1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: % Spatial and null infinity via advanced and retarded conformal factors
3: % Sean A. Hayward
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5:
6: \documentclass[twocolumn,showpacs,preprintnumbers]{revtex4}
7: \usepackage{graphicx}
8: \usepackage{amssymb}
9:
10: %Note to publisher:
11: %this is a crude macro for an Anglo-Saxon character; improvements welcome.
12: \def\thorn{I\kern-0.4em\raise0.35ex\hbox{\it o}}
13:
14: \begin{document}
15: \title{Spatial and null infinity via advanced and retarded conformal factors}
16: \author{Sean A. Hayward}
17: \affiliation{Department of Science Education, Ewha Womans University,
18: Seodaemun-gu, Seoul 120-750, Korea\\
19: {\tt hayward@mm.ewha.ac.kr}}
20: \date{14th September 2003}
21:
22: \begin{abstract}
23: A new approach to space-time asymptotics is presented, refining Penrose's idea
24: of conformal transformations with infinity represented by the conformal
25: boundary of space-time. Generalizing examples such as flat and Schwarzschild
26: space-times, it is proposed that the Penrose conformal factor be a product of
27: {\em advanced and retarded conformal factors}, which asymptotically relate
28: physical and conformal null (light-like) coordinates and vanish at future and
29: past null infinity respectively, with both vanishing at spatial infinity. A
30: correspondingly refined definition of asymptotic flatness at both spatial and
31: null infinity is given, including that the conformal boundary is locally a
32: light cone, with spatial infinity as the vertex. It is shown how to choose the
33: conformal factors so that this asymptotic light cone is locally a metric light
34: cone. The theory is implemented in the spin-coefficient (or null-tetrad)
35: formalism by a simple joint transformation of the spin-metric and spin-basis
36: (or metric and tetrad). The advanced and retarded conformal factors may be used
37: as expansion parameters near the respective null infinity, together with a
38: dependent expansion parameter for both spatial and null infinity, essentially
39: inverse radius. Asymptotic regularity conditions on the spin-coefficients are
40: proposed, based on the conformal boundary locally being a smoothly embedded
41: metric light cone. These conditions ensure that the Bondi-Sachs energy-flux
42: integrals of ingoing and outgoing gravitational radiation decay at spatial
43: infinity such that the total radiated energy is finite, and that the
44: Bondi-Sachs energy-momentum has a unique limit at spatial infinity, coinciding
45: with the uniquely rendered ADM energy-momentum.
46: \end{abstract}
47: \pacs{04.20.Ha, 04.20.Gz, 04.30.Nk}
48: \maketitle
49:
50: \section{Introduction}
51: Space-time asymptotics, the study of isolated gravitational systems at large
52: distances, is one of the theoretical pillars of General Relativity, whereby one
53: can define physically important quantities which are difficult to capture in
54: general, such as gravitational radiation, the energy flux of gravitational
55: radiation and the active gravitational mass-energy of the system. The discovery
56: of the Bondi-Sachs energy-loss equation \cite{B,BBM,S}, relating the change in
57: total mass-energy to the energy flux of outgoing gravitational radiation, just
58: as for electromagnetic radiation, marked a transition from an epoch where some
59: argued that Einstein's original prediction of gravitational radiation rested on
60: mathematical artefacts, to the present epoch where there is significant
61: investment in expectations of detecting gravitational radiation from
62: astrophysical sources.
63:
64: The early work on space-time asymptotics was brilliantly reformulated by
65: Penrose \cite{P,PR}, who realized how to handle infinite distances and times in
66: a mathematically finite way. The space-time metric is mapped to a multiple of
67: itself, the factor tending to zero in such a way that the new metric is finite
68: and well behaved at the physical space-time infinity. The mapping is described
69: as conformal since it preserves space-time angles and therefore causal
70: relations, while changing distances and durations. In the conformal picture,
71: infinity becomes a finite boundary where one can derive exact formulas,
72: encapsulating physical laws which were otherwise approximate or limiting.
73:
74: Penrose's conformal theory rapidly became the standard framework for space-time
75: asymptotics at null (light-like) infinity, and was quite thoroughly implemented
76: using the spin-coefficient or null-tetrad formalism \cite{PR,NP,NU,GHP,NT}.
77: However, it does not cover spatial infinity, described initially by the ADM
78: method \cite{ADM} and by a spatial version of Penrose's method by Geroch
79: \cite{G}. The essential problem is to describe spatial and null infinity in a
80: unified way. Ashtekar and coworkers \cite{AH,AM,A} proposed and investigated
81: such a definition, realizing that spatial infinity is generally a directional
82: singularity of the conformal metric. However, there still seems to be no
83: practical calculational formalism. Some physically important issues are whether
84: the gravitational radiation decays near spatial infinity such that the
85: Bondi-Sachs energy exists and coincides with the ADM energy at spatial
86: infinity, and whether initial data on an asymptotically flat spatial
87: hypersurface (or past null infinity) determines final data at future null
88: infinity. For a more recent perspective, see e.g.\ Friedrich \cite{F}.
89:
90: This article presents a quite simple, natural unification of spatial and null
91: infinity, based on a re-examination of Penrose's original insights. The key new
92: idea is that Penrose's conformal factor should be a product of advanced and
93: retarded conformal factors, which do the work of the conformal factor at future
94: and past null infinity respectively, while cooperating at spatial infinity so
95: that it is the vertex of a light cone, generating null infinity. These factors,
96: squared, differentially relate physical and conformal null coordinates near the
97: respective null infinity, up to the relevant coordinate freedom.
98:
99: The basic idea is most easily apprehended in examples, described in \S II. \S
100: III gives a definition of asymptotic flatness intended to encapsulate the new
101: refinements, and shows that conformal infinity is locally a metric light cone,
102: thereby determining standard coordinates. \S IV shows how to implement the
103: theory in the spin-coefficient (or null-tetrad) formalism, using a simple joint
104: transformation of the spin-metric and spin-basis (or metric and null tetrad).
105: \S V identifies appropriate expansion parameters and proposes asymptotic
106: regularity conditions at both null and spatial infinity, based on the
107: asymptotic light cone being smoothly embedded. \S VI uses the Hawking
108: quasi-local energy \cite{H} to study the Bondi-Sachs energy flux (of ingoing
109: and outgoing gravitational radiation) and total energy at both null and spatial
110: infinity, and the ADM energy. \S VII concludes.
111:
112: \section{Advanced and retarded conformal factors}
113: For flat space-time, the line element may be written as
114: \begin{equation}
115: ds^2=r^2dS^2+dr^2-dt^2
116: \end{equation}
117: where $dS^2$ is a line element for the unit sphere. In terms of dual-null (or
118: characteristic) coordinates
119: \begin{equation}
120: 2\xi^\pm=t\pm r
121: \end{equation}
122: it becomes
123: \begin{equation}
124: ds^2=(\xi^+-\xi^-)^2dS^2-4d\xi^+d\xi^-.
125: \end{equation}
126: To approach infinity, the physical null coordinates $\xi^\pm$ are transformed
127: to conformal null coordinates $\psi^\pm$ by
128: \begin{equation}
129: \xi^\pm=\tan\psi^\pm.
130: \end{equation}
131: This describes sterographic projection, used to map the Earth on flat paper: if
132: one projects from the north pole of a unit sphere onto the equatorial plane,
133: with $\psi$ the angle from the vertical, then $\xi=\tan\psi$ is the distance
134: from the centre of a point in the plane. Thus the infinite plane is mapped into
135: the compact sphere, with the north pole representing infinity; the sphere
136: describes the whole plane, plus its infinity. The formula also occurs in
137: complex analysis as a way of using the Riemann sphere to represent the Argand
138: plane, plus a point at infinity.
139:
140: The novel point concerns the derivatives occurring in the null coordinate
141: transformations,
142: \begin{equation}
143: (\omega^\pm)^2=\frac{d\psi^\pm}{d\xi^\pm}=\cos^2\psi^\pm
144: \end{equation}
145: which vanish at the relevant infinity. The line element written in $\psi^\pm$
146: coordinates therefore contains a term $4d\psi^+d\psi^-/(\omega^+\omega^-)^2$.
147: To obtain a metric which is regular at infinity, one can multiply by the
148: discrepancy, the square of
149: \begin{equation}\label{Omega}
150: \Omega=\omega^+\omega^-.
151: \end{equation}
152: Then
153: \begin{equation}
154: \Omega^2ds^2=\sin^2(\psi^+-\psi^-)dS^2-4d\psi^+d\psi^-.
155: \end{equation}
156: A final transformation
157: \begin{equation}\label{rhotau}
158: 2\psi^\pm=\tau\mp\rho
159: \end{equation}
160: puts this in the form
161: \begin{equation}
162: \Omega^2ds^2=\sin^2\rho\,dS^2+d\rho^2-d\tau^2
163: \end{equation}
164: which is the standard line element of the Einstein static universe, a metric
165: $S^3\times R$. The physical space-time metric $g$ has been mapped to the
166: conformal metric $\Omega^2g$ and the physical space-time manifold has been
167: mapped to the conformal coordinate range $\psi^\pm\in(-\pi/2,\pi/2)$. Thus
168: physical infinity has been mapped to the boundary of this region. This
169: conformal boundary consists of two null hypersurfaces connected by three
170: points: future null infinity $\Im^+$, $\psi^+=\pi/2$,
171: $\psi^-\in(-\pi/2,\pi/2)$; past null infinity $\Im^-$, $\psi^-=-\pi/2$,
172: $\psi^+\in(-\pi/2,\pi/2)$; future temporal infinity $i^+$,
173: $\psi^+=\psi^-=\pi/2$; past temporal infinity $i^-$, $\psi^+=\psi^-=-\pi/2$;
174: and spatial infinity $i^0$, $\psi^+=-\psi^-=\pm\pi/2$. The two null
175: hypersurfaces locally have the structure of light cones, with the three points
176: each being vertices. This initially bizarre sixties spacewarp led Penrose
177: \cite{P,PR} to the definition of conformal infinity as essentially $\Omega=0$,
178: $\nabla\Omega\not=0$, the physical metric $g$ admitting a well behaved
179: conformal metric $\Omega^2g$, $\Omega$ being called the conformal factor.
180:
181: Nevertheless, there is more structure here. Firstly, spatial infinity is the
182: vertex of a conformal light cone consisting of null infinity. Secondly,
183: $\Im^\pm$ are given by $\omega^\pm=0$, $\omega^\mp\not=0$, with
184: $\omega^+=\omega^-=0$ at spatial infinity. Clearly there is more information in
185: the sub-factors $\omega^\pm$ than in the Penrose factor $\Omega$ alone. This
186: structure will be used to refine Penrose's definition of asymptotic flatness.
187: Henceforth $\omega^+$ and $\omega^-$ will be called the {\em advanced and
188: retarded conformal factors} respectively.
189:
190: As a second example, the Schwarzschild space-time with mass $M$ in standard
191: coordinates is
192: \begin{equation}
193: ds^2=r^2dS^2+(1-2M/r)^{-1}dr^2-(1-2M/r)dt^2.
194: \end{equation}
195: This can be written in dual-null coordinates
196: \begin{equation}
197: 2\xi^\pm=t\pm r_*
198: \end{equation}
199: where
200: \begin{equation}
201: r_*=r+2M\ln(r-2M)
202: \end{equation}
203: rescales $r$ such that $dr/dr_*=1-2M/r$. Then
204: \begin{equation}
205: ds^2=r^2dS^2-(1-2M/r)4d\xi^+d\xi^-
206: \end{equation}
207: where $r$ is implicitly determined as a function of $r_*=\xi^+-\xi^-$. As
208: before, the physical null coordinates $\xi^\pm$ may be transformed to conformal
209: null coordinates $\psi^\pm$ by, for instance, simple inversion:
210: \begin{equation}
211: \xi^\pm=-1/\psi^\pm.
212: \end{equation}
213: This is valid only in a neighbourhood of spatial infinity, but is simple enough
214: to constitute a local canonical transformation. The derivatives occurring in
215: the null coordinate transformations are
216: \begin{equation}
217: (\omega^\pm)^2=\frac{d\psi^\pm}{d\xi^\pm}=(\psi^\pm)^2
218: \end{equation}
219: and one can fix $\omega^\pm=\mp\psi^\pm$ so that $\omega^\pm>0$ in the physical
220: space-time, $\xi^\pm$ and $\psi^\pm$ being future-pointing. As before, the
221: metric in $\psi^\pm$ coordinates can be regularized by multiplying by the
222: square of $\Omega=\omega^+\omega^-$. Then
223: \begin{equation}
224: \Omega^2ds^2=(\psi^+\psi^-r)^2dS^2-4(1-2M/r)d\psi^+d\psi^-.
225: \end{equation}
226: Transforming as before by (\ref{rhotau}) puts this in the form
227: \begin{equation}
228: \Omega^2ds^2=(\rho r/r_*)^2dS^2+(1-2M/r)(d\rho^2-d\tau^2).
229: \end{equation}
230: Since $r_*/r\to1$ as $r\to\infty$, the conformal metric becomes flat at
231: infinity, with $\rho$ playing the role of conformal radius:
232: \begin{equation}
233: \Omega^2ds^2\to\rho^2dS^2+d\rho^2-d\tau^2.
234: \end{equation}
235: Thus infinity consists of a locally metric light cone with spatial infinity as
236: the vertex.
237:
238: Space-time has been turned inside out: spatial infinity has become a point and
239: the physical space-time lies outside its light cone, as depicted in
240: Fig.~\ref{fig}. In more detail, the original region $\xi^+>0$, $\xi^-<0$ has
241: been mapped to $\psi^+<0$, $\psi^->0$, or $\omega^\pm>0$. The conformal
242: boundary again locally consists of $\Im^\pm$, where $\omega^\pm=0$,
243: $\omega^\mp\not=0$, and spatial infinity $i^0$, where $\omega^+=\omega^-=0$.
244: The case $M=0$ was not excluded, so this also provides a conformal
245: transformation of flat space-time such that infinity is locally a metric light
246: cone at spatial infinity. The key points are the identification of the advanced
247: and retarded conformal factors $\omega^\pm$ as differentially relating physical
248: and conformal null coordinates, and their behaviour at null and spatial
249: infinity.
250:
251: \begin{figure}
252: \includegraphics[width=5cm,height=7cm,angle=0]{inf1.eps}
253: \caption{The light cone at infinity, depicting the advanced and retarded
254: conformal factors $\omega^\pm$. The physical space-time ($\omega^+>0$,
255: $\omega^->0$) lies outside this asymptotic light cone. Radiation propagates in
256: from past null infinity $\Im^-$ ($\omega^-=0$) and out to future null infinity
257: $\Im^+$ ($\omega^+=0$). Spatial infinity $i^0$ ($\omega^+=\omega^-=0$) is the
258: vertex. For non-zero total mass, beware the singularity beyond infinity.}
259: \label{fig}
260: \end{figure}
261:
262: \section{The light cone at infinity}
263: The above insights are now converted into a definition of asymptotic flatness
264: at both null and spatial infinity. In this section alone, $g$ will be used for
265: the conformal metric and $\tilde g$ for the physical metric. Quantities without
266: tildes, e.g.\ $\nabla$, similarly refer to the conformal metric.
267:
268: In this article, a space-time $(\tilde M,\tilde g)$ will be said to be
269: asymptotically flat if the following conditions hold. (i) There exists a
270: space-time $(M,g)$ with boundary $\Im=\partial M$ and functions $\omega^\pm>0$
271: on $\tilde M$ such that $M=\tilde M\cup\Im$ and $g=(\omega^+\omega^-)^2\tilde
272: g$ in $\tilde M$. (ii) $\Im$ is locally a light cone with respect to $g$, with
273: vertex denoted by $i^0$ and future and past open cones denoted by $\Im^+$ and
274: $\Im^-$ respectively. (iii) $\omega^\pm=0$, $\omega^\mp\not=0$,
275: $\nabla\omega^\pm\not=0$ on $\Im^\pm$, and
276: $g^{-1}(\nabla\omega^+,\nabla\omega^-)\not=0$ on $\Im$. (iv) ($g,\omega^\pm$)
277: are $C^k$ on $M-i^0$, with $k=3$ sufficing for the purposes of this article.
278: Differentiability at $i^0$ is deferred to \S V.
279:
280: One may call $\Im=\Im^+\cup\Im^-\cup i^0$ (locally) the asymptotic light cone,
281: or more prosaically, the {\em light cone at infinity} \cite{AH}. The definition
282: can be easily seen to recover the basic conditions of Penrose's definition of
283: asymptotic simplicity \cite{P,PR}, namely $\Omega=0$ and $\nabla\Omega\not=0$
284: on $\Im^\pm$, with the Penrose factor $\Omega$ as before (\ref{Omega}). The
285: causal restriction on (plain or weak) asymptotic simplicity has been replaced
286: with the light-cone condition. Note by continuity that $\omega^+=\omega^-=0$ at
287: $i^0$. The differentiability conditions at spatial infinity are notoriously
288: awkward: if $g$ is $C^1$ then the total (ADM) energy vanishes, while if $g$ is
289: only $C^0$, the total energy is generally ill-defined \cite{AH,AM,A}. Readers
290: unable to proceed otherwise can provisionally assume the Ashtekar-Hansen
291: differentiability conditions at $i^0$. Essentially, a non-zero mass forces
292: radial derivatives of the conformal metric to be discontinuous at spatial
293: infinity, even for the Schwarzschild metric, while angular derivatives should
294: be much smoother. A geometrical way to understand this stems from the fact that
295: the positive-mass Schwarzschild space-time can be conformally extended through
296: $\Im^\pm$ to smoothly include a negative-mass Schwarzschild space-time whose
297: $(\Im^\mp,i^\mp)$ coincide with the physical $(\Im^\pm,i^0)$ \cite{PR}. Since
298: the negative-mass Schwarzschild space-time has a naked curvature singularity
299: extending from its $i^-$ to $i^+$, this {\em singularity beyond infinity}
300: (Fig.~\ref{fig}) piercing physical $i^0$ can be understood as forcing spatial
301: infinity to be a directional singularity of the conformal metric.
302:
303: The analysis of Geroch \cite{G} for $\Im^\pm$, summarized together with the
304: Ashtekar-Hansen analysis by Wald \cite{W}, can be taken as a guide to the
305: following refinements. Firstly, the definition allows considerable gauge
306: freedom in the conformal factors $\omega^\pm$, namely
307: $\omega^\pm\mapsto\alpha_\pm\omega^\pm$ for any $C^k$ functions $\alpha_\pm>0$
308: on $M-i^0$. This gauge freedom allows the conformal metric of $\Im$ to be
309: locally fixed as that of a metric cone, as follows. The starting point is a
310: standard expression for the physical Ricci tensor $\tilde R$ in terms of the
311: conformal Ricci tensor $R$:
312: \begin{equation}\label{Ricci}
313: \Omega\tilde R=\Omega R+2\nabla\otimes\nabla\Omega
314: +(\nabla^2\Omega-3\Omega^{-1}g^{-1}(\nabla\Omega,\nabla\Omega))g.
315: \end{equation}
316: Applying either the vacuum Einstein equation $\tilde R=0$ or the full Einstein
317: equation with suitable fall-off of the matter fields, one sees that
318: $\Omega^{-1}g^{-1}(\nabla\Omega,\nabla\Omega)$ is smooth ($C^{k-2}$) at
319: $\Im^\pm$ and therefore that $\nabla\Omega$ is null there. Further, expanding
320: \begin{eqnarray}
321: &&\frac{g^{-1}(\nabla\Omega,\nabla\Omega)}{\Omega}
322: =2g^{-1}(\nabla\omega^+,\nabla\omega^-)+{}\\&&\qquad\qquad\quad
323: \frac{\omega^-}{\omega^+}g^{-1}(\nabla\omega^+,\nabla\omega^+)
324: +\frac{\omega^+}{\omega^-}g^{-1}(\nabla\omega^-,\nabla\omega^-)\nonumber
325: \end{eqnarray}
326: and using the above definition, one finds that
327: $(\omega^\mp/\omega^\pm)g^{-1}(\nabla\omega^\pm,\nabla\omega^\pm)$ are smooth
328: at $\Im^\pm$ and therefore that the 1-forms
329: \begin{equation}
330: n^\pm=\pm2\nabla\omega^\pm
331: \end{equation}
332: are null at $\Im^\pm$, where the signs are chosen so that the vectors
333: $g^{-1}(n^\pm)$ are future pointing, taking the choice $\omega^\pm>0$ in
334: $\tilde M$ and the $({-}{+}{+}{+})$ metric convention. The gauge freedom
335: $\omega^\pm\mapsto\alpha_\pm\omega^\pm$ can then be used to set
336: $g^{-1}(\nabla\omega^\pm,\nabla\omega^\pm)/\omega^\pm=0$ on $\Im^\pm$, by an
337: argument given by Geroch for $\Omega$. This still leaves $\omega^\mp$ free on
338: $\Im^\pm$. The normalization condition, which now reads
339: $g^{-1}(\nabla\omega^+,\nabla\omega^-)>0$ on $\Im$, expresses that the
340: hypersurfaces of constant $\omega^\pm$ intersect $\Im^\pm$ transversely; then
341: the gauge freedom allows one to choose them to intersect in a null direction,
342: $g^{-1}(\nabla\omega^\mp,\nabla\omega^\mp)=0$ on $\Im^\pm$. Thus we have two
343: independent null directions $g^{-1}(n^\pm)$ at $\Im^+$ and $\Im^-$. The local
344: null rescaling freedom of a null hypersurface can then be used to further
345: adjust $\omega^\mp$ on $\Im^\pm$ so that the null directions are relatively
346: normalized: $g^{-1}(\nabla\omega^+,\nabla\omega^-)=1/2$ on $\Im^\pm$. All the
347: above conditions extend to $i^0$ by continuity. Using $\approx$ to denote
348: equality on $\Im$ in a neighbourhood of $i^0$, the gauge conditions are
349: \begin{eqnarray}\label{gauge}
350: g^{-1}(\nabla\omega^\pm,\nabla\omega^\pm)/\omega^\pm&\approx&0\\
351: g^{-1}(\nabla\omega^+,\nabla\omega^-)&\approx&1/2.
352: \end{eqnarray}
353: Together they imply that
354: \begin{equation}
355: g^{-1}(\nabla\Omega,\nabla\Omega)/\Omega\approx1
356: \end{equation}
357: which shows that $\Omega$ is spatial in a neighbourhood of $\Im$. The gauge
358: condition usually chosen for $\Im^+$ is that this quantity vanishes.
359:
360: Taking the trace and rearranging, the curvature relation (\ref{Ricci}) then
361: implies
362: \begin{equation}
363: 2\nabla\otimes\nabla\Omega\approx g
364: \end{equation}
365: which again differs from the usual relation $\nabla\otimes\nabla\Omega=0$ on
366: $\Im^+$, which would imply that $\Im^+$ be a metric cylinder. Instead it
367: implies that $\Im$ is a metric cone, as follows. First note that there are now
368: preferred spatial sections of $\Im^\pm$, given by constant $\omega^\mp$.
369: Expanding the above relation,
370: \begin{equation}
371: g\approx2\omega^+\nabla\otimes\nabla\omega^-
372: +2\omega^-\nabla\otimes\nabla\omega^++4\nabla\omega^+\otimes\nabla\omega^-.
373: \end{equation}
374: The last term gives the part of the metric normal to the spatial sections of
375: $\Im$, while the first two terms are related to the expansions $\theta_\mp$ and
376: shears $\sigma_\mp$ of $\Im^\pm$ by $\bot2\nabla\otimes n^\pm=\theta_\mp
377: h+\sigma_\mp$ on $\Im^\pm$, where $h$ is the metric of the spatial sections and
378: $\bot$ denotes projection by $h$. Since
379: \begin{equation}
380: g\approx h+4\nabla\omega^+\otimes\nabla\omega^-
381: \end{equation}
382: one reads off $\theta_\mp=\pm2/\omega^\mp$, $\sigma_\mp=0$ on $\Im^\pm$.
383: However, this describes a metric light cone, as follows. The flat metric
384: \begin{equation}
385: ds^2=\rho^2dS^2+d\rho^2-d\tau^2
386: \end{equation}
387: in dual-null coordinates (\ref{rhotau}) is
388: \begin{equation}
389: ds^2=(\psi^--\psi^+)^2dS^2-4d\psi^+d\psi^-
390: \end{equation}
391: where the signs of $\psi^\pm$ have been chosen so that they are
392: future-pointing. Then $\sigma_\mp=0$ and
393: $\theta_\mp=2\partial_\mp\rho/\rho=2/\psi^\mp$ on $\Im^\pm$. Since the
394: expansions and shears, including vertex conditions, characterize the intrinsic
395: geometry of the light cone, this identifies the conformal metric with the flat
396: metric at $\Im$ if
397: \begin{equation}
398: \psi^\pm\approx\mp\omega^\pm.
399: \end{equation}
400: In general, $\psi^\pm$ can be taken as null coordinates with this gauge choice
401: at $\Im$. In summary, the conformal metric at $\Im$ has been locally fixed as
402: that of a metric light cone.
403:
404: \section{Formalisms}
405: The most developed implementation of conformal infinity involves the
406: spin-coefficient formalism, also due to Penrose and coworkers \cite{NP,GHP},
407: which provides the most elegant formulation of various issues such as the Sachs
408: ``peeling'' of the gravitational field near $\Im^+$ \cite{S,P,PR}. Henceforth
409: this formalism is employed, though an attempt is made to render the description
410: accessible to the uninitiated. Alternatively, the null-tetrad formalism
411: suffices for most purposes.
412:
413: Henceforth equation numbers in triples will indicate equations in Penrose \&
414: Rindler \cite{PR}. It is convenient to modify some notation and summarize, as
415: follows. The basic geometrical object is a 2-form $\varepsilon$ (2.5.2) acting
416: on complex 2-spinors (spin-vectors), here called the {\em spin-metric}. It can
417: be regarded as a square root of the metric, which is expressed as a direct
418: product $g=\varepsilon\circ\bar\varepsilon$, where the bar denotes the complex
419: conjugate and the exact meaning of the product is given in abstract index
420: notation (3.1.9). An ordered pair of non-parallel spin-vectors $(o,\iota)$ is
421: here called a {\em spin-basis}. Expressing vectors as spin-vector dyads, this
422: defines a null tetrad, a vector basis (3.1.14):
423: \begin{equation}
424: (l,m,\bar m,l')=(o\bar o,o\bar\iota,\iota\bar o,\iota\bar\iota).
425: \end{equation}
426: Then $(l,l')$ are real null vectors, whereas $m$ is a complex vector encoding
427: transverse spatial vectors. In terms of the complex normalization factor
428: (2.5.46)
429: \begin{equation}
430: \chi=\varepsilon(o,\iota)
431: \end{equation}
432: one finds
433: \begin{equation}
434: g(l,l')=-g(m,\bar m)=\chi\bar\chi
435: \end{equation}
436: with the other eight independent (symmetric) contractions of the tetrad vectors
437: vanishing. Here the metric convention has switched to $({+}{-}{-}{-})$, as is
438: regrettably standard in the spin-coefficient formalism. Some conventional
439: factors of $\sqrt2$ are also entailed compared to the previous sections, e.g.\
440: physical null coordinates $(\xi,\xi')=\sqrt2(\xi^+,\xi^-)$ and conformal null
441: coordinates $(\psi,\psi')=\sqrt2(\psi^+,\psi^-)$.
442:
443: The inverse metric can be written
444: \begin{equation}
445: g^{-1}=\frac2{\chi\bar\chi}(l\otimes l'-m\otimes\bar m)
446: \end{equation}
447: where $\otimes$ denotes the symmetric tensor product. The dual basis of 1-forms
448: (4.13.32) will be denoted by
449: \begin{equation}
450: (n,w,\bar w,n')=g(l',-\bar m,-m,l)
451: \end{equation}
452: so that
453: \begin{equation}
454: n(l)=w(m)=\bar w(\bar m)=n'(l')=1
455: \end{equation}
456: with the other twelve such contractions vanishing. Then the metric can be
457: written
458: \begin{equation}
459: g=2\chi\bar\chi(n\otimes n'-w\otimes\bar w).
460: \end{equation}
461: The tetrad covariant derivative operators (4.5.23) are denoted by
462: \begin{equation}
463: (D,\delta,\delta',D')=(\nabla_l,\nabla_m,\nabla_{\bar m},\nabla_{l'}).
464: \end{equation}
465: The complex spin-coefficients
466: $(\kappa,\sigma,\rho,\tau,\varepsilon,\beta,\alpha,\gamma)$ and
467: $(\kappa',\sigma',\rho',\tau',\varepsilon',\beta',\alpha',\gamma')$ encode the
468: Ricci rotation coefficients and can be defined by (4.5.26--27) as
469: \begin{eqnarray}
470: D(o,\iota)&=&(\varepsilon o-\kappa\iota,\gamma'\iota-\tau'o)\\
471: \delta(o,\iota)&=&(\beta o-\sigma\iota,\alpha'\iota-\rho'o)\\
472: \delta'(o,\iota)&=&(\alpha o-\rho\iota,\beta'\iota-\sigma'o)\\
473: D'(o,\iota)&=&(\gamma o-\tau\iota,\varepsilon'\iota-\kappa'o).
474: \end{eqnarray}
475: Other notation will be taken as in Penrose \& Rindler \cite{PR}.
476:
477: It is convenient to choose null coordinates $(\xi,\xi')$ such that
478: $(n,n')=(d\xi,d\xi')$, which implies the dual-null gauge conditions \cite{bhs}
479: \begin{eqnarray}\label{dualnull0}
480: &&0=\kappa=\rho-\bar\rho=\varepsilon+\bar\varepsilon=\tau-\bar\alpha-\beta\\
481: \label{dualnull1} &&0=\kappa'=\rho'-\bar\rho'=\varepsilon'+\bar\varepsilon'
482: =\tau'-\bar\alpha'-\beta'.
483: \end{eqnarray}
484: It is also convenient to use Penrose's complex stereographic coordinate
485: (1.2.10)
486: \begin{equation}
487: \zeta=e^{i\phi}\cot(\theta/2)
488: \end{equation}
489: where $(\theta,\phi)$ are standard spherical polar coordinates. Then $(w,\bar
490: w)=(d\bar\zeta/P,d\zeta/\bar P)$, (4.14.29), where (4.14.31)
491: \begin{equation}
492: P=\delta\bar\zeta.
493: \end{equation}
494: In summary, we have preferred coordinates $(\xi,\zeta,\bar\zeta,\xi')$ such
495: that the 1-form basis is
496: \begin{equation}
497: (n,w,\bar w,n')=(d\xi,d\bar\zeta/P,d\zeta/\bar P,d\xi').
498: \end{equation}
499: The above quantities will refer to the physical metric $g$ and corresponding
500: quantities for the conformal metric $\hat g$ will be denoted by hats.
501:
502: The physical coordinates are now to be transformed to conformal coordinates
503: $(\psi,\zeta,\bar\zeta,\psi')$, such that the angular coordinates
504: $\hat\zeta=\zeta$ are preserved, while the null coordinates transform to
505: $(\hat\xi,\hat\xi')=(\psi(\xi),\psi'(\xi'))$, with derivatives
506: \begin{equation}
507: (\omega^2,\omega'^2)=\left(\frac{d\psi}{d\xi},\frac{d\psi'}{d\xi'}\right).
508: \end{equation}
509: Taking $\xi$ as an advanced (outgoing) null coordinate and $\xi'$ as a retarded
510: (ingoing) null coordinate, this means that $\omega$ and $\omega'$ are the
511: advanced and retarded conformal factors respectively, assumed positive in the
512: space-time. For the canonical example of inversion,
513: $(\psi,\psi')=(-2/\xi,-2/\xi')$ with the current convention, yielding
514: $(\omega,\omega')=(-\psi,\psi')/\sqrt2$. The Penrose conformal factor is
515: \begin{equation}
516: \Omega=\omega\omega'
517: \end{equation}
518: and the physical spin-metric $\varepsilon$ is correspondingly transformed to
519: the conformal spin-metric
520: \begin{equation}\label{spin}
521: \hat\varepsilon=\omega\omega'\varepsilon
522: \end{equation}
523: and the physical metric $g$ to the conformal metric $\hat
524: g=\hat\varepsilon\circ\hat{\bar\varepsilon}$:
525: \begin{equation}\label{metric}
526: \hat g=(\omega\omega')^2g.
527: \end{equation}
528: By comparing the explicit forms
529: \begin{eqnarray}\label{g}
530: g&=&2|\chi|^2\left(d\xi\otimes d\xi'-\frac{d\zeta\otimes
531: d\bar\zeta}{|P|^2}\right)\\
532: \hat g&=&2|\hat\chi|^2\left(d\psi\otimes d\psi'-\frac{d\zeta\otimes
533: d\bar\zeta}{|\hat P|^2}\right)
534: \end{eqnarray}
535: one reads off
536: \begin{eqnarray}\label{chi}
537: \hat\chi&=&\chi\\\label{P}\hat P&=&P/\omega\omega'.
538: \end{eqnarray}
539: Then the conformal 1-form basis
540: \begin{equation}
541: (\hat n,\hat w,\hat{\bar w},\hat n')=(d\psi,d\bar\zeta/\hat P,d\zeta/\hat{\bar
542: P},d\psi')
543: \end{equation}
544: is related to the physical 1-form basis by
545: \begin{equation}
546: (\hat n,\hat w,\hat{\bar w},\hat
547: n')=(\omega^2n,\omega\omega'w,\omega\omega'\bar w,\omega'^2n').
548: \end{equation}
549: Since
550: \begin{equation}
551: \hat g^{-1}=g^{-1}/(\omega\omega')^2
552: \end{equation}
553: the physical null tetrad
554: \begin{equation}
555: (l,m,\bar m,l')=g^{-1}(n',-\bar w,-w,n)
556: \end{equation}
557: and the conformal null tetrad
558: \begin{equation}
559: (\hat l,\hat m,\hat{\bar m},\hat l')=\hat g^{-1}(\hat n',-\hat{\bar w},-\hat
560: w,\hat n)
561: \end{equation}
562: are then related by
563: \begin{equation}\label{tetrad}
564: (\hat l,\hat m,\hat{\bar m},\hat
565: l')=\left(\frac{l}{\omega^2},\frac{m}{\omega\omega'},\frac{\bar
566: m}{\omega\omega'},\frac{l'}{\omega'^2}\right).
567: \end{equation}
568: Remarkably, this is just the behaviour of the null tetrad under a
569: transformation of spin-basis $(o,\iota)$ to
570: \begin{equation}\label{frame}
571: (\hat o,\hat\iota)=(o/\omega,\iota/\omega').
572: \end{equation}
573: Such transformations $(o,\iota)\mapsto(\lambda o,\lambda'\iota)$, with
574: $(\lambda,\lambda')$ generally complex, (4.12.2), are well understood and led
575: to the concept of weighted scalars (4.12.9) and the compacted spin-coefficient
576: formalism \cite{GHP}. In summary, it has been shown that {\em the desired
577: conformal transformation is given by a simultaneous transformation of the
578: spin-metric and spin-basis}, (\ref{spin}) and (\ref{frame}). This can be so
579: without the conformal factors $(\omega,\omega')$ exactly relating physical and
580: conformal null coordinates as above, due to the considerable freedom in the
581: conformal factors.
582:
583: The physical and conformal tetrad derivative operators are related by
584: \begin{equation}\label{deriv}
585: (\hat D,\hat\delta,\hat\delta',\hat D')=
586: \left(\frac{D}{\omega^2},\frac{\delta}{\omega\omega'},
587: \frac{\delta'}{\omega\omega'},\frac{D'}{\omega'^2}\right).
588: \end{equation}
589: The relations between the weighted spin-coefficients are obtained
590: straightforwardly from (5.6.15) as
591: \begin{eqnarray}\label{sc0}
592: \hat\kappa&=&\kappa\omega'/\omega^3\\
593: \hat\sigma&=&\sigma/\omega^2\\
594: \hat\rho&=&(\rho-D\log\omega\omega')/\omega^2\\
595: \hat\tau&=&(\tau-\delta\log\omega\omega')/\omega\omega'\\
596: \hat\tau'&=&(\tau'-\delta'\log\omega\omega')/\omega\omega'\\
597: \hat\rho'&=&(\rho'-D'\log\omega\omega')/\omega'^2\\
598: \hat\sigma'&=&\sigma'/\omega'^2\\\label{sc1}
599: \hat\kappa'&=&\kappa'\omega/\omega'^3.
600: \end{eqnarray}
601: They are much simpler than the expressions (5.6.25) or (5.6.27) used previously
602: to describe conformal transformations, due to the separation of the Penrose
603: conformal factor into advanced and retarded conformal factors. The components
604: (4.11.6)
605: \begin{eqnarray}
606: \Psi_0&=&\Psi(o,o,o,o)/|\chi|^2\\
607: \Psi_1&=&\Psi(o,o,o,\iota)/|\chi|^2\\
608: \Psi_2&=&\Psi(o,o,\iota,\iota)/|\chi|^2\\
609: \Psi_3&=&\Psi(o,\iota,\iota,\iota)/|\chi|^2\\
610: \Psi_4&=&\Psi(\iota,\iota,\iota,\iota)/|\chi|^2
611: \end{eqnarray}
612: of the conformally invariant Weyl curvature spinor $\Psi=\hat\Psi$ (4.6.35)
613: transform even more simply as
614: \begin{eqnarray}\label{weyl0}
615: \hat\Psi_0&=&\Psi_0/\omega^4\\
616: \hat\Psi_1&=&\Psi_1/\omega^3\omega'\\
617: \hat\Psi_2&=&\Psi_2/\omega^2\omega'^2\\
618: \hat\Psi_3&=&\Psi_3/\omega\omega'^3\\\label{weyl1}
619: \hat\Psi_4&=&\Psi_4/\omega'^4.
620: \end{eqnarray}
621: One can also generalize the concept of {\em conformal density} (5.6.32) to a
622: quantity $\eta$ which transforms as
623: \begin{equation}
624: \hat\eta=\eta/\omega^k\omega'^{k'}
625: \end{equation}
626: where $\{k,k'\}$ is the {\em conformal weight}, with signs chosen for
627: convenience. Then the shears $(\sigma,\sigma')$ have conformal weights
628: $\{2,0\}$ and $\{0,2\}$ respectively. Also $\tau-\bar\tau'$ has conformal
629: weight $\{1,1\}$. Generally $\tau+\bar\tau'$ is not a conformal density, but it
630: is so in the dual-null gauge (\ref{dualnull0}--\ref{dualnull1}), which yields
631: \cite{bhs}
632: \begin{equation}\label{dchi}
633: \tau+\bar\tau'=\delta\log\chi\bar\chi.
634: \end{equation}
635: Then $(\tau,\tau')$ both have conformal weight $\{1,1\}$.
636:
637: It should also be remarked that this implementation can be achieved purely in
638: the null-tetrad formalism, without ever mentioning spinors. For instance, the
639: weighted spin-coefficients can be expressed as (4.5.22)
640: \begin{eqnarray}
641: (\kappa,\sigma,\rho,\tau)&=&(Dn',\delta n',\delta'n',D'n')(m)/|\chi|^2\\
642: (\tau',\rho',\sigma',\kappa')&=&(Dn,\delta n,\delta'n,D'n)(\bar m)/|\chi|^2.
643: \end{eqnarray}
644: The components of the Weyl and Ricci spinors can also be expressed in terms of
645: the null tetrad (4.11.9--10), as can the weighted derivative operators
646: $(\thorn,\eth,\eth',\thorn')$ (4.12.15) acting on tensors, and consequently all
647: of the compacted spin-coefficient equations (4.12.32). The only spinorial
648: remnant is the phase of $\chi$, which does not enter tensorial expressions. In
649: this purely tensorial view, {\em the desired conformal transformation is given
650: by a simultaneous transformation of the metric and null tetrad}, (\ref{metric})
651: and (\ref{tetrad}).
652:
653: \section{Asymptotic regularity}
654: It is now possible to investigate the asymptotic behaviour of the space-time
655: near both spatial and null infinity. While a detailed set of asymptotic
656: expansions is not derived in the present article, a leading-order analysis
657: suffices to derive some key results.
658:
659: Fixing the null coordinates $(\psi^+,\psi^-)=(\psi,\psi')/\sqrt2$ on $\Im$ as
660: in \S III so that it is locally a metric light cone, the conformal metric
661: spheres are propagated into $M$ by the vectors $(\hat l,\hat l')$, forming a
662: preferred two-parameter family of transverse spatial surfaces, close to metric
663: spheres, in a neighbourhood of $\Im$. It is also convenient to fix the
664: conformal freedom by
665: \begin{equation}
666: (\omega,\omega')=(-\psi,\psi')/\sqrt2
667: \end{equation}
668: so that the advanced and retarded conformal factors themselves can be used as
669: conformal null coordinates, and as expansion parameters near $(\Im^+,\Im^-)$
670: respectively. For expansions near both spatial and null infinity, a useful
671: combination is
672: \begin{equation}
673: u=\frac{\omega\omega'}{\omega+\omega'}.
674: \end{equation}
675: This parameter has the properties of being linear in spatial linear
676: combinations of $(\omega,\omega')$ near $i^0$, with $u\sim\omega$ at $\Im^+$
677: and $u\sim\omega'$ at $\Im^-$, and that the constant-$u$ hypersurfaces are
678: hyperboloids wrapping the asymptotic light cone. The expansion parameter $u$ is
679: also asymptotically the inverse radius $1/r$ of the transverse surfaces, as
680: follows. This in turn implies $\omega\sim1/r$ at $\Im^+$ and $\omega'\sim1/r$
681: at $\Im^-$.
682:
683: Inspecting the form of the metric (\ref{g}), a radius function can be defined
684: by
685: \begin{equation}
686: r=|\chi||P_0|/|P|
687: \end{equation}
688: where (4.15.116)
689: \begin{equation}
690: P_0=(\zeta\bar\zeta+1)/\sqrt2
691: \end{equation}
692: refers to the unit sphere. The conformal radius
693: \begin{equation}
694: \hat r=|\hat\chi||P_0|/|\hat P|
695: \end{equation}
696: is then related by (\ref{chi})--(\ref{P}) as
697: \begin{equation}
698: \hat r=\omega\omega'r.
699: \end{equation}
700: However, for the metric spheres at $\Im$ this is just $\rho$ of \S II--III,
701: \begin{equation}
702: \hat r\approx\omega+\omega'
703: \end{equation}
704: and so one finds
705: \begin{equation}
706: u\sim1/r.
707: \end{equation}
708: Here and henceforth, $a\sim b$ means $a/b\approx1$, i.e.\ they have the same
709: leading-order asymptotic behaviour at $\Im$. When considering spatial or null
710: infinity separately, one can always use $u$ as the sole expansion parameter,
711: remembering that $(\omega,\omega')$ count as $(u,O(1))$ at $\Im^+$, $(O(1),u)$
712: at $\Im^-$ and $(O(u),O(u))$ at $i^0$ from spatial directions. When treating
713: the whole of $\Im$, it is convenient to use all three expansion parameters
714: explicitly.
715:
716: Now one wishes to develop asymptotic expansions valid near both spatial and
717: null infinity. In the usual treatment of null infinity, one can derive the
718: asymptotic expansions from the definition of asymptotic simplicity \cite{PR}.
719: In the current context, this is not so, since the differentiability at $i^0$
720: was deliberately left open. This issue will now be examined in terms of the
721: spin-coefficients.
722:
723: Using (\ref{deriv}), it is straightforward to calculate the useful expressions
724: \begin{eqnarray}
725: \hat D(\omega,\omega',u)&=&(-1,0,-u^2/\omega^2)/\sqrt2\\
726: \hat D'(\omega,\omega',u)&=&(0,1,u^2/\omega'^2)/\sqrt2\\
727: D(\omega,\omega',u)&=&(-\omega^2,0,-u^2)/\sqrt2\\
728: D'(\omega,\omega',u)&=&(0,\omega'^2,u^2)/\sqrt2.
729: \end{eqnarray}
730: With the asymptotic gauge choice
731: \begin{equation}
732: \chi\approx1
733: \end{equation}
734: the conformal convergences are found as
735: \begin{eqnarray}\label{rho0}
736: \hat\rho&=&-\frac{\hat D\hat r}{\hat r}\sim1/\sqrt2(\omega+\omega')\\
737: \label{rho1} \hat\rho'&=&-\frac{\hat D'\hat r}{\hat
738: r}\sim-1/\sqrt2(\omega+\omega').
739: \end{eqnarray}
740: Putting these results together and using the spin-coefficient transformations
741: (\ref{sc0})--(\ref{sc1}), the physical convergences are straightforwardly
742: calculated to have the leading-order behaviour
743: \begin{eqnarray}
744: \rho&\sim&-u/\sqrt2\\
745: \rho'&\sim&u/\sqrt2
746: \end{eqnarray}
747: which agree with standard expressions at $\Im^+$ \cite{PR,NU,NT}, but with a
748: symmetric relative normalization of the null tetrad.
749:
750: For the remaining weighted spin-coefficients, which are all conformal densities
751: in the dual-null gauge (\ref{dualnull0}--\ref{dualnull1}), one can apply the
752: generalized peeling theorem (9.7.4). In the current context, this shows that a
753: scalar conformal density $\eta$ of weight $\{k,k'\}$ satisfies
754: $\eta=O(\omega^k)$ at $\Im^+$ and $\eta=O(\omega'^{k'})$ at $\Im^-$. This
755: constrains the possible asymptotic behaviour at $\Im^\pm$ but does not
756: determine it, nor the behaviour at $i^0$. It is tempting to conjecture the
757: minimal asymptotic behaviour $\eta=O(\omega^k\omega'^{k'})$ for the whole of
758: $\Im$, which turns out to be consistent with the following.
759:
760: In any case, asymptotic behaviour of the conformal spin-coefficients is more
761: precisely determined by geometrical arguments, essentially concerning the
762: intrinsic and extrinsic curvature of a smoothly embedded metric light cone.
763: This has already been done above for the conformal convergences
764: (\ref{rho0}--\ref{rho1}). For the conformal shears, which vanish at the
765: respective null infinity, as shown in \S III or as (9.6.28), this suggests
766: $\hat\sigma'=O(\omega)$ at $\Im^+$. On the other hand, there is no reason for
767: $\hat\sigma'$ to vanish at $\Im^-$ or $i^0$, though it should be finite. With
768: corresponding conditions for $\hat\sigma$, assuming dependence on integral
769: powers of $(\omega, \omega',u)$ uniquely yields
770: \begin{eqnarray}
771: \hat\sigma&=&O(u/\omega)\\
772: \hat\sigma'&=&O(u/\omega').
773: \end{eqnarray}
774: To be more precise, a function $f(\omega,\omega',\theta,\phi)$ is said to be
775: {\em regular at} $\Im$ if its limits at $\Im^+$ ($\omega\to0$) and $\Im^-$
776: ($\omega'\to0$) exist, together with the limits
777: \begin{equation}\label{regular}
778: f_0(\theta,\phi)=\lim_{\omega\to0}f(\omega,0,\theta,\phi)
779: =\lim_{\omega'\to0}f(0,\omega',\theta,\phi).
780: \end{equation}
781: To economize on notation, $a=O(b)$ applied at $\Im$ is here and henceforth
782: taken to mean that $a/b$ tends to a function regular at $\Im$. This allows the
783: limits at $i^0$ to depend on the angular direction, as is necessary e.g.\ for
784: $\Psi_2/u^3$ in the Kerr solution \cite{AH}.
785:
786: For $(\hat\tau,\hat\tau')$, one can use the geometrical interpretation of
787: $\hat\tau-\hat{\bar\tau}'$ as the conformal twist of the dual-null foliation,
788: measuring the lack of commutativity of the null normal vectors $(l,l')$
789: \cite{mon}. This suggests that it should be $O(\omega\omega')$. Similarly
790: $\hat\tau+\hat{\bar\tau}'$ (\ref{dchi}) measures transverse derivatives of the
791: relative normalization of the null normals, suggesting similar behaviour. Then
792: \begin{eqnarray}
793: \hat\tau&=&O(\omega\omega')\\
794: \hat\tau'&=&O(\omega\omega').
795: \end{eqnarray}
796: In terms of the physical spin-coefficients, this yields
797: \begin{eqnarray}\label{shear0}
798: \sigma&=&O(u\omega)\\\label{shear1}
799: \sigma'&=&O(u\omega')\\\label{twist0}
800: \tau&=&O((\omega\omega')^2)\\\label{twist1}
801: \tau'&=&O((\omega\omega')^2).
802: \end{eqnarray}
803: The practical test of these geometrically motivated conditions is whether they
804: describe a class of space-times with desired physical properties, as considered
805: in the next section.
806:
807: Summarizing this section: leading-order behaviour of the conformal weighted
808: spin-coefficients has been proposed according to the geometrical nature of
809: $\Im$ as a smoothly embedded metric light cone with respect to the conformal
810: metric. This directly determines the leading-order behaviour of the physical
811: weighted spin-coefficients
812: $(\kappa,\sigma,\rho,\tau,\tau',\rho',\sigma',\kappa')$. One might translate
813: the conditions back to differentiability conditions on the metric, which would
814: be stricter than previous suggestions \cite{AH,AM,A}, but such metric-level
815: requirements are not particularly illuminating.
816:
817: Since the asymptotic regularity conditions necessarily allow angular dependence
818: at spatial infinity, it may sometimes be useful to expand $i^0$ from a point to
819: a sphere, defined by $\omega\to0$, $\omega'\to0$ and parametrized by
820: $(\theta,\phi)$. This is in contrast to previous descriptions of $i^0$ as a
821: hyperboloid \cite{G,AH,A}, depending also on the boost direction, or as a
822: cylinder \cite{F}. Here there is no detailed dependence on the boost direction
823: $\omega/\omega'$; to use Geroch's terminology of universal versus physical
824: structure \cite{G}, the boost dependence at $i^0$ is treated universally here,
825: described using $(\omega,\omega')$ in the given gauge, while the physical
826: dependence at $i^0$ is purely angular, encoded in the next section in functions
827: $(\chi_1,\rho_1,\rho'_1,\sigma_1,\sigma'_1)$ which are regular at $\Im$,
828: defined as above (\ref{regular}) to allow only angular dependence at $i^0$.
829: This seems to reflect the intuitive nature of spatial infinity as a large
830: time-independent sphere.
831:
832: \section{Energy}
833: It is widely agreed that the most physically important discovery in space-time
834: asymptotics is the Bondi-Sachs energy-loss equation \cite{B,BBM,S}, whereby the
835: total mass-energy $E_{BS}$ decreases at $\Im^+$ according to \cite{PR,NT,mon}
836: $\sqrt2D'E_{BS}=-\oint{\tilde*}|N'|^2/4\pi$, where the integral is over a
837: sphere at $\Im^+$, ${\tilde*}1$ denotes the area form of a unit sphere and the
838: complex function $N'$ corresponds to what Bondi dubbed the news. In more
839: physical terminology, $|N'|^2$ is the conformal energy flux of the
840: gravitational radiation, meaning that the physical energy flux is $|N'|^2/r^2$
841: near $\Im^+$, the above integral giving the integrated energy flux through a
842: sphere near $\Im^+$. Now the news is $N'=\sigma'/u$ \cite{mon}, as will be
843: verified below. The asymptotic regularity condition (\ref{shear1}) then implies
844: $N'=O(\omega')$, so that it is not only finite at $\Im^+$, but decays near
845: $i^0$. Transforming the Bondi-Sachs energy-loss equation to use the conformal
846: null derivative (\ref{deriv}) yields $\hat D'E_{BS}=D'E_{BS}/\omega'^2=O(1)$.
847: Thus the gravitational radiation decays near spatial infinity in such a way
848: that the total energy flux, integrated over time, is finite. Moreover, the
849: prescribed fall-off of $\sigma'$ is the weakest which would achieve this
850: result. This property has been presented first because it follows so simply
851: from the regularity conditions, thereby verifying that the asymptotic behaviour
852: of the shears, proposed on geometrical grounds, is exactly right on physical
853: grounds.
854:
855: It is also independent even of the existence of the total energy, for which one
856: needs to expand some spin-coefficients further, specifically
857: \begin{eqnarray}\label{exp0}
858: \chi-1&\sim&\chi_1u\\
859: -\sqrt2\rho/u-1&\sim&\rho_1u\\\label{exp1}
860: \sqrt2\rho'/u-1&\sim&\rho'_1u
861: \end{eqnarray}
862: where $(\chi_1,\rho_1,\rho'_1)$ are regular at $\Im$. Here the proposal is that
863: $u$ is the natural expansion parameter near both spatial and null infinity,
864: even though leading-order behaviour may depend on $(\omega,\omega')$
865: independently
866:
867: To study energy near $\Im$, a useful quantity is the Hawking quasi-local
868: mass-energy \cite{H}, which can be written as \cite{mon}
869: \begin{equation}
870: E=\frac{A^{1/2}}{(4\pi)^{3/2}}\oint{*}\frac{K+\rho\rho'}{\chi\bar\chi}
871: \end{equation}
872: where the integral is over a transverse surface, ${*}1$ denotes the area form
873: of a transverse surface,
874: \begin{equation}
875: A=\oint{*}1
876: \end{equation}
877: is the area of a transverse surface and (4.14.20)
878: \begin{equation}\label{K}
879: K=\sigma\sigma'-\rho\rho'-\Psi_2+\Phi_{11}+\Pi
880: \end{equation}
881: is the complex curvature, satisfying a complex generalization of the
882: Gauss-Bonnet theorem (4.14.42--43):
883: \begin{equation}
884: \oint{*}\frac{K}{\chi\bar\chi}=\oint{*}\frac{\bar
885: K}{\chi\bar\chi}=2\pi(1-\gamma)
886: \end{equation}
887: where $\gamma$ is the genus of the transverse surface, vanishing in this
888: context. For the Schwarzschild space-time considered in \S II, one finds
889: $\chi=(1-2M/r)^{1/2}$ (phase irrelevant), $\rho=-Dr/r=-(1-2M/r)/\sqrt2r$,
890: $\rho'=-D'r/r=(1-2M/r)/\sqrt2r$, $K=1/2r^2$ and therefore $E=M$.
891:
892: The Bondi-Sachs energy \cite{B,BBM,S} at $\Im^+$ can be expressed as
893: \cite{P,PR,NT,mon}
894: \begin{equation}
895: E_{BS}=\lim_{\omega\to0}\frac{A^{1/2}}{(4\pi)^{3/2}}
896: \oint{*}\frac{\sigma\sigma'-\Psi_2}{\chi\bar\chi}.
897: \end{equation}
898: In vacuum or with suitable fall-off of the matter terms $\Phi_{11}$ and $\Pi$,
899: (\ref{K}) shows that it can be written as the limit of the Hawking energy:
900: \begin{equation}\label{BS}
901: E_{BS}=\lim_{u\to0}E.
902: \end{equation}
903: Since
904: \begin{eqnarray}
905: {*}1&\sim&{\tilde*}1/u^2\\
906: A&\sim&4\pi/u^2
907: \end{eqnarray}
908: it is straightforward to see from the expansions (\ref{exp0}--\ref{exp1}) that
909: it is finite, given specifically by
910: \begin{equation}
911: E_{BS}=\frac1{8\pi}\oint{\tilde*}(\chi_1+\bar\chi_1-\rho_1-\rho'_1).
912: \end{equation}
913: Again one may check the Schwarzschild case: $\chi_1=-M$, $\rho_1=\rho'_1=-2M$,
914: yielding $E_{BS}=M$. Note that generally {\em the limit of $E_{BS}$ at spatial
915: infinity exists uniquely} and is given by the same formula (\ref{BS}). This
916: follows from the definition of functions regular at $\Im$ (\ref{regular}). Any
917: such transverse surface integral of functions regular at $\Im$ has a unique
918: limit at spatial infinity.
919:
920: On the other hand, the ADM energy \cite{ADM} at spatial infinity can be
921: expressed as \cite{G,AH,AM,A,qle}
922: \begin{equation}
923: E_{ADM}=-\lim_{\omega\to0}\lim_{\omega'\to0}\frac{A^{1/2}}{(4\pi)^{3/2}}
924: \oint{*}\frac{\Re\Psi_2}{\chi\bar\chi}
925: \end{equation}
926: where, in the original treatment, it was unclear whether the limit depended on
927: the boost direction $\omega/\omega'$, i.e.\ on the choice of spatial
928: hypersurface. As above, this can be written
929: \begin{equation}
930: E_{ADM}=\lim_{\omega\to0}\lim_{\omega'\to0}\frac{A^{1/2}}{(4\pi)^{3/2}}
931: \oint{*}\frac{K+\rho\rho'-\Re(\sigma\sigma')}{\chi\bar\chi}
932: \end{equation}
933: which is similar to the expression (\ref{BS}) for the Bondi-Sachs energy, but
934: apparently differs by the term $\Re(\sigma\sigma')$ in the shears. The
935: discrepancy in the two energies is found from the asymptotic regularity
936: conditions (\ref{shear0}--\ref{shear1}) to be $O(\omega+\omega')$. This
937: discrepancy is generally non-zero at null infinity ($\omega=0$ or $\omega'=0$),
938: so that the ADM energy, if extended to null infinity by the same formula, would
939: generally not agree with the Bondi-Sachs energy. However, the discrepancy
940: vanishes at spatial infinity ($\omega=\omega'=0$) from any direction. Thus {\em
941: the ADM energy is the limit of the Bondi-Sachs energy at spatial infinity} in
942: this context, and also the limit of the Hawking energy from any spatial or null
943: direction:
944: \begin{equation}
945: E_{ADM}=\lim_{\omega\to0}\lim_{\omega'\to0}E=\lim_{\omega+\omega'\to0}E_{BS}.
946: \end{equation}
947: This provides a remarkably simple resolution of the long-standing questions
948: over the relation of the Bondi-Sachs and ADM energies, and the uniqueness of
949: the latter. In the current framework, one may simply use the Bondi-Sachs energy
950: on the entire asymptotic light cone $\Im$. The resolution explicitly rests on
951: the additional structure at spatial infinity provided by the advanced and
952: retarded conformal factors $(\omega,\omega')$.
953:
954: Returning to the energy flux of gravitational radiation: the propagation
955: equations for the Hawking energy can be found from the compacted
956: spin-coefficient equations (4.12.32), using $\thorn{*}1=-{*}2\rho$ and
957: $\thorn'{*}1=-{*}2\rho'$, as \cite{mon}
958: \begin{eqnarray}
959: \thorn E&=&\frac{A^{1/2}}{(4\pi)^{3/2}}\left[\oint{*}\rho
960: \frac{K+\rho\rho'-\tau'\bar\tau'+\eth\tau'-\Phi_{11}-3\Pi}{\chi\bar\chi}
961: \right.\nonumber\\
962: &&{}+\left.\oint{*}\rho'\frac{\sigma\bar\sigma+\Phi_{00}}{\chi\bar\chi}
963: -\frac1A\oint{*}\rho\oint{*}\frac{K+\rho\rho'}{\chi\bar\chi}\right]\\
964: \thorn'E&=&\frac{A^{1/2}}{(4\pi)^{3/2}}\left[\oint{*}\rho'
965: \frac{K+\rho\rho'-\tau\bar\tau+\eth'\tau-\Phi_{11}-3\Pi}{\chi\bar\chi}
966: \right.\nonumber\\
967: &&{}+\left.\oint{*}\rho\frac{\sigma'\bar\sigma'+\Phi_{22}}{\chi\bar\chi}
968: -\frac1A\oint{*}\rho'\oint{*}\frac{K+\rho\rho'}{\chi\bar\chi}\right].
969: \end{eqnarray}
970: Rewriting these equations using the conformal spin-weighted derivatives
971: $(\hat\thorn,\hat\eth,\hat\eth',\hat\thorn')$, assuming either vacuum or
972: suitable fall-off of the matter fields, and using the asymptotic regularity
973: conditions and expansions (\ref{shear0}--\ref{exp1}), it can be shown that
974: $(\hat DE,\hat D'E)$ are $O(1)$, as expected on physical grounds. The general
975: argument is non-trivial, subtly confirming the asymptotic regularity
976: conditions, and uses the fact that $u$ is constant on the transverse surfaces
977: to cancel terms in $K+\rho\rho'$, and $\oint{*}\eth'\tau=0$ (4.14.69).
978: Fortunately the cases of most interest are $(\hat DE,\hat D'E)$ at
979: $(\Im^-,\Im^+$) respectively, where it is easier to see that most of the terms
980: disappear, in the vacuum case leaving just
981: \begin{eqnarray}\label{massgain}
982: \sqrt2\hat DE&=&\frac1{4\pi}\oint{\tilde*}|\sigma_1|^2 \quad\hbox{at $\Im^-$
983: ($\omega'=0$)}\\\label{massloss} \sqrt2\hat
984: D'E&=&-\frac1{4\pi}\oint{\tilde*}|\sigma'_1|^2 \quad\hbox{at $\Im^+$
985: ($\omega=0$)}
986: \end{eqnarray}
987: where
988: \begin{eqnarray}
989: \sigma_1&=&\sigma/u\omega\\
990: \sigma'_1&=&\sigma'/u\omega'
991: \end{eqnarray}
992: are the leading-order terms in the shears (\ref{shear0}--\ref{shear1}), regular
993: at $\Im$. In terms of the retarded and advanced news functions
994: \begin{eqnarray}
995: N&=&\sigma/u\\
996: N'&=&\sigma'/u
997: \end{eqnarray}
998: this implies
999: \begin{eqnarray}
1000: \sqrt2DE_{BS}&=&\frac1{4\pi}\oint{\tilde*}|N|^2 \quad\hbox{at $\Im^-$}\\
1001: \sqrt2D'E_{BS}&=&-\frac1{4\pi}\oint{\tilde*}|N'|^2 \quad\hbox{at $\Im^+$.}
1002: \end{eqnarray}
1003: The second equation is the usual Bondi-Sachs energy-loss equation, showing that
1004: the outgoing gravitational radiation carries energy away from the system. The
1005: first equation similarly shows that ingoing gravitational radiation supplies
1006: energy to the system. The new energy propagation equations
1007: (\ref{massgain}--\ref{massloss}) improve on these equations by applying also in
1008: the limit at $i^0$. Thus the change in energy from $i^0$ to a section of
1009: $\Im^\pm$ is finite, as mentioned above. Physically this means that {\em the
1010: ingoing and outgoing gravitational radiation decays near spatial infinity such
1011: that its total energy is finite}. This property, assumed in addition to the
1012: Ashtekar-Hansen definition of asymptotic flatness \cite{AH}, is known to imply
1013: the coincidence of Bondi-Sachs and ADM energies at spatial infinity \cite{AM},
1014: but here the property has been derived from the new definition of asymptotic
1015: flatness.
1016:
1017: It is straightforward to generalize to an asymptotic energy-momentum vector
1018: \begin{equation}
1019: P=\lim_{u\to0}\frac{A^{1/2}}{(4\pi)^{3/2}}
1020: \oint{*}\ell\frac{\sigma\sigma'-\Psi_2}{\chi\bar\chi}
1021: \end{equation}
1022: where the vector $\ell$ in Cartesian coordinates is (1.4.11)
1023: \begin{equation}
1024: \ell=(t,x,y,z)=\left(1,\frac{\zeta+\bar\zeta}{\zeta\bar\zeta+1},
1025: \frac{i(\bar\zeta-\zeta)}{\zeta\bar\zeta+1},
1026: \frac{\zeta\bar\zeta-1}{\zeta\bar\zeta+1}\right).
1027: \end{equation}
1028: Then $P$ coincides with the Bondi-Sachs energy-momentum \cite{BBM,S} at null
1029: infinity and with the ADM energy-momentum \cite{ADM} at spatial infinity. The
1030: corresponding energy-momentum propagation equations are
1031: \begin{eqnarray}
1032: \sqrt2\hat DP&=&\frac1{4\pi}\oint{\tilde*}\ell|\sigma_1|^2 \quad\hbox{at
1033: $\Im^-$ ($\omega'=0$)}\\ \sqrt2\hat
1034: D'P&=&-\frac1{4\pi}\oint{\tilde*}\ell|\sigma'_1|^2 \quad\hbox{at $\Im^+$
1035: ($\omega=0$)}
1036: \end{eqnarray}
1037: cf.\ \cite{NT}.
1038:
1039: Finally, it should be noted that, while the gravitational radiation at
1040: $(\Im^-,\Im^+)$ is respectively encoded as above in complex functions
1041: $(\sigma_1,\sigma'_1)$ or $(N,N')$, a more tensorial formulation would encode
1042: such shear terms in transverse traceless tensors, exactly as in the
1043: conventional treatment of linearized gravitational radiation.
1044:
1045: \section{Conclusion}
1046: A new framework for space-time asymptotics has been proposed and developed,
1047: refining Penrose's conformal framework by introducing advanced and retarded
1048: conformal factors. This allows a relatively simple definition of asymptotic
1049: flatness at both spatial and null infinity. It has been shown how to fix the
1050: light cone at infinity so that it is locally a metric light cone. Assuming
1051: smooth embedding of the light cone, asymptotic regularity conditions have been
1052: proposed and asymptotic expansions developed. These conditions ensure that the
1053: Bondi-Sachs energy-momentum is finite and tends uniquely to the uniquely
1054: rendered ADM energy-momentum at spatial infinity, that the ingoing and outgoing
1055: gravitational radiation has the expected two modes as encoded in gravitational
1056: news functions, and that the energy-flux integrals decay at spatial infinity
1057: such that the total energy of the gravitational radiation is finite. The most
1058: basic physical properties of isolated gravitational systems have thereby been
1059: included.
1060:
1061: Mathematically the new structure proposed for infinity is unprecedentedly
1062: rigid, allowing physical fields to have very simple behaviour at spatial
1063: infinity. Currently there are no indications that this is too simple to be
1064: physically realistic, though open issues remain. The asymptotic symmetry group
1065: presumably simplifies accordingly.
1066:
1067: The framework, implemented in the spin-coefficient or null-tetrad formalism, is
1068: quite practical, allowing explicit calculations as for null infinity alone.
1069: While this article has presented only a leading-order analysis, asymptotic
1070: expansions can now be developed to higher orders. This would presumably allow
1071: insights into angular momentum and multipole moments of the gravitational
1072: field. A related issue is the asymptotic behaviour of the conformal curvature
1073: near spatial infinity. It may even be possible to address the long-standing
1074: question of finding a conformal form of the Einstein equations which is regular
1075: at both null and spatial infinity. In any case, it is to be hoped that the
1076: refined conformal picture will stimulate a renewal of interest in space-time
1077: asymptotics.
1078:
1079: \begin{references}
1080: \bibitem{B}H Bondi, Nature {\bf 186}, 535 (1960).
1081: \bibitem{BBM}H Bondi, M G J van der Burg \& A W K Metzner,
1082: {Proc. Roy. Soc. Lond.} {\bf A269}, 21 (1962).
1083: \bibitem{S}R K Sachs, {Proc. Roy. Soc. Lond.} {\bf A270}, 103 (1962);
1084: {Phys. Rev.} {\bf 128}, 2851 (1962);
1085: in Relativity, Groups and Topology: the 1963 Les Houches Lectures,
1086: ed.\ B S de Witt \& C M de Witt (Gordon \& Breach 1963).
1087: \bibitem{P}R Penrose, {Phys. Rev. Lett.} {\bf 10}, 66 (1963);
1088: in Relativity, Groups and Topology: the 1963 Les Houches Lectures,
1089: ed.\ B S de Witt \& C M de Witt (Gordon \& Breach 1963);
1090: {Proc. Roy. Soc. Lond.} {\bf A284}, 159 (1965).
1091: \bibitem{PR}R Penrose \& W Rindler, Spinors and Space-Time Vols.\ 1 \& 2
1092: (Cambridge University Press 1984 \& 1988).
1093: \bibitem{NP}E T Newman \& R Penrose, {J. Math. Phys.} {\bf 3}, 566 (1962).
1094: \bibitem{NU}E T Newman \& T Unti, {J. Math. Phys.} {\bf 3}, 891 (1962).
1095: \bibitem{GHP}R Geroch, A Held \& R Penrose,
1096: {J. Math. Phys.} {\bf 14}, 874 (1973).
1097: \bibitem{NT}E T Newman \& K P Tod, in General Relativity and Gravitation,
1098: ed.\ A Held (Plenum 1980).
1099: \bibitem{ADM}R Arnowitt, S Deser \& C W Misner, {Phys. Rev.} {\bf 118}, 1100
1100: (1960); {\bf 121}, 1556 (1961); {\bf 122}, 997 (1961); in Gravitation: An
1101: Introduction to Current Research, ed.\ L Witten (Wiley 1962).
1102: \bibitem{G}R Geroch, {J. Math. Phys.} {\bf 13}, 956 (1972);
1103: in Asymptotic Structure of Space-Time, ed.\ F P Esposito \& L Witten (Plenum
1104: 1977)
1105: \bibitem{AH}A Ashtekar \& R O Hansen, {J. Math. Phys.} {\bf 19}, 1542 (1978).
1106: \bibitem{AM}A Ashtekar \& A Magnon-Ashtekar, {Phys. Rev. Lett.} {\bf 43}, 181
1107: (1979).
1108: \bibitem{A}A Ashtekar, in General Relativity and Gravitation,
1109: ed.\ A Held (Plenum 1980).
1110: \bibitem{F}H Friedrich, Einstein's equation and geometric asymptotics
1111: (gr-qc/9804009).
1112: \bibitem{H}S W Hawking, {J. Math. Phys.} {\bf 9}, 598 (1968).
1113: \bibitem{W}R M Wald, General Relativity (University of Chicago Press 1984).
1114: \bibitem{bhs}S A Hayward, {Class. Quantum Grav.} {\bf 11}, 3025 (1994).
1115: \bibitem{mon}S A Hayward, {Class. Quantum Grav.} {\bf 11}, 3037 (1994).
1116: \bibitem{qle}S A Hayward, {Phys. Rev.} {\bf D49}, 831 (1994).
1117: \end{references}
1118: \end{document}
1119: