose existence is the result of the balance
210: between the attractive gravitational force and the dispersive nature of
211: the wave function.  By the uncertainty principle, if the bosons are confined in a region of
212: size $R$, we have $p R \sim \hbar$, where $p$ is the typical momentum of the
213: bosons.  For a moderately relativistic boson we have $p \sim mc$ 
214: and hence $R \sim \hbar/(mc)$.  Equating this with the Schwarzschild radius
215: we have $\hbar/(mc) \sim 2GM_{\mathrm{max}}/c^2$.  Hence
216: $M_{\mathrm{max}}\sim (\hbar c)/(2Gm) = 0.5 M_{\rm pl}^2/m$.  Numerical
217: calculation shows that $M_{\mathrm{max}} \approx 0.633 M_{\rm pl}^2/m$ (see
218: Sec.~\ref{family_statsol}),
219: surprisingly close to the above estimate.
220: 
221: The maximum mass of neutron stars can be estimated in a similar way.~\footnote{
222: \cite{shapiro_teukolsky:BHWD} gives another heuristic argument due to Landau
223: (1932), which applies to both neutron stars and white dwarfs.  The current argument applies to neutron stars only; for white
224: dwarfs the mass of the star is dominated by baryons, while the pressure is
225: provided by electrons.}
226: The existence of these stars is the result of the balance
227: between the attractive gravitational force and the pressure due to
228: degenerate neutrons (fermions).  Suppose there
229: are $N$ fermions confined in a region of size $R$. Then by Pauli's exclusion
230: principle, each particle occupies a volume $1/n$, where $n \eq N/R^3$ is the
231: number density.  Effectively, each particle has a size of $R/N^{1/3}$.
232: Again, by the uncertainty principle we have $ pR/N^{1/3} \sim \hbar$.  Following
233: the same argument as for the boson star case, we have $R \sim \hbar N^{1/3}/(mc)$, and  hence
234: $2GM_{\mathrm{max}}/c^2 \sim \hbar N^{1/3}/(mc) \sim \hbar
235: {M_{\mathrm{max}}}^{1/3}/(m^{4/3}c)$.  Thus we have $M_{\mathrm{max}} \sim
236: 0.35 {M_{\rm pl}}^3/m^2$.  
237: In contrast to the bosonic case, then, the maximum mass of fermionic stars scales as
238: ${M_{\mathrm{max}}} \sim {M_{\rm pl}}^3/m^2$.
239: 
240: \subsection{Rotating Boson Stars}
241: 
242: Although spherically symmetric boson star solutions were found 
243: as early as 1968,
244: for many years it was unclear whether solutions describing time-independent boson 
245: stars with angular momentum existed or not.
246: The first attempt to construct such stars was due to Kobayashi, Kasai \&
247: Futamase in 1994~\cite{kobayashi:1994}.  These authors followed the same approach as 
248: Hartle and others~\cite{hartle:1967,hartle:1968}, in which 
249: the {\em slow} rotation of a general relativistic star is treated 
250: via perturbation of a 
251: spherically symmetric equilibrium configuration. 
252: In their study they found that slowly
253: rotating boson stars solutions coupled with a $U(1)$ gauge field (charged
254: boson stars) do {\em not} exist, at least perturbatively.  
255: Shortly thereafter, however, it was demonstrated that rotating boson stars
256: {\em could} be constructed on the basis of an ansatz which leads to a quantized
257: (or perhaps more properly, discretized) angular momentum.
258: It was later understood that rotating boson stars have quantized angular
259: momenta and that the concomitantly discrete nature of the on-axis regularity 
260: conditions prohibit a continuous (perturbative) change from 
261: non-rotating to rotating configurations. 
262: 
263: A year later, Silveira \& de~Sousa~\cite{sousa_silveira:1995}, following the
264: approach of Ferrell \& Gleiser~\cite{ferrell_gleiser:1989}, succeeded in
265: obtaining equilibrium solutions of rotating boson stars within the framework of
266: Newtonian gravity.  Specifically, they adopted the ansatz
267: \footnote{Note that (\ref{2dansatz}) is clearly not the most general ansatz we could make
268: for stationary solutions of the Einstein-Klein-Gordon system. For example, we could consider
269: $p$ independent bosonic fields $\ph_i(t,{\bf x}), i = 1,2,\cdots, p$, each satisfying ansatz
270: (\ref{2dansatz}) for specific values of $(k,\om) = (k_i, \om_i)$, i.e.\ $\ph_i(t,{\bf x}) =
271: \ph_i(r,\te) e^{i \left( \om_i t + k_i \vp\right)}$, with a total stress energy tensor
272: given by $T_{\mu \nu} = \sum_{i=1}^p T_{\mu \nu}^{\,i}$, with each of the $T_{\mu \nu}^{\,i}$ given
273: by  (\ref{Ctmunu}).  The stationary solutions $\ph_i(t,{\bf x}; k_i, \om_i(k_j))$ would then
274: be labelled by $p$ quantum numbers ($k_i$) with $p$ associated eigenvalues $\left(
275: \om_i(k_j)\right)$, and the spectrum would still be discrete.}
276: 
277: \beq \lab{2dansatz}
278:  \ph(t,r,\te,\vp) = \ph_0(r,\te) e^{i \left( \om t + k \vp\right)}\,,
279: \eeq
280: \noi
281: where $k$ is an integer (we use the symbol $k$, instead of the symbol $m$,
282: that is commonly used in quantum mechanics, to avoid confusion with the
283: particle mass, $m$), so that the stress-energy tensor $T^\phi_{\mu \nu}$
284: is independent of both time $t$ and the azimuthal angle $\vp$.
285: In the non-relativistic limit~\footnote{
286: By which we mean that {\em both} the gravitational and matter fields are treated using 
287: non-relativistic equations of motion.
288: } 
289: the governing field equations constitute a 
290: coupled Poisson-Schr\"{o}dinger (PS) system 
291: (for simplicity, we have dropped the subscript ``0'' so that $\ph_0(r,\te) \to \ph(r,\te)$)
292: 
293: \beq \lab{V_eq}
294:  \na^2 V = 8 \pi m^2 \ph \ph^{\ast}\,,
295: \eeq
296: \beq \lab{ph_eq}
297: -\fr{1}{2m} \na^2 \ph + m V \ph = E \ph\,,
298: \eeq
299: \noi
300: where $V$ is the Newtonian gravitational potential, and $E$ is an
301: energy eigenvalue.  
302: The scalar field $\ph_0(r,\te)$ is then expanded in associated Legendre functions,
303: $P_l^k(\te)$:
304: \beq \lab{ph_exp}
305:  \ph_0(r,\te) = \fr{1}{\sr{4 \pi}} \sum^{\infty}_{l=k} R_l(r) P^k_l(\te)\,,
306: \eeq 
307: Similarly, the potential $V(r,\te)$ is assumed to have no $\vp$-dependence, and thus can be expanded
308: as
309: \beq\lab{V_exp}
310:  V(r,\te) = \sum_{l=0}^{\infty} V_l(r) P_l(\te)\,.
311: \eeq
312: Eqs.~(\ref{ph_exp}) and (\ref{V_exp}) are then substituted into (\ref{V_eq})
313: and (\ref{ph_eq}), and multiplied by $P_{l_0}(\te)$ to obtain a system of
314: equations for any particular value of $l_0$
315: (using the orthogonality of the associated Legendre functions).
316: Using this strategy, Silveira and de~Sousa were 
317: able to obtain solutions for each specific combination of $l_0$ and $k$.  
318: A main difference of these solutions relative to the
319: spherically symmetric ones, is that the scalar field vanishes at
320: the origin, and hence the rotating star solutions have toroidal level
321: surfaces of the matter field, 
322: rather than spheroidal level surfaces as in the spherical case. 
323: 
324: 
325: In 1996 Schunck \& 
326: Mielke~\cite{schunck_mielke:1996} used the ansatz (\ref{2dansatz}) to 
327: construct rotating boson star solutions.
328: Specifically, 
329: they chose some particular values of $k$, specifically $k=0,1\cdots,10$ and $k=500$
330: and showed that solutions to the 
331: fully general relativistic equations {\em did} exist for those cases.
332: They also showed that 
333: the angular momentum, $J$, and the total particle number, $N$, of the stars are related by
334: \beq
335:  J = k N\,,
336: \eeq
337: \noi
338: where $k$ is the integer defined in (\ref{2dansatz}).  An important implication of
339: the above equation is that if we consider equilibrium configurations with
340: the same total particle number $N$, then the total angular momentum
341: has to be quantized.  This property is in clear contrast with that of a
342: perfect fluid star.
343: 
344: Later, in the work most relevant to the study of rotating boson stars described 
345: in this thesis,
346: Yoshida \& Eriguchi~\cite{yoshida_eriguchi:1997b} 
347: used a self-consistent-field method~\cite{yoshida_eriguchi:1997} 
348: to obtain the whole family of solutions
349: for $k=1$, as well as part of the family for $k=2$.  The maximum mass they found
350: for the $k=1$ case was $1.314M_{\rm pl}^2/m$.  However, 
351: their code broke down before they could compute the star with maximum mass 
352: for the $k=2$ case.
353: 
354: In Table~\ref{fermionboson} we summarize
355: some similarities and differences between rotating fermion stars and
356: rotating boson stars.  For further background information on boson 
357: stars we suggest that readers consult the review by Jetzer~\cite{jetzer}, or
358: the more up-to-date survey by Schunck \& Mielke~\cite{schunck:2003}.
359: 
360: \vspace{0.5cm}
361: \begin{table}[htbp]
362: \begin{center}
363: \begin{tabular}[l]{cc}
364: \hline
365:  {\bf Relativistic Rotating Fermion Stars }& {\bf Relativistic Rotating Boson Stars }\\ \hline
366: \hline
367: \multicolumn{2}{c}{{\em  Similarities }}  \\
368: \multicolumn{2}{c}{\small Come in families of solutions parametrized by a
369: single value} \\
370:              \multicolumn{2}{c}{\small Each family has a maximum possible mass} \\
371: \hline
372: \multicolumn{2}{c}{{\em  Differences}}  \\
373: {\small Parametrized by $p(0)$} & {\small Parametrized by 
374: $|\partial^{(k)}_r\phi(0)|$} \\
375:   {\small Spheroidal level surfaces of rotating matter field} &   {\small Toroidal level surfaces
376: of matter field for $k=1,2,\cdots$}\\
377:  {\small Finite size, abrupt change in $\partial_r p, \partial_r \rh$ at surface} & {\small
378: Exponential decay to infinity} \\
379:  {\small Angular momentum can vary continuously} & {\small Angular momentum
380: quantized: $k=1,2,\cdots$} \\
381: \hline
382: \end{tabular}
383: \caption[Similarities and differences between relativistic rotating fermion
384: stars and relativistic rotating boson stars]
385: {Similarities and differences between relativistic rotating fermion
386: stars and relativistic rotating boson stars.}
387: \end{center}
388: \label{fermionboson}
389: \end{table}
390: 
391: \section{Critical Phenomena in Gravitational Collapse} 
392: 
393: Over the past decade, intricate and unexpected phenomena related to 
394: black holes have been discovered through the detailed numerical 
395: study of various models for gravitational collapse,
396: starting with Choptuik's investigation of the spherically symmetric 
397: collapse of a massless scalar field~\cite{choptuik}.
398: These studies generally concern the {\em threshold} of black hole 
399: formation (a concept described below), and the phenomena observed
400: near threshold are collectively called (black hole) critical phenomena,
401: since they share many of the features associated with critical phenomena
402: in statistical mechanical systems.  The study of critical phenomena 
403: continues to be an active area of research in numerical relativity, 
404: and we refer the interested reader to the recent review article
405: by Gundlach~\cite{gundlach:2003} for full details on the subject. 
406: Here we will simply summarize some key points that are most germane 
407: to the work in this thesis.
408: 
409: To understand black hole critical phenomena, one must understand
410: the notion of the ``threshold of black hole formation".  
411: The basic idea is to consider {\em families} of solutions of
412: the coupled dynamical equations for the gravitational field 
413: and the matter field that is undergoing collapse (the complex 
414: scalar field, $\phi$, in our case).  Since we are considering a dynamical problem,
415: and since we assume that the overall dynamics is uniquely determined 
416: by the initial conditions, we can view the families as being 
417: parametrized by the initial conditions---variations in one or more 
418: of the parameters that fix the initial values will then generate 
419: various solution families.  We also emphasize that we are considering 
420: {\em collapse} problems.  This means that we will generically 
421: be studying the dynamics of systems that have length scales 
422: comparable to their Schwarzschild radii, {\em for at least some period of time 
423: during the dynamical evolution}.  We also note that we will often
424: take advantage of the complete freedom we have as numerical
425: experimentalists to choose initial conditions that lead to collapse, but which may 
426: be highly unlikely to occur in an astrophysical setting.
427: 
428: We now focus attention on {\em single parameter} families of 
429: data, so that the specification of the initial data is fixed 
430: up to the value of {\em the} family parameter, $p$.  We will generally 
431: view $p$ as a non-linear {\em control parameter} that will
432: be used to govern how strong the gravitational field becomes
433: in the subsequent evolution of the initial data, and in particular,
434: whether a black hole forms or not. Specifically, we will always 
435: demand that any one-parameter family of solutions has the 
436: following properties:
437: \begin{enumerate}  
438: \item For sufficiently small values of $p$ the dynamics remains regular 
439:       for all time, and no black hole forms.  
440: \item For sufficiently large values of $p$, complete gravitational collapse 
441:       sets in at some point during the dynamical development of the initial
442:       data, and a black hole forms.
443: \end{enumerate}
444: From the point of view of simulation, it turns out to be a relatively 
445: easy task for many models of collapse to construct such families,
446: and then to identify 2 specific parameter values, $p^-$ ($p^+$) which do not (do)
447: lead to black hole formation.  Once such a  ``bracket'' $[p^-,p^+]$ has been
448: found, it is straightforward in principle to use a technique such as 
449: binary search to hone in on a {\em critical parameter value}, $p^\star$, such 
450: that all solutions with $p<p^\star$ ($p>p^\star$) do not (do) contain
451: black holes. A solution corresponding to $p=p^\star$ thus sits at the 
452: threshold of black hole formation, and is known as a {\em critical solution}.
453: It should be emphasized that underlying the existence of critical solutions 
454: are the facts that (1) the end states (infinite-time behaviour)  corresponding to
455: properties 1.~and 2.~above are {\em distinct} (a spacetime containing a black hole 
456: {\em vs} a spacetime not containing a black hole) and (2) the process 
457: characterizing the black hole threshold (i.e.\ gravitational collapse) 
458: is {\em unstable}. We also note that we will term evolutions with $p<p^\star$
459: {\em subcritical}, while those with $p>p^\star$ will be called {\em supercritical}.
460: 
461: Having discussed the basic concepts underlying black hole critical phenomena,
462: we now briefly describe the features of critical collapse that are most 
463: relevant to the work in this thesis.  
464: 
465: First, critical solutions {\em do}
466: exist for all matter models that have been studied to date, and for 
467: any given matter model, almost certainly constitute discrete sets.  In
468: fact, for some models, there may be only {\em one} critical solution,
469: and we therefore have a form of universality.  
470: 
471: Second, critical solutions 
472: tend to have additional symmetry beyond that which has been adopted in the 
473: specification of the model (e.g. we will impose spherical and axial 
474: symmetry in our calculations).  
475: 
476: Third, the critical solutions known
477: thus far, and the black hole thresholds associated with them, come 
478: in two broad classes.  The first, dubbed Type I, is characterized 
479: by static or periodic critical solutions (i.e.\ the additional symmetry 
480: is a continuous or discrete time-translational symmetry), and by
481: the fact that the black hole mass just above threshold is {\em finite}
482: (i.e.\ so that there is a minimum black hole mass that can be 
483: formed from the collapse). 
484: The second class, called Type II, is characterized by continuously or 
485: discretely self-similar critical solutions (i.e.\ the additional
486: symmetry is a continuous or discrete scaling symmetry), and by the 
487: fact that the black hole mass just above threshold is {\em infinitesimal}
488: (i.e.\ so that there is {\em no} minimum for the black hole mass that 
489: can be formed).    The nomenclature Type I and Type II is by analogy 
490: with first and second order phase transitions in statistical mechanics,
491: and where the black hole mass is viewed as an order parameter.
492: 
493: Fourth, solutions close to criticality exhibit various scaling laws.
494: For example, in the case of Type I collapse, where the critical solution is an
495: unstable, time-independent (or periodic) compact object, the amount 
496: of time, $\tau$, that the dynamically evolved configuration is well 
497: approximated by the critical solution {\em per se} satisfies a scaling law of 
498: the form
499: \beq
500: 	\label{tau-scaling}
501: 	\tau(p) \sim -\gamma \ln | p - p^\star | \,,
502: \eeq
503: where $\gamma$ is a {\em universal} exponent in the sense of not 
504: depending on which particular family of initial data is used to 
505: generate the critical solution, and $\sim$ indicates that the relation 
506: (\ref{tau-scaling}) is expected to hold in the limit $p \to p^\star$.
507: 
508: Fifth, and finally, much insight into critical phenomena comes 
509: from the observation that although unstable, critical solutions 
510: tend to be {\em minimally} unstable, in the sense that they 
511: tend to have only a few, and perhaps only one, unstable modes 
512: in perturbation theory.  In fact, if one assumes that a Type I
513: solution, for example, has only a single unstable mode, then 
514: the growth factor (Lyapunov exponent) associated with that mode can 
515: be immediately related to the scaling exponent~$\gamma$ defined by
516: (\ref{tau-scaling}).
517: 
518: In this thesis we will be exclusively concerned with Type I critical 
519: phenomena, where the threshold solutions will generally turn
520: out to be unstable boson stars.  Previous work relevant to ours
521: includes studies by (1) Hawley \cite{shawley:phd} and Hawley \& Choptuik~\cite{scott_matt:2000} of 
522: boson stars in spherically symmetry, (2) Noble~\cite{scn:phd} of 
523: fluid stars in spherical symmetry and (3) Rousseau~\cite{rousseau:master} of 
524: axisymmetric boson stars within the context of the  conformally flat approximation
525: to general relativity.  Evidence for Type I transitions have been found in 
526: all three cases.
527: 
528: \section{Layout}
529: 
530: The remaining chapters of this thesis are organized as follows.  In
531: Chap.~\ref{MathForm} we summarize the mathematical formalism used in the
532: work of this thesis.  This includes a brief summary of the mathematical
533: model of spacetime, in which the key ingredient to be used is the Einstein
534: field equation.  We then summarize the ADM (3+1) formalism, which will be used in
535: the study of boson stars in spherical symmetry, as well as  the (2+1)+1 formalism,
536: which is used in the study of boson stars in axisymmetry.  We also describe
537: the Einstein-Klein-Gordon system, which is the fundamental set of PDEs 
538: underlying all of our studies.
539: 
540: In Chap.~\ref{num_method} we summarize the numerical methods used in the
541: thesis.  This includes finite differencing techniques which are central to
542: all the calculations shown;  the multigrid method, which is used in the construction
543: of rotating boson stars, as well as in the solution of the 
544: constraint equations in the
545: dynamical study of boson stars in axisymmetry;  adaptive mesh refinement,
546: which is essential in the study of critical phenomena in axisymmetry;
547: excision techniques which are used in the study of boson stars in spherical
548: symmetry; and the technique of spatial compactification, which is used in the 
549: construction of rotating boson stars.
550: 
551: In Chap.~\ref{bs1d} we study Type I critical phenomena of boson stars in spherical symmetry.
552: This research can be viewed as an extension of the work reported by Hawley \&
553: Choptuik in~\cite{scott_matt:2000}.  Our principal new result is 
554: compelling numerical evidence for the existence of oscillatory final states of
555: subcritical evolutions.  We also perform perturbation analyses and show 
556: that the simulation results agree very well with those obtained from perturbation
557: theory.  We then present a rudimentary, but stable and convergent implementation of the black hole excision
558: method for the model.  Supercritical simulations using excision show that the spacetimes 
559: approach static black holes at late time, so there is no impediment to very long run
560: times (in physical time).
561: 
562: In Chap.~\ref{bs2d} we describe a study of boson stars in axisymmetry.
563: We first present an algorithm to construct the equilibrium configurations of rotating
564: boson stars that is based on the multigrid technique.  We argue that our method
565: is more computationally efficient than methods previously used and reported 
566: in the literature.  
567: More importantly, we obtain 
568: numerical solutions in the highly relativistic regime for an angular momentum parameter
569: $k=2$.  We then discuss studies of the
570: dynamics of boson stars in axisymmetry.  
571: Following Choi's~\cite{dale} work in the Newtonian limit, we show that 
572: solitonic behaviour occurs in the head-on collision of boson stars with 
573: sufficiently large relative initial velocities.
574: We also present a study of Type I critical phenomena in the model.
575: The two classes of simulations (collisions of boson stars, and boson stars gravitationally
576: perturbed by a massless real scalar field) provide evidence for Type I 
577: black hole transitions, as well as scaling laws for the lifetime of near-critical 
578: configurations of the form~(\ref{tau-scaling}).  This represents the first 
579: time that Type I behaviour has been 
580: observed in the context of {\em axisymmetric} collapse.
581: 
582: The final chapter summarizes the results, gives overall conclusions and points to some 
583: directions of future work.  Several appendices providing various technical details 
584: are also included.
585: 
586: \section{Conventions, Notation and Units}\lab{notation}
587: Throughout this thesis the signature of 
588: metric is taken to be ($-$ + + +).  Spacetime indices of four dimensional (1 temporal +
589: 3 spatial) tensors are
590: labeled by lower case Greek letters ($\al, \bt, \ga, \cdots$). 
591: Spatial indices of three dimensional (3 spatial) tensors are labeled by lower case Latin 
592: letters starting from $i$ ($i, j, k, \cdots$).
593: Spacetime indices of three dimensional (1 temporal + 2 spatial) tensors are
594: labeled by 
595: the first few Latin indices ($a, b, c,\cdots, h$), and
596: spatial indices of two dimensional (2 spatial) tensors
597: are labelled by upper case Latin letter ($A, B, C,\cdots$).
598: The Einstein summation convention is implied for all types of indices. 
599: That is, repeated (one upper and one
600: lower) indices are automatically summed over the appropriate range.
601: Covariant derivatives are denoted by $\na$ or by a semi-colon ``;".
602: Ordinary partial derivatives are denoted by $\pa$ or by a comma ``,". 
603: Conventions
604: for the Riemann and Ricci tensors are $\na_{[\al} \na_{\bt]} V_{\ga} = \ha R_{\al
605: \bt \ga \de} V^{\de}, R_{\mu \nu} = R^{\al}{}_{\mu \al \nu}$.  
606: 
607: We adopt a system of ``natural" units 
608: in which $G=c=\hbar=1$.  
609: In this system the unit time, unit length and unit mass are known
610: as the Planck time, Planck length and Planck mass, respectively.
611: Specifically, we have
612: 
613: \bea
614:  T_{\rm{pl}} &=& \sr{\fr{\hbar G}{c^5}} = 5.39\times 10^{-44}\mbox{s}
615: \,,\\ 
616:  L_{\rm{pl}}&=&\sr{\fr{\hbar G}{c^3}} = 1.62\times10^{-35}\mbox{m}\,, \\
617:  M_{\rm{pl}} &=& \sr{\fr{\hbar c}{G}} = 2.18 \times 10^{-8}\mbox{kg}\,.
618: \eea
619: 
620: \noi
621: The corresponding energy scale is $M_{\rm{pl}}\,c^2 \approx
622: 10^{19}\mbox{GeV}$.  
623: For the scalar field, the action has dimension $[\hbar]$.  Hence
624: $\left[ d^4x (\pa \ph)^2 \right] = [\hbar]$. Therefore, $[\ph] =
625: [\sr{\hbar}]/L=\sr{M/T}$.  Thus, $\ph$ is measured in units of
626: $\sr{M_{\rm{pl}}/T_{\rm{pl}}} = 6.36\times 10^{17} \sr{\mbox{kg/s}}$.
627: 
628: Finally, we adopt a nomenclature commonly used in numerical work, whereby 
629: we refer to calculations requiring the solution of 
630: partial differential equations in $n$ {\em spatial} dimensions
631: as ``$n$D''.  In particular, we will often refer to spherically symmetric 
632: computations (whether time-dependent or not) as ``1D'', and axisymmetric
633: calculations (again, irrespective of any time-dependence) as ``2D''.
634: