1: \chapter{Lorentz spaces} \label{lsp}
2: \section{Introduction}
3: In section \ref{construction}, we constructed the unique, limit space up to isometry (cfr. theorem \ref{four}), which is compact in the strong topology, of a sequence of compact, interpolating cobordisms. The natural question which arises is whether one can find a maximal class of spaces, which is complete and Hausdorff separated in the natural extension of the GGH-uniformity. The answer to this question will be affirmative, and such space will be called a Lorentz space. \\* \\*
4: The main goal of chapter \ref{lsp} is to further study the properties of the moduli
5: space of isometry classes of Lorentz spaces. In particular, in section \ref{GHvGGH}, we try to find out if the GH-metric and the GGH-uniformity are equivalent in the sense that
6: they have the same Cauchy sequences. In section \ref{strD}, we touch upon the question whether the strong
7: metric determines the Lorentz distance uniquely up to time reversal. This
8: would be particulary interesting since, if it were true, then Lorentzian interpolating spacetimes would be a subclass of ``Riemannian'', non-path metric, compact spaces modulo $\mathbb{Z}_2$. Furthermore, in section \ref{cau}, we study the definition of a suitable causal relation and causal curves on limit spaces of compact, globally
9: hyperbolic interpolating spacetimes. For example, if we knew how to define a causal relation
10: between two points in the ``degenerate area'' of a limit space, then one could
11: raise the question about the physical meaning of such ``causal relationships''. Finally, we deal with the moduli space and some matters of precompactness in section \ref{comp}. \\*
12: \\*
13: For notational convenience, in the sequel, the subscript $\mathcal{M}$ will be dropped in the notation of the strong metric $D_{\mathcal{M}}$.
14:
15: \section{Definition of Lorentz spaces}
16:
17: The note following definition \ref{Str} and theorem \ref{four} strongly suggest the following definition of a Lorentz space.
18:
19: \begin{deffie}
20:
21: A Lorentz space is a pair $(\mathcal{M},d)$, where $\mathcal{M}$ is a set and $d$ is a Lorentz distance
22: on $\mathcal{M}$, such that $(\mathcal{M}, D)$ is a compact metric space (with $D$ the strong metric induced by $d$).
23:
24: \end{deffie}
25: On the space of all Lorentz spaces $\aleph_{c}$, we can
26: introduce an equivalence relation $\sim$ by defining $(\mathcal{M}_{1},d_{1})
27: \sim (\mathcal{M}_{2},d_{2})$ iff there exists a bijection $\psi$ such that
28: $d_{2}(\psi(x), \psi(y)) = d_{1}(x,y)$ for all $x,y \in \mathcal{M}_{1}$.
29: Such a bijection is automatically a homeomorphism.
30: \begin{deffie}
31:
32: The moduli space of all isometry classes of Lorentz spaces is the space
33: $\mathcal{LS} = \aleph _{c}/\sim$, equipped with the Hausdorff, quantitative,
34: GGH-uniformity.
35:
36: \end{deffie}
37:
38: \noindent Studying the proofs of theorems \ref{three} and \ref{four} (Ref. \cite{Noldus2}), the reader can see that $\mathcal{LS}$ is a complete, contractible space in which the finite spaces form a dense subset. It is also
39: easily seen that it is not a locally compact space.\\*
40: \\*
41: \textbf{Note}: The results in section \ref{GHvGGH}, in particular
42: theorem \ref{isom}, imply that the obvious extension of $d_{GH}$
43: to the moduli space of isometry classes is also a metric. In the
44: above definition, we prefer to equip this space with the GGH
45: uniformity, since herewith $\mathcal{LS}$ is complete\footnote{The
46: author is unaware of any proof or counterexample of the fact that the moduli
47: space of isometry classes equipped with $d_{GH}$ would be complete.}.
48: $\square$ \\* \\* These results are surprisingly easy and
49: analogous to the ones obtained by Gromov for the metric case. Let
50: me elaborate now a bit why this is so. The main reason is that we
51: made a complete switch from the ``local'' (Alexandrov) viewpoint
52: to the more global perspective provided by the strong metric. The
53: reader shall become even more aware of this after having read
54: section \ref{GHvGGH} in which, amongst other results, all results
55: proven in sections \ref{ld} and \ref{GHD} are generalised.
56:
57: \section{$d_{\rm GH}$ versus the GGH-uniformity} \label{GHvGGH}
58:
59: In this section, we examine the relationship between the GH-distance and the GGH-uniformity for Lorentz spaces $(\mathcal{M},d)$. Along this study, some questions raised in section \ref{prop} will be solved. Since the difference
60: between GH-closeness and GGH-closeness lies in the condition that the mappings
61: used in the definition be approximate inverses of each other, we find it
62: useful to introduce the concepts of $\epsilon$-isometry and $\epsilon$-surjection.
63:
64: \begin{deffie}
65:
66: Let $\epsilon >0$ and $(\mathcal{M},d)$ be a Lorentz space. A mapping $f:\mathcal{M} \rightarrow \mathcal{M}$ is
67:
68: \begin{itemize}
69:
70: \item an $\epsilon$-isometry iff for all $x,y \in\mathcal{M}$
71: $$
72: \left| d(f(x),f(y)) - d(x,y) \right|
73: < \epsilon \;.
74: $$
75: \item an $\epsilon$-surjection iff for all $p \in \mathcal{M}$ there
76: exists a $q \in \mathcal{M}$ such that
77: $$
78: D(p, f(q)) < \epsilon \;.
79: $$
80:
81: \end{itemize} \hfill$\square$
82:
83: \end{deffie}
84: We start with the following theorem.
85: \begin{theo}
86:
87: Let $(\mathcal{M},g)$ be a compact, globally hyperbolic cobordism. Then for any $\eta > 0$, there exists an $\epsilon > 0$ such that for each
88: $\epsilon$-isometry $f$, there
89: exists an isometry $h$ of $\mathcal{M}$ for which the following holds:
90: $$
91: D(f(x), h(x)) < \eta \quad \forall x \in \mathcal{M}\;.
92: $$
93: \end{theo}
94:
95: \noindent\textsl{Proof:}\\*
96:
97: \noindent Suppose that the statement is false. Then there exists an $\eta > 0$ such that
98: for each $n \in \mathbb{N}_{0}$ there exists a $\frac{1}{n}$-isometry $f_{n}$, such
99: that for any isometry $h$ we can find a point $x(n,h)$ in $\mathcal{M}$ such that
100: $$
101: D(f_{n}(x(n,h)), h(x(n,h))) \geq \eta\;.
102: $$
103: The proof of theorem \ref{GHD1} in section \ref{GHD} reveals that we can then find a subsequence $(f_{n_{k}})_{k \in \mathbb{N}}$ and an isometry $f$, such that $f_{n_{k}}
104: \stackrel{k \rightarrow \infty}{\rightarrow} f$ pointwise. We now show that
105: this convergence is uniform in the strong metric, which provides the necessary
106: contradiction. We restrict ourselves to proving that for any interior point
107: $p$ and $\epsilon >0$, there exists a $\delta >0$ such that $q \in B_{D}(p,
108: \delta)$ implies that $D(f(q),f_{n}(q)) < \epsilon$ for $n$ big enough. The
109: rest of the statement is easy (but tedious) and is left as an exercise to the
110: courageous reader. Choose $s,r \in B_{D}(p, \frac{\epsilon}{2})$ such that $s
111: \ll p \ll r$ with, say, $d(s,p)=d(p,r)$ as large as possible. Let $\delta =
112: \frac{1}{3}d(p,r)$; then for $n > \frac{1}{\delta}$ such that $f_{n}(r) \in
113: B_{D}(f(r), \delta)$ and $f_{n}(s) \in B_{D}(f(s), \delta)$ we have that
114: $$
115: f_{n}(B_{D}(p, \delta)) \subset A(f(r),f(s)) \subset B_{D}(f(p),
116: {\textstyle\frac{\epsilon}{2}})\;.
117: $$
118: Hence,
119: $$D(f_{n}(q), f(q)) \leq \frac{\epsilon}{2} + \delta < \epsilon $$
120: for all $q \in B_{D}(p, \delta)$. \hfill$\square$
121: \\*
122: \\*
123: This result reveals that for any $\eta > 0$, there exists an $\epsilon >0$
124: such that any $\epsilon$-isometry is an $\eta$-surjection. This, however, is
125: a fairly weak result and we would like to know if $\eta$ could be bounded by some universal
126: function of $\epsilon$, the timelike diameter and dimension of $\mathcal{M}$,
127: which goes to zero when either $\epsilon$ or the timelike diameter goes to
128: zero. However, the following example shows that such a function cannot exist.
129: \begin{exie} \label{ex11}
130: \end{exie}
131: Consider the 2-dimensional flat cylinder $\mathcal{CYL} = ({\rm S}^1
132: \times [0,1], -d t^{2} + d\theta^2)$ with the region $\mathcal{R}_\delta
133: = \{(\theta,t) \mid \theta \in [0,\pi],\ t > T(\theta)\}$ removed, where $T(\theta) \geq
134: 1 - \delta$ for all $\theta \in [0,\pi]$, $T(0)=T(\pi)=1$ and
135: $\left|T'(\theta)\right| < 1$.
136: \begin{figure}[h]
137: \begin{center}
138: \scalebox{0.53}{\includegraphics{grliph5-1.eps}}
139: \end{center}
140:
141: \caption{Illustration of example \ref{ex11}}
142: \label{fig:grliph5-1}
143: \end{figure}
144: We will construct an approximate isometry $\psi$ which is far from
145: any isometry. Since the only isometry is the identity, this
146: simplifies our analysis. $\psi$ is constructed as the composition
147: of a rotation by $\pi$ times a retraction
148: $R_{\mathcal{R}_{\delta}}$ which maps a point $(\theta,t)$ to the
149: unique, closest point $(\theta,\tilde{t}) \in \mathcal{R}_{\delta}^{c}$.
150: It is not difficult to check that $\psi$ is a $\sqrt{2
151: \delta}$-isometry. However, the point $p$ defined by
152: $p=(\frac{3\pi}{2},1)$ gets mapped to a point which is strong
153: distance $1=tdiam(R_{\delta})$ apart. This shows that for
154: $\delta$ arbitrarily small, one can construct universes which
155: allow $\delta$-isometries to be a distance $1$ apart from any
156: isometry. $\square$
157: \begin{exie} \label{teg}
158: \end{exie}
159: In this example, we show that a near isometry can be arbitrarily far from
160: being a surjection. The following picture shows a sequence of $N$ ``bumps'' with
161: a fixed width $L>1$. Let $0 < \epsilon < \frac{1}{2}$ and consider the
162: function $g_{\epsilon} : \left[ -\epsilon , \epsilon \right] \rightarrow
163: \mathbb{R}^{+} : x \rightarrow \epsilon + x^{2}$. Define a sequence of functions
164: $\Omega^{i}_{\epsilon} : \left[ (i-1)L , iL \right] \rightarrow \mathbb{R}^{+}$, $i=1
165: \ldots N$, which satisfy the following properties:
166: \begin{itemize}
167: \item $ 0 \leq \Omega^{i+1}_{\epsilon}(x+L) -
168: \Omega^{i}_{\epsilon}(x) \leq \frac{L}{\sqrt{2}N}$ for all $x \in
169: \left[ (i-1)L , iL \right]$ and $i:1 \ldots N-1$.
170: \item $\Omega^{i}_{\epsilon}$ is symmetric around $x = (i-\frac{1}{2})L$.
171: \item $\max_{x \in \left[ (i-1)L , iL \right] }
172: \Omega^{i}_{\epsilon}(x) = \frac{iL}{\sqrt{2}N}$
173: \item $\Omega^{i}_{\epsilon} (x) = g_{\epsilon}(x - (i-1)L)$ for $x
174: \in \left[ (i-1)L, (i-1)L + \epsilon \right]$
175: \item $ \left| \frac{d\Omega^{i}_{\epsilon} (x)}{dx} \right| < 1$ for
176: all $x \in \left[ (i-1)L , iL \right]$
177: \end{itemize}
178: Let $\Omega_{\epsilon}$ be the concatenation of all $\Omega^{i}_{\epsilon}$.
179: By identifying $0$ and $NL$, we obtain that $\Omega_{\epsilon}$ is a smooth function on
180: the circle of radius $\frac{NL}{2 \pi}$. Define $\mathcal{A}$ as
181: $$
182: \mathcal{A} = \left\{ (x,t) | x \in \left[0 , NL \right] \textrm{ and }
183: t \in \left[ 0 , \Omega_{\epsilon} (x)\right] \right\}.
184: $$
185: Then, $(\mathcal{A}, -d t^{2} + d x^{2})$ is a globally hyperbolic
186: cobordism cut out of the cylinder universe with radius
187: $\frac{NL}{2 \pi}$. Define $\psi : \mathcal{A} \rightarrow
188: \mathcal{A}$ as the composition of a rotation to the left over an
189: angle of $\frac{2 \pi}{N}$ with a retraction $R_{\mathcal{A}}:
190: S^{1}_{NL} \times \mathbb{R} \rightarrow S^{1}_{NL} \times
191: \mathbb{R}$ which maps every point $(x,t)$ to the closest point
192: $(x,\tilde{t}) \in \mathcal{A}$. Clearly, $\psi$ is a
193: $\frac{L}{\sqrt{N}}$-isometry which is not a
194: $\frac{(N-1)L}{\sqrt{2} N}$-surjection. The figure is called the
195: carousel for obvious reasons. \hfill$\square$
196: \begin{figure}[]
197: \scalebox{0.45}{\includegraphics{grliph5-2.eps}}
198:
199: \caption{The carousel}
200: \label{fig:grliph5-2}
201: \end{figure}
202: \\*
203: \\*
204: Let $f,g : \mathcal{LS} \times \mathcal{LS} \times \mathbb{R}^{+}
205: \rightarrow \mathbb{R}^{+}$ be functions depending only upon the
206: timelike diameters of the respective Lorentz spaces, $f(x,y, 0) = g(x,y,0) = 0$ and $f,g$ are
207: continuous in the third element in $(x,y,0)$ for all $x,y \in
208: \mathcal{LS}$. Do functions satisfying the above conditions exist
209: such that if $x$ and $y$ are $\epsilon$-close, then they are
210: $(f(x,y, \epsilon) , g(x,y, \epsilon))$-close? It is not
211: difficult to prove that if there exist two mappings $\psi, \zeta$
212: which make $x$ and $y$ $\epsilon$-close, such that either $\psi$
213: or $\zeta$ is \emph{surjective}, then $x$ and $y$ are $(\epsilon,
214: 2\epsilon)$ close. Hence, if we want to find a counterexample
215: then we have to look for mappings which are ``far off'' being a
216: surjection. The carousel hints the following counterexample.
217: \begin{exie} \label{teggie}
218: \end{exie}
219: Suppose $L=4m$, $m \in \mathbb{N} \setminus \left\{ 0,1 \right\}$ and let $\mathcal{P}^{L}_{1}, \mathcal{P}^{L}_{2}$ be causal sets given by the Hasse diagrams below. In the pictures, it is understood that the fatter dots are identified. On a locally finite causal set $\mathcal{P}$, the maximum number of links between two timelike related points $p \ll q$ determines a Lorentz distance. Obviously, if $\mathcal{P}$ is finite, then it has a natural interpretation as a Lorentz space. In Appendix E, the following theorem is proven.
220: \begin{theo} \label{conGHvGGH}
221: $\mathcal{P}^{L}_{1}, \mathcal{P}^{L}_{2}$ are $1$-close and for every pair of mappings $\psi : \mathcal{P}^{L}_{1} \rightarrow \mathcal{P}^{L}_{2}$, $\zeta :\mathcal{P}^{L}_{2} \rightarrow \mathcal{P}^{L}_{1}$ which make $\mathcal{P}^{L}_{1}, \mathcal{P}^{L}_{2}$ $k$-close, with $k < L/4$, there exists a $p \in \mathcal{P}^{L}_{2}$ such that $$D(p , \psi \circ \zeta (p) ) = L.$$ $\square$
222: \end{theo}
223: \begin{figure}[h]
224: \begin{center}
225: \setlength{\unitlength}{0.7cm}
226: \begin{picture}(10,10)
227: \thicklines
228: \put(0,0){\line(0,1){1}}
229: \put(0,1){\line(1,-1){1}}
230: \put(1,0){\line(0,1){1}}
231: \put(1,1){\line(-1,-1){1}}
232: \put(1,1){\line(1,-1){1}}
233: \put(2,0){\line(0,1){}}
234: \put(2,1){\line(-1,-1){1}}
235: \put(2,1){\line(1,-1){1}}
236: \put(3,0){\line(0,1){1}}
237: \put(3,1){\line(-1,-1){1}}
238: \put(3,1){\line(1,-1){1}}
239: \put(3,0){\line(1,1){1}}
240: \put(5,0){\ldots}
241: \put(5,1){\ldots}
242: \put(7,0){\line(-1,1){1}}
243: \put(7,1){\line(-1,-1){1}}
244: \put(7,0){\line(1,1){1}}
245: \put(7,1){\line(1,-1){1}}
246: \put(8,0){\line(0,1){1}}
247: \put(8,0){\line(1,1){1}}
248: \put(8,1){\line(1,-1){1}}
249: \put(7,0){\line(0,1){1}}
250: \put(8,0){\line(0,1){1}}
251: \put(9,0){\line(0,1){1}}
252: \put(0,1){\line(0,1){1}}
253: \put(0,2){\line(0,1){1}}
254: \put(0,4){\vdots}
255: \put(0,5){\line(0,1){1}}
256: \put(0,6){\line(0,1){1}}
257: \put(0,7){\line(0,1){1}}
258: \put(0,8){\line(0,1){1}}
259: \put(1,1){\line(0,1){1}}
260: \put(1,2){\line(0,1){1}}
261: \put(1,4){\vdots}
262: \put(1,5){\line(0,1){1}}
263: \put(1,6){\line(0,1){1}}
264: \put(1,7){\line(0,1){1}}
265: \put(2,1){\line(0,1){1}}
266: \put(2,2){\line(0,1){1}}
267: \put(2,4){\vdots}
268: \put(2,5){\line(0,1){1}}
269: \put(2,6){\line(0,1){1}}
270: \put(3,1){\line(0,1){1}}
271: \put(3,2){\line(0,1){1}}
272: \put(3,4){\vdots}
273: \put(3,5){\line(0,1){1}}
274: \put(4,1){\line(0,1){1}}
275: \put(4,2){\line(0,1){1}}
276: \put(7,1){\line(0,1){1}}
277: \put(7,2){\line(0,1){1}}
278: \put(7,3){\line(0,1){1}}
279: \put(8,1){\line(0,1){1}}
280: \put(8,2){\line(0,1){1}}
281: \put(9,1){\line(0,1){1}}
282: \put(4,4){\vdots}
283: \put(0,0){\circle*{0.2}}
284: \put(0,1){\circle*{0.3}}
285: \put(1,1){\circle*{0.1}}
286: \put(1,0){\circle*{0.1}}
287: \put(2,0){\circle*{0.1}}
288: \put(2,1){\circle*{0.1}}
289: \put(3,0){\circle*{0.1}}
290: \put(3,1){\circle*{0.1}}
291: \put(7,0){\circle*{0.1}}
292: \put(7,1){\circle*{0.1}}
293: \put(8,0){\circle*{0.1}}
294: \put(8,1){\circle*{0.1}}
295: \put(9,0){\circle*{0.1}}
296: \put(9,1){\circle*{0.1}}
297: \put(0,2){\circle*{0.1}}
298: \put(0,3){\circle*{0.1}}
299: \put(0,5){\circle*{0.1}}
300: \put(0,6){\circle*{0.1}}
301: \put(0,7){\circle*{0.1}}
302: \put(0,8){\circle*{0.1}}
303: \put(0,9){\circle*{0.1}}
304: \put(1,2){\circle*{0.1}}
305: \put(1,3){\circle*{0.1}}
306: \put(1,5){\circle*{0.1}}
307: \put(1,6){\circle*{0.1}}
308: \put(1,7){\circle*{0.1}}
309: \put(1,8){\circle*{0.1}}
310: \put(2,2){\circle*{0.1}}
311: \put(2,3){\circle*{0.1}}
312: \put(2,5){\circle*{0.1}}
313: \put(2,6){\circle*{0.1}}
314: \put(2,7){\circle*{0.1}}
315: \put(3,2){\circle*{0.1}}
316: \put(3,3){\circle*{0.1}}
317: \put(3,5){\circle*{0.1}}
318: \put(3,6){\circle*{0.1}}
319: \put(7,2){\circle*{0.1}}
320: \put(7,3){\circle*{0.1}}
321: \put(7,4){\circle*{0.1}}
322: \put(8,2){\circle*{0.1}}
323: \put(8,3){\circle*{0.1}}
324: \put(9,2){\circle*{0.1}}
325: \put(9,1){\line(1,-1){1}}
326: \put(9,0){\line(1,1){1}}
327: \put(10,1){\circle*{0.3}}
328: \put(10,0){\circle*{0.2}}
329: \put(7,8){{\Large $ \mathcal{P}_2^{L}$}}
330: \put(0,10){$L+1$}
331: \put(1,9){$L$}
332: \put(2,8){$L-1$}
333: \put(3,7){$L-2$}
334: \put(7,5){$4$}
335: \put(8,4){$3$}
336: \put(9,3){$2$}
337: \end{picture}
338: \caption{Example \ref{teggie}, Lorentz spaces which are GH but not GGH close.}
339: \label{figure3}
340: \end{center}
341:
342: \end{figure}
343: \begin{figure}[h]
344: \begin{center}
345: \setlength{\unitlength}{0.7cm}
346: \begin{picture}(10,13)
347: \thicklines
348: \put(0,0){\line(0,1){1}}
349: \put(0,1){\line(1,-1){1}}
350: \put(1,0){\line(0,1){1}}
351: \put(1,1){\line(-1,-1){1}}
352: \put(1,1){\line(1,-1){1}}
353: \put(2,0){\line(0,1){}}
354: \put(2,1){\line(-1,-1){1}}
355: \put(2,1){\line(1,-1){1}}
356: \put(3,0){\line(0,1){1}}
357: \put(3,1){\line(-1,-1){1}}
358: \put(3,1){\line(1,-1){1}}
359: \put(3,0){\line(1,1){1}}
360: \put(5,0){\ldots}
361: \put(5,1){\ldots}
362: \put(7,0){\line(-1,1){1}}
363: \put(7,1){\line(-1,-1){1}}
364: \put(7,0){\line(1,1){1}}
365: \put(7,1){\line(1,-1){1}}
366: \put(8,0){\line(0,1){1}}
367: \put(8,0){\line(1,1){1}}
368: \put(8,1){\line(1,-1){1}}
369: \put(7,0){\line(0,1){1}}
370: \put(8,0){\line(0,1){1}}
371: \put(9,0){\line(0,1){1}}
372: \put(0,1){\line(0,1){1}}
373: \put(0,2){\line(0,1){1}}
374: \put(0,4){\vdots}
375: \put(0,5){\line(0,1){1}}
376: \put(1,1){\line(0,1){1}}
377: \put(1,2){\line(0,1){1}}
378: \put(1,4){\vdots}
379: \put(1,5){\line(0,1){1}}
380: \put(1,7){\vdots}
381: \put(1,8){\line(0,1){1}}
382: \put(1,9){\line(0,1){1}}
383: \put(1,10){\line(0,1){1}}
384: \put(1,11){\line(0,1){1}}
385: \put(2,1){\line(0,1){1}}
386: \put(2,2){\line(0,1){1}}
387: \put(2,4){\vdots}
388: \put(2,5){\line(0,1){1}}
389: \put(2,7){\vdots}
390: \put(2,8){\line(0,1){1}}
391: \put(2,9){\line(0,1){1}}
392: \put(3,1){\line(0,1){1}}
393: \put(3,2){\line(0,1){1}}
394: \put(3,4){\vdots}
395: \put(3,5){\line(0,1){1}}
396: \put(3,7){\vdots}
397: \put(3,8){\line(0,1){1}}
398: \put(4,1){\line(0,1){1}}
399: \put(4,2){\line(0,1){1}}
400: \put(4,4){\vdots}
401: \put(7,1){\line(0,1){1}}
402: \put(7,2){\line(0,1){1}}
403: \put(7,3){\line(0,1){1}}
404: \put(8,1){\line(0,1){1}}
405: \put(8,2){\line(0,1){1}}
406: \put(9,1){\line(0,1){1}}
407: \put(4,4){\vdots}
408: \put(0,0){\circle*{0.2}}
409: \put(0,1){\circle*{0.3}}
410: \put(1,1){\circle*{0.1}}
411: \put(1,0){\circle*{0.1}}
412: \put(2,0){\circle*{0.1}}
413: \put(2,1){\circle*{0.1}}
414: \put(3,0){\circle*{0.1}}
415: \put(3,1){\circle*{0.1}}
416: \put(7,0){\circle*{0.1}}
417: \put(7,1){\circle*{0.1}}
418: \put(8,0){\circle*{0.1}}
419: \put(8,1){\circle*{0.1}}
420: \put(9,0){\circle*{0.1}}
421: \put(9,1){\circle*{0.1}}
422: \put(0,2){\circle*{0.1}}
423: \put(0,3){\circle*{0.1}}
424: \put(0,5){\circle*{0.1}}
425: \put(0,6){\circle*{0.1}}
426: \put(1,2){\circle*{0.1}}
427: \put(1,3){\circle*{0.1}}
428: \put(1,5){\circle*{0.1}}
429: \put(1,6){\circle*{0.1}}
430: \put(1,8){\circle*{0.1}}
431: \put(1,9){\circle*{0.1}}
432: \put(1,10){\circle*{0.1}}
433: \put(1,11){\circle*{0.1}}
434: \put(1,12){\circle*{0.1}}
435: \put(2,2){\circle*{0.1}}
436: \put(2,3){\circle*{0.1}}
437: \put(2,5){\circle*{0.1}}
438: \put(2,6){\circle*{0.1}}
439: \put(2,8){\circle*{0.1}}
440: \put(2,9){\circle*{0.1}}
441: \put(2,10){\circle*{0.1}}
442: \put(3,2){\circle*{0.1}}
443: \put(3,3){\circle*{0.1}}
444: \put(3,5){\circle*{0.1}}
445: \put(3,6){\circle*{0.1}}
446: \put(7,2){\circle*{0.1}}
447: \put(7,3){\circle*{0.1}}
448: \put(7,4){\circle*{0.1}}
449: \put(8,2){\circle*{0.1}}
450: \put(8,3){\circle*{0.1}}
451: \put(9,2){\circle*{0.1}}
452: \put(9,1){\line(1,-1){1}}
453: \put(9,0){\line(1,1){1}}
454: \put(10,1){\circle*{0.3}}
455: \put(10,0){\circle*{0.2}}
456: \put(3,8){\circle*{0.1}}
457: \put(3,9){\circle*{0.1}}
458: \put(4,5){\line(0,1){1}}
459: \put(4,5){\circle*{0.1}}
460: \put(4,6){\circle*{0.1}}
461: \put(4,7){$L-3$}
462: \put(7,8){{\Large $ \mathcal{P}_1^{L}$}}
463: \put(0,7){$\frac{L}{2}$}
464: \put(1,13){$L+1$}
465: \put(2,11){$L-1$}
466: \put(3,10){$L-2$}
467: \put(7,5){$4$}
468: \put(8,4){$3$}
469: \put(9,3){$2$}
470: \end{picture}
471: \caption{Example \ref{teggie}, Lorentz spaces which are GH but not GGH close.}
472: \label{figure4}
473: \end{center}
474: \end{figure}
475: Choose $\alpha > 0$ and let $\epsilon_{m} = \frac{4\alpha}{4m + 1}$, $L_{m} = 4m$, $m \in \mathbb{N} \setminus \left\{ 0,1 \right\}$. Define for $i=1,2$ the discrete Lorentz spaces $\mathcal{P}_{i}^{\epsilon_{m}, L_{m}}$ by the same Hasse diagrams, but now suppose that every link has length $\epsilon_{m}$. The previous theorem teaches us that $\mathcal{P}_{1}^{\epsilon_{m},L_{m}}$ and $\mathcal{P}_{2}^{\epsilon_{m},L_{m}}$ are $\epsilon_{m}$-close, but for every pair of mappings $\psi_{m} : \mathcal{P}^{\epsilon_{m},L_{m}}_{1} \rightarrow \mathcal{P}^{\epsilon_{m},L_{m}}_{2}$, $\zeta_{m} :\mathcal{P}^{\epsilon_{m},L_{m}}_{2} \rightarrow \mathcal{P}^{\epsilon_{m},L_{m}}_{1}$ which make $\mathcal{P}^{\epsilon_{m},L_{m}}_{1}, \mathcal{P}^{\epsilon_{m},L_{m}}_{2}$ $\epsilon$- close, with $\epsilon < \frac{4m\alpha}{4m+1}$, there exists a $p \in \mathcal{P}^{\epsilon_{m},L_{m}}_{2}$ such that $$D_{m}(p , \psi_{m} \circ \zeta_{m} (p) ) = \frac{16m\alpha}{4m+1}.$$
476: However, $tdiam(\mathcal{P}_{i}^{\epsilon_{m}, L_{m}}) = 4 \alpha$ for $i=1,2$ and $m >1$. This proves the claim that universal functions $f$ and $g$ satisfying the above conditions do not exist. $\square$
477: \\*
478: \\*
479: Hence, the qualitative uniformities defined by GH and GGH are inequivalent. However, this does not prove yet that there exist GH Cauchy sequences which are not GGH. In fact, we show now that if a Lorentz space $(\mathcal{M},d)$ is a limit space of a GH Cauchy sequence $(\mathcal{M}_{i},d_{i})_{i \in \mathbb{N}}$, then this sequence is GGH Cauchy and converges to the same limit (up to isometry). An intermediate result is the following.
480: \begin{theo} \label{bij}
481: Any isometry $\psi$ on a Lorentz space $(\mathcal{M},d)$ is a bijection.
482: \end{theo}
483: \textsl{Proof}: \\*
484: Evidently, $D(\psi(p), \psi(q)) \geq D(p,q)$ for all $p,q \in \mathcal{M}$. Hence, we only have to show that $\psi$ is a surjection since, obviously, it is an injection. Suppose we can find an open ball $B_{D}(r, \epsilon)$ which is not in $\psi(\mathcal{M})$, then $\psi^{k}(r) \notin B_{D}(\psi^{l}(r), \epsilon)$ for all $k > l$. Since $\mathcal{M}$ is compact, we may, by passing to a subsequence if necessary, assume that $\psi^{l}(r) \stackrel{l \rightarrow \infty}{\rightarrow} \psi^{\infty}(r)$. Hence, we arrive at the contradiction that $\psi^{\infty}(r) \notin B_{D}(\psi^{\infty}(r), \epsilon)$. But then we have that $$D(\psi(p), \psi(q)) = \sup_{r \in \mathcal{M}} \left| d(\psi(r),\psi(p)) + d(\psi(p),\psi(r)) - d(\psi(r), \psi(q)) - d(\psi(q) , \psi(r)) \right| $$
485: The rhs. of this equation equals $D(p,q)$. This shows that $\psi$ is surjective since all isometries of compact metric spaces are. $\square$ \\*
486: \\*
487: Before we prove the main result, we still need the following.
488: \begin{theo} \label{cov}
489: Let $\left\{\psi_{i} | i \in \mathbb{N}_{0} \right\}$ be a set of $\frac{1}{i}$-isometries on $\mathcal{M}$. Then, there exists a subsequence $(\psi_{i_{n}})_{n \in \mathbb{N}}$ which uniformly converges in the strong sense to an isometry $\psi$.
490: \end{theo}
491: \textsl{Proof}: \\*
492: As usual, let $\mathcal{C}$ be a countable dense subset of $\mathcal{M}$ and let $(\psi_{i_{n}})_{n \in \mathbb{N}}$ be a subsequence such that $\psi_{i_{n}}(p) \stackrel{n \rightarrow \infty}{\rightarrow} \psi(p)$ for all $p \in \mathcal{C}$. It is easy to see that $\psi$ has a unique extension to a $D$-isometry (and $d$-isometry) using Theorem \ref{bij}. The proof of Theorem \ref{bij} also implies that $\psi(\mathcal{C})$ is dense in $\mathcal{M}$. As a consequence, we have that for any $\epsilon > 0$ there exists a $k(\epsilon) > 0$ such that $\psi_{i_{k}} ( \mathcal{C} )$ is $\epsilon$-dense in $\mathcal{M}$ for $k > k(\epsilon)$. Hence, $$\left| D(\psi_{i_{k}}(p), \psi_{i_{k}}(q)) - D(p,q) \right| < \epsilon + \frac{2}{i_{k}}$$ for $k > k(\epsilon)$ and for all $p,q \in \mathcal{M}$. This implies that
493: $$ D(\psi(r), \psi_{i_{k}}(r)) \leq D(\psi(p) , \psi_{i_{k}}(p)) + 2D(p,r) + \frac{2}{i_{k}} + \epsilon $$ for $k > k(\epsilon)$ and $p \in \mathcal{C}$. Since $\epsilon$ and $p$ can independently be chosen arbitrarly close to $0$ and $r$ respectively, the result follows. $\square$ \\*
494: \\*
495: We are now in position to prove the main result.
496: \begin{theo} \label{isom}
497: Let $(\mathcal{M}_{i},d_{i})_{i \in \mathbb{N}}$ be a GH Cauchy sequence of Lorentz spaces converging to a Lorentz space $(\mathcal{M},d)$, then this sequence is GGH Cauchy and converges to the same limit space.
498: \end{theo}
499: \textsl{Proof}: \\*
500: Choose $\delta > 0$, then Theorem \ref{cov} implies that there exists a $\gamma >0$, such that if $f$ is a $\gamma$-isometry, then there exists an isometry $g$ such that $$D(f(x),g(x)) < \frac{\delta}{2} \, \forall x \in \mathcal{M}$$
501: Let $\psi_{i} : \mathcal{M}_{i} \rightarrow \mathcal{M}$ and $\zeta_{i} : \mathcal{M} \rightarrow \mathcal{M}_{i}$ which make $(\mathcal{M}_{i},d_{i})$ and $(\mathcal{M},d)$ $\epsilon_{i}$-close, where $\epsilon_{i} \stackrel{i \rightarrow \infty}{\rightarrow} 0$. Then, the previous remark implies that for $i$ sufficiently large such that $2 \epsilon_{i} < \min \left\{\gamma , \frac{\delta}{2} \right\}$, there exists an isometry $\beta_{i}$ such that
502: $$D(\beta_{i}(x), \psi_{i} \circ \zeta_{i} (x)) < \frac{\delta}{2} \, \forall x \in \mathcal{M}$$
503: or,
504: $$D(x, \psi_{i} \circ \zeta_{i} \circ \beta_{i}^{-1}(x)) < \frac{\delta}{2} \, \forall x \in \mathcal{M}.$$
505: Hence,
506: $$ D_{i}(p_{i} , \zeta_{i} \circ \beta_{i}^{-1} \circ \psi_{i} ( p_{i})) \leq 2 \epsilon_{i} + D(\psi_{i}(p_{i}), \psi_{i} \circ \zeta_{i} \circ \beta_{i}^{-1} \circ \psi_{i} ( p_{i})) $$
507: which implies that
508: $$ D_{i}(p_{i} , \zeta_{i} \circ \beta_{i}^{-1} \circ \psi_{i} ( p_{i})) \leq 2 \epsilon_{i} + \frac{\delta}{2} < \delta $$
509: Hence, for $i$ sufficiently large, $\psi_{i}$ and $\zeta_{i} \circ \beta_{i}^{-1}$ make $(\mathcal{M}_{i},d_{i})$ and $(\mathcal{M},d)$, $(\epsilon_{i}, \delta)$-close. $\square$ \\*
510: \\*
511: This does not show yet that every GH Cauchy sequence is GGH, but a GH Cauchy sequence which is not GGH has either no sensible limit, or a limit which is not ``spatially compact''. The last theorem has for consequence that the trivial extension of $d_{GH}$ to the moduli space of isometry classes of Lorentz spaces is a metric.
512: \section{The strong metric D} \label{strD}
513: In this section, we study some properties of the strong metric. Particular questions of interest are:
514: \begin{itemize}
515: \item What is the ``shape'' of the balls in the strong metric for spacetimes?
516: \item Does the strong metric determine the Lorentz metric up to time reversal?
517: \end{itemize}
518: We shall treat the first question in considerable detail, the second one is only answered partially. \\*
519: \\*
520: In what follows, $\mathcal{M}$ is assumed to be a compact, interpolating spacetime. To start with, we ``split'' the strong metric $D$ into two
521: pseudodistances $D^\pm$, which will be useful later on, and then study
522: properties of the open balls $B_D(p,\epsilon)$ of $D$-radius $\epsilon$
523: around $p$. We start by defining
524: $$
525: D^+( p,q) = \max_{r \in \mathcal{M}}
526: \left| d_{g}(p,r) - d_{g}(q,r) \right|
527: $$
528: and
529: $$
530: D^-( p,q)
531: = \max_{r \in \mathcal{M}} \left| d_{g}(r,p) - d_{g}(r,q) \right|.
532: $$
533: Then $D$ can be recovered from $D^\pm$ as \cite{Noldus2}
534: $$
535: D(p,q) = \max \left\{ D^+(p,q) , D^-(p,q) \right\},
536: $$
537: although, separately, $D^+$ and $D^-$ are \emph{pseudo} metrics ($D^{\pm}(p,q)
538: = 0$ does not necessarily imply that $p=q$). However, this limitation of $D^+$
539: ($D^-$) arises only for $p$ and $q$ both belonging to the future (past)
540: boundary of $\mathcal{M}$. For example, clearly $D^+(p,q) = 0$ for all $p,q \in
541: \partial_{\rm F}\mathcal{M}$, but if $p \not\in \partial_{\rm F}\mathcal{M}$ and $q\in
542: \partial_{\rm F}\mathcal{M}$, any $r \in I^+(p)$ gives $d_g(p,r) = |d_g(p,r)-d_g(q,r)|
543: > 0$, and if both $p,q\not\in \partial_{\rm F}\mathcal{M}$, the same holds for any
544: $r \in I^+(p)\,\triangle\,I^+(q)$\footnote{For any two sets $A$
545: and $B$, $A\,\triangle\,B$ stands for the symmetric difference $(A
546: \setminus B) \cup (B \setminus A)$.}. \\* \\* These remarks show
547: that both $D^\pm$ are true distances on the interior of
548: $\mathcal{M}$, and they also motivate us to try to locate the
549: ``distance-maximising points'', i.e., points which realise the
550: maximum in the definition of both functions for given $p$ and
551: $q$. \\* \\*
552:
553: \noindent\textbf{Property}: {\em Given any two points $p$ and $q$ not both
554: belonging to $\partial_{\rm F} \mathcal{M}$, a point $r$ such that $D^+(p,q)
555: = |d_{g}( p,r) - d_{g}(q,r)|$ is an element of $I^{+} (p) \,\triangle\,
556: I^{+}(q)$, and $I^+(r)
557: \subset I^+(p) \cap I^+(q)$. A dual property holds for $D^-$.}\\*
558: \\*
559: \noindent\textsl{Proof:}\\* \noindent Obviously, the distance-maximising point $r$ belongs to $I^+(p)
560: \cup I^+(q)$. Suppose $r \in I^{+}(p) \cap I^{+}(q)$ and, without
561: loss of generality, assume that $d_{g}(p,r) > d_{g}(q,r)$. Let
562: $\gamma$ be a distance maximising geodesic from $p$ to $r$; then
563: $\gamma$ cuts $E^{+}(q)$ in a point $s$. But then, the reverse
564: triangle inequality implies that
565: \begin{eqnarray*}
566: d_g(p,r) - d_g(q,r) &=& d_g(p,s) + d_g(s,r) - d_g(q,r) \\
567: &<& d_g(p,s)\ \ =\ \ d_g(p,s) - d_g(q,s)\;,
568: \end{eqnarray*}
569: which is a contradiction. Hence, $r \in I^{+}(p)\,\triangle\,
570: I^{+}(q)$ and without loss of generality we may assume that $r \in I^+(p)
571: \setminus I^+(q)$, which means that $p \notin \partial_{F} \mathcal{M}$. Either $r \in \partial_{F} \mathcal{M}$ which implies that $I^{+}(r) = \emptyset$ or $I^{+}(r) \neq \emptyset$. The latter implies that $I^{+}(r) \subset I^{+}(p) \cap I^{+}(q)$ since otherwise there exists a point $s$ such that $$d_g(p,r) < d_g(p,s) = d_g(p,s) - d_g(q,s) $$ which is a contradiction. \hfill$\square$\\*
572: \\*
573: Now, if $(\mathcal{M},g)$ contains no cut points\footnote{The definition given in \cite{Beem} is more general but reduces to the following in the globally hyperbolic case: a future oriented causal geodesic $\gamma$ starting at $p$ ($\gamma(0) = p$) has a future cut point $\gamma(t_{0})$ iff $\gamma$ is distance maximising between $p$ and $\gamma(t_{0})$, i.e., $d_{g}(\gamma(s),\gamma(t)) = L(\gamma_{\left[s,t\right]})$ for all $0 \leq t < s \leq t_{0}$ and $t_{0}$ is the largest (affine) parameter with this property; a past cut point for a (past oriented) geodesic is defined similarly. A Lorentzian equivalent of an earlier result by Poincar\'e shows that a causal geodesic with initial endpoint $p$ has a cut point $\gamma(t_{0})$ iff there exists a second causal geodesic starting at $p$ which contains $\gamma(t_{0})$ or if $\gamma(t_{0})$ is \emph{conjugate} point for $\gamma$ (i.e., there exists a Jacobi field along $\gamma$ which vanishes in $p$ and $\gamma(t_{0})$). A spacetime has no cut points iff any causal geodesic with initial (final) endpoint contains no future (past) cut points.}, then the distance-maximising point $r$ must
574: belong to $\partial_{\rm F} \mathcal{M}$. Suppose that $r$ does not belong to
575: $\partial_{\rm F}\mathcal{M}$, then $I^{+}(r) \subset I^{+}(p)
576: \cap I^{+} (q)$ implies that $r$ belongs to $E^{+}(q)$. Let
577: $\gamma$ be the unique null geodesic from $q$ to $r$, then moving
578: $r$ to the future along this null geodesic up to $\partial_{\rm F}
579: \mathcal{M}$ keeps $r$ out of $I^{+}(q)$, otherwise the geodesic
580: would have a cut point, which is contrary to the assumption. \\* \\*
581: The next theorem is also valid when there are cut points, but then the
582: statement can be made sharper. As was remarked before in theorem \ref{amal}, the $\epsilon$-balls $B_D(p,\epsilon)$
583: are causally convex, in the sense that if $x,y \in B_D(p,\epsilon)$, then the Alexandrov set
584: $A(x,y) \subset B_D(p,\epsilon)$. We now wish to find out more about those
585: sets. To begin with, notice that
586: $$
587: B_D(p,\epsilon) = B_{D^+}(p,\epsilon) \cap B_{D^-}(p,\epsilon)\;.
588: $$
589: Then, we have:
590:
591: \begin{theo} \label{ball}
592:
593: Let $(\mathcal{M},g)$ be a spacetime with no cut points and choose a point
594: $p \in \mathcal{M} \setminus \partial_{\rm F} \mathcal{M}$ and an $\epsilon >
595: 0$ such that $K^{+} ( p , \epsilon ) \neq \emptyset$. Then the open ``sphere''
596: $B_{D^+} (p,\epsilon)$ of radius $\epsilon$, centered at $p$ with respect to
597: the pseudometric $D^{+}$, satisfies
598: $$
599: B_{D^+} (p , \epsilon) \subseteq \left[ \bigcap_{x \in \mathcal{H}^{+}
600: (p)} \left( \mathcal{O}^{-} (x, \epsilon ) \right)^{\rm c} \right]
601: \bigcap \left[
602: \bigcap_{x \in \mathcal{F}^{+} (p , \epsilon )} I^{-} (x) \right],
603: $$
604: where
605: \begin{itemize}
606:
607: \item $K^{+} (x , \epsilon) = \left\{ y \in \mathcal{M} \mid d_{g}
608: (x,y) = \epsilon \right\}$, i.e. the future $\epsilon$-sphere centered at $x$,
609:
610: \item $\mathcal{O}^{-} (x , \epsilon) = \left\{ y \in \mathcal{M}
611: \mid d_{g} (y,x) \geq \epsilon \right\}$, i.e. the closed outer past
612: $\epsilon$-ball around $x$,
613:
614: \item $\mathcal{H}^{+}(p) = E^{+}(p) \cap \partial_{\rm F} \mathcal{M}$,
615:
616: \item $\mathcal{F}^{+}(p , \epsilon) = K^{+}( p ,\epsilon) \cap
617: \partial_{\rm F} \mathcal{M}$.
618:
619: \end{itemize}
620:
621: \noindent The open sphere $B_{D^-} ( p, \epsilon )$ defined with respect to
622: the pseudometric $D^{-}$, satisfies a similar inclusion property with all
623: pasts and futures interchanged.
624: \end{theo}
625:
626: \noindent\textsl{Proof:}\\*
627: \noindent Let $x \in B_{D^+} (p , \epsilon)$; then $x$ must be chronologically
628: connected to all points in $\mathcal{F}^{+}(p , \epsilon)$. For, suppose there
629: exists a point $y \in \mathcal{F}^{+}(p , \epsilon)$ such that $x \notin
630: I^{-}(y)$ then $d_{g}(p,y) - d_{g}(x,y) = \epsilon$, which is a contradiction.
631: On the other hand, $x$ cannot belong to $\mathcal{O}^-(z,\epsilon)$ for any $z
632: \in \mathcal{H}^+(p)$, since otherwise
633: $$
634: d_{g} ( x , z ) - d_{g}(p,z) \geq \epsilon\;,
635: $$
636: which is impossible. \hfill$\square$
637: \\*
638:
639: \noindent Figure \ref{figball} shows that the above inclusion can be an
640: equality. The universe is $(S^{1} \times \left[ 0 , 1\right], - dt^{2} +
641: d\theta^{2})$ and the shaded area represents $B_{D^+} (p, \epsilon)$ for
642: $\epsilon$ sufficiently small.
643: \begin{figure}[h]
644:
645: \begin{center}
646:
647: \scalebox{0.50}{\includegraphics{grliph5-3.eps}}
648:
649: \end{center}
650:
651: \caption{Illustration of theorem \ref{ball} }
652:
653: \label{figball}
654: \end{figure}
655: Obviously, a dual statement holds for $B_{D^-}(p,\epsilon)$. For $B_D$, notice
656: that we can write down a simpler, but weaker bound
657: $$
658: B_D(p,\epsilon) \subseteq
659: \bigcap_{x\in\mathcal{F}^-(p,\epsilon),\,y\in\mathcal{F}^+(p,\epsilon)}
660: A(x,y)\;.
661: $$
662: \\*
663: Concerning the second question posed at the beginning of this section, we notice that the strong metric $D$
664: does not determine the Lorentz distance $d$ up to time reversal for discrete Lorentz spaces.
665: \begin{exie}
666: \end{exie}
667: Consider the causal sets $\mathcal{P}_{1}$ and $\mathcal{P}_{2}$ defined by the Hasse diagrams below.
668: \begin{figure}[h]
669: \begin{center}
670: \setlength{\unitlength}{0.7cm}
671: \begin{picture}(3.5,2)
672: \thicklines
673: \put(0,0){\line(0,1){1}}
674: \put(0,1){\line(1,-1){1}}
675: \put(1,0){\line(0,1){1}}
676:
677: \put(0,0){\circle*{0.1}}
678: \put(0,1){\circle*{0.1}}
679: \put(1,1){\circle*{0.1}}
680: \put(1,0){\circle*{0.1}}
681: \put(2,0){\circle*{0.1}}
682:
683: \put(1, 1.5){{\Large $ \mathcal{P}_1$}}
684: \end{picture}
685: \begin{picture}(2,2)
686: \thicklines
687: \put(0,0){\line(0,1){1}}
688: \put(0,1){\line(1,-1){1}}
689: \put(1,0){\line(0,1){1}}
690: \put(1,1){\line(1,-1){1}}
691:
692: \put(0,0){\circle*{0.1}}
693: \put(0,1){\circle*{0.1}}
694: \put(1,1){\circle*{0.1}}
695: \put(1,0){\circle*{0.1}}
696: \put(2,0){\circle*{0.1}}
697:
698: \put(1, 1.5){{\Large $\mathcal{P}_2$}}
699: \end{picture}
700: \end{center}
701: \caption{Hasse diagrams}
702: \label{fig7}
703: \end{figure}
704: Clearly, $d_{GH}((\mathcal{P}_{1},d_{1}),(\mathcal{P}_{2},d_{2}))
705: = 1$ while
706: $d_{GH}((\mathcal{P}_{1},D_{1}),(\mathcal{P}_{2},D_{2})) = 0$.
707: Moreover, $D_{1}^{+} \neq D_{2}^{+}$ while $D_{1}^{-} =
708: D_{2}^{-}$. This clearly shows that convergence in the strong
709: metric is not sufficient to guarantee convergence of the Lorentz
710: metrics, as far as discrete Lorentz spaces is concerned.
711: $\square$
712:
713: \section{Causality} \label{cau}
714:
715: In this section, we study further causal properties of limit
716: spaces. The most daunting problem is how to define a causal
717: relation on the degenerate regions. First of all, one might ask
718: if this is physically meaningful in the sense that on may wonder whether
719: ``particles'' could travel on such causal curves? The answer to
720: this question \emph{appears} to be negative, since one would like particle motion in a spacetime region to
721: be determined by the distribution of matter and initial
722: data in the past region of the considered events, while any
723: definition of causality in the degenerate region \emph{will} be
724: based upon \emph{global (future)} data, such as specific differences
725: in chronological futures. We cannot stress the word \emph{global}
726: enough, since, if there were a difference in the future
727: lightcones, then this can already be spotted by a tiny Alexandrov
728: neighbourhood which is ``boost equivalent'' to a neighbourhood
729: which might look more localised to a specific class of observers,
730: such as defined by the strong distance\footnote{Pick a point $p$
731: of the past boundary, note then that $x \rightarrow D^{-}(p,x)$ is
732: a time function.}. Hence, it is not entirely clear whether
733: defining a causal relation is physically meaningful or not. It is
734: for sure a quite interesting mathematical problem and it shall be
735: treated as such in the rest of this section. \\* \\* In the
736: sequel, we speak about a \emph{good} proposal for the causal
737: relation if, roughly speaking, it coincides on limit spaces of
738: arbitrary GGH Cauchy sequences of conformally equivalent
739: spacetimes with the causal relations defined by the elements of
740: these sequences. To start with, we give an example which shows
741: that the closure of the space of discrete Lorentz spaces contains
742: spaces, whose causal behaviour differs significantly from the kind
743: of limit spaces already considered before\footnote{A similar kind
744: of limit space was communicated to me by R. Sorkin.}.
745: \begin{exie} \label{ex15}
746: \end{exie}
747: Suppose $L > 2$ and let $(\mathcal{P}^{L}_{n},d_{n})$ be a discrete Lorentz space defined by the set $\mathcal{P}_{n}^{L} = \left\{ \frac{iL}{n} | i=0 \ldots n \right\}$ and $d_{n}(p,q) = \max \left\{ 0 , q - p - 1 \right\}$. It is easy to check that $d_{n}$ defines a Lorentz distance. Moreover, $(\mathcal{P}^{L}_{n},d_{n}) \stackrel{n \rightarrow \infty}{\rightarrow} (\left[0,L \right], d)$ where, evidently, $d(p,q) = \max \left\{ 0 , q - p - 1 \right\}$. Obviously, we want the causal relation to be the ordinary order relation on $\left[0,L\right]$. Hence, any pair of timelike related points can \emph{only} be connected by a causal curve $\gamma$, which is nowhere timelike in the sense that for any $\gamma(t)$, $0 < \gamma(s) - \gamma(t) < 1$ implies that $d(\gamma(t),\gamma(s)) = 0$. The strong metric $D(t,s)$ between points $t<s$ equals $s-t$ unless $0 < t < 1$ and $L-1 < s < L$, then it equals $\max \left\{L-t,s\right\} -1$. Hence, locally, $D$ is the path metric defined by the the standard line element $dt^{2}$. Conclusion: although the limit space has a manifold structure, the Lorentz distance is far from being derived from a tensor. $\square$ \\* \\*
748: In the previous example, we have defined the causal relation using our intuition. Since we are looking for a general prescription for the causal relation, we might postulate something like: $p \prec q$ iff $I^{+}(q) \subset I^{+}(p)$ and $I^{-}(p) \subset I^{-}(q)$, although, as mentioned in section \ref{prop}, this is not sufficient. Therefore, let us start by defining the causal relation on this subset of $\mathcal{M}$ which we are most familiar with namely $\overline{\mathcal{TCON}}$. A good candidate for $\prec$ on $\overline{\mathcal{TCON}}$ is the $K^{+}$-causal relation defined by Sorkin and Woolgar
749: \cite{Sorkin2}, i.e.,
750:
751: \begin{deffie}
752: $K^{+}$ is the smallest, topologically closed, partial order in
753: $\mathcal{M} \times
754: \mathcal{M}$ containing $I^{+}$. $\square$
755: \end{deffie}
756: \textbf{Remarks:}
757: As mentioned in \cite{Sorkin2}, $K^{+}$ can be build by transfinite induction as follows:
758: \begin{itemize}
759: \item $\prec^{0} = I^{+}$
760: \item $\prec^{\alpha} = \bigcup_{\beta < \alpha} \prec^{\beta}$ if $\alpha$ is a limit ordinal
761: \item $\prec^{\beta + 1}$ is constructed from $\prec^{\beta}$ by adding pairs which are implied either by transitivity or closure.
762: \end{itemize}
763: Since $\mathcal{M} \times \mathcal{M}$ has at most
764: $2^{\aleph_{0}}$ elements,the procedure has to terminate at an
765: ordinal\footnote{For more information about ordinals and
766: transitive induction see \cite{Kelley}.} with cardinality less or equal to $2^{\aleph_{0}}$. The
767: following example illustrates that the procedure can run up to an
768: ordinal with cardinality $\aleph_{0}$. $\square$
769: \begin{exie}
770: \end{exie}
771: Let $\mathbb{N} = \aleph_{0}$ and construct the discrete Lorentz space $(\mathcal{P},d)$ as follows:
772: \begin{itemize}
773: \item the set of points is $\mathcal{P} = \left\{w^{i}, x^{j}_{k}, y^{j}_{k}, z^{j} | i \in \mathbb{N}+1 \textrm{ and } j,k \in \mathbb{N} \right\}$.
774: \item the Lorentz distance if given by $d(w^{i},z^{j}) = d(x^{i}_{k},z^{j}) = \frac{1}{(j+1)^{2}}$ and $d(y^{i}_{k}, z^{j+1}) = \frac{1}{(j+2)^{2}}$ for all $k$ and $i \leq j$ in $\mathbb{N}$. Moreover, $d(x^{i}_{j},y^{i}_{j}) = \frac{1}{(j+1)(i+1)^{2}}$. All other distances are calculated from these values by taking the maximum over all ``timelike'' chains.
775: \item from these data, it is easy to calculate the strong distance: \\* $D(w^{i},x^{i}_{j}),D(w^{i+1},y^{i}_{j}) = \frac{1}{(j+1)(i+1)^{2}}$ and $D(x^{i}_{j},y^{i}_{j}) = \frac{1}{(i+1)^{2}}$.
776: \end{itemize}
777: We now start our program: $\prec^{1}$ is the closure of $I^{+}$. Obviously, the new relations induced by this procedure are $w^{i} \prec^{1} w^{i+1}$. $\prec^{2}$ is constructed from $\prec^{1}$ by adding pairs which are implied by transitivity and closure: this results in $w^{i} \prec^{2} w^{i+2}$ for all $i \in \mathbb{N}$ and $w^{\mathbb{N}} \prec^{2} w^{\mathbb{N}}$. The reader may easily check that at stage $n>2$ : $\prec^{n} = \prec^{n-1} \cup \left\{(w^{i}, w^{j}) | i+2^{n-1} < j \leq i + 2^{n} \in \mathbb{N} \right\}$. Hence, $$ \prec^{\mathbb{N}} = I^{+} \cup \left\{(w^{i}, w^{j}) | i < j \in \mathbb{N} \right \} \cup \left\{(w^{\mathbb{N}} ,w^{\mathbb{N}}) \right\}.$$ But this relation is not closed yet and
778: $$ \prec^{\mathbb{N}+1} = I^{+} \cup \left\{(w^{i}, w^{j}) | i < j \in \mathbb{N}+1 \right \} \cup \left\{(w^{\mathbb{N}} ,w^{\mathbb{N}}) \right\}.$$ So the procedure stops at the $\mathbb{N}+1$'th step and the cardinality of $\mathbb{N}+1$ is $\aleph_{0}$. $\square$ \\* \\*
779: Since $K^{+}$ gives in general more information\footnote{One can construct Lorentz spaces, where $K^{+}(q) \subseteq K^{+}(p)$ and $K^{-}(p) \subseteq K^{-} (q)$ do \emph{not} imply that $I^{+}(q) \subseteq I^{+}(p)$ and $I^{-}(p) \subseteq I^{-}(q)$ and vice versa. However, $K^{+}(q) \subseteq K^{+}(p)$ and $K^{-}(p) \subseteq K^{-} (q)$ does imply that $I^{+}(q) \subseteq I^{+}(p)$ and $I^{-}(p) \subseteq I^{-}(q)$ for Lorentz spaces $(\mathcal{M},d)$ satisfying the following division property:
780: $$\forall p \ll q, \exists r : p \ll r \ll q$$} than $I^{+}$, we might hope that adding the conditions $K^{+}(q) \subset K^{+}(p)$ and $K^{-}(p) \subset K^{-}(q)$ in order for $p \prec q$ leads to a satisfying definition. Unfortunately, it does not, as illustrated in the following example.
781: \begin{exie} \label{ex16}
782: \end{exie}
783: To simplify the discussion, define a relation $\mathcal{R}$ between $p,q \in \mathcal{M} \setminus \overline{\mathcal{TCON}}$ as follows:
784: $$ p\mathcal{R}q \Leftrightarrow K^{+}(q) \subseteq K^{+}(p), K^{-}(p) \subseteq K^{-} (q), I^{+}(q) \subseteq I^{+}(p) \textrm{ and } I^{-}(p) \subseteq I^{-}(q). $$
785: Picture \ref{seven} shows a Lorentz space with points $p$ and $q$ such that $p\mathcal{R}q$ holds. Clearly, we do not want that $p \prec q$. However, there exists no curve $\gamma$ between $p$ and $q$ satisfying the condition that $\gamma(t) \mathcal{R} \gamma (s)$ for all $t \leq s$. The picture shows a part of the cylinder universe with degenerate regions, which are indicated by the shading. $\square$
786: \begin{figure}[h]
787: \begin{center}
788: \setlength{\unitlength}{1.7cm}
789: \begin{picture}(8,3.5)
790:
791: \put(1,1){\line(1,0){6}}
792: \put(1,1){\line(0,1){1.91}}
793: \put(1,2.91){\line(1,0){6}}
794: \put(7,1){\line(0,1){1.91}}
795: \put(1,0.5){$0$}
796: \put(7,0.5){$2 \pi$}
797: \put(0.5,3){$2$}
798: \put(3.53,1.96){\circle*{0.05}}
799: \put(4.45,1.96){\circle*{0.05}}
800: \thicklines
801: \put(3.53,1.96){\line(1,1){0.95}}
802: \put(3.53,1.96){\line(1,-1){0.95}}
803: \put(3.53,1.96){\line(-1,1){0.95}}
804: \put(3.53,1.96){\line(-1,-1){0.95}}
805: \put(4.45,1.96){\line(1,1){0.95}}
806: \put(4.45,1.96){\line(1,-1){0.95}}
807: \put(4.45,1.96){\line(-1,1){0.95}}
808: \put(4.45,1.96){\line(-1,-1){0.95}}
809:
810: %arcering onderste deel
811: \thinlines
812: \put(3.53,2.06){\Blue{\line(1,-1){0.5}}}
813: \put(3.43,1.96){\Blue{\line(1,-1){0.5}}}
814: \put(3.38,1.91){\Blue{\line(1,-1){0.5}}}
815: \put(3.33,1.86){\Blue{\line(1,-1){0.5}}}
816: \put(3.28,1.81){\Blue{\line(1,-1){0.5}}}
817: \put(3.23,1.76){\Blue{\line(1,-1){0.5}}}
818: \put(3.18,1.71){\Blue{\line(1,-1){0.5}}}
819: \put(3.13,1.66){\Blue{\line(1,-1){0.5}}}
820: \put(3.08,1.61){\Blue{\line(1,-1){0.5}}}
821: \put(3.03,1.56){\Blue{\line(1,-1){0.5}}}
822: \put(2.98,1.51){\Blue{\line(1,-1){0.5}}}
823: \put(2.93,1.46){\Blue{\line(1,-1){0.45}}}
824: \put(2.88,1.41){\Blue{\line(1,-1){0.40}}}
825: \put(2.83,1.36){\Blue{\line(1,-1){0.35}}}
826: \put(2.78,1.31){\Blue{\line(1,-1){0.30}}}
827: \put(2.73,1.26){\Blue{\line(1,-1){0.25}}}
828: \put(2.68,1.21){\Blue{\line(1,-1){0.20}}}
829: \put(2.63,1.16){\Blue{\line(1,-1){0.15}}}
830: \put(2.58,1.11){\Blue{\line(1,-1){0.10}}}
831: \put(2.53,1.06){\Blue{\line(1,-1){0.10}}}
832: \put(2.48,1.01){\Blue{\line(1,-1){0.10}}}
833: \put(2.43,0.96){\Blue{\line(1,-1){0.10}}}
834:
835: \put(3.93,2.36){\Blue{\line(1,-1){0.5}}}
836: \put(4.03,2.46){\Blue{\line(1,-1){0.5}}}
837: \put(4.08,2.51){\Blue{\line(1,-1){0.5}}}
838: \put(4.13,2.56){\Blue{\line(1,-1){0.5}}}
839: \put(4.18,2.61){\Blue{\line(1,-1){0.5}}}
840: \put(4.23,2.66){\Blue{\line(1,-1){0.5}}}
841: \put(4.28,2.71){\Blue{\line(1,-1){0.5}}}
842: \put(4.33,2.76){\Blue{\line(1,-1){0.5}}}
843: \put(4.38,2.81){\Blue{\line(1,-1){0.5}}}
844: \put(4.43,2.86){\Blue{\line(1,-1){0.5}}}
845: \put(4.48,2.91){\Blue{\line(1,-1){0.5}}}
846: \put(4.58,2.91){\Blue{\line(1,-1){0.45}}}
847: \put(4.68,2.91){\Blue{\line(1,-1){0.40}}}
848: \put(4.78,2.91){\Blue{\line(1,-1){0.35}}}
849: \put(4.88,2.91){\Blue{\line(1,-1){0.30}}}
850: \put(4.98,2.91){\Blue{\line(1,-1){0.25}}}
851: \put(5.08,2.91){\Blue{\line(1,-1){0.20}}}
852: \put(5.18,2.91){\Blue{\line(1,-1){0.15}}}
853: \put(5.28,2.91){\Blue{\line(1,-1){0.10}}}
854: \put(3.24,1.96){$p$}
855: \put(4.65,1.96){$q$}
856: \end{picture}
857: \caption{Example \ref{ex16}, cylinder universe with degenerate regions.}
858: \label{seven}
859: \end{center}
860: \end{figure}
861: \\* \\*
862: The above example is very unfortunate in the sense that it shows
863: that using all imaginable relations between points derived from
864: the chronological partial order, using zero dimensional objects
865: only, is not sufficient for obtaining a satisfactory causal
866: relation. However, it also suggests that the following definition
867: might have more success.
868: \begin{deffie}
869: Define a partial order $\mathcal{P}$ on a Lorentz space $(\mathcal{M},d)$ by putting $p\mathcal{P}q$ iff there exists a continuous curve $\gamma : \left[0,1\right] \rightarrow \mathcal{M}$ from $p$ to $q$ such that $\gamma(t)\mathcal{R}\gamma(s)$ for all $0 \leq t\leq s \leq 1$.
870: Finally, define $\prec_{d}$ on $\mathcal{M} \setminus \overline{\mathcal{TCON}}$ as the smallest topologically closed transitive relationship containing $\mathcal{P}$ and $I^{+}$. It is easy to see that $\prec_{d}$ is compatible with $K^{+}$, i.e., $p \prec_{d} q$ and $q \in K^{-}(r)$ imply that $p \in K^{-}(r)$ and vice versa. $\square$
871: \end{deffie}
872: The following example in three dimensions shows that also this definition has its limitations. However, in two dimensions, it does work as is proven in Theorem \ref{causr}.
873: \begin{exie} \label{ex17}
874: \end{exie}
875: Consider the three dimensional cylinder universe $(S^{2} \times
876: \left[-1,1\right], -dt^{2} + d\theta^{2} + \sin^{2}\theta
877: d\phi^{2})$. Consider the spacelike geodesic
878: $\gamma:\left[0,\frac{1}{4} \right] \rightarrow S^{2} \times
879: \left[-1,1\right]:s \rightarrow \gamma(s)=( \theta_{0}+s
880: ,\phi_{0}, 0)$. Take the limit $(S^{2} \times
881: \left[-1,1\right],d)$ over a suitable sequence of conformally
882: equivalent metrics, with conformal factors which converge to zero
883: on thin, specific, (see picture \ref{eight} below) open
884: neighbourhoods of $J^{+}(\gamma(s))\setminus J^{+}(\gamma(t)) $
885: and $J^{-}(\gamma(t)) \setminus J^{-}(\gamma(s))$, which are
886: subsets of $J^{+}(\gamma(t))^{c}$ and $J^{-}(\gamma(s))^{c}$
887: respectively for all $t<s$\footnote{The relation $J^{+}$ denotes
888: here the usual causal relation defined by the metric tensor
889: $-dt^{2} + d\theta^{2} + \sin^{2}\theta d\phi^{2}$.}.
890: \begin{figure}[h]
891: \begin{center}
892: \setlength{\unitlength}{1.7cm}
893: \begin{picture}(8,3.5)
894:
895: \put(1,1){\line(1,0){6}}
896: \put(1,1){\line(0,1){1.91}}
897: \put(1,2.91){\line(1,0){6}}
898: \put(7,1){\line(0,1){1.91}}
899: \put(1,0.5){$0$}
900: \put(7,0.5){$2 \pi$}
901: \put(3.53,1.96){\circle*{0.05}}
902: \put(4.45,1.96){\circle*{0.05}}
903: \thicklines
904: \put(3.53,1.96){\line(1,0){0.94}}
905: \put(3.53,1.96){\circle{1}}
906: \put(4.45,1.96){\circle{1}}
907: \put(3.53,2.37){\line(1,0){0.94}}
908: \put(3.53,1.54){\line(1,0){0.94}}
909: %arcering onderste deel
910: \thinlines
911: \put(3.53,2.42){\Blue{\line(1,0){0.99}}}
912: \put(3.64,2.38){\Blue{\line(1,0){0.99}}}
913: \put(3.75,2.32){\Blue{\line(1,0){0.99}}}
914: \put(3.80,2.27){\Blue{\line(1,0){0.99}}}
915: \put(3.85,2.22){\Blue{\line(1,0){0.99}}}
916: \put(3.89,2.17){\Blue{\line(1,0){0.99}}}
917: \put(3.91,2.12){\Blue{\line(1,0){0.99}}}
918: \put(3.92,2.07){\Blue{\line(1,0){0.99}}}
919: \put(3.94,2.02){\Blue{\line(1,0){0.99}}}
920: \put(3.95,1.97){\Blue{\line(1,0){0.99}}}
921: \put(3.94,1.92){\Blue{\line(1,0){0.99}}}
922: \put(3.92,1.87){\Blue{\line(1,0){0.99}}}
923: \put(3.91,1.82){\Blue{\line(1,0){0.99}}}
924: \put(3.89,1.77){\Blue{\line(1,0){0.99}}}
925: \put(3.87,1.72){\Blue{\line(1,0){0.99}}}
926: \put(3.83,1.67){\Blue{\line(1,0){0.99}}}
927: \put(3.79,1.62){\Blue{\line(1,0){0.99}}}
928: \put(3.73,1.57){\Blue{\line(1,0){0.99}}}
929: \put(3.63,1.52){\Blue{\line(1,0){0.99}}}
930: \put(3.52,1.47){\Blue{\line(1,0){0.99}}}
931: \put(3.24,1.96){$p$}
932: \put(4.65,1.96){$q$}
933: \end{picture}
934: \caption{Example \ref{ex17}, intersection of the lightcones with a constant time hypersurface. The shading indicates the degenerate regions.}
935: \label{eight}
936: \end{center}
937: \end{figure}
938: It is not difficult to see that for any point $q$ belonging to the degenerate region, either $I_{d}^{+}(q) \cap I_{d}^{+}(\gamma(0)) \neq \emptyset$ or $I_{d}^{-}(q) \cap I_{d}^{-}(\gamma(1/4)) \neq \emptyset$. Using this, it is easy to see that for any two points $p$ and $q$ belonging to the degenerate region, we have that $I_{d}^{-}(p) \neq I_{d}^{-}(q)$ or $I_{d}^{+}(p) \neq I_{d}^{+}(q)$. $\square$ \\*
939:
940: \begin{theo} \label{causr} Let $(\mathcal{M},g)$ be a \emph{two dimensional}, globally hyperbolic, interpolating spacetime which is isometrically embeddable in the interior of an interpolating spacetime \emph{without cut points} and suppose $\Omega_{i}$ is a sequence of positive $C^{\infty}$ functions on $\mathcal{M}$ such that $\left| d_{\Omega_{i}^{2}g}(p,q) - d_{\Omega_{j}^{2}g}(p,q) \right| < \frac{1}{i}$ for all $j > i > 0$ and $p,q \in \mathcal{M}$. Denote by $(\mathcal{N},d)$ the GGH limit space. Suppose that $\mathcal{M}=\mathcal{N}$, i.e., no points get identified. Then, one has that $p \prec_{g} q$ iff $p \prec_{d} q$ for all $p,q \in \mathcal{M}$.
941: \end{theo}
942: \textbf{Note}: If $(\mathcal{M},g)$ were
943: allowed to have cut points, then the theorem would not be valid
944: anymore. It is possible to construct a counterexample by making a
945: drawing on the two dimensional cylinder universe. \\* \\*
946: \textsl{Proof}: \\* First notice that the strong topology defined
947: by $D$ coincides with the manifold topology\footnote{Clearly, $D$
948: is continuous in the manifold topology on $\mathcal{M} \times
949: \mathcal{M}$ since it is the uniform limit of a sequence of
950: continuous functions. On the other hand, let $p \in \mathcal{M}$
951: and $p \in \mathcal{V} \subset \overline{\mathcal{V}} \subset
952: \mathcal{U}$ where $\mathcal{U}$ and $\mathcal{V}$ are open
953: neighbourhoods of $p$ in the manifold topology. Suppose,
954: moreover, that there exists a sequence $(p_{n})_{n \in
955: \mathbb{N}}$ such that $p_{n} \notin \mathcal{U}$ and $D(p,p_{n})
956: < \frac{1}{n}$. By passing to a subsequence if necessary, we may
957: assume that $p_{n} \stackrel{n \rightarrow \infty}{\rightarrow} q
958: \in \overline{V}^{c}$ in the manifold topology. Hence, $D(p,q)=0$
959: which contradicts $\mathcal{M} = \mathcal{N}$.}. \\*
960: $\Rightarrow)$ We show that \emph{any} $g$ causal curve $\gamma$
961: is a $\mathcal{R}$ causal curve. Clearly, $I^{+}_{d}(\gamma(s))
962: \subseteq I^{+}_{d}(\gamma(t))$ and $I^{-}_{d}(\gamma(t))
963: \subseteq I^{-}_{d}(\gamma(s))$ for all $t < s$ which proves the
964: basis of induction. Let $\alpha = \beta + 1$ and suppose that $t
965: < s$ implies that $\gamma(s) \prec^{\beta} y \Longrightarrow
966: \gamma(t) \prec^{\beta} y$. Obviously, if $\gamma(s)
967: \prec^{\beta} y \prec^{\beta} z$ then $\gamma(t) \prec^{\beta} y
968: \prec^{\beta} z$. So suppose that there exist sequences
969: $(q_{n})_{n \in \mathbb{N}}$, $(y_{n})_{n \in \mathbb{N}}$
970: converging to $\gamma(s)$ and $y$ respectively such that $q_{n}
971: \prec^{\beta} y_{n}$ for any $n$, then there exists a sequence $(p_{n})_{n \in
972: \mathbb{N}}$ converging to $\gamma(t)$ with $p_{n} \prec_{g}
973: q_{n}$. The induction hypothesis then implies that $p_{n}
974: \prec^{\beta} y_{n}$ for all $n$ which proves the claim. \\*
975: $\Leftarrow)$ We have to show that $g$-spacelike events cannot be
976: connected by an $\mathcal{R}$-causal curve. Suppose $p$ and $q$
977: are such events, and suppose $\gamma: \left[0,1\right] \rightarrow
978: \mathcal{M}$ is an $\mathcal{R}$-causal curve connecting them.
979: Without loss of generality, we may assume that $\gamma$ is
980: spacelike to $p$ and $q$ in the sense that $\gamma(t)$, $p$ and
981: $q$ are $g$-spacelike events for all $t \in
982: \left(0,1\right)$\footnote{Note that $\gamma$ cannot intersect
983: $E^{-}(p)$, nor $E^{+}(q)$, because this would violate the
984: assumption that the limit space equals $\mathcal{M}$. By
985: continuity, there exists a $t$, such that $\gamma(t) \in E^{+}(p)$
986: but $\gamma(u) \notin E^{+}(p)$ for all $u > t$, and a minimal
987: $s>t$ such that $\gamma(s) \in E^{-}(q)$.}. Moreover, we may
988: assume that $\gamma$ is a subset of a convex open neighborhood
989: $\mathcal{U}$ on which $g$ is conformally flat. By using the
990: nonexistence of cut points, one can deduce that the set
991: $\mathcal{S} \subset \mathcal{M}$ bounded by the two right $g$
992: null geodesics containing $\gamma$ is entirely degenerate as is
993: shown in picture \ref{nine}\footnote{Choose $\gamma(t), t \in
994: \left(0,1\right)$. Then, there exists an open neighborhood
995: $\mathcal{O}$ of $\gamma(t)$ such that for all $r \in \mathcal{O}$
996: which are $g$-spacelike to the left or in the $g$ chronological
997: past of $\gamma(t)$, we have that the left, future oriented, null
998: geodesic starting at $r$ does \emph{not} intersect the future
999: oriented, right null geodesic starting at $\gamma(t)$. Otherwise,
1000: $\gamma(t)$ would have a cut point in any extension of
1001: $(\mathcal{M},g)$. Hence, for all $s < t$ such that $\gamma(s)
1002: \in \mathcal{O}$ is such point $r$, we have that any point in
1003: $J^{+}(\gamma(t))$ to the right of the right null geodesic
1004: emanating from $\gamma(s)$ belongs to the degenerate area. A
1005: similar argument is valid for the past, with left and right
1006: switched. Using this for all $t$ leads to picture \ref{nine}.}.
1007: Any two points in $\mathcal{S} \cap \mathcal{U}$ belonging to any
1008: left $g$ null geodesic have the same chronological relations.
1009: \begin{figure}[h]
1010: \begin{center}
1011: \setlength{\unitlength}{1.7cm}
1012: \begin{picture}(7,3)
1013: \put(1,1){\line(1,0){6}}
1014: \put(1,1){\line(0,1){1.91}}
1015: \put(1,2.91){\line(1,0){6}}
1016: \put(7,1){\line(0,1){1.91}}
1017: \put(3.53,1.96){\circle*{0.05}}
1018: \put(4.45,1.96){\circle*{0.05}}
1019: \thicklines
1020: \put(3.53,1.96){\line(1,1){0.95}}
1021: \put(3.53,1.96){\line(1,-1){0.95}}
1022: \put(3.53,1.96){\line(-1,1){0.95}}
1023: \put(3.53,1.96){\line(-1,-1){0.95}}
1024: \put(4.45,1.96){\line(1,1){0.95}}
1025: \put(4.45,1.96){\line(1,-1){0.95}}
1026: \put(4.45,1.96){\line(-1,1){0.95}}
1027: \put(4.45,1.96){\line(-1,-1){0.95}}
1028: \put(3.53,1.96){\line(1,0){0.95}}
1029: \put(3.8,1.3){\line(-1,1){0.3}}
1030: %arcering onderste deel
1031: \thinlines
1032: \put(3.58,2.01){\Blue{\line(1,-1){0.45}}}
1033: \put(3.48,1.91){\Blue{\line(1,-1){0.45}}}
1034: \put(3.43,1.86){\Blue{\line(1,-1){0.45}}}
1035: \put(3.38,1.81){\Blue{\line(1,-1){0.45}}}
1036: \put(3.33,1.76){\Blue{\line(1,-1){0.45}}}
1037: \put(3.28,1.71){\Blue{\line(1,-1){0.45}}}
1038: \put(3.23,1.66){\Blue{\line(1,-1){0.45}}}
1039: \put(3.18,1.61){\Blue{\line(1,-1){0.45}}}
1040: \put(3.13,1.56){\Blue{\line(1,-1){0.45}}}
1041: \put(3.08,1.51){\Blue{\line(1,-1){0.45}}}
1042: \put(3.03,1.46){\Blue{\line(1,-1){0.45}}}
1043: \put(2.98,1.41){\Blue{\line(1,-1){0.4}}}
1044: \put(2.93,1.36){\Blue{\line(1,-1){0.35}}}
1045: \put(2.88,1.31){\Blue{\line(1,-1){0.30}}}
1046: \put(2.83,1.26){\Blue{\line(1,-1){0.25}}}
1047: \put(2.78,1.21){\Blue{\line(1,-1){0.20}}}
1048: \put(2.73,1.16){\Blue{\line(1,-1){0.15}}}
1049: \put(2.68,1.11){\Blue{\line(1,-1){0.10}}}
1050: \put(2.63,1.06){\Blue{\line(1,-1){0.10}}}
1051: \put(2.58,1.01){\Blue{\line(1,-1){0.10}}}
1052:
1053: \put(3.48,1.91){\Blue{\line(1,-1){0.45}}}
1054: \put(3.53,1.96){\Blue{\line(1,-1){0.45}}}
1055: \put(3.58,2.01){\Blue{\line(1,-1){0.45}}}
1056: \put(3.63,2.06){\Blue{\line(1,-1){0.45}}}
1057: \put(3.68,2.11){\Blue{\line(1,-1){0.45}}}
1058: \put(3.73,2.16){\Blue{\line(1,-1){0.45}}}
1059: \put(3.78,2.21){\Blue{\line(1,-1){0.45}}}
1060: \put(3.83,2.26){\Blue{\line(1,-1){0.45}}}
1061: \put(3.88,2.31){\Blue{\line(1,-1){0.45}}}
1062: \put(3.93,2.36){\Blue{\line(1,-1){0.45}}}
1063: \put(4.03,2.46){\Blue{\line(1,-1){0.45}}}
1064: \put(4.08,2.51){\Blue{\line(1,-1){0.45}}}
1065: \put(4.13,2.56){\Blue{\line(1,-1){0.45}}}
1066: \put(4.18,2.61){\Blue{\line(1,-1){0.45}}}
1067: \put(4.23,2.66){\Blue{\line(1,-1){0.45}}}
1068: \put(4.28,2.71){\Blue{\line(1,-1){0.45}}}
1069: \put(4.33,2.76){\Blue{\line(1,-1){0.45}}}
1070: \put(4.38,2.81){\Blue{\line(1,-1){0.45}}}
1071: \put(4.43,2.86){\Blue{\line(1,-1){0.45}}}
1072: \put(4.48,2.91){\Blue{\line(1,-1){0.45}}}
1073: \put(4.58,2.91){\Blue{\line(1,-1){0.40}}}
1074: \put(4.68,2.91){\Blue{\line(1,-1){0.35}}}
1075: \put(4.78,2.91){\Blue{\line(1,-1){0.30}}}
1076: \put(4.88,2.91){\Blue{\line(1,-1){0.25}}}
1077: \put(4.98,2.91){\Blue{\line(1,-1){0.20}}}
1078: \put(5.08,2.91){\Blue{\line(1,-1){0.15}}}
1079: \put(5.18,2.91){\Blue{\line(1,-1){0.10}}}
1080: \put(5.28,2.91){\Blue{\line(1,-1){0.10}}}
1081: \put(3.24,1.96){$p$}
1082: \put(4.65,1.96){$q$}
1083: \end{picture}
1084: \caption{Proof of theorem \ref{causr}, conflict with the $T_{0}$ property.}
1085: \label{nine}
1086: \end{center}
1087: \end{figure}
1088: $\square$
1089: \\* \\* The results of example \ref{ex17} and theorem \ref{causr} are quite discouraging, since any good definition of a causal relation seems to depend upon some notion of \emph{dimension} of the Lorentz space. One could try to make the definition more restrictive, so that it would be possible to reproduce a result analogous to theorem \ref{causr} in all dimensions. It seems to me that ``local''\footnote{``Local'' in the sense that one studies properties of local congruences of curves between neighbourhoods of points. One such idea would be to construct the following kind of definition: define the relation $p\mathcal{Q}q$ iff there exist neighbourhoods $\mathcal{U}$, $\mathcal{V}$ of $p$ and $q$ respectively and a mapping $\psi : \mathcal{U} \times \left[0,1 \right] \rightarrow \mathcal{M}$ such that
1090: \begin{itemize}
1091: \item $\psi_{t} : \mathcal{U} \rightarrow \mathcal{M}: r \rightarrow \psi(r,t)$ is a homeomorphism for any $t$ and $\psi_{1} (\mathcal{U}) = \mathcal{V}$.
1092: \item $\psi_{r}: \left[0,1\right] \rightarrow \mathcal{M} : t \rightarrow \psi(r,t)$ defines a $\mathcal{R}$-causal curve for any $r \in \mathcal{U}$.
1093: \end{itemize}
1094: $\prec_{d}$ is then defined as the smallest topologically closed transitive relation encompassing $\mathcal{Q}$ and $I^{+}$. Again it is not difficult to construct a counterexample similar to example \ref{ex17}.} ideas won't work. \\* \\*
1095: The rest of this paragraph is devoted to proving that the limit space $(\mathcal{M},d)$ of a $\mathcal{C}^{+}_{\alpha}$ and $\mathcal{C}^{-}_{\alpha}$ sequence $(\mathcal{M}_{i},d_{i})_{i \in \mathbb{N}}$ of path metric\footnote{For a precise definition of a path metric Lorentz space, see definition \ref{path}.} Lorentz spaces is a path metric Lorentz space. Strictly speaking, we should still define the $\mathcal{C}^{\pm}_{\alpha}$ properties for general Lorentz spaces $(\mathcal{M},d)$. Looking at the definition in the intermezzo of section \ref{prop}, the reader can see that this boils down to defining the future and past boundaries of $(\mathcal{M},d)$. Obviously, the past boundary $\partial_{P}\mathcal{M}$ is the set of points $p$ such that $I^{-}(p) = \emptyset$, the future boundary $\partial_{F}\mathcal{M}$ is defined dually. \\* \\*
1096: \textbf{Property}: The $\mathcal{C}^{+}_{\alpha}$ property implies that the interior of $\partial_{P} \mathcal{M}$ is empty and, likewise, the $\mathcal{C}^{-}_{\alpha}$ property implies that the interior of $\partial_{F} \mathcal{M}$ is empty. \\*
1097: \\*
1098: \textsl{Proof}: \\*
1099: I shall only prove the former. First, note that $\partial_{P} \mathcal{M} \cap \partial_{F} \mathcal{M}$ contains at most one point. Let $p \in \partial_{P} \mathcal{M} \setminus \partial_{F} \mathcal{M}$ and $\epsilon > 0$ be such that $B_{D}(p, \epsilon) \subset \partial_{P} \mathcal{M} \setminus \partial_{F} \mathcal{M}$. Then, $d(p,r) = 0$ for all $r \in B_{D}(p, \epsilon)$ which is impossible by the $\mathcal{C}^{+}_{\alpha}$ property. $\square$ \\* \\*
1100: As a consequence, we have that for a Lorentz space $(\mathcal{M},d)$ satisfying the $\mathcal{C}^{+}_{\alpha}$ and $\mathcal{C}^{-}_{\alpha}$ property, $\overline{\mathcal{TCON}} \cup \left( \partial_{P} \mathcal{M} \cap \partial_{F} \mathcal{M} \right) = \mathcal{M}$\footnote{For example, let $p \in \partial_{P} \mathcal{M} \setminus \partial_{F} \mathcal{M}$, then for $\epsilon > 0$ sufficiently small, we have that $B_{D}(p , \epsilon) \cap \partial_{F} \mathcal{M} = \emptyset$. By the $\mathcal{C}^{+}_{\alpha}$ property, there exists an $r \in \overline{B_{D}(p, \frac{\epsilon}{2})}$ such that $d(p,r) = \alpha(\frac{\epsilon}{2})$. Hence $$D(r, \partial_{P} \mathcal{M}), D(r, \partial_{F} \mathcal{M}) \geq \alpha \left(\frac{\epsilon}{2} \right)$$ which implies, by the $\mathcal{C}^{+}_{\alpha}$ and $\mathcal{C}^{-}_{\alpha}$ properties, that $r \in \mathcal{TCON}$.}. Note that the second term on the left-hand side of this equality only needs to be accounted for iff $\partial_{P} \mathcal{M} \cap \partial_{F} \mathcal{M}$ is an \emph{isolated} point. Hence, the causal relation on such space is the $K^{+}$ relation. \\* \\*
1101: First, we give an example of spacetime with the $\mathcal{C}^{\pm}_{x^{2}/2}$ property.
1102: \begin{exie} \label{ex18}
1103: \end{exie}
1104: Consider again the cylinder universe $\mathcal{CYL} = ({\rm S}^1
1105: \times [0,1], -d t^{2} + d\theta^2)$. We argue that $\mathcal{CYL}$
1106: belongs to the category defined by $\alpha : \mathbb{R}^{+} \rightarrow \mathbb{R}^{+} : x
1107: \rightarrow \frac{x^{2}}{2}$. Since SO(2) is the isometry group of $d_g$, it is
1108: sufficient to prove the assertion for points with a fixed spatial coordinate,
1109: say, $\theta = \pi$. Let $1 \geq \tilde{t} > t \geq 0$; then it is easy to prove that
1110: $$
1111: D((\pi , t) ,(\pi ,\tilde{t})) = \sqrt{\tilde{t} - t} \,
1112: \max \left\{\sqrt{\tilde{t} + t} , \sqrt{2 - (t + \tilde{t})} \right\},
1113: $$
1114: where $D$ is the strong metric on $\mathcal{CYL}$. Hence
1115: $$
1116: d_g ((\pi,t) , (\pi, \tilde{t})) \leq D((\pi , t) ,(\pi ,\tilde{t}))^{2}
1117: \leq 2\, d_g ((\pi,t) , (\pi, \tilde{t}))\;,
1118: $$
1119: which proves the assertion. \hfill$\square$ \\* \\*
1120: Before we proceed, we should define causal curves $\gamma$ and lengths thereof\footnote{The reader can find similar definitions in \cite{Busemann}.}.
1121: \begin{deffie}
1122: Let $(\mathcal{M},d)$ be a Lorentz space. Assume $a<b$ and let $\gamma : [a, b] \rightarrow \mathcal{M}$ be a continuous (w.r.t. the strong topology) mapping such that for all $a \leq t < s \leq b$ : $ \gamma (t) \prec \gamma
1123: (s)$ ($ \gamma (t) \ll \gamma (s)$); then $\gamma$ is a basic, causal (timelike)
1124: curve. Let $a < b$, $c < d$ and $\gamma_{1}: [a, b] \rightarrow \mathcal{M}$,
1125: $\gamma_{2}: [c ,d] \rightarrow \mathcal{M}$ be basic causal curves such that
1126: $\gamma_{2}(c) = \gamma_{1}(b)$. We define the concatenation $\gamma_{2}
1127: \circ \gamma_{1}$ of $\gamma_{2}$ with $\gamma_{1}$ as the basic causal curve
1128: $\gamma_{2} \circ \gamma_{1}: [a, b + d - c] \rightarrow \mathcal{M}$ such that
1129: $$
1130: %\begin{equation*}
1131: \gamma_{2} \circ \gamma_{1} (t)
1132: = \begin{cases} \gamma_{1} (t) & {\rm if } \quad a \leq t \leq b \cr
1133: \gamma_{2} (t + c -b) & {\rm if } \quad b \leq t \leq b + d - c. \end{cases}
1134: %\end{equation*}
1135: $$
1136: A (countably infinite) concatenation of basic, causal curves is a causal
1137: curve. \\*
1138: The length $L(\gamma)$ of a basic, causal curve $\gamma : [a , b] \rightarrow
1139: \mathcal{M}$ is defined as
1140: $$
1141: L(\gamma) = \inf_{ \Delta } \sum_{i = 0}^{ \left| \Delta \right| - 1}
1142: d(\gamma(t_{i}) , \gamma(t_{i+1}))\;,
1143: $$
1144: where, $\Delta = \left\{t_{i} | a = t_{0} < t_{1} < \ldots < t_{n-1} < t_{n}
1145: = b \right\}$ is a partition of $\left[a,b \right]$. Obviously,
1146: $$
1147: L(\gamma_{2} \circ \gamma_{1}) = L(\gamma_{1}) + L(\gamma_{2})\;.
1148: $$
1149: \hfill$\square$
1150:
1151: \end{deffie}
1152: Now, we are able to give the definition of a path metric Lorentz space.
1153: \begin{deffie} \label{path}
1154: $(\mathcal{M},d)$ is a path metric Lorentz space iff for any $p \prec q$, there exists a causal curve $\gamma$ from $p$ to $q$, such that $L(\gamma) = d(p,q)$.
1155: \end{deffie}
1156: In case the causal relation coincides with the $K^{+}$ relation, I will prove that $(\mathcal{M},d)$ is a path metric space iff for any $p \ll q$, there exists a distance realising ($K^{+}$) causal curve from $p$ to $q$. We need to introduce the Vietoris topology on the set $2^{(\mathcal{M},D)}$ of all closed, non-empty subsets of $(\mathcal{M},D)$ for which a \emph{sub-basis} is given by the sets $\mathcal{B}(\mathcal{M},\mathcal{O})$ and $\mathcal{B}(\mathcal{O}, \mathcal{M})$. The former are sets with as members closed sets which meet the open set $\mathcal{O}$, the latter consists of the closed subsets of $\mathcal{O}$. It is known that $2^{(\mathcal{M},D)}$ equipped with the Vietoris topology is compact \cite{Sorkin2}. Also, it is proven in this paper that the Vietoris limit of a sequence of $K^{+}$ causal curves is a $K^{+}$ causal curve using topological arguments only.
1157: \begin{theo} \label{theo27} Let $(\mathcal{M},d)$ be a Lorentz space, then $(\mathcal{M},d)$ is a path metric space with respect to the $K^{+}$ relation iff for any $p \ll q$, there exists a distance realising $K^{+}$ causal curve from $p$ to $q$.
1158: \end{theo}
1159: \textsl{Proof}: \\* We only have to prove that the latter implies
1160: the former, the other way around being obvious. We shall once
1161: more proceed by transfinite induction with as induction hypothesis
1162: $H_{\alpha}$ the statement that $p \prec^{\alpha} q$ implies that
1163: there exists a distance realising $K^{+}$ causal curve from $p$ to
1164: $q$. The basis of induction is nothing else but our assumption.
1165: Hence, let $\alpha = \beta + 1$ and assume $H_{\beta}$ is valid.
1166: If $p \prec^{\beta} q \prec^{\beta} r$ then there exists a $K^{+}$
1167: causal curve from $p$ to $r$, by the induction hypothesis and
1168: concatenation, which is obviously distance maximising. So assume
1169: that there exist sequences $(p_{n})_{n \in \mathbb{N}}$,
1170: $(q_{n})_{n \in \mathbb{N}}$ converging to $p$ and $q$
1171: respectively and that $p_{n} \prec^{\beta} q_{n}$ for all $n \in
1172: \mathbb{N}$. Then, $H_{\beta}$ implies that there exist $K^{+}$
1173: causal curves $\gamma_{n}$ from $p_{n}$ to $q_{n}$. We may assume
1174: that, by passing to a subsequence if necessary, $(\gamma_{n})_{n
1175: \in \mathbb{N}}$ converges in the Vietoris topology to a (distance
1176: maximising) $K^{+}$ causal curve connecting $p$ with $q$.
1177: $\square$ \\* \\* Before I prove the main result, it is useful to
1178: study some properties of causal curves with respect to the strong
1179: metric $D$.
1180: \begin{exie} \label{ex19}
1181: \end{exie}
1182: We show that the $D$-length of a compact, basic causal curve is in general
1183: infinite. Obviously, the way to define the $D$-length, $DL( \gamma )$, of a
1184: basic causal curve $\gamma: [a , b] \rightarrow \mathcal{M}$ is
1185: $$
1186: DL( \gamma ) = \sup_{\Delta} \sum_{i = 0}^{\left| \Delta \right| - 1 }
1187: D(\gamma (t_{i}) , \gamma(t_{i+1}))\;,
1188: $$
1189: where, as before, $\Delta = \left\{t_{i} \mid a = t_{0} < t_{1} <
1190: \ldots < t_{n-1} < t_{n} = b \right\}$ is a partition of $[a,b]$.
1191: Returning to example \ref{ex18}, we prove that the length of the interval $\left\{
1192: (\pi , t) \mid 0 \leq t \leq 1 \right\}$ equals $\infty$. Choose $\Delta_{n}
1193: = \left\{ 1 - \frac{1}{k} \mid k = 1 \ldots n \right\} \cup \{1\}$, then
1194: $$
1195: \sum_{k = 0}^{ n } D((\pi , t_{k}) , (\pi , t_{k+1}))
1196: > \sum_{k = 2}^{n+2} \frac{1}{k}\;,
1197: $$
1198: which proves the claim. \hfill$\square$
1199: \\*
1200: \\*
1201: Since example \ref{ex19} shows that the $D$-length of a causal
1202: curve is a meaningless concept, we have to come up with some other
1203: way to divide a causal curve into smaller pieces. As a starter,
1204: we mention the following result.
1205: \begin{theo}
1206: Let $(\mathcal{M},g)$ be an interpolating spacetime and let $\gamma : \left[a , b \right] \rightarrow \mathcal{M}$ be any
1207: basic, causal curve. Then, there exists no $t \in \left( a , b
1208: \right)$ such that $D(\gamma(a), \gamma(t)) = D(\gamma(t) ,\gamma(b))
1209: = \frac{D(\gamma(a), \gamma(b))}{2}$, i.e., $\gamma$ has no
1210: $D$-midpoint.
1211: \end{theo}
1212: \textsl{Proof:} \\* We show that for all $t \in \left( a ,b
1213: \right)$: $D(\gamma(a), \gamma(b)) < D(\gamma(a) , \gamma(t)) +
1214: D(\gamma(t), \gamma(b))$. Assume that the point $r$, which
1215: realises $D(\gamma(a), \gamma(b))$, belongs to $I^{+} (\gamma(a))
1216: \setminus I^{+} (\gamma(b))$. The case where $r$ belongs to
1217: $I^{-}(\gamma(b)) \setminus I^{-} (\gamma(a))$ is identical and is
1218: left as an exercise to the reader. Then,
1219: $$ D(\gamma(a), \gamma(b)) = d(\gamma(a),r) = \left( d(\gamma(a),r)
1220: - d(\gamma(t),r) \right) + d(\gamma(t),r) $$ Both terms on the
1221: rhs. are nonnegative and bounded by $D(\gamma(a) , \gamma(t))$ and
1222: $D(\gamma(t) , \gamma(b))$ respectively. The first term can only
1223: realise $D(\gamma(a), \gamma(t))$ if $r \in I^{+}(\gamma(a))
1224: \setminus I^{+} (\gamma(t))$. But, in that case $d(\gamma(t),r) =
1225: 0$ and this concludes the proof. $\square$ \\*
1226: \\*
1227: Hence, we define the following concept of division of a causal curve.
1228: \begin{deffie}
1229: Let $\gamma : \left[ a ,b \right] \rightarrow \mathcal{M}$ be a basic
1230: causal curve and
1231: denote $\delta = D (\gamma(a) , \gamma(b))$. The division,
1232: $\gamma^{1/2}$, of $\gamma$ is defined as the set of points $\left\{p_{i} | i=0 \ldots k \right\}$ such
1233: that $\gamma (a) = p_{0} \prec p_{1} \prec \ldots \prec p_{k-1} \prec p_{k} = \gamma (b)$, $D (p_{i} , p_{i+1}) = \textstyle{\frac{\delta}{2}},\; \forall i: 0 \ldots k-2$, $\textstyle{\frac{\delta}{4}} \leq D(p_{k-1}, p_{k}) < {\textstyle{\frac{3\delta}{4}}}$ and there exists no point $q$ such that $p_{k-1} \prec q \prec \gamma(b)$ with $D(p_{k-1},q) =\textstyle{\frac{\delta}{2}}$ and $\textstyle{\frac{\delta}{4}} \leq D(q, \gamma(b)) < {\textstyle{\frac{3\delta}{4}}}$. Such finite number $2 \leq k = \left|\gamma^{1/2} \right| - 1$ exists, since $\gamma$ is continuous with respect to the strong topology.
1234: \end{deffie}
1235: This new concept facilitates the proof of the final theorem.
1236: \begin{theo}
1237: The limit space $(\mathcal{M},d)$ of a GGH $\mathcal{C}^{+}_{\alpha}$ and
1238: $\mathcal{C}^{-}_{\alpha}$ Cauchy sequence $(\mathcal{M}_{i},d_{i})_{i \in
1239: \mathbb{N}}$ of path metric Lorentz spaces, is a path metric Lorentz space.
1240: \end{theo}
1241: \noindent\textsl{Proof}:
1242: According to theorem \ref{theo27} and arguments preceding it, we only have to prove that $p \ll q$, $p,q \in \mathcal{M}$, implies that there exists a $K^{+}$ causal curve connecting $p$ with $q$. \\*
1243: Let $\psi_{i}:\mathcal{M}_{i} \rightarrow \mathcal{M}$ and $\zeta_{i}: \mathcal{M} \rightarrow \mathcal{M}_{i}$ be mappings which make $(\mathcal{M}_{i},d_{i})$ and $(\mathcal{M},d)$ $(\epsilon_{i},\epsilon_{i})$-close where $\epsilon_{i} \stackrel{i \rightarrow \infty}{\rightarrow} 0$. Choose $p,q \in \mathcal{M}$ such that $d(p,q) > 0$. Let $\epsilon < \frac{d(p,q)}{8} $ and choose $i$ sufficiently large such that $\epsilon_{i} < \alpha(\epsilon)$. Hence, $\left| d_{i}(\zeta_{i}(p), \zeta_{i}(q)) - d(p,q) \right| < \epsilon_{i}$ and $\left| D_{i}(\zeta_{i}(p), \zeta_{i}(q)) - D(p,q) \right| < 4 \epsilon_{i}$. Let $\gamma_{i}$, be a geodesic from $\zeta_{i}(p)$ to $\zeta_{i}(q)$ and consider $\gamma_{i}^{\frac{1}{2}} = \left\{p^{i}_{s} | s=0 \ldots k_{i} \right\}$. Assume that $s^{i}$ is the largest number such that $d_{i}(p^{i}_{s^{i}+1}, \zeta_{i}(q)) > \frac{d_{i}(\zeta_{i}(p),\zeta_{i}(q))}{2}$. Then, for all $s \leq s^{i}$, pick $r^{i}_{s+1}$ such that $\zeta_{i}(q) \gg_i r^{i}_{s+1} \gg_i p^{i}_{s+1}$, $D_{i}(p^{i}_{s+1},r^{i}_{s+1}) \leq \epsilon$ and $d_{i}(p^{i}_{s+1},r^{i}_{s+1}) = \alpha (\epsilon )$. This is possible, since the $\mathcal{C}^{+}_{\alpha}$ property is valid and since $\frac{d_{i}(\zeta_{i}(p),\zeta_{i}(q))}{2} > \frac{d(p,q)-\alpha(\epsilon)}{2} > \frac{7d(p,q)}{16}$. If $d_{i}(p^{i}_{s^{i}+1},p^{i}_{s^{i}+2}) < \alpha(\epsilon)$, then construct in a similar way $r^{i}_{s^{i}+2}$; this is possible since $\frac{5d(p,q)}{16} > \epsilon$. For all $s > s^{i}+1$, define $t^{i}_{s} \ll_i p^{i}_{s}$ such that $D_{i}(t^{i}_{s},p^{i}_{s+1}) \leq \epsilon$ and $d_{i}(t^{i}_{s},p^{i}_{s+1}) = \alpha (\epsilon)$. Obviously, $d_{i}(\zeta_{i}(p),t^{i}_{s}),d_{i}(r^{i}_{s},\zeta_{i}(q)) > \frac{3d(p,q)}{16}$ and $d_{i}(p^{i}_{s},r^{i}_{s+1}), d_{i}(t^{i}_{s}, p^{i}_{s+1}) \geq \alpha \left( \epsilon \right)$. Hence, one can uniquely define sequences of the types $$\left\{\zeta_{i}(p),r^{i}_{1},p^{i}_{1}, r^{i}_{2}, p^{i}_{2}, \ldots , p^{i}_{s^{i}}, r^{i}_{s^{i}+1}, p^{i}_{s^{i}+1}, r^{i}_{s^{i}+2}, t^{i}_{s^{i}+2}, p^{i}_{s^{i}+3}, t^{i}_{s^{i}+3}, \ldots , t^{i}_{k_{i}-1}, \zeta_{i}(q) \right\}$$ and $$\left\{\zeta_{i}(p),r^{i}_{1},p^{i}_{1}, r^{i}_{2}, p^{i}_{2}, \ldots , p^{i}_{s^{i}}, r^{i}_{s^{i}+1}, p^{i}_{s^{i}+1}, p^{i}_{s^{i}+2}, t^{i}_{s^{i}+2}, p^{i}_{s^{i}+3}, t^{i}_{s^{i}+3}, \ldots , t^{i}_{k_{i}-1}, \zeta_{i}(q) \right\}$$ depending on the fact if $d_{i}(p^{i}_{s^{i}+1},p^{i}_{s^{i}+2}) < \alpha(\epsilon)$ or $d_{i}(p^{i}_{s^{i}+1},p^{i}_{s^{i}+2}) \geq \alpha(\epsilon)$ respectively. In general, we have constructed a sequence of the form $(z^{i}_{s})_{s=0}^{2k_{i}-1}$, with the following useful properties:
1244: \begin{itemize}
1245: \item $z^{i}_{0} = \zeta_{i}(p)$ and $z^{i}_{2k_{i}-1} = \zeta_{i}(q)$
1246: \item $\frac{D_{i}(\zeta_{i}(p),\zeta_{i}(q))}{2} \leq D_{i}(z^{i}_{2s}, z^{i}_{2s+1}) \leq \frac{D_{i}(\zeta_{i}(p),\zeta_{i}(q))}{2} + \epsilon$ for $s \leq k^{i}-2$,\\* $\frac{D_{i}(\zeta_{i}(p),\zeta_{i}(q))}{4} \leq D_{i}(z^{i}_{2k^{i}-2}, z^{i}_{2k^{i}-1}) < \frac{3D_{i}(\zeta_{i}(p),\zeta_{i}(q))}{4} + \epsilon$ and \\* $d_{i}(z^{i}_{2s}, z^{i}_{2s+1}) \geq \alpha(\epsilon)$ for all $s \leq k^{i}-1$.
1247: \item $D_{i}(z^{i}_{2s-1}, z^{i}_{2s}) < 2 \epsilon$
1248: \end{itemize}
1249: Hence, the sequence $(\psi_{i}(z^{i}_{s}))_{s=0}^{2k_{i}-1}$ satisfies:
1250: \begin{itemize}
1251: \item $D(\psi_{i}(z^{i}_{0}),p), D(\psi_{i}(z^{i}_{2k_{i}-1}),q) < \alpha(\epsilon)$
1252: \item $\frac{D(p,q)}{2} - 7\epsilon \leq D(\psi_{i}(z^{i}_{2s}), \psi_{i}(z^{i}_{2s+1})) \leq \frac{D(p,q)}{2} + 8\epsilon$ for $s \leq k^{i}-2$, \\* $\frac{D(p,q)}{4} -6 \epsilon < D(\psi_{i}(z^{i}_{2k^{i}-2}), \psi_{i}(z^{i}_{2k^{i}-1})) < \frac{3D(p,q)}{4} + 10\epsilon$ and \\* $d_{i}(\psi_{i}(z^{i}_{2s}), \psi_{i}(z^{i}_{2s+1})) > 0$ for all $s \leq k^{i}-1$
1253: \item $D(\psi_{i}(z^{i}_{2s-1}), \psi_{i}(z^{i}_{2s})) < 6 \epsilon$
1254: \end{itemize}
1255: Hence, for every $n$ such that $\epsilon_{n} < \alpha(d(p,q)/8)$ we can find a
1256: sequence $(\alpha^{n}_{s})_{s=0}^{2k_{n}-1}$ in $\mathcal{M}$
1257: satisfying the above properties\footnote{In the sequel, the reader should keep in
1258: mind that these finite sequences can be extended to infinite ones
1259: by putting everything after $2k_{n}-1$ equal to $q$.}. By using a
1260: diagonalisation argument, we can find a subsequence (which we
1261: label with the same index) such that $k_{n+1} \geq k_{n}$ for all
1262: $n \in \mathbb{N}$ and a sequence
1263: $(\alpha_{s})_{s=0}^{2\sup_{n}k_{n} -1}$ such that $\alpha^{n}_{s}
1264: \stackrel{n \rightarrow \infty}{\rightarrow} \alpha_{s}$ for all
1265: $s \leq 2\sup_{n}k_{n} -1$. Obviously $\sup_{n}k_{n}$ must be
1266: finite, since otherwise we found an infinite sequence of points
1267: which are all a distance greater or equal than $\frac{D(p,q)}{2}$
1268: apart, which is impossible by compactness\footnote{To see this,
1269: notice that the strong distance is increasing along
1270: causal paths.}. Hence, we have found a finite sequence of points
1271: $\beta_{s} \leq \beta_{s+1}$, $s=0 \ldots k$, such that:
1272: \begin{itemize}
1273: \item $\beta_{0}=p$ and $\beta_{k}=q$
1274: \item $\sum_{s=0}^{k-1}d(\beta_{s}, \beta_{s+1}) = d(p,q)$
1275: \item $D(\beta_{s},\beta_{s+1}) = \frac{D(p,q)}{2}$, $s \leq k-2$ and \\* $\frac{D(p,q)}{4} \leq D(\beta_{k-1},q) \leq \frac{3D(p,q)}{4}$.
1276: \end{itemize}
1277: It is possible that for some $s$, $d(\beta_{s}, \beta_{s+1}) = 0$ but these are limits of timelike intervals as follows from the construction. Subdividing each of these approximating timelike intervals and using a compactness argument together with the continuity of $K^{+}$ in the strong topology, one obtains that every two timelike related points are connected by a causal geodesic. $\square$ \\* \\*
1278: This result is, in the author's viewpoint, very encouraging since it shows that all concepts fit nicely together. Notice also that the proof is considerably more difficult than the one in the metric case, where it suffices to use the existence of a midpoint for path metrics.
1279: \section{Compactness of classes of Lorentz spaces.} \label{comp}
1280: To end this chapter, we give some criteria for a collection of Lorentz spaces to
1281: be precompact with respect to the GGH-uniformity.
1282: The ideas presented here can be traced back to Gromov and proofs of the
1283: results at hand can be found in Petersen \cite{Petersen}. Let $(\mathcal{M},d)$
1284: be a Lorentz space, and define (as in Gromov \cite{Gromov})
1285:
1286: \begin{itemize}
1287:
1288: \item $\textrm{Cap}_{\mathcal{M}} ( \epsilon) = $ the maximum number
1289: of disjoint $\frac{\epsilon}{2}$-balls in $(\mathcal{M},D)$.
1290:
1291: \item $\textrm{Cov}_{\mathcal{M}} (\epsilon) = $ the minimum number of
1292: $\epsilon$-balls needed to cover $\mathcal{M}$.
1293:
1294: \end{itemize}
1295: Clearly, $\textrm{Cov}_{\mathcal{M}} (\epsilon ) \leq
1296: \textrm{Cap}_{\mathcal{M}}
1297: (\epsilon)$ and both are decreasing functions of $\epsilon$. What do these
1298: definitions mean? $\textrm{Cov}_{\mathcal{M}} (\epsilon)$ tells us that one
1299: can choose $\textrm{Cov}_{\mathcal{M}} (\epsilon)$ points $p_{i}$ in
1300: $\mathcal{M}$ such that the pair $\left( \left\{ p_{i} \mid i = 1 \ldots
1301: \textrm{Cov}_{\mathcal{M}} (\epsilon)\right\},d\right)$ is $(2 \epsilon,
1302: \epsilon)$-close in the Gromov-Hausdorff metric to $(\mathcal{M},d)$. On the
1303: other hand, suppose that $(\mathcal{M}_{1},d_{1})$ and $(\mathcal{M}_{2} ,
1304: d_{2})$ are $(\epsilon, \delta)$ GGH-close, then we know that
1305: $$
1306: d_{\rm GH} ( (\mathcal{M}_{1},D_{1}),
1307: (\mathcal{M}_{2},D_{2})) \leq \epsilon + {\textstyle\frac{3\delta}{2}}\;,
1308: $$
1309: and therefore one obtains from the triangle inequality that
1310: $$
1311: \textrm{Cov}_{\mathcal{M}_{1}} ( \gamma + 2 \epsilon + 3 \delta)
1312: \leq \textrm{Cov}_{\mathcal{M}_{2}} ( \gamma )
1313: $$
1314: and
1315: $$
1316: \textrm{Cap}_{\mathcal{M}_{1}} (\gamma )
1317: \geq \textrm{Cap}_{\mathcal{M}_{2}} ( \gamma + 4 \epsilon + 6 \delta )
1318: $$
1319: for all $\gamma > 0$. Since we have a quantitative Hausdorff uniformity on the moduli space
1320: $\mathcal{LS}$ with a countable basis around every point, the following two
1321: criteria for compactness are equivalent:
1322:
1323: \begin{itemize}
1324:
1325: \item Every open cover has a finite subcover.
1326:
1327: \item Every sequence has a subsequence which converges to a limit point.
1328:
1329: \end{itemize}
1330:
1331: \begin{theo} For a class $\mathcal{C} \subset \mathcal{LS}$, the
1332: following statements
1333: are equivalent:
1334:
1335: \begin{enumerate}
1336:
1337: \item $\mathcal{C}$ is precompact in $\mathcal{LS}$, i.e., every sequence
1338: has a subsequence that is convergent in $\mathcal{LS}$
1339:
1340: \item There is a function $N: (0 ,\alpha ) \rightarrow
1341: (0, \infty)$ such that $\textrm{Cap}_{\mathcal{M}} ( \epsilon ) \leq N(
1342: \epsilon )$ for all $(\mathcal{M},d) \in \mathcal{C}$.
1343:
1344: \item There is a function $N: (0 ,\alpha ) \rightarrow
1345: (0, \infty)$ such that $\textrm{Cov}_{\mathcal{M}} ( \epsilon ) \leq N(
1346: \epsilon )$ for all $(\mathcal{M},d) \in \mathcal{C}$.
1347:
1348: \end{enumerate}
1349:
1350: \end{theo}
1351:
1352: \noindent\textsl{Proof}: \\*
1353: \noindent $1 \Rightarrow 2)$ If $\mathcal{C}$ is precompact, then
1354: for any $\epsilon > 0$ there exist points $(\mathcal{M}_{1},d_{1})$, ...,
1355: $(\mathcal{M}_{k},d_{k}) \in \mathcal{C}$ such that any $(\mathcal{M},d)$ is
1356: $(\frac{\epsilon}{16}, \frac{\epsilon}{24})$ close to some
1357: $(\mathcal{M}_{i},d_{i})$. Hence, $\textrm{Cap}_{\mathcal{M}}(\epsilon) \leq
1358: \textrm{Cap}_{\mathcal{M}_{i}} (\frac{\epsilon}{2}) \leq \max_{j}
1359: \textrm{Cap}_{\mathcal{M}_{j}} (\frac{\epsilon}{2})$, which clearly proves a
1360: bound for $\textrm{Cap}_{\mathcal{M}}(\epsilon)$ for any $\epsilon > 0$. \\*
1361:
1362: \noindent $2 \Rightarrow 3)$ is obvious. \\*
1363:
1364: \noindent $3 \Rightarrow 1)$ Because of the generalised triangle
1365: inequality, it suffices to show that for any $\epsilon >0$, there
1366: exists a finite collection $\mathcal{A}$ of spaces in
1367: $\mathcal{LS}$ such that any pair $(\mathcal{M},d) \in
1368: \mathcal{C}$ is $(\epsilon, \epsilon)$-close to one of the
1369: elements in $\mathcal{A}$. Observe that for any $(\mathcal{M},d)$
1370: and $\delta > 0$: $tdiam(\mathcal{M}) \leq 2 \delta\,
1371: \textrm{Cov}_{\mathcal{M}} (\delta)$, since $D_{\mathcal{M}}(p,q)
1372: \geq d(p,q)$ for all $p,q \in \mathcal{M}$. The hypothesis implies
1373: the existence of a function $N(\epsilon)$ such that
1374: $\textrm{Cov}_{\mathcal{M}}(\frac{\epsilon}{8}) \leq
1375: N(\frac{\epsilon}{8})$. Hence, every space in $\mathcal{C}$ is
1376: $(\frac{\epsilon}{4}, \frac{\epsilon}{8})$-close to a finite space
1377: with $N(\frac{\epsilon}{8})$ elements, such that the timelike
1378: distance between any two points does not exceed the value
1379: $\frac{\epsilon}{4}\, N(\frac{\epsilon}{8})$. The Lorentz metric
1380: on such a finite space consists of a square matrix $(d_{ij})_{1
1381: \leq i,j \leq N(\epsilon/8)}$ such that $0 \leq d_{ij} \leq
1382: \frac{\epsilon}{4}\, N(\frac{\epsilon}{8})$. Obviously, one can
1383: find a finite collection $\mathcal{A}$ of Lorentz spaces with
1384: $N(\frac{\epsilon}{8})$ elements such that any of the $(d_{ij})_{1
1385: \leq i,j \leq N(\epsilon/8)}$ is $(\frac{\epsilon}{4}, 0)$-close
1386: to some element of $\mathcal{A}$. Hence, all spaces
1387: $(\mathcal{M},d) \in \mathcal{C}$ are $(\frac{\epsilon}{2},
1388: \frac{5\epsilon}{8})$-close to some element of $\mathcal{A}$,
1389: which concludes the proof. \hfill$\square$ \\* \\* We show that
1390: the covering property with covering function $N$ is stable under
1391: GGH-convergence provided that $N$ is continuous (cfr.\ the
1392: $\mathcal{C}^{\pm}_{\alpha}$ properties in section \ref{prop}).
1393:
1394: \begin{theo}
1395:
1396: Let $\mathcal{C}(N(\epsilon))$ be the collection of pairs $(\mathcal{M},d) \in
1397: \mathcal{LS}$ such that $\textrm{Cov}_{\mathcal{M}}(\epsilon) \leq
1398: N(\epsilon)$ for all $\epsilon > 0$; suppose $N$ is continuous. Then,
1399: $\mathcal{C}(N(\epsilon))$ is compact.
1400:
1401: \end{theo}
1402:
1403: \noindent\textsl{Proof}:\\*
1404: We already know that $\mathcal{C}(N(\epsilon))$ is precompact, hence
1405: suppose $(\mathcal{M}_{i},d_{i}) \stackrel{i \rightarrow \infty}{\rightarrow}
1406: (\mathcal{M},d)$ in the GGH-uniformity, then with
1407: $\alpha_{i} \stackrel{i \rightarrow \infty}{\rightarrow} 0$ such that
1408: $(\mathcal{M}_{i},d_{i})$ and $(\mathcal{M},d)$ are $(\alpha_{i},
1409: \alpha_{i})$-close, we obtain that
1410: $$
1411: \textrm{Cov}_{ \mathcal{M}}(\epsilon) \leq \textrm{Cov}_{\mathcal{M}_{i}}
1412: ( \epsilon - 5 \alpha_{i} ) \leq N( \epsilon - 5 \alpha_{i})\;.
1413: $$
1414: The continuity of $N$ concludes the proof. \hfill$\square$
1415:
1416: