0707.3632/ks.tex
1: \documentclass[12pt, reqno]{amsart}
2: \usepackage{amsmath, amssymb, amsthm, epsfig}
3: %\usepackage{showkeys}
4: 
5: %% Page dimensions
6: \setlength{\textheight}{8.7in}
7: \setlength{\textwidth}{6.7in}
8: \setlength{\topmargin}{-0.25in}
9: \setlength{\leftmargin}{-1.0in}
10: \setlength{\evensidemargin}{0.0in}
11: \setlength{\oddsidemargin}{0.0in}
12: 
13: %% Theorems and Equation numbering
14: \newtheorem{thm}{Theorem}[section]
15: \newtheorem{corollary}[thm]{Corollary}
16: \newtheorem{lemma}[thm]{Lemma}
17: \newtheorem{theorem}[thm]{Theorem}
18: \newtheorem{prop}[thm]{Proposition}
19: \newtheorem{idea}[thm]{Idea of proof}
20: \newtheorem{conj}[thm]{Conjecture}
21: \numberwithin{equation}{section}
22: 
23: \theoremstyle{definition}
24: \newtheorem{rem}[thm]{Remark}
25: \newtheorem{ex}[thm]{Example}
26: 
27: %% Greek letters
28: \newcommand{\al}{\alpha}
29: \renewcommand{\b}{\beta}
30: \renewcommand{\c}{\gamma}
31: \newcommand{\de}{\delta}
32: \newcommand{\e}{\varepsilon}
33: \newcommand{\la}{\lambda}
34: \renewcommand{\phi}{\varphi}
35: 
36: %% Abbreviations
37: \renewcommand{\d}{\partial}
38: \newcommand{\R}{{\mathbb R}}
39: \newcommand{\Case}[1]{\noindent \textbf{Case #1.}}
40: \newcommand{\Step}[1]{\noindent \textbf{Step #1.}}
41: 
42: 
43: %% Proof environments
44: \renewcommand{\qed}{\rule{3mm}{3mm}}
45: \renewenvironment{proof}
46: {\vspace{1mm}\noindent\textbf{Proof.}} {\hspace*{\fill} $\qed$\vspace{1mm}}
47: 
48: \begin{document}
49: \title[Sharp bounds on $2m/r$]{Sharp bounds on $2m/r$ for static spherical objects}
50: \author{Paschalis Karageorgis}
51: \author{John G. Stalker}
52: \address{School of Mathematics, Trinity College, Dublin 2, Ireland.}
53: \email{pete@maths.tcd.ie}
54: \email{stalker@maths.tcd.ie}
55: 
56: \begin{abstract}
57: Sharp bounds are obtained, under a variety of assumptions on the eigenvalues of the Einstein tensor, for the ratio of the
58: Hawking mass to the areal radius in static, spherically symmetric space-times.
59: \end{abstract}
60: \maketitle
61: 
62: \section{Introduction}
63: All of the space-times considered in this paper are connected, four-dimensional and satisfy the following conditions.
64: \begin{itemize}
65: \item \emph{Spherical Symmetry.}\footnote{We follow Synge~\cite{Synge} in calling this assumption
66: ``spherical symmetry'' for brevity.  This is a bit misleading, as we are assuming more than just spherical symmetry.  Our
67: assumption excludes, for example, Schwarzschild space, which lacks a time axis.} There is a time-like curve, called the
68: \emph{time axis}, with the property that at any point all normal directions are equivalent, \textit{i.e.}, that for any
69: two space-like normal unit vectors there is an isometry of the space-time which fixes the point and takes the first
70: vector to the second.  This defines an action of $SO(3)$ on the space-time whose orbits are called \emph{spheres}.
71: 
72: \item \emph{Staticity.}
73: There is a one-parameter group of isometries, called time translations, whose generating vector field is everywhere
74: time-like.
75: 
76: \item \emph{Regularity.}
77: The space-time, together with its metric, is of class $\mathcal{C}^3$, except possibly on 3-surfaces of discontinuity,
78: where the second derivatives of the metric are allowed to have jump discontinuities.
79: \end{itemize}
80: 
81: The somewhat odd looking regularity assumption is borrowed from Lichnerowicz~\cite{Lich}. It is meant to allow such
82: discontinuities as one expects to find at the interface between two different materials, but nothing worse.
83: 
84: The areal radius~$r$ is defined by the requirement that the area of a sphere be $4\pi r^2$. In terms of the radius~$r$
85: and metric tensor~$g$, we may then define the Hawking mass~$m$ by the relation
86: \begin{equation} \label{Hm}
87: g^{jk} \partial_j r \partial_ k r = 1 - \frac{2m}{r} \,.
88: \end{equation}
89: The purpose of this paper is to prove sharp bounds on the ratio $2m/r$ under various hypotheses on the eigenvalues of the
90: Einstein tensor. The particular hypotheses considered, their history and the resulting bounds are discussed in
91: Section~\ref{matter}.
92: 
93: Three general comments should be made at this stage.  First, the method employed is quite general and can be used to
94: obtain sharp bounds on $2m/r$ for any matter model, not just those described below. Second, obtaining sharp bounds is, in
95: each case, relatively easy.  Proving sharpness, while not conceptually difficult, requires considerably more effort.
96: Third, we carefully avoid the assumption, made tacitly by previous authors, that $2 m < r$. This point is discussed in
97: more detail in the next section.
98: 
99: Section~\ref{coordinates} is devoted to a discussion of coordinates and the components of the Einstein tensor in our
100: chosen coordinate system. Section~\ref{matter} introduces the various assumptions on this tensor which are needed for the
101: statement of our theorem. Our main result, Theorem~\ref{main}, appears in Section~\ref{omr}, while its proof is given in
102: Section~\ref{pro}.
103: 
104: \section{Geometry and Coordinates}\label{coordinates}
105: A space-time of the class considered above has coordinates $r$, $\theta$, $\phi$ and $t$, known as
106: \emph{curvature coordinates}, in which the metric takes the form
107: \begin{equation}\label{met}
108: g_{jk} \,dx^j dx^k = e^\al\,dr^2 + r^2(d\theta^2 + \sin^2\theta \,d\phi^2) - e^\c \,dt^2.
109: \end{equation}
110: Here, $\al$ and $\c$ are functions of $r$. As shown in Synge~\cite{Synge}, the Einstein tensor in curvature coordinates
111: is of the form
112: \begin{align}
113: G_r^r &= r^{-2} - r^{-2}e^{-\al} (1+r\c'), \label{eq1} \\
114: G_\theta^\theta = G_\phi^\phi &= e^{-\al} \left( -\frac{1}{2} \,\c'' -\frac{1}{4} \,\c'\c' - \frac{1}{2r} \,\c' +
115: \frac{1}{2r} \,\al' + \frac{1}{4} \,\al'\c' \right), \label{eq2}\\
116: G_t^t &= r^{-2} - r^{-2}e^{-\al} (1-r\al'). \label{eq3}
117: \end{align}
118: Here, primes denote derivatives with respect to $r$, while the off-diagonal entries are all zero.  The formulae become a
119: bit cleaner when one uses derivatives with respect to
120: \begin{equation*}
121: \b = 2\log r,
122: \end{equation*}
123: instead. Denoting such derivatives by dots, one obtains the equivalent system
124: \begin{align}
125: G_r^r &= r^{-2} - r^{-2}e^{-\al} (1+2\dot{\c}), \label{eq4} \\
126: G_\theta^\theta = G_\phi^\phi &= r^{-2} e^{-\al} (-2\ddot{\c} -\dot{\c}^2 +\dot{\al} + \dot{\al}\dot{\c}), \label{eq5}\\
127: G_t^t &= r^{-2} - r^{-2}e^{-\al} (1-2\dot{\al}). \label{eq6}
128: \end{align}
129: The corresponding Einstein tensor, given by Einstein's field equations, has diagonal entries
130: \begin{equation}\label{diag}
131: G_r^r = -8\pi p_R, \quad\quad G_\theta^\theta = G_\phi^\phi = -8\pi p_T, \quad\quad G_t^t = 8\pi \mu
132: \end{equation}
133: and all off-diagonal entries equal to zero.  Here, $p_R$ and $p_T$ are interpreted as the radial and tangential
134: pressures, respectively, while $\mu$ is interpreted as the energy density.
135: 
136: There are two annoying points about curvature coordinates.
137: \begin{itemize}
138: \item
139: The functions $\alpha$ and $\gamma$ may be of lower regularity than the metric, since $r$ itself is of lower regularity
140: than the metric.  This is discussed in more detail by Israel~\cite{Israel}.  For our purposes it suffices to note that
141: regularity is, in the presence of the other assumptions, equivalent to the statement that $\alpha$ and $\gamma$ are
142: $\mathcal{C}^3$ functions of~$r$, except possibly at certain points, where the radial pressure~$p_R$ is continuous and
143: the tangential pressure~$p_T$ and energy density $\mu$ are allowed to have jump discontinuities.  At $r = 0$, the correct
144: condition is that $\al'(0)= \c'(0) = 0$.
145: 
146: \item
147: The coordinates may fail to cover the whole space-time.  In fact, they cover the region from the time axis out to the
148: first marginally trapped sphere, \emph{i.e.}, the first sphere where $r = 2 m$.  If we were to assume, as most authors
149: do, that curvature coordinates cover the whole space-time, then we would, in effect, be making the very strong additional
150: assumption that $2 m / r < 1$ everywhere.  This we wish to avoid.  For the classes of space-times we consider it is, in
151: fact, the case that $ 2 m / r < 1$ everywhere, but this belongs to the conclusion of our theorem, not to its hypotheses.
152: 
153: The simplest example of a space-time that satisfies our spherical symmetry, staticity and regularity assumptions but has
154: a marginally trapped surface is de Sitter space, for which $e^\al = - e^{-\c} = 1 - r ^ 2 / R ^ 2$.  In this case, the
155: coordinates cover a region where $r < R$ but break down at the boundary.  Outside this region, there is another which is
156: isometric to the first, and it is easy to check that $2 m = r$ at $r = R$.  However, de Sitter space does not satisfy the
157: hypotheses of our theorem because it has negative pressures everywhere.
158: \end{itemize}
159: 
160: \section{Matter Models}\label{matter}
161: Various conditions on the three functions $p_R$, $p_T$ and $\mu$ are of interest:
162: \begin{itemize}
163: \item \textit{Non-negative Isotropic Pressure:}
164: For fluids, one expects $p_R = p_T \geq 0$.  The sharp bound
165: \begin{equation*}
166: 2 m / r \leq 12\sqrt 2 - 16 \approx 0.9706
167: \end{equation*}
168: under this assumption, and no others, was derived by Bondi~\cite{Bondi}.  His method of proof is closely related to ours
169: but is not rigorous.
170: 
171: \item \textit{Buchdahl Assumption:}
172: For static stars with constant density, one has the bound
173: \begin{equation*}
174: 2 m / r \leq 8 / 9
175: \end{equation*}
176: derived by Buchdahl \cite{Buch}.  More generally, this bound holds when $p_R= p_T\geq 0$, as long as $\mu\geq 0$ is
177: decreasing; see \cite{Buch}.  The isotropy assumption was relaxed in \cite{GuMu}, where the case $p_R\geq p_T\geq 0$ was
178: treated, still the monotonicity assumption remains crucial.
179: 
180: \item \textit{Dominant Energy Condition:}
181: For almost any reasonable matter model, one expects $|p_R| \leq \mu$ and $|p_T| \leq \mu$.  In the special case that
182: $p_R,p_T\geq 0$, the bound
183: \begin{equation*}
184: 2 m / r \leq 48 / 49 \approx 0.9796
185: \end{equation*}
186: is provided by \cite{Hak}. Our bound for this special case is roughly $0.963$, which we show to be sharp.
187: 
188: \item \textit{Vlasov-Einstein:}
189: For Vlasov-Einstein matter, the stress energy tensor is an integral of those of individual particles, each of which has
190: rank one and satisfies the dominant energy condition.  This implies that $p_R\geq 0$, $p_T \geq 0$ and $p_R + 2p_T \leq
191: \mu$. Under these assumptions, Andr\'easson \cite{Hak} has recently shown that the sharp\footnote{A somewhat unfortunate
192: feature of our argument is that the sharpness of the estimate $\frac{2m}{r} \leq \frac{8}{9}$ is proved only within the class
193: of space-times satisfying the pressure conditions above, without considering whether such space-times arise from solutions of
194: the full Vlasov-Einstein system. Andr\'easson's argument, on the other hand, does provide solutions to the full system.} bound
195: is
196: \begin{equation*}
197: 2 m / r \leq 8 / 9.
198: \end{equation*}
199: Our method provides a new, and considerably shorter, proof of this result.
200: 
201: \item \textit{Zero Radial Pressure:}
202: The case $p_T \geq p_R = 0$ was studied by Florides~\cite{Flo} who obtained the sharp bound
203: \begin{equation*}
204: 2 m / r \leq 2 / 3.
205: \end{equation*}
206: This can also be proved using our method, but the resulting proof is neither shorter nor clearer than the original, so we
207: do not consider this case further.
208: \end{itemize}
209: 
210: \section{Our main result}\label{omr}
211: \begin{theorem}\label{main}
212: Consider a space-time satisfying the regularity, staticity and spherical symmetry conditions described in the
213: introduction. Suppose that the corresponding Hawking mass \eqref{Hm} is finite and that the pressures $p_R,p_T$ and
214: energy density $\mu$ are all non-negative.
215: 
216: \begin{itemize}
217: \item[(1)] \textbf{Vlasov-Einstein case}. Assuming that $p_R +2p_T\leq \mu$, one has
218: \begin{equation}\label{es1}
219: \left( 4 - \frac{6m}{r} + 8\pi r^2 p_R\right)^2 \leq 16\left( 1- \frac{2m}{r} \right),
220: \quad\quad \frac{2m}{r} \leq \frac{8}{9} \,.
221: \end{equation}
222: 
223: \item[(2)] \textbf{Isotropic case}. Assuming that $p_R= p_T$, one has
224: \begin{equation}\label{es2}
225: \left( 2 - \frac{2m}{r} + 8\pi r^2 p_R\right)^2 \leq 36\left( 1- \frac{2m}{r} \right),
226: \quad\quad \frac{2m}{r} \leq 12\sqrt 2-16.
227: \end{equation}
228: 
229: \item[(3)] \textbf{Isotropic case with dominant energy}. Assuming that $p_R = p_T\leq \mu$, one has
230: \begin{equation}\label{es3}
231: \left( 2 - \frac{2m}{r} + 8\pi r^2 p_R\right)^2 \leq 9.551 \left( 1- \frac{2m}{r} \right),
232: \quad\quad \frac{2m}{r} \leq 0.865.
233: \end{equation}
234: 
235: \item[(4)] \textbf{Dominant energy in tangential direction}. Assuming that $p_T \leq \mu$, one has
236: \begin{equation}\label{es4}
237: \left( 6 - \frac{10m}{r} + 8\pi r^2 p_R\right)^2 \leq 40\left( 1- \frac{2m}{r} \right),
238: \quad\quad \frac{2m}{r} \leq \frac{2\sqrt 2+2}{5} \,.
239: \end{equation}
240: 
241: \item[(5)] \textbf{Dominant energy case}. Assuming that $p_R, p_T \leq \mu$, one has
242: \begin{equation}\label{es5}
243: \left( 6 - \frac{10m}{r} + 8\pi r^2 p_R\right)^2 \leq 37.924\left( 1- \frac{2m}{r} \right),
244: \quad\quad \frac{2m}{r} \leq 0.963.
245: \end{equation}
246: \end{itemize}
247: Moreover, these ten estimates are all sharp for the class of space-times considered in each case.  As for the numerical
248: values that appear in \eqref{es3} and \eqref{es5}, these can be described in terms of a system of ODEs which arises in
249: the course of the proof; see \eqref{ode}.  The values given here are accurate up to three decimal places.
250: \end{theorem}
251: 
252: \begin{rem}
253: There is no assumption on the behavior of the space-time as $r$ tends to infinity.  In fact, we do not even assume that
254: $r$ is unbounded. This point is crucial.  It allows us to apply the theorem to the interior of a finite sphere and, in
255: particular, to the interior of the first marginally trapped surface, if there is such a surface.
256: 
257: More precisely, suppose we can prove the theorem in the region where the curvature coordinates are defined, namely in the
258: region where $2 m < r$. For each matter model, we may then deduce that $2m/r \leq c < 1$ for some constant $c$ which
259: depends on the matter model considered. Since $2 m/r$ is continuous and our space-time is connected, this actually
260: implies that $2m/r \leq c$ throughout the space-time. In other words, the marginally trapped surface that we allowed is
261: not, in fact present, and the curvature coordinates, which might \emph{a priori} have covered only part of the
262: space-time, cover the whole space-time. We therefore obtain the full theorem from the special case where the whole
263: space-time is covered by curvature coordinates.  In particular, we may, and do, use curvature coordinates throughout the
264: proof without further comment.
265: \end{rem}
266: 
267: \begin{rem}
268: Our proof for the isotropic case $p_R = p_T$ applies verbatim in the more general case $p_R \geq p_T$, while the
269: estimates \eqref{es2} are sharp for that case as well.
270: \end{rem}
271: 
272: \begin{rem}
273: The \textit{assumptions} of Theorem \ref{main} can be slightly improved in the sense that we do not use our hypothesis
274: $p_T\geq 0$ to establish the given estimates.  This hypothesis is merely included to improve the \textit{conclusions} of
275: Theorem \ref{main}, as sharpness is now shown over a smaller class of space-times. In fact, the examples we construct in
276: order to prove sharpness belong to the even smaller class of space-times which are vacuum outside a sphere.
277: \end{rem}
278: 
279: The proof of Theorem \ref{main} is based on the following elementary fact, which is essentially due to Bondi
280: \cite{Bondi}.
281: 
282: \begin{lemma}\label{cur}
283: Let the assumptions of Theorem \ref{main} hold. Then the variables
284: \begin{equation} \label{c}
285: x \equiv 1-e^{-\al} = \frac{2m}{r} \,,\quad\quad y \equiv -r^2 G_r^r = 8\pi r^2 p_R
286: \end{equation}
287: give rise to a parametric curve which lies in $[0,1)\times [0,\infty)$ and satisfies the equations
288: \begin{align}
289: 8\pi r^2 p_R &= y, \label{cur1} \\
290: 8\pi r^2 p_T &= \frac{x+y}{2(1-x)} \:\dot{x} + \dot{y} + \frac{(x+y)^2}{4(1-x)} \,, \label{cur2} \\
291: 8\pi r^2 \mu &= 2\dot{x} + x, \label{cur3}
292: \end{align}
293: where the dots denote derivatives with respect to $\b= 2\log r$.
294: \end{lemma}
295: 
296: \begin{proof}
297: First of all, we combine equations \eqref{eq3} and \eqref{diag} to write
298: \begin{equation*}
299: 8\pi r^2 \mu = 1 - \d_r (re^{-\al}).
300: \end{equation*}
301: Integrating over $[0,r]$ and using the definition of the Hawking mass \eqref{Hm}, we then get
302: \begin{equation} \label{nd1}
303: \frac{2m}{r} = 1 -e^{-\al} = x.
304: \end{equation}
305: This implies $x\geq 0$ because $m\geq 0$ whenever $\mu\geq 0$.  Next, we use \eqref{diag} to get
306: \begin{equation*}
307: y \equiv - r^2 G_r^r = 8\pi r^2 p_R \geq 0.
308: \end{equation*}
309: To establish our assertion \eqref{cur3}, we combine \eqref{diag}, \eqref{eq6} and \eqref{nd1} to find that
310: \begin{equation*}
311: 8\pi r^2 \mu = r^2 G_t^t = 1 - e^{-\al} (1-2\dot{\al}) = 1 - (1-x) \left( 1- \frac{2\dot{x}}{1-x} \right) = x+2\dot{x}.
312: \end{equation*}
313: To establish our remaining assertion \eqref{cur2}, we first use \eqref{eq4} and \eqref{nd1} to get
314: \begin{equation*}
315: y\equiv -r^2 G_r^r = -1 + (1-x)(1+2\dot{\c}), \quad\quad \dot{\al} = \frac{\dot{x}}{1-x} \,.
316: \end{equation*}
317: Solving the leftmost equation for $\dot{\c}$ and differentiating, we conclude that
318: \begin{equation*}
319: 2\dot{\c} = \frac{x+y}{1-x} \,,\quad\quad 2\ddot{\c} = \frac{(1+y) \dot{x} + (1-x) \dot{y}}{(1-x)^2} \,.
320: \end{equation*}
321: On the other hand, equations \eqref{diag}, \eqref{eq5} and \eqref{nd1} combine to give
322: \begin{equation*}
323: 8\pi r^2 p_T = (1-x)(2\ddot{\c} +\dot{\c}^2 -\dot{\al} - \dot{\al}\dot{\c}).
324: \end{equation*}
325: Using these facts and a simple computation, one may thus easily deduce \eqref{cur2}.
326: \end{proof}
327: 
328: 
329: \section{Proof of Theorem \ref{main}} \label{pro}
330: To prove the desired estimates, we study the curve \eqref{c} provided by Lemma \ref{cur}.  In each case, we are seeking
331: an upper bound for $x= 2m/r$ and also an upper bound for
332: \begin{equation}\label{wn}
333: w_n(x,y) = \frac{(n(1-x)+1+y)^2}{1-x}\,,
334: \end{equation}
335: where the exact value of $n$ varies from case to case. Differentiating \eqref{wn}, we get
336: \begin{equation}\label{wd}
337: \dot{w}_n = \frac{n(1-x)+1+y}{(1-x)^2} \cdot \Bigl[ (1+y-n(1-x)) \dot{x} + 2(1-x) \dot{y} \Bigr]
338: \end{equation}
339: throughout the curve \eqref{c}, where dots denote derivatives with respect to $\b= 2\log r$.  In the special case that
340: $n=1$, this formula reads
341: \begin{equation}\label{w1d}
342: \dot{w}_1 = \frac{2-x+y}{(1-x)^2} \cdot \Bigl[ (x+y) \dot{x} + 2(1-x) \dot{y} \Bigr]
343: \end{equation}
344: and it is closely related to the tangential pressure $p_T$; see \eqref{cur2}.  Let us also recall that
345: \begin{equation*}
346: 0\leq x < 1, \quad\quad y\geq 0
347: \end{equation*}
348: throughout the curve \eqref{c}, a fact we shall frequently need to use in what follows.
349: 
350: \subsection{Vlasov-Einstein case}
351: In this case, we are assuming that $p_R + 2p_T\leq \mu$.  According to Lemma \ref{cur}, the corresponding curve \eqref{c}
352: must thus satisfy
353: \begin{align}\label{1-as}
354: (3x+y-2) \dot{x} + 2(1-x) \dot{y} \leq -\frac{z_3(x,y)}{2}\,,\quad\quad z_3 = 3x^2 - 2x + y^2 + 2y.
355: \end{align}
356: Combining the last equation with our computation \eqref{wd}, we now find
357: \begin{align*}
358: \dot{w}_3 &= \frac{4-3x+y}{(1-x)^2} \cdot \Bigl[ (3x+y-2) \dot{x} + 2(1-x) \dot{y} \Bigr] \\
359: &\leq -\frac{4-3x+y}{2(1-x)^2} \cdot z_3(x,y).
360: \end{align*}
361: In particular, $w_3$ is decreasing whenever $z_3> 0$, so it must be the case that
362: \begin{equation*}
363: w_3 \leq \max_{\substack{0\leq x\leq 1\\ z_3\leq 0\leq y}} w_3(x,y) = w_3(0,0) = 16
364: \end{equation*}
365: throughout the curve.  This proves the first inequality in \eqref{es1}, which also implies the second inequality because
366: the maximum value of $x$ over the region $0\leq x\leq 1$, $y\geq 0$, $w_3\leq 16$ is attained at $(8/9,0)$, namely at the
367: point at which the curve $w_3=16$ intersects the $x$-axis. We refer the reader to Fig. 1 for a sketch of the curves
368: $z_3=0$ and $w_3=16$.
369: 
370: To show that the estimates in \eqref{es1} are sharp, we need to construct a space-time such that the corresponding curve
371: of Lemma \ref{cur} intersects a small neighbourhood of $(8/9,0)$.  Let us now temporarily assume that we have a
372: parametric curve
373: \begin{equation*}
374: x= x(\tau), \quad\quad y= y(\tau); \quad\quad \tau\in (0,\infty)
375: \end{equation*}
376: which passes near the point $(8/9,0)$ and also satisfies the following properties:
377: \begin{itemize}
378: \item[(A1)] $\frac{1}{z_3} \cdot \frac{dw_3}{d\tau}$ is both negative and integrable;
379: \item[(A2)] $0\leq x(\tau) < 1$ and $y(\tau)\geq 0$ for each $\tau > 0$;
380: \item[(A3)] $y(\tau)=0$ for all large enough $\tau$ and $x(\tau)\to 0$ as $\tau\to\infty$;
381: \item[(A4)] the curve is $\mathcal{C}^1$ except for finitely many points.
382: \end{itemize}
383: Given such a curve, we can easily construct a space-time as follows. First, we define
384: \begin{equation}\label{def1}
385: \kappa(\tau) = -\frac{dw_3/d\tau}{z_3(x,y)} \cdot \frac{2(1-x)^2}{4-3x+y}
386: \end{equation}
387: and we note that $\kappa$ is both positive and integrable by (A1)-(A2). Next, we define
388: \begin{equation}\label{def2}
389: \b = \int \kappa\:d\tau, \quad\quad r= \exp(\b/2)
390: \end{equation}
391: and finally, we define the metric coefficients in \eqref{met} by
392: \begin{equation}\label{def3}
393: \al(r) = -\log(1-x), \quad\quad \c(r)= \int \frac{x+y}{2(1-x)} \cdot \kappa\,d\tau.
394: \end{equation}
395: Letting dots denote derivatives with respect to $\b= 2\log r$, as usual, we then get
396: \begin{equation*}
397: \dot{w}_3 = \frac{1}{\kappa} \cdot \frac{dw_3}{d\tau} = -\frac{4-3x+y}{2(1-x)^2} \cdot z_3(x,y)
398: \end{equation*}
399: using our definitions \eqref{def2} and \eqref{def1}.  In view of our computation \eqref{wd}, this gives
400: \begin{equation}\label{1-def}
401: (3x+y-2) \dot{x} + 2(1-x) \dot{y} = -\frac{z_3(x,y)}{2} \,,
402: \end{equation}
403: which is equivalent to the equation $p_R + 2p_T = \mu$ because of Lemma \ref{cur}.
404: 
405: To finish the proof for this case, it thus remains to construct the curve whose existence we assumed in the previous
406: paragraph. We have to ensure that the curve satisfies (A1)-(A4), that it passes arbitrarily close to $(8/9,0)$ and that
407: the corresponding quantities $p_R,p_T, \mu$ provided by Lemma \ref{cur} are non-negative.  Let us then fix some small
408: $\e>0$ and consider the curve
409: \begin{equation}\label{1-A}
410: w_{3-3\e}(x,y) = \Bigl[ \e \sqrt{1+3x} + 4(1-\e) \Bigr]^2.
411: \end{equation}
412: When $\e=0$, this reduces to the curve $w_3=16$ which passes through the origin and $(8/9,0)$.  When $\e=1$, it reduces
413: to the curve $z_3=0$ which passes through the origin and $(2/3,0)$.  In the more general case $0<\e<1$, it describes a
414: curve that lies between these two curves.  We start out at the origin and follow this curve until we hit the $x$-axis,
415: and then we return to the origin along the $x$-axis.  Let us henceforth denote by $C_1$ the curve obtained in this
416: manner; we refer the reader to Fig. 1 for a typical sketch of this curve.
417: 
418: 
419: \begin{figure}[t]
420: \centerline{\psfig{figure=c1.eps,height=6cm}\psfull}
421: \caption{The curve $C_1$ for the Vlasov-Einstein case.}
422: \end{figure}
423: 
424: 
425: The fact that $C_1$ satisfies (A2)-(A4) is trivial.  To check that it satisfies (A1) along the part defined by
426: \eqref{1-A}, we recall that this part lies between the curves $w_3=16$ and $z_3=0$.  Thus, it is easy to see that $z_3>0$
427: along this part, and we need only check that
428: \begin{equation} \label{ch1}
429: \frac{d w_3}{d \tau} < 0
430: \end{equation}
431: as one follows the curve \eqref{1-A} in the positive $x$-direction. Differentiation of \eqref{1-A} gives
432: \begin{align*}
433: \frac{dw_{3-3\e}}{d\tau} &= \frac{3\e \sqrt{w_{3-3\e}}}{\sqrt{1+3x}} \cdot \frac{dx}{d\tau} \\
434: &= \frac{3\e}{\sqrt{1+3x}} \cdot \frac{3(1-\e)(1-x)+1+y}{\sqrt{1-x}} \cdot \frac{dx}{d\tau}
435: \end{align*}
436: along the curve \eqref{1-A}, and we may compare this equation with \eqref{wd} to find that
437: \begin{equation} \label{1-Ad}
438: 2(1-x) \cdot \frac{dy}{d\tau} = \left[ \frac{3\e(1-x)^{3/2}}{\sqrt{1+3x}} + 3(1-\e)(1-x) -1-y \right] \cdot
439: \frac{dx}{d\tau}
440: \end{equation}
441: along the curve \eqref{1-A}.  Employing our computation \eqref{wd} once again, we deduce that
442: \begin{equation*}
443: \frac{dw_3}{d\tau} = \frac{3\e(4-3x+y)}{1-x} \cdot
444: \frac{\sqrt{1-x}-\sqrt{1+3x}}{\sqrt{1+3x}} \cdot \frac{dx}{d\tau} \,.
445: \end{equation*}
446: Since $\frac{dx}{d\tau}>0$ here, the desired \eqref{ch1} follows.  To show that (A1) also holds for the remaining part of
447: the curve $C_1$, we need only note that
448: \begin{equation*}
449: \frac{1}{z_3} \cdot \frac{d w_3}{d \tau} = \frac{1}{x} \cdot \frac{4-3x}{(1-x)^2} \cdot \frac{dx}{d\tau} < 0
450: \end{equation*}
451: along the line $y=0$ because this line is traversed in the direction of decreasing $x$.
452: 
453: Finally, we check that $p_R,p_T,\mu\geq 0$ throughout the curve $C_1$. The fact that $p_R\geq 0$ follows by (A2) because
454: $8\pi r^2 p_R = y$ by definition.  Since \eqref{1-def} ensures that $\mu = p_R + 2p_T$, we need only check that $p_T\geq
455: 0$ as well. Let us now write
456: \begin{align}
457: 8\pi r^2 p_T &= \frac{x+y}{2(1-x)} \:\dot{x} + \dot{y} + \frac{(x+y)^2}{4(1-x)} \notag \\
458: &= \frac{1-x}{2(2-x+y)} \:\dot{w}_1 + \frac{(x+y)^2}{4(1-x)} \label{pT}
459: \end{align}
460: using equations \eqref{cur2} and \eqref{w1d}. Along the part of $C_1$ defined by \eqref{1-A}, we have
461: \begin{equation*}
462: \dot{w}_1 = \frac{2-x+y}{1-x} \cdot \left[ \frac{3\e\sqrt{1-x}}{\sqrt{1+3x}} + 2-3\e \right] \cdot \dot{x}
463: \end{equation*}
464: by \eqref{w1d} and \eqref{1-Ad}.  In view of our definition \eqref{def2}, we thus have
465: \begin{equation*}
466: \dot{w}_1 = \frac{2-x+y}{1-x} \cdot \left[ \frac{3\e\sqrt{1-x}}{\sqrt{1+3x}} + 2-3\e \right]
467: \cdot \frac{dx/d\tau}{\kappa} \,.
468: \end{equation*}
469: Since $\e>0$ is small and $\kappa$ is positive by above, this implies $\dot{w}_1>0$, hence $p_T>0$ by \eqref{pT}.  For
470: the remaining part of $C_1$ along the $x$-axis, Lemma \ref{cur} and \eqref{1-def} give
471: \begin{equation*}
472: p_R= 0, \quad\quad p_T = \frac{x\mu}{4(1-x)} \,,\quad\quad \mu = 2p_T,
473: \end{equation*}
474: so it easily follows that $p_R= p_T = \mu = 0$ throughout this part of the curve.
475: 
476: \subsection{Isotropic case} In this case, our assumption that $p_R =p_T$ is equivalent to
477: \begin{align}\label{2-as}
478: (x+y) \dot{x} + 2(1-x) \dot{y} = - \frac{z_1(x,y)}{2} \,,\quad\quad  z_1 = (x+y)^2 -4y(1-x).
479: \end{align}
480: Proceeding as before, we use our computation \eqref{w1d} to find that
481: \begin{align} \label{2-m}
482: \dot{w}_1 = \frac{2-x+y}{(1-x)^2} \cdot \Bigl[ (x+y) \dot{x} + 2(1-x) \dot{y} \Bigr]= -\frac{2-x+y}{2(1-x)^2}
483: \cdot z_1(x,y).
484: \end{align}
485: Once again, $w_1$ is decreasing as soon as $z_1> 0$, so it must be the case that
486: \begin{equation*}
487: w_1 \leq \max_{\substack{0\leq x\leq 1\\ z_1\leq 0\leq y}} w_1(x,y) = w_1(0,4) = 36
488: \end{equation*}
489: throughout the curve.  This proves the first inequality in \eqref{es2}, while the second inequality follows because the
490: maximum value of $x$ over the region $0\leq x\leq 1$, $y\geq 0$, $w_1\leq 36$ is attained at $(12\sqrt 2-16,0)$.
491: 
492: To show that the estimates in \eqref{es2} are sharp, we argue as in the previous case.  Suppose we have a curve which
493: passes near the point $(12\sqrt 2-16,0)$ and satisfies
494: \begin{itemize}
495: \item[(B1)] $\frac{1}{z_1} \cdot \frac{dw_1}{d\tau}$ is both negative and integrable
496: \end{itemize}
497: as well as (A2)-(A4).  Then we can follow our previous approach with
498: \begin{equation}\label{def4}
499: \kappa(\tau) = -\frac{dw_1/d\tau}{z_1(x,y)} \cdot \frac{2(1-x)^2}{2-x+y} >0
500: \end{equation}
501: instead of \eqref{def1}. Our definitions \eqref{def2}-\eqref{def3} are still applicable, however they now imply
502: \begin{equation} \label{def5}
503: \dot{w}_1 = \frac{1}{\kappa} \cdot \frac{dw_1}{d\tau} = -\frac{2-x+y}{2(1-x)^2} \cdot z_1(x,y).
504: \end{equation}
505: In view of our computation \eqref{w1d}, they thus imply
506: \begin{equation}\label{2-def}
507: (x+y) \dot{x} + 2(1-x) \dot{y} = -\frac{z_1(x,y)}{2} \,,
508: \end{equation}
509: which is equivalent to the equation $p_R = p_T$ because of Lemma \ref{cur}.
510: 
511: To finish the proof for this case, it thus remains to construct the curve whose existence we assumed in the previous
512: paragraph. Fix some small $\e>0$ and set
513: \begin{equation}\label{2-xy}
514: x_\e= \e, \quad\quad y_\e=2-3\e + 2\sqrt{(1-\e)(1-2\e)}
515: \end{equation}
516: for convenience.  Then $(x_\e,y_\e)$ is a point on the curve $z_1=0$ which is close to $(0,4)$. To define the first part
517: of the desired curve, we use the equation
518: \begin{align}\label{2-A}
519: \sqrt{w_1(x,y)} &= \sqrt{w_1(0,0)} + 2A_\e x_\e x - A_\e x^2,
520: \end{align}
521: where $A_\e$ is determined by requiring that the curve passes through $(x_\e, y_\e)$, namely
522: \begin{equation}\label{2-AB}
523: A_\e = \frac{ \sqrt{w_1(x_\e,y_\e)} - 2}{x_\e^2} \,.
524: \end{equation}
525: We start out at the origin and we follow the curve \eqref{2-A} until we reach the point $(x_\e,y_\e)$, then we follow the
526: curve
527: \begin{equation} \label{2-B}
528: \sqrt{w_1(x,y)} = \sqrt{w_1(x_\e,y_\e)} - \frac{\e (x-x_\e)^2}{\sqrt{1-x}} \,,\quad\quad x\geq x_\e
529: \end{equation}
530: until we hit the $x$-axis, and finally we return to the origin along the $x$-axis.  Let $C_2$ denote the curve obtained
531: in this manner; a typical sketch of this curve appears in Fig. 2.
532: 
533: 
534: \begin{figure}[t] \centerline{\psfig{figure=c2.eps,height=6cm}\psfull}
535: \caption{The curve $C_2$ for the isotropic case.}
536: \end{figure}
537: 
538: The fact that $C_2$ satisfies (A2)-(A4) is trivial; we now check that it satisfies (B1).  When it comes to the part of
539: $C_2$ defined by \eqref{2-A}, we have $z_1\leq 0$, $x\leq x_\e$ and
540: \begin{equation}\label{2-Ad}
541: \frac{dw_1/d\tau}{2\sqrt{w_1}} = 2A_\e (x_\e - x) \cdot \frac{dx}{d\tau} \,.
542: \end{equation}
543: Since $x$ is increasing along this part of $C_2$, it thus suffices to check that $A_\e$ is positive.  In view of
544: \eqref{2-AB}, this is certainly the case for all small enough $\e>0$ because
545: \begin{align*}
546: \lim_{\e \to 0} \: \e^2 A_\e &= \sqrt{w_1(0,4)} - 2 = 4.
547: \end{align*}
548: When it comes to the part of $C_2$ defined by \eqref{2-B}, we have $z_1\geq 0$, $x\geq x_\e$ and
549: \begin{equation*}
550: \frac{dw_1/d\tau}{2\sqrt{w_1}} = - \e (x-x_\e) \cdot \frac{4(1-x)+x-x_\e}{2(1-x)^{3/2}} \cdot
551: \frac{dx}{d\tau} \leq 0,
552: \end{equation*}
553: as needed.  When it comes to the remaining part of $C_2$ along the $x$-axis, we have
554: \begin{equation*}
555: \frac{dw_1/d\tau}{z_1} = \frac{2-x}{x(1-x)^2} \cdot \frac{dx}{d\tau} <0,
556: \end{equation*}
557: and this shows that the desired property (B1) holds throughout the curve $C_2$.
558: 
559: Finally, we check that $p_R,p_T,\mu\geq 0$ throughout the curve $C_2$. The fact that $p_R\geq 0$ follows trivially as
560: before, hence $p_T=p_R\geq 0$ by \eqref{2-def} and we need only worry about $\mu$.  Since
561: \begin{equation*}
562: 8\pi r^2 \mu = 2\dot{x} + x
563: \end{equation*}
564: by \eqref{cur3}, we have $\mu\geq 0$ as long as $x$ is increasing along the curve, so we need only check the part of
565: $C_2$ along the $x$-axis.  As in the previous case, however, Lemma \ref{cur} and \eqref{2-def} combine to give $p_R= p_T
566: = \mu = 0$ throughout this part, so the proof for this case is complete.
567: 
568: \subsection{Isotropic case with dominant energy} Our assumption that $p_R = p_T\leq \mu$ gives
569: \begin{equation} \label{3-as}
570: (x+y) \dot{x} + 2(1-x) \dot{y} = - \frac{z_1(x,y)}{2} \,,\quad\quad 2\dot{x} \geq y-x
571: \end{equation}
572: with $z_1$ as in \eqref{2-as}.  Due to the isotropy condition, \eqref{2-m} remains valid, so $w_1$ is increasing if and
573: only if $z_1 \leq 0$.  Since the curve of Lemma \ref{cur} starts out at the origin, where $z_1=0$, it may only attain the
574: largest possible value of $w_1$ at a point along the curve $z_1=0$. It is easy to check that higher values of $w_1$ occur
575: at higher points on this curve.  To attain the largest possible value of $w_1$, the curve of Lemma \ref{cur} must thus
576: ascend as fast as possible within the region $z_1\leq 0$.  Since it starts out at the origin, it must satisfy
577: \begin{equation*}
578: (x+y) \dot{x} + 2(1-x) \dot{y} = - \frac{z_1(x,y)}{2} \,,\quad\quad 2\dot{x} = y-x
579: \end{equation*}
580: until it exits the region $z_1\leq 0$.  This gives rise to the system of ODEs
581: \begin{equation}\label{ode}
582: 2\dot{x} = y-x, \quad\quad 2\dot{y} = \frac{y(2-3x-y)}{1-x}
583: \end{equation}
584: which has a saddle point at the origin.  The solution of interest is the one corresponding to the unstable manifold
585: associated with the origin.  Using numerical integration, we find that it intersects the curve $z_1=0$ at
586: the point $(x_1,y_1) = (0.4927,0.6939)$; see Fig. 3.  This makes %% x = 0.492707181 or 182
587: \begin{equation*}
588: w_1(x_1,y_1) \approx 9.551
589: \end{equation*}
590: the largest possible value of $w_1$, and then we can use the fact that $w_1\leq 9.551$ to deduce that the largest
591: possible value of $x$ is attained at $(0.865,0)$.
592: 
593: To show that our results for this case are sharp, we need to find a curve which passes near the point $(0.865,0)$ and
594: satisfies (B1) as well as (A2)-(A4).  Given such a curve, one can use our approach in the previous case to obtain a
595: space-time for which $p_R = p_T$. We start out at the origin and we follow the solution to the ODE
596: \begin{equation}\label{ode2}
597: \frac{dy}{dx} =\frac{y(2-3x-y)}{(1-x)(y-x)}
598: \end{equation}
599: corresponding to the associated unstable manifold; we do so until we reach the point $(x_1,y_1)$ that lies on the curve
600: $z_1=0$, then we follow the curve
601: \begin{equation} \label{3-B}
602: \sqrt{w_1(x,y)} = \sqrt{w_1(x_1,y_1)} - \frac{\e (x-x_1)^2}{\sqrt{1-x}} \,,\quad\quad x\geq x_1
603: \end{equation}
604: until we hit the $x$-axis, and finally we return to the origin along the $x$-axis. We refer the reader to Fig. 3 for a
605: typical sketch of the curve $C_3$ obtained in this manner.
606: 
607: \begin{figure}[t]
608: \centerline{\psfig{figure=c3.eps,height=6cm}\psfull}
609: \caption{The curve $C_3$ for the isotropic case with dominant energy.}
610: \end{figure}
611: 
612: The only nontrivial properties we need to verify are (B1) and the fact that $p_R\leq \mu$.  When it comes to the part of
613: $C_3$ defined by \eqref{ode2}, we have $p_R= p_T= \mu$ and also
614: \begin{equation} \label{3def}
615: \frac{1}{z_1} \cdot \frac{d w_1}{d \tau} = -\frac{2-x+y}{(1-x)^2(y-x)} \cdot \frac{dx}{d\tau} \leq 0,
616: \end{equation}
617: so the desired properties are easily seen to hold.  The same is true for the part of $C_3$ along the $x$-axis because
618: $p_R= p_T = \mu = 0$ and since
619: \begin{equation*}
620: \frac{1}{z_1} \cdot \frac{d w_1}{d \tau} = \frac{1}{x} \cdot \frac{2-x}{(1-x)^2} \cdot \frac{dx}{d\tau} < 0
621: \end{equation*}
622: along this part.  For the remaining part defined by \eqref{3-B}, we have
623: \begin{equation}\label{3-dw1}
624: \frac{dw_1/d\tau}{2\sqrt{w_1}} = - \e (x-x_1) \cdot
625: \frac{4(1-x)+x-x_1}{2(1-x)^{3/2}} \cdot \frac{dx}{d\tau} \leq 0,
626: \end{equation}
627: which implies property (B1) because $z_1\geq 0$ for this part.  Writing \eqref{def5} in the form
628: \begin{equation*}
629: \dot{w}_1 = -\frac{2-x+y}{2(1-x)^2} \cdot z_1(x,y) = -\frac{\sqrt{w_1(x,y)}}{2(1-x)^{3/2}} \cdot z_1(x,y),
630: \end{equation*}
631: we now combine the last two equations to deduce that
632: \begin{equation*}
633: 2\dot{x} = \frac{z_1(x,y)}{\e(x-x_1)(4-3x-x_1)}
634: \end{equation*}
635: throughout the curve \eqref{3-B}.  According to Lemma \ref{cur}, the condition $p_R\leq \mu$ we need to verify is
636: equivalent to the condition $2\dot x \geq y-x$, so we need to check that
637: \begin{equation} \label{ch2}
638: \frac{z_1(x,y)}{x-x_1} \geq \e(4-3x-x_1) (y-x)
639: \end{equation}
640: throughout the curve \eqref{3-B}.  Write equation \eqref{3-B} in the equivalent form
641: \begin{equation*}
642: y = f(x) \equiv \sqrt{w_1(x_1,y_1)} \sqrt{1-x} + x- 2 - \e (x-x_1)^2.
643: \end{equation*}
644: Then $z_1(x_1,f(x_1))= z_1(x_1,y_1)= 0$ by construction, so one easily finds
645: \begin{align*}
646: \lim_{x\to x_1} \frac{z_1(x,f(x))}{x-x_1} &= 8(x_1+y_1) - 4 - (3x_1+y_1-2)
647: \cdot \frac{\sqrt{w_1(x_1,y_1)}}{\sqrt{1-x_1}} \approx 4.746
648: \end{align*}
649: using \eqref{2-as}.  Thus, the left hand side of \eqref{ch2} is bounded away from zero near $x=x_1$.  Since the same is
650: true away from $x=x_1$, where $z_1$ itself is bounded away from zero, we can always find a small enough $\e>0$ so that
651: \eqref{ch2} holds throughout the curve \eqref{3-B}.
652: 
653: \subsection{Dominant energy in tangential direction} Our assumption that $p_T\leq \mu$ gives
654: \begin{align}\label{4-as}
655: (5x+y-4) \dot{x} + 2(1-x) \dot{y} \leq -\frac{z_5(x,y)}{2}\,,\quad\quad z_5 = (x+y)^2 - 4x(1-x).
656: \end{align}
657: Proceeding as before, we use our computation \eqref{wd} to find that
658: \begin{align} \label{4-m}
659: \dot{w}_5 = \frac{6-5x+y}{(1-x)^2} \cdot \Bigl[ (5x+y-4) \dot{x} + 2(1-x) \dot{y} \Bigr] \leq
660: -\frac{6-5x+y}{2(1-x)^2} \cdot z_5(x,y).
661: \end{align}
662: Once again, $w_5$ is decreasing as soon as $z_5> 0$, so it must be the case that
663: \begin{equation*}
664: w_5 \leq \max_{\substack{0\leq x\leq 1\\ z_5\leq 0\leq y}} w_5(x,y) = w_5(1/10,1/2) = 40
665: \end{equation*}
666: throughout the curve. This proves the first inequality in \eqref{es4}, and the second inequality follows as before.
667: 
668: To show that the estimates in \eqref{es4} are sharp, we need to find a curve which satisfies
669: \begin{itemize}
670: \item[(C1)] $\frac{1}{z_5} \cdot \frac{dw_5}{d\tau}$ is both negative and integrable
671: \end{itemize}
672: as well as (A2)-(A4).  Given such a curve, one can use our previous approach to obtain a space-time for which $p_T =
673: \mu$.  To define the first part of the curve, we use the equation
674: \begin{align} \label{4-A}
675: \sqrt{w_5(x,y)} = \sqrt{w_5(0,0)} + \frac{Ax}{5} - Ax^2,
676: \end{align}
677: where $A$ is chosen so that the curve passes through $(1/10,1/2)$, namely
678: \begin{equation*}
679: A= 100(\sqrt{40} -6) >0.
680: \end{equation*}
681: We start out at the origin and we follow the curve \eqref{4-A} until we reach the point $(1/10,1/2)$, then we follow the
682: curve
683: \begin{equation*}
684: \sqrt{w_5(x,y)} = \sqrt{w_5(1/10,1/2)} - \frac{\e (x-1/10)^2}{\sqrt{1-x}} \,,\quad\quad x\geq 1/10
685: \end{equation*}
686: until we hit the $x$-axis, and finally we return to the origin along the $x$-axis.  Since this curve is almost identical
687: with the one for the isotropic case, our previous approach applies with minor changes; we shall not bother to include the
688: details here.
689: 
690: \subsection{Dominant energy case} In this case, our assumption that $p_R,p_T\leq \mu$ gives
691: \begin{equation} \label{5-as}
692: (5x+y-4) \dot{x} + 2(1-x) \dot{y} \leq -\frac{z_5(x,y)}{2}\,,\quad\quad 2\dot{x} \geq y-x
693: \end{equation}
694: with $z_5$ as in \eqref{4-as}.  Since \eqref{4-m} remains valid, $w_5$ is decreasing when $z_5>0$, so its maximum value
695: is attained in the region $z_5\leq 0$. To obtain the largest possible value of $w_5$, we need to ensure that $\dot{w}_5$
696: is as large as possible in this region. In view of \eqref{4-m}, this simply means that equality must hold in the first
697: inequality in \eqref{5-as}.  We are thus faced with a situation which is almost identical with \eqref{3-as}.  Arguing as
698: before, we find that the curve must satisfy
699: \begin{equation*}
700: 2\dot{x} = y-x, \quad\quad (5x+y-4) \dot{x} + 2(1-x) \dot{y} = -\frac{z_5(x,y)}{2}
701: \end{equation*}
702: until it exits the region $z_5\leq 0$.  This is the same system of ODEs that we had in \eqref{ode}, and the rest of our
703: argument applies almost verbatim.  The solution associated with the unstable manifold at the origin intersects the curve
704: $z_5=0$ at the point $(0.2746,0.6180)$ and so %% x = 0.2746345049 or 050
705: \begin{equation*}
706: w_5(0.2746, 0.6180) \approx 37.924
707: \end{equation*}
708: is the largest possible value of $w_5$. Using this fact, we get the upper bound on $x$ which is stated in the theorem.
709: To show that our results for this case are sharp, we follow our approach in the isotropic case with dominant energy.  As
710: there are only minor changes that need to be made, we are going to omit the details.
711: 
712: \section*{Acknowledgements}
713: We would like to thank Aur\'{e}lien Decelle to whom we are indebted for both the numerical analysis and the figures which
714: appear in this paper.  We would also like to thank Petros Florides for his encouragement and H\aa{}kan Andr\'{e}asson
715: whose recent papers \cite{Hak2, Hak3, Hak} have revived interest in this important problem.
716: 
717: \begin{thebibliography}{10}
718: 
719: \bibitem{Hak2}
720: {\sc H.~Andr{\'e}asson}, {\em On static shells and the {B}uchdahl inequality
721:   for the spherically symmetric {E}instein-{V}lasov system}.
722: \newblock To appear in Comm. Math. Phys.
723: 
724: \bibitem{Hak3}
725: \leavevmode\vrule height 2pt depth -1.6pt width 23pt, {\em On the {B}uchdahl
726:   inequality for spherically symmetric static shells}.
727: \newblock To appear in Comm. Math. Phys.
728: 
729: \bibitem{Hak}
730: \leavevmode\vrule height 2pt depth -1.6pt width 23pt, {\em Sharp bounds on
731:   {$2m/r$} of general spherically symmetric static objects}.
732: \newblock Preprint gr-qc/0702137.
733: 
734: \bibitem{Bondi}
735: {\sc H.~Bondi}, {\em Massive spheres in general relativity}, Proc. R. Soc.
736:   Lond. Ser. A, 282 (1964), pp.~303--317.
737: 
738: \bibitem{Buch}
739: {\sc H.~A. Buchdahl}, {\em General relativistic fluid spheres}, Phys. Rev., 116
740:   (1959), pp.~1027--1034.
741: 
742: \bibitem{Flo}
743: {\sc P.~S. Florides}, {\em A new interior {S}chwarzschild solution}, Proc. R.
744:   Soc. Lond. Ser. A, 337 (1974), pp.~529--535.
745: 
746: \bibitem{GuMu}
747: {\sc J.~Guven and N.~{\'O Murchadha}}, {\em Bounds on {$2m/R$} for static
748:   spherical objects}, Phys. Rev. D, 60 (1999), p.~084020.
749: 
750: \bibitem{Israel}
751: {\sc W.~Israel}, {\em Discontinuities in spherically symmetric gravitational
752:   fields and shells of radiation}, Proc. Roy. Soc. London. Ser. A, 248 (1958),
753:   pp.~404--414.
754: 
755: \bibitem{Lich}
756: {\sc A.~Lichnerowicz}, {\em Th\'eories relativistes de la gravitation et de
757:   l'\'electromagn\'etisme. {R}elativit\'e g\'en\'erale et th\'eories
758:   unitaires}, Masson et Cie, Paris, 1955.
759: 
760: \bibitem{Synge}
761: {\sc J.~L. Synge}, {\em Relativity: {T}he general theory}, Series in Physics,
762:   North-Holland Publishing Co., Amsterdam, 1960.
763: 
764: \end{thebibliography}
765: 
766: \end{document}
767: