1: \documentclass[11pt]{article}
2: \usepackage{citesort,bibmods}
3: %\usepackage{santheorem}
4:
5: %\usepackage{drafthead}
6: %\usepackage{amsmath2000}
7: \usepackage{url}
8: \usepackage{amsmath}
9: %\usepackage{xypic}
10: \usepackage{amscd}
11: \usepackage{amssymb,latexsym}
12: %\usepackage[final]{graphicx}
13: \usepackage{float}
14: \usepackage{deleq}
15:
16: %\usepackage{srcltx}
17: %\SRCOKfalse
18: \usepackage[mathscr]{eucal}
19: \usepackage{a4}
20: %\usepackage[notcite]{showkeys}
21: %\usepackage[first,bottomafter,light]{draftcopy}
22: %\usepackage{xypic}
23: %\input xy % this loads in xypic, for diagrams
24: %\xyoption{all}
25:
26: \emergencystretch=15pt %%% Does away with most overfull \hboxs
27: %%% Remove for final version
28: \sloppy
29:
30: \DeclareFontFamily{OT1}{rsfs}{}
31: \DeclareFontShape{OT1}{rsfs}{m}{n}{ <-7> rsfs5 <7-10> rsfs7 <10-> rsfs10}{}
32: \DeclareMathAlphabet{\mycal}{OT1}{rsfs}{m}{n}
33: \def\scri{{\mycal I}}%
34: \def\scrip{\scri^{+}}%
35: \def\scrp{{\mycal I}^{+}}%
36: \def\Scri{\scri}
37:
38: %\renewcommand{\cal}{\mycal}
39: \newcommand{\nn}{\nonumber}
40: \newcommand{\ncO}{{\cal O}}
41: \newcommand{\ncU}{{\cal U}}
42: \newcommand{\mcU}{{\mycal U}}
43: \newcommand{\cOp}{\widehat{\ncO}_{\hat p}}
44: \newcommand{\hOp}{\cOp}
45:
46: \def\s{\_\!\imath}
47: \def\th{{e}}
48: \def\tH{{H}}
49: \def\tV{\tilde{V}}
50: \def\tK{{K}}
51: %\def\ext{\mbox{{\footnotesize ext}}}
52: \newtheorem{theorem}{Theorem}
53: \newtheorem{proposition}{Proposition}
54: \newtheorem{definition}{Definition}
55: \newcommand{\oeps}{o(r^{-2\alpha})}
56: \newcommand{\oepsb}{o(r^{\beta-2\alpha})}
57: \newcommand{\myvareps}{}
58: \newcommand{\newa}{c}
59:
60: \newcommand{\macro}[1]{\vphantom{A}
61: \raisebox{-0.085ex}{$\stackrel{{\scriptscriptstyle (#1)}}{x}$}
62: \vphantom{A}}
63:
64: \newcommand{\two}[1]{\vphantom{#1}\stackrel{(2)}{#1}\!\vphantom{#1}}
65: %\newcommand{\zero}[1]{\vphantom{#1}\stackrel{(0)}{#1}\!\vphantom{#1}}
66: %\newcommand{\one}[1]{\vphantom{#1}\stackrel{(1)}{#1}\!\vphantom{#1}}
67: \newcommand{\yone} {\macro{1}}%{{}^{(1)}x} %{\one {y}}
68: \newcommand{\yzero} {\macro{0}}%{{}^{(0)}x} %{{\zero {y}}
69: \newcommand{\ytwo}{\macro{2}}% {{}^{(2)}x} %{{\two {y}}
70:
71: \newcommand{\mnoteprzelicz}[1]{}%{\mnote{#1}}
72: \newcommand{\przelicz}{}%{\mnote{ja tego nie przeliczalem; zrobie to po przyjeciu pracy do druku i/lub jak bedziemy cos pisac o momencie pedu}}
73: \renewcommand{\lrcorner}{\rfloor} %redefinicja dla zwezenia form
74:
75: \newcounter{mnotecount}[section]
76:
77: \renewcommand{\themnotecount}{\thesection.\arabic{mnotecount}}
78: \newcommand{\mnote}[1]{}
79: %{\protect{\stepcounter{mnotecount}}$^{\mbox{\footnotesize $%\!\!\!\!\!\!\,
80: % \bullet$\themnotecount}}$ \marginpar{\raggedright\tiny\em
81: % $\!\!\!\!\!\!\,\bullet$\themnotecount: #1} }
82:
83:
84: \newcommand{\bA}{\,\,{\stackrel{\circ}{\!\!A}}\vphantom{A}}
85:
86: \newcommand{\FS} %{F_1} %
87: {F}
88: %{F_{\mbox{\scriptsize volume}}}
89:
90: \newcommand{\HS} %{F_2}
91: {H_{\mbox{\scriptsize volume}}}
92:
93: \newcommand{\ren}{\mbox{\scriptsize normalised}}
94: \newcommand{\HSren} %{F_2}
95: {{\HS^{\ren}}}
96:
97: %
98: % nowe makra do grawitacyjnych przestrzeni fazowych
99: %
100:
101: \newcommand{\calPx}{\mycal P} %{\hcalPo}% generic phase space
102: \newcommand{\calPo}{{\mycal P}_{[-1,0]}} % hatted phase space; the
103: % hat needs fixing!
104: \newcommand{\hcalPo}{{\widehat{\!\mycal P}}_{[-1,0]}} % hatted phase space; the
105: % hat needs fixing!
106: \newcommand{\barSigma}{\;\bar{\!\Sigma}} %conformaly completed three
107: %dimensional hypersurface
108: \newcommand{\lpsi}{\zeta} % the maps which map the hypersurfaces to
109: % each other, an other symbol is badly needed
110: \newcommand{\uomega}{u} % the variable u on scri
111: \newcommand{\gtaum}{\tau_-} % the lower end point of the u interval on
112: % Scri
113: \newcommand{\gSigma}{\mycal S} % the hypersurfaces in the physical
114: % space--time
115:
116:
117: %
118: %
119: %
120:
121: %\newcommand{\Freud}{{\scriptsize\mbox{Freud}}}
122: \newcommand{\Freud}{{\mbox{\scriptsize boundary}}}
123: \newcommand{\HF}{H_{\Freud}}
124: \newcommand{\HFren}{H_{\Freud}^{\ren}}
125: \newcommand{\hype}{{\hyp^{\epsilon}}}
126: \newcommand{\hypext}{{\hyp_{\ext}}}
127:
128: \newcommand{\pblurb}{(X,{\partial\hyp})}
129: \newcommand{\blurb}{(X,{\hyp})} % {(X,{\hyp},\Jped)}
130: \newcommand{\blurbe}{(X,{\hype})} %{(X,{\hype},\Jped)}
131: \newcommand{\HFh}{H_{\Freud}\blurb}
132: \newcommand{\HFhe}{H_{\Freud}\blurbe}
133: \newcommand{\Hh}{H\blurb}
134: \newcommand{\Hhe}{H\blurbe}
135:
136:
137: \newcommand{\hamiltonian}{Hamiltonian } %decide lower or upper case?
138: \newcommand{\Lagrangian}{Lagrangian } %decide lower or upper case?
139:
140: \newcommand{\reword}{\mnote{rewordings}}
141: \newcommand{\ptc}[1]{\mnote{\textbf{PTC:} #1}}
142: \newcommand{\marc}[1]{\mnote{\textbf{PTC:} #1}}
143: % allow equations to be reset within sections.
144:
145: {\catcode `\@=11 \global\let\AddToReset=\@addtoreset}
146: \AddToReset{equation}{section}
147: \renewcommand{\theequation}{\thesection.\arabic{equation}}
148:
149:
150: \newcommand{\lie}{{\pounds}} %Lie derivative
151: %\newcommand{\lie}{{\mycal L}}
152: %\def\varOmega{{\it\Omega}}
153: \def\varOmega{{\mathit{\Omega}}}
154: %\def\varOmega{{\Omega}}
155: \newcommand{\rdphi}{\rd \varphi}
156: \newcommand{\rdS}{\rd S} % formy objetosciowe trzeba zmienic na dS
157:
158: \newcommand{\cU}{{\mycal U}}
159: \newcommand{\cP}{{\mycal P}}
160: %\newcommand{\cDX}{{\mycal D}_X}
161: \newcommand{\cDX}{{\mathscr D}_X}
162: \newcommand{\dg}{\ \!\!\vartriangle\!\!\Kp}%{\Delta\Kp}{\blacktriangle\Kp}%
163: %roznica metryki a metryka tla
164:
165: \newtheorem{Theorem} {Theorem} [section] \newtheorem{Corollary}
166: [Theorem] {Corollary} \newtheorem{Lemma} [Theorem] {Lemma}
167: \newtheorem{Proposition} [Theorem] {Proposition}
168: \newtheorem{Conjecture}[Theorem]{Conjecture}
169: \newtheorem{Problem}[Theorem]{Problem}
170: \newtheorem{Definition}[Theorem]{Definition}
171: \newtheorem{example}[Theorem]{Example}
172: % Scri
173: % \font\svtnscr=rsfs10 scaled1700
174: % \font\tenscr=rsfs10 scaled1200
175: % \font\sevenscr=rsfs7 % scaled \magstep1
176: % \font\fivescr=rsfs5 % scaled \magstep1
177: % \skewchar\tenscr='177 \skewchar\sevenscr='177 \skewchar\fivescr='177
178: % \newfam\scrfam
179: % \textfont\scrfam=\tenscr
180: % \scriptfont\scrfam=\sevenscr
181: % \scriptscriptfont\scrfam=\fivescr
182: % \def\scr{\fam\scrfam}
183:
184: % \def\scri{{\fam\scrfam I}}
185:
186: % \newcommand{\SCRI}{{\svtnscr I}}
187: % %\newcommand{\Scri}{{\tenscr I}}
188: %\input{scrifont}
189:
190:
191: % \def\scri{\hbox{${\mycal J}$\kern -.645em {\raise
192: % .57ex\hbox{$\scriptscriptstyle (\ $}}}}
193: % \newcommand{\Scri}{\scri}
194: % \newcommand{\scrip}{\scri^{+}}
195: \newcommand{\scrim}{\scri^-}
196: \newcommand{\eq}[1]{(\ref{#1})}
197: \newcommand{\bh}{{\breve h}} % standard metric on S^2
198: \newcommand{\taum}{{\tau_{-}}}
199: \newcommand{\taumn}{-1}
200: \newcommand{\phield}{{f}}
201: \newcommand{\phieldone}{{h}}
202: \newcommand{\phieldtwo}{{g}}
203: \newcommand{\tauz}{{\tau_{0}}}
204: \newcommand{\TB}{Trautman--Bondi\ }
205: \newcommand{\be}{\begin{equation}}
206: \newcommand{\ee}{\end{equation}}
207: \newcommand{\eea}{\end{eqnarray}}
208: \newcommand{\bea}{\begin{eqnarray}}
209: \newcommand{\dtwo}{\Delta_2}
210: \newcommand{\kolo}[1]{\vphantom{#1}\stackrel{\circ}{#1}\!\vphantom{#1}}
211: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% %% Additional fonts and characters
212: \newfont{\msa}{msam10 scaled\magstep1}
213: \font\SYMA=msam10 % For some AMS symbols%
214: %\font\SYM=msbm10 % For blackboard bold
215:
216:
217: \def\Complex{{\mathbb C}} \def\Rationals{{\mathbb Q}}
218: \def\Reals{{\mathbb R}}
219: \def\T{{\mathbb T}}
220: \def\Integers{{\mathbb Z}} \def\Naturals{{\mathbb N}}
221: %\def\lrcorner{{\mathbb y}}
222:
223: \def\cosec{\mathop{\rm cosec}\nolimits}
224: \newcommand{\comment}{\mnote{comment added}}
225: \newcommand{\dx}[1]{{\mbox{\rm d}#1}}
226: \newcommand{\mH}{{\mycal H}}%{{\mathscr H}}
227: \newcommand{\mHS}{{\mH}_{\Sigma}}
228: \newcommand{\bp}{p}%{\mathbf{p}}
229: \newcommand{\bd}{\mathbf{d}}
230: \newcommand{\bL}{\mathbf{L}}
231: %\newcommand{\bpi}{{\pi}\!\!\!\pi}
232: \newcommand{\bpi}{\pi}%{\underline{\pi}}
233: %\newcommand{\bpi}{{\mathscr p}}
234: \newcommand{\bbpi}{{\mbox{\boldmath$ \pi$}}}
235: %\newcommand{\bbpi}{\underline{\pi}}%{\underline{\underline{\pi}}}
236: %\newcommand{\bcL}{{\mathbf{\mycal L}}}
237: \newcommand{\bcL}{{\mbox{\boldmath$ {\mycal L}$}}}%{{\underline{\mycal %L}}}
238: %\newcommand{\bcL}{{\mbox{\boldmath$ {\mathscr L}$}}}%{{\underline{\mycal L}}}
239: \newcommand{\caL}{{{\mycal L}}}
240: \newcommand{\cLX}{{\lie_X}}
241: \newcommand{\caD}{{{\mycal D}}}
242: \newcommand{\rd}{\,{ d}} % exterior differential
243: %\newcommand{\rd}{\,{\rm d}} % exterior differential
244: \newcommand{\N}{\Naturals}
245: \newcommand{\R}{\Reals} \newcommand{\Z}{\Integers}
246: \newcommand{\C}{\Complex}
247: \newcommand{\Jphi}{f}%{{\tilde f}}%{\phi} % pole skalarne fizyczne
248: \newcommand{\dotphi}{\dot\phi}
249: \newcommand{\dotJphi}{\dot\Kphi}
250: \newcommand{\dotpi}{\dot\pi}
251: \newcommand{\dotJpi}{\dot\Jpi}
252: \newcommand{\Jpsi}{\tilde f}%{\psi} % pole skalarne niefizyczne
253: \newcommand{\Jpi}{{\tilde p}}%{\pi} %ped konforemnie przetransformowany
254: \newcommand{\Jped}{p}%{\pi} % ped na przestrzeni fizycznej
255: \newcommand{\tphi}{{\tilde\phi}}
256: \newcommand{\tpi}{{\tilde\pi}}
257:
258: \newcommand{\aalpha}{\lambda} % funkcja wyznaczajaca supertranslacje
259: \newcommand{\przecinekJped}{}
260: \newcommand{\const}{\mbox{\scriptsize const}}
261: \newcommand{\bM}{{\widetilde {\! M}}} %conformally completed Minkowski
262: \newcommand{\bF}{{\widetilde F}}
263: \newcommand{\bi}{i_{\bSigma}}%{{\tilde \imath}}
264: \newcommand{\bhyp}{{\widetilde \hyp}}
265: \newcommand{\bSigma}{{\widetilde \Sigma}}
266: \newcommand{\Sscr}{\Sigma_{\scrip}}
267: \newcommand{\iscr}{i_{\scrip}}
268: \newcommand{\iS}{i_{\Sigma}}
269: \newcommand{\iSs}{i_{\Sscr}}
270: \newcommand{\PhiS}{\Phi_{\Sigma}}
271: \newcommand{\PhiSscr}{\Phi_{\Sscr}}
272: \newcommand{\PhibS}{\Phi_{\bSigma}}
273: \newcommand{\PsiS}{\Psi_{I\times\Sigma}}
274: \newcommand{\psiS}{\psi_{I\times\Sigma}}
275: \newcommand{\PsiScr}{\Psi_{I\times\Sscr}}
276: \newcommand{\psiScr}{\psi_{I\times\Sscr}}
277: %\newcommand{\hyp}{{\mathscr S}}
278: \newcommand{\hyp}{{\mycal S}}
279: \newcommand{\tg}{{\tilde g}}
280: \newcommand{\hamX}{{\mathscr{X}}}
281: \newcommand{\jky}{{ v}}
282: \newcommand{\scrimap}{{\sigma}} %odwzorowanie opisujace dynamike na Scri
283: \newcommand{\fscri}{{\chi}} % wartosc pola skalarnego na Scri
284:
285: \newcommand{\Bgamma}{{B}} % koneksja tla
286: \newcommand{\bmetric}{{b}} % metryka tla
287:
288: \newcommand{\Kp}{{\mathfrak g}} %{\pi}} % kontrawariantna gestosc metryki
289: %na czasoprzestrzeni
290: \newcommand{\KA}{p} % ped sprzezony do kontrawariantnej gestosci
291: % metryki (A duze w innych pracach JK i JJ)
292: \newcommand{\E}[1]{{\rm e}^{#1}}
293: %\newcommand{\kolo}[1]{\stackrel{\circ}{#1}}
294: \newcommand{\ve}{\varepsilon}
295: \newcommand{\gthreeup}{\,{}^3g} %three dimensional metric in
296: %space--time, used for the inverse
297: \newcommand{\cO}{{\mathscr O}}
298: \newcommand{\arcsh}{{\rm argsh}}
299: \newcommand{\threeg}{{\gamma}} %three dimensional metric pulled back
300: %to the model space
301: \newcommand{\detthreeg}{{\det(\gamma_{ij})}} % determinant of the
302: % three dimensional
303: % metric pulled back
304: %to the model space
305: \newcommand{\lapse}{{\nu}} % the lapse function on the model space
306: \newcommand{\shift}{{\beta}} % the shift vector field on the model
307: % space
308: \newcommand{\threeP}{{\pi}} %three dimensional ADM momentum pulled back
309: %to the model space
310: \newcommand{\tildeg}{{\tilde g}} % conformally rescaled metric
311: \newcommand{\hst}{{\breve{h}}} % standard round metric on the two
312: % sphere
313: \newcommand{\ext}{{\mbox{\rm\scriptsize ext}}}
314: \newcommand{\Gtwo}{\two\Gamma}
315: \newcommand{\Gst}{\breve\Gamma}
316: \newcommand{\bel}[1]{\be\label{#1}}
317: \newcommand{\ourU}{\mathbb U}
318: \newcommand{\ourW}{\mathbb W}
319:
320: \newcommand{\hzome}{\widehat{\mathring{\omega}}}
321: \newcommand{\dvaromega}{\dot\varomega} %time derivative of the conformal factor
322: \newcommand{\ohyp}{\,\,\overline{\!\!\hyp}} % closure of the initial data hypersurface
323: \newcommand{\thyp}{\tilde\hyp}
324: \newcommand{\hhyp}{\,\,\widehat{\!\!\hyp}\,} % extension of the initial data
325: % hypersurface beyond Scri
326: \newcommand{\tmet}{\tilde\met} % conformally rescaled metric on a
327: % three dimensional manifold
328: \newcommand{\hmet}{\hat\met}
329: \newcommand{\met}{\gamma} % metric induced on a three dimensional hypersurface
330: \newcommand{\extK}{K} % extrinsic curvature of a three dimensional hypersurface
331: \newcommand{\tiext}{\tilde\extK } % conformally rescaled extrinsic
332: % curvature of a three
333: % dimensional hypersurface (with a tilde)
334: \newcommand{\hext}{\hat\extK } % extrinsic curvature of a three
335: % dimensional hypersurface with a hat
336: \newcommand{\hM}{\,\widehat{\!M}} % four dimensional extended spacetime
337: \newcommand{\hfg}{\hat g} % four dimensional extended metric
338: \newcommand{\calD}{{\mathcal{D}}} %domain of dependence
339: \newcommand{\tn}{{\tilde n}} %unit normal with respect to the rescaled
340: %metric
341: \newcommand{\varomega}{\omega} % conformal factor on a fixed hypersurface
342: \newcommand{\hvaromega}{\hat{\varomega}} % conformal factor on a fixed hypersurface,
343: % extended beyond the boundary
344: \newcommand{\hdotvaromega}{\widehat{\dot{\varomega}}} % conformal factor on a fixed hypersurface,
345: % extended beyond the boundary
346: \newcommand{\toto}{\R}% \toto used to be \R^*_+, replaced by \R
347: \newcommand{\tLambda}{\tilde\Lambda} % tensor densities
348: \newcommand{\newH}{\widehat H}
349:
350: \newcommand{\pgp}{P_{\hM}\gamma_p}
351: \newcommand{\pgq}{P_{\hM}\gamma_q}
352:
353: \newcommand{\conffour}{\varOmega} % four dimensional conformal factor
354: % here
355: \newcommand{\confthree}{\varomega} % three dimensional conformal factor
356: % here
357: \newcommand{\dconfthree}{\dvaromega} % time derivative of the three
358: % dimensional conformal factor
359: \newcommand{\extL}{L} % trace free part of the extrinsic curvature of
360: % a three dimensional hypersurface
361: \newcommand{\textL}{\tilde L} % conformally rescaled trace free part
362: % of the extrinsic curvature of
363: % a three dimensional hypersurface
364: %\newcommand{\textK}{\tilde K} % conformally rescaled extrinsic curvature of
365: % a three dimensional hypersurface
366: %\newcommand{\tK}{\tilde K} % trace of the conformally rescaled
367: % extrinsic curvature of
368: % a three dimensional hypersurface
369: \newcommand{\tlambda}{\tilde \lambda} % extrinsic curvature of
370: % a two dimensional boundary in a conformal picture
371: \newcommand{\ccalD}{\;\breve{\!\mycal D}}
372: \newcommand{\tilh}{\tilde h} % metric induced by the conformal metric on S^2
373: \newcommand{\DG}{D} % zbior odwzorowan konforemnych
374: \newcommand{\zR} {\zRs} % Ricci itd tla
375: \newcommand{\zRs} {\mathring{R}} % Ricci itd tla
376: \newcommand{\zRm} {\mathring{R}} % Ricci itd tla
377: \newcommand{\zo} {\mathring{\omega}} % koneksja metryki tla
378: \newcommand{\hzo} {\widehat{\zo}{}} % koneksja metryki tla
379: \newcommand{\znabla} {\mathring{\nabla}} % koneksja metryki tla
380: \newcommand{\sRz} {\mathring{\sigma}_R} % odleglosc geodezyjna metryki tla
381: \newcommand{\sR} {\sigma_R} % odleglosc geodezyjna metryki
382:
383: \newcommand{\frg}{{\mathfrak g}} % determinant of the metric
384: \newcommand{\fb}{{\mathfrak b}} % determinant of the background metric
385:
386: \newcommand{\zn} {\mathring{\nabla}} % pochodna metryki tla
387: \newcommand{\mn} {M} %{{}^{n-1}M} %the n-1 dimensional manifold
388: \newcommand{\cM}{\mycal M} % space-time
389: \newcommand{\cK}{\mycal K} % Killing vectors
390: \newcommand{\hcK} {\,\,\widehat{\!\!\cK}} % hatted Killing vectors
391: \newcommand{\cKS}{{\mycal K}_{\hyp^{\perp}}} % normal Killing vectors
392: \newcommand{\cKt}{{\mycal K}_{\hyp^{\parallel}}} % tangential Killing vectors
393: \newcommand{\cKST}{\cKS^\perp}%{{\mycal K}_{\mycal B}} % remainging Killing
394: % vectors
395: \newcommand{\Eq}[1]{Equation~\eq{#1}}
396: \newcommand{\Eqsone}[1]{Equations~\eq{#1}}
397: \newcommand{\Eqs}[2]{Equations~\eq{#1}-\eq{#2}}
398: \newcommand{\hr}{{\hat r}}
399: \newcommand{\hbeta}{{\hat \beta}}
400: \newcommand{\hx}{{\hat x}}
401: \newcommand{\hphi}{{\hat \phi}}
402: \newcommand{\hth}{{\hat \theta}}
403: \newcommand{\htheta}{{\hth}}
404: \newcommand{\home}{{\hat \omega}}
405: \newcommand{\hh}{{\hat h}}
406: \newcommand{\hf}{{\hat f}}
407: \newcommand{\hb}{{\hat b}}
408: \newcommand{\hv}{{\hat v}}
409: \newcommand{\he}{{\hat e}}
410: \newcommand{\proof}{\noindent {\sl Proof:} }
411: \newcommand{\remark}{\noindent {\bf Remark:} }
412: \newcommand{\remarks}{\noindent {\bf Remarks:} }
413: \newcommand{\qed}{\hfill $\Box$\\ \medskip}
414: \newcommand{\ds}{\; ds}
415: %\newcommand{\xymatrix}[1]{}
416: %\newcommand{\ar}{}
417: %\input{scrif1}
418: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
419: \begin{document}\newcommand{\Sext}{\hyp_{{\text{\scriptsize\rm ext}}}}
420: \newcommand{\fg}{{}^4{g}}
421: \newcommand{\ac}{\zeta}
422:
423:
424: \title{The mass of spacelike hypersurfaces in asymptotically anti-de
425: Sitter space-times}
426:
427: \author{Piotr T.\ Chru\'sciel\thanks{ Supported in part by the Polish
428: Research Council grant KBN 2 P03B 073 15, and by the A. von Humboldt foundation. E--mail:
429: \texttt{chrusciel@univ-tours.fr}} \\ Gabriel Nagy\thanks{
430: Current address: Max-Planck-Institut f\"ur Gravitationsphysik,
431: Albert-Einstein-Institut,
432: Am M\"uhlenberg 1,
433: D-14476 Golm,
434: Germany. Supported by a grant from R\'egion Centre. E--mail:
435: \texttt{nagy@aei-potsdam.mpg.de}} \\
436: D\'epartement de
437: Math\'ematiques \\ Facult\'e des Sciences\\
438: Parc de Grandmont \\F-37200 Tours, France}
439:
440:
441: \maketitle
442:
443: \begin{abstract}
444: We give a Hamiltonian definition of mass for spacelike hypersurfaces
445: in space-times with metrics which are asymptotic to the anti-de
446: Sitter one, or to a class of generalizations thereof. We show that
447: our definition provides a geometric invariant for a spacelike
448: hypersurface embedded in a space-time.
449: \end{abstract}
450: \tableofcontents
451:
452: \newcommand{\mSigma}{\cal S}
453: \section{Introduction}
454: \label{Si} Let $\hyp$ be an $n$-dimensional spacelike hypersurface
455: in a $n+1$-dimensional Lorentzian space-time $(\cM,g)$, $n\ge 2$.
456: Suppose that $\cM$ contains an open set $\cU$ with a global time
457: coordinate $t$ (with range not necessarily equal to $\R$), as well
458: as a global ``radial'' coordinate $r\in[R,\infty)$, leading to
459: local coordinate systems $(t,r,v^A)$, with $(v^A)$ --- local
460: coordinates on some compact $n-1$ dimensional manifold $\mn$. We
461: further require that $\hyp\cap \cU=\{t=0\}$. Assume that the
462: metric $g$ approaches (as $r$ tends to infinity, in a sense which
463: is made precise in Section~\ref{S2} below) a background metric $b$
464: of the form
465: \begin{equation}
466: \label{eq:a1}
467: b= -{\textstyle{\left({r^2\over\ell^2}+k\right)}} dt^2 + {1\over{r^2\over\ell^2}+k } dr^2
468: +r^2 h\;,
469: \end{equation}
470: where $h$ is an $r$-independent Riemannian metric on $\mn$, while
471: $k$ and $\ell$ are constants\footnote{\label{footWoolgar}A warped
472: product form of the metric, with the factor in front of $h$ not
473: being constant, together with the Einstein equations~\eq{Ee},
474: force $g_{rr}$ and $g_{tt}$ to have the form \eq{eq:a1} in an
475: appropriate coordinate system \cite{CadeauWoolgar}, with $k$ being
476: a function of $r$ which approaches a constant as $r$ tends to
477: infinity. Further $h$ itself has to satisfy the Einstein
478: equation~\eq{Ee} with $\Lambda$ replaced by an appropriate
479: constant. Some metrics slightly more general than \eq{eq:a1} will
480: be considered in the body of the paper. }.
481: Suppose further that $g$ satisfies the vacuum Einstein equations with
482: a cosmological constant
483: \begin{equation}
484: \label{Ee}
485: R_{\mu\nu}-\frac{g^{\alpha \beta} R_{\alpha \beta}}{2} g_{\mu\nu} =
486: -\Lambda g_{\mu\nu} \;, \qquad\Lambda = -\textstyle{n(n-1) \over
487: 2\ell^2}\;,
488: \end{equation}
489: similarly for $b$. (The existence of a large family of such $g$'s
490: follows from the work in \cite{Friedrich:adS,Kannar:adS}.) A
491: Hamiltonian analysis (following \cite{ChAIHP}, and discussed in
492: some more detail in Appendix \ref{AHam}; see also \cite[Section
493: 5]{ChruscielSimon}) leads to the following expression for the
494: Hamiltonian associated to the flow of a vector field $X$, assumed
495: to be a Killing vector field of the background $b$:\footnote{The
496: integral over $\partial \hyp$ should be understood by a limiting
497: process, as the limit as $R$ tends to infinity of integrals over
498: the sets $t=0$, $r=R$. $d S_{\alpha\beta}$ is defined as
499: % $\frac{\partial}{\partial x^\mu}\lrcorner
500: % \rd x^0 \wedge\rd x^1 \wedge\rd x^2 \wedge\rd x^3 $,
501: $\frac{\partial}{\partial x^\alpha}\lrcorner
502: \frac{\partial}{\partial x^\beta}\lrcorner \rd x^0 \wedge\cdots
503: \wedge\rd x^{n} $, with $\lrcorner$ denoting contraction; $g$ stands
504: for the space-time metric unless explicitly indicated
505: otherwise. Further, a semicolon denotes covariant differentiation
506: \emph{with respect to the background metric $b$}.}
507: \begin{eqnarray}
508: m(\hyp,g,b, X)&= &\frac 12 \int_{\partial\hyp}
509: \ourU^{\alpha\beta}dS_{\alpha\beta}\;,
510: \label{toto}
511: \end{eqnarray}\begin{eqnarray}
512: \ourU^{\nu\lambda}&= &
513: {\ourU^{\nu\lambda}}_{\beta}X^\beta + \frac 1{8\pi}
514: \left(\sqrt{|\det g_{\rho\sigma}|}~g^{\alpha[\nu}-\sqrt{|\det
515: b_{\rho\sigma}|}~ b^{\alpha[\nu}\right) {X^{\lambda]}}_{;\alpha} \
516: ,\label{Fsup2new}
517: \\ {\ourU^{\nu\lambda}}_\beta &= & \displaystyle{\frac{2|\det
518: \bmetric_{\mu\nu}|}{ 16\pi\sqrt{|\det g_{\rho\sigma}|}}}
519: g_{\beta\gamma}(e^2 g^{\gamma[\nu}g^{\lambda]\kappa})_{;\kappa}
520: \;,\label{Freud2.0} %\nonumber
521: \\
522: \label{mas2}
523: e&=& %\frac
524: {\sqrt{|\det g_{\rho\sigma}|}}/{\sqrt{|\det\bmetric_{\mu\nu}|}}\;
525: .
526: \end{eqnarray}
527: (The question of convergence of the right-hand-side of \eq{toto}
528: is considered in Section~\ref{S2} below.) The hypersurface $\hyp$
529: singles out a set of Killing vectors $X$ for the metric $b$ which
530: are normal to $\hyp$,
531: \begin{equation} \label{KVF}
532: X\Big|_{\hyp}=N n\;,
533: \end{equation}
534: where $N$ is a function and $n=e_0
535: =({r^2\over\ell^2}+k)^{-1/2}\partial_t$ is the future-directed
536: $b$-unit normal to $\hyp$. We shall use the symbol $\cKS$ to
537: denote this set of Killing vectors. The question then arises
538: whether one can extract out of \eq{toto}, with $X\in\cKS$, one or
539: more geometric invariants associated to $g$ along $\hyp$. Another
540: way of stating this question is, essentially, whether the
541: integrals \eq{toto} are background independent. As discussed in
542: more detail below, every metric $g$ asymptotes many different
543: backgrounds of the form \eq{eq:a1} whenever it asymptotes one, and
544: it is not at all clear how these backgrounds relate to each other:
545: if the geometry of space-time does not sufficiently constrain the
546: set of allowed backgrounds \eq{eq:a1}, then the numbers obtained
547: from \eq{toto} could be completely unrelated to each other when
548: different backgrounds are chosen. If this were the case, it would
549: appear questionable to associate physical meaning to the integrals
550: \eq{toto}. The purpose of this paper is to prove that, in several
551: cases of interest, geometric invariants can indeed be extracted
552: out of the integrals \eq{toto}.
553:
554: The model problem of interest are space-times which are asymptotic
555: to anti-de Sitter space-time. In this context there exist several
556: alternative methods of defining mass --- using coordinate systems
557: \cite{BGH,GHHP83}, preferred foliations \cite{GibbonsGPI},
558: generalized Komar integrals \cite{Magnon}, conformal techniques
559: \cite{AshtekarDas,AshtekarMagnonAdS,Balasubramanian}, or
560: \emph{ad-hoc} methods \cite{AbbottDeser}; an extended discussion
561: can be found in \cite[Section 5]{ChruscielSimon}. We wish to
562: stress that the key advantage of the Hamiltonian approach is the
563: uniqueness of the candidate expression for the energy (which
564: follows from the fact that Hamiltonians are uniquely defined up to
565: a constant on each path connected component of the phase space),
566: and that no such uniqueness properties are known in the
567: alternative approaches mentioned above (\emph{cf.}, however,
568: \cite{WaldZoupas,SilvaJulia:2000} for some partial results in the
569: ``Noether charges'' framework). Now, independently of the question
570: of what is the ``correct'' candidate expression for the energy,
571: each of the expressions proposed in the existing literature
572: suffers from some ambiguities, so that the question of
573: well-posedness of the definition of mass as defined in those
574: papers arises as well. For instance, the Abbott-Deser
575: mass~\cite{AbbottDeser}, or the Hamiltonian mass of \cite{HT},
576: both suffer from precisely the same potential ambiguities as the
577: Hamiltonian mass analyzed in this paper. As shown in
578: Appendix~\ref{AAD}, under the asymptotic conditions considered in
579: our well-posedness results, the Hamiltonian mass defined by
580: \eq{toto} coincides with the Abbott-Deser one. Thus, one way of
581: interpreting our results is that we prove the existence of a
582: geometric invariant which can be calculated using Abbott-Deser
583: type integrals. As another example, we note the potential
584: ambiguity in the mass defined by the conformal methods
585: in~\cite{AshtekarDas,AshtekarMagnonAdS}, related to the
586: possibility of existence of conformal completions which are
587: \emph{not} smoothly conformally equivalent. The results proved
588: here can be used to show~\cite{ChHerzlich} that no such
589: completions exist, establishing the invariant character of the
590: definitions of~\cite{AshtekarDas,AshtekarMagnonAdS}.
591:
592: We note that a similar problem for the ADM mass of asymptotically
593: flat initial data sets has been solved in
594: \cite{Bartnik:mass,ChErice} (see also \cite{Chmass}). Our
595: treatment here is a non-trivial adaptation to the current setup of
596: the methods of \cite{ChErice}. Some of the results proved here
597: have been independently observed in \cite{Wang}.
598:
599: The detailed statements of our results in a general context are to
600: be found in the sections below, and will not be reproduced here.
601: We shall, instead, discuss the application of our results to two
602: families of examples:
603: \begin{enumerate}
604: \item Let $M$ be the $n-1$ dimensional sphere $^{n-1}S$ with the
605: round metric $h$ of scalar curvature $(n-1)(n-2)$, while $b$ is the
606: $n+1$ dimensional anti-de Sitter metric:
607: \begin{equation}
608: \label{eq:a1a}
609: b= -\left({r^2\over\ell^2}+1\right) dt^2 + {1\over{r^2\over\ell^2}+1 } dr^2
610: +r^2 h\;,
611: \end{equation}
612: The space $\cKS$ of $b$-Killing vector fields normal to
613: $\hyp\cap\cU$ consists of vector fields $X(\lambda)$,
614: $\lambda=(\lambda^{(\mu)})\in \R^{n+1}$, which\footnote{We stress
615: that the index ${(\mu)}$ on
616: $\lambda$ does not have anything to do with space-time;
617: $\lambda^{(\mu)}$ is simply a coordinate on the $n+1$ dimensional
618: vector space $\cKS$. Similarly the Lorentz metric
619: $\eta_{{(\mu)}{(\nu)}}$, which arises naturally on $\cKS$ from the
620: construction here, does not have anything to do with the space-time
621: metric $g$. To emphasize this we put brackets around the $\mu$'s.}
622: on $\hyp$ take the form \eq{KVF} with $N=\lambda^{(\mu)}
623: N_{(\mu)}$, where $$N_{(0)}=\sqrt{{r^2\over\ell^2}+1}\;, \qquad
624: N_{(i)}={x^i\over\ell}\;,$$ and $x^i = r n^i$, $r$ being the
625: coordinate which appears in \eq{eq:a1}, while
626: $n^i\in{}^{n-1}S\subset\R^n$. The group $\mathit{Iso}(\hyp,b)$ of
627: isometries $\Phi$ of $b$ which map $\hyp$ into $\hyp$ acts on $\cKS$
628: by push-forward; in Appendix \ref{Aai1} we show for completeness the
629: well known fact that for every such $\Phi$ there exists a Lorentz
630: transformation $M:\R^{n+1}\to\R^{n+1}$ so that we have $$\Phi_*
631: X(\lambda)= X( M\lambda)\;.$$ Letting ${\mathfrak g}_{(\mu)}$ be the
632: $\mu$'th basis vector of $\cKS\approx\R^{n+1}$, $ {\mathfrak
633: g}_{(\mu)}:=X(\lambda^{(\alpha)} =\delta^\alpha_\mu)$, we set
634: $$m_{(\mu)}\equiv m\left(\hyp,g,b,{\mathfrak g}_{(\mu)}\right)\;;$$
635: it follows that the number $$m^2(\hyp,g) = |\eta^{{(\mu)}{(\nu)}}
636: m_{(\mu)} m_{(\nu)}|\;,$$ where
637: $\eta^{{(\mu)}{(\nu)}}=\mathit{diag}(-1,+1,\cdots,+1)$ is the
638: Minkowski metric on $\R^{n+1}$, is an invariant of the action of
639: $\mathit{Iso}(\hyp,b)$.\footnote{This has been observed
640: independently by X.~Wang~\cite{Wang} in, however, a considerably
641: less general setting.} Further, if we define $m(\hyp,g)$ to
642: be positive if $m^\mu$ is spacelike, while we take the sign of
643: $m(\hyp,g)$ to coincide with that of $m_0=-m^0$
644: if $m^\mu$ is timelike or
645: null, then $m(\hyp,g)$ so defined is invariant under the action of
646: the connected component $\mathit{Iso}_0(\hyp,b)$ of the identity in
647: $\mathit{Iso}(\hyp,b)$. We show in detail in Section~\ref{Sgc} that
648: $m(\hyp,g)$ is independent of the background metric chosen to
649: calculate the integrals \eq{toto}, provided that the fall-off
650: conditions of Theorem~\ref{T0} and Theorem~\ref{TP1} hold, which
651: justifies the notation. The number $m(\{t=0\},g)$ so defined
652: coincides with the mass parameter $m$ of the Kottler
653: metrics\footnote{The Kottler metrics, published in 1918
654: \cite{Kottler}, are also known as the ``Schwarzschild -- de Sitter
655: metrics''.} \begin{eqnarray}\label{Kottler} g & = & -\left(1 -
656: \frac {2m}r + {r^2\over \ell^2}\right) dt^2 + \left(1 - \frac {2m}r
657: + {r^2\over \ell^2}\right)^{-1} dr^2 + r^2 h \;,
658: \end{eqnarray}
659: where $h=d\theta^2+\sin^2\theta d\varphi^2$. Similarly
660: $m(\{t=0\},g)$ is proportional to the parameter $\eta$ which
661: occurs in the $(n+1)$-dimensional generalizations of the Kottler
662: metrics ({\em cf., e.g.,} \cite{HorowitzMyers})
663: \begin{eqnarray}\label{Kottler2}
664: g & = & -\left(1 - \frac {2\eta}{r^{n-2}} + {r^2\over \ell^2}\right) dt^2 +
665: \left(1 - \frac
666: {2\eta}{r^{n-2}} + {r^2\over \ell^2}\right)^{-1} dr^2 + r^2 h \;,
667: \end{eqnarray}
668: with $h$ --- a round metric on a $n-1$ dimensional sphere of
669: scalar curvature $(n-1)(n-2)$.
670:
671: Some further global geometric invariants of the metrics asymptotic
672: to the backgrounds \eq{Kottler2} are discussed in
673: Section~\ref{Sgc}.
674: \item Let $M$ be a compact $n-1$ dimensional manifold
675: with a metric $h$ of constant scalar curvature and with non-positive
676: Ricci tensor, and let $b$ take the form
677: \begin{equation}
678: \label{eq:a1c}
679: %b= -\textstyle\left({r^2\over\ell^2}-1\right) dt^2 +
680: %\displaystyle{1\over{r^2\over\ell^2}-1 } dr^2
681: % +r^2 h\;,
682: b= -{1\over a^2(r)} dt^2 + a^2(r) dr^2 +r^2 h\;,
683: \end{equation}
684: $h$ being $r$-independent, as before. We show (see
685: Proposition~\ref{PA1}, Appendix~\ref{Aai2}) that for such metrics
686: the space of $b$-Killing vector fields normal to $\hyp$ consists
687: of vector fields of the form
688: \begin{equation}
689: \label{eq:kvh}
690: X(\lambda)= \lambda\partial_t\;, \quad\lambda\in\R\;.
691: \end{equation}
692: The discussion in Section~\ref{Sgc} shows that $$m(\hyp,g)\equiv
693: m(\hyp,g,b,X(1))$$ is background independent, hence a geometric
694: invariant. Some other geometric invariants can be obtained from
695: the integrals \eq{toto} when Killing vectors which are not
696: necessarily normal to $\hyp$ exist, using invariants of the action
697: of the isometry group of $b$ on the space of Killing vectors. If
698: the Ricci tensor of $\mn$ is strictly negative no other Killing
699: vectors exist, \emph{cf.\/} Appendix~\ref{Aai2}. On the other
700: hand, if $h$ is a flat torus, then each $h$--Killing vector
701: provides a geometric invariant via the integrals \eq{toto},
702: provided that those converge and that the fall-off conditions of
703: Theorem~\ref{TP1} are met (this will be the case if,
704: \emph{e.g.,\/} Equations~\eq{C4}-\eq{C5} hold).
705: \end{enumerate}
706:
707: The number $m(\hyp,g)$ defined in each case above is our proposal
708: for the geometric definition of total mass of $\hyp$ in $(\cM,g)$.
709:
710: The results described above can be reformulated in a purely
711: Riemannian context, this will be discussed elsewhere
712: \cite{ChHerzlich}. The extension of the results proved here to
713: hyperboloidal hypersurfaces in Minkowski space-time, that leads
714: to a geometric definition of the Trautman-Bondi mass, requires a
715: considerable amount of work and will be discussed
716: elsewhere~\cite{ChJTB}. Let us simply mention that if the metric
717: of a hyperboloidal hypersurface in asymptotically Minkowskian
718: space-times satisfies the fall-off conditions here then its
719: Trautman-Bondi mass coincides with the Hamiltonian one. More
720: general statements require care.
721:
722: It is natural to study the behaviour of the mass when $\hyp$ is
723: allowed to move in $\cM$. A partial answer to this question is
724: given in Theorem~\ref{TP1} below. A complete answer would require
725: establishing an equivalent of Theorem~\ref{T1} in a space-time
726: setting. The difficulties that arise in the corresponding problem
727: for asymptotically Minkowskian metrics \cite{Chmass} suggest that
728: this might be a considerably more delicate problem, which we plan
729: to analyze in the future. It should be stressed that this problem
730: mixes two different issues, one being the potential background
731: dependence of \eq{toto}, another one being the possibility of
732: energy flowing in or out through the timelike conformal boundary
733: of space-time.
734:
735: This paper is organized as follows: In Section~\ref{S2} we present
736: conditions which guarantee convergence of the mass integrals
737: \eq{toto}, see Theorem~\ref{T0}. We also show that the integrals
738: \eq{toto} are invariant (Theorem~\ref{TP1}) or covariant
739: (Lemma~\ref{L1}) under a class of well-controlled coordinate
740: transformations, consisting of symmetries of the background, and
741: of certain generalizations of the usual ``supertranslations'' that
742: occur in the asymptotically flat case. Section~\ref{Sai} contains
743: the proof of the asymptotic symmetries theorem, Theorem~\ref{T1},
744: which is the key result in this work. In this theorem we show
745: that the coordinate transformations allowed by our conditions are
746: compositions of those considered in Section~\ref{S2}. In
747: Section~\ref{Sgc} we apply the previous results to the
748: construction of global geometric invariants in a reasonably
749: general setting. In Appendix~\ref{AHam} the Hamiltonian approach
750: to the definition of mass is examined in our context.
751: Appendix~\ref{Aiso} contains some results on Killing vectors which
752: are needed in the body of the paper.
753:
754:
755:
756: \section{Convergence, covariance under well behaved coordinate
757: transformations}\label{S2}
758: \newcommand{\zero}{0} \newcommand{\one}{1} \newcommand{\backg}{b}
759: \newcommand{\A}{A} \newcommand{\ta}{a} \newcommand{\ha}{a}
760: %\newcommand{\hb}{b}
761: \newcommand{\hc}{c} \newcommand{\gf}{f} \newcommand{\ce}{\beta} Let us
762: start by establishing convergence of the mass integrals \eq{toto}
763: --- this involves setting up appropriate boundary conditions on
764: $g$. Let, thus, $g$ and $b$ be two metrics on a set $\{R_0\le r
765: <\infty\;, \; (v^A)\in \mn\}$, let $e_a$ be an orthonormal frame
766: for $b$, set
767: \begin{equation}
768: \label{mas3}
769: e^{\mu\nu}\equiv g^{\mu\nu}-b^{\mu\nu}%\;, \qquad %\textstyle
770: \; ,
771: \end{equation} and let $e^{\ha b}\equiv g(\theta^a,\theta^b)-\eta^{ab}$
772: denote the coefficients of $e^{\mu\nu}$ with respect to the frame
773: $\theta^a$ dual to the $e_a$'s:
774: $$e^{\mu\nu}\partial_\mu\otimes \partial_\nu=e^{\ha\hc}e_{\ha}\otimes
775: e_{\hc}\;.$$ Here $\eta^{ab}=\mathrm{diag}(-1,+1,\cdots,+1)$. We
776: stress that we do {\em not} assume existence of global frames on
777: the asymptotic region: when $M$ is not parallelizable, then any
778: conditions on the $e^{ab}$'s, {\em etc.} assumed below should be
779: understood as the requirement of {\em existence of a covering of
780: $M$ by a finite number of open sets ${\mycal O}_i$ together with
781: frames defined on $[R_0,\infty)\times {\mycal O}_i$ satisfying the
782: conditions spelled out above.} The ``matter energy-momentum
783: tensor'' $T^\lambda{}_\kappa$ is defined as \be\label{N.5g2} 8\pi
784: T^\lambda{}_\kappa:= R^\lambda{}_\kappa -\frac 12
785: g^{\alpha\beta}R_{\alpha\beta} \delta ^\lambda_\kappa +\Lambda
786: \delta ^\lambda_\kappa\;. \ee In our first result we assume for
787: simplicity\footnote{Using \Eq{C3}, Appendix~\ref{AHam}, it is
788: straightforward to obtain results similar to Theorem~\ref{T0}
789: without the hypothesis that $b$ is Einstein. Similarly, the
790: hypothesis that $X$ is a Killing vector field can be relaxed using
791: the calculations of \cite[Appendix~B]{ChAIHP}; {\em cf.\/} also
792: \cite[Section~5.1]{CJK}.} that $b$ is Einstein, that is, $b$
793: satisfies \Eq{N.5g2} with $T^\lambda{}_\kappa=0$, with a
794: cosmological constant $\Lambda$ the sign of which is irrelevant
795: for the theorem that follows:
796: \begin{Theorem}\label{T0}
797: Let $X$ be a Killing vector of an Einstein metric $b$,
798: set
799: %$$|X|^2\equiv \sum_a
800: % |X^a|^2\right\;, \quad |\znabla X|\equiv \left(\sum_{a,b}
801: % |\znabla _bX^a|^2\right)^{1/2}\;,\quad |\zR|= \left(\sum_{a,b}
802: % |\zR_{ab}|^2\right)^{1/2}\;,$$
803: \be\label{c0}|X|^2\equiv \sum_a |X^a|^2\;, \quad |\znabla X|^2\equiv
804: \sum_{a,b} |\znabla _bX^a|^2\;,\quad |J|^2=
805: \sum_{b}|T^0{}_{b}|^2\;
806: \ee %\quad |\zR|^2= \sum_{a,b}|\zRm_{ab}|^2\;,$$
807: where $\znabla$ is the covariant derivative of $b$; the indices
808: here refer to a $b$-orthonormal frame such that $e_0$ is normal to
809: the hypersurface $t=0$.
810: %and $\zR=\zRm_{ab}\theta^a\otimes\theta^b$ is the Ricci tensor of $b$.
811: Suppose that $\lim_{r\to\infty} e^{ab}=0$ and that
812: \begin{eqnarray}\nonumber
813: \int_{\{r\ge R_0\}} \left\{|X|\left( |J|+{|\Lambda|} |e^{ab}b_{ab}| +
814: \sum_{a,b,c} e_a(e^{bc})^2 +\sum_{b,c}
815: |e^{bc}|^2 \right) \right.
816: \\ \left. + |\znabla X| \sum_{a,b,c,d,e}
817: |e_a(e^{bc})| \; |e^{de}| \right\} \; d\mu_b < \infty\;,
818: \label{c1}\end{eqnarray}
819: where $d\mu_b\equiv \sqrt{\det b_{ij}}dr\; dv^2\ldots dv^{n}$ is
820: the Riemannian measure induced on $\{t=0\}$ by $b$. Then the
821: right-hand-side of \Eq{toto}, understood as the limit as $R\to
822: \infty$ of integrals over the sets $\{r=R, t=0\}$, exists and is
823: finite.
824: \end{Theorem}
825:
826: \remark We note that the somewhat unexpected restriction on
827: integrability (for $\Lambda \ne 0$) of $e^{ab}b_{ab}$ arises also
828: in the requirement of a well defined generalized Komar mass for
829: static asymptotically anti-de Sitter metrics
830: \cite{Magnon,ChruscielSimon}.
831:
832: \medskip
833:
834: \proof We have
835: \begin{eqnarray}
836: \int_{\{r=R\}}
837: \ourU^{\alpha\beta}dS_{\alpha\beta}= 2\int_{\{R_0\le r\le R\}}
838: \znabla_\beta \ourU^{\alpha\beta}dS_{\alpha} + \int_{\{r=R_0\}}
839: \ourU^{\alpha\beta}dS_{\alpha\beta}\;.
840: \label{C2}
841: \end{eqnarray}
842: A formula for the volume integrand in \Eq{C2} is given in \Eq{C3},
843: Appendix~\ref{AHam}. Clearly conditions \eq{c1} guarantee
844: convergence of that volume integral to a finite value when $R$
845: tends to infinity. \qed
846:
847: In the remainder of the paper we will only consider background
848: metrics of the form
849: \begin{equation}
850: \label{A1}
851: b= -{ a^{-2}(r)} dt^2 + a^2(r) dr^2 +r^2 h\;, \qquad
852: h=h_{AB}(v^C)dv^A dv^B\;,
853: \end{equation}
854: where $h$ is a Riemannian metric on a $(n-1)$-dimensional compact
855: manifold $\mn$. The condition that $b$ is Einstein will not be
856: made unless explicitly stated otherwise. Let
857: $$\theta^0 = {1\over a} dt\;, \quad \theta^1= a dr\;, \quad \theta^A = r
858: \alpha^A\;,$$ where $\alpha^A$ is an $h$-orthonormal coframe. We
859: let $e_a$ be the frame dual to $\theta^a$,
860: \begin{eqnarray}
861: \label{frame}
862: e_0 = a \partial_t\;, \qquad e_1 = {1 \over a } \partial _r
863: % e_{\zero}= \frac{1}{\sqrt{k + \frac{r^2}{\ell^2}}}\partial_t\;,
864: % \quad e_{\one}= {\sqrt{k + \frac{r^2}{\ell^2}}} \partial_r\;,
865: \quad
866: e_{\A} = \frac 1r \ce_{\A}\;,
867: \end{eqnarray}
868: so that $\ce_A$ is a $h$-orthonormal frame on $(M,h)$. As an
869: application of Theorem~\ref{T0}, consider, first, the Killing
870: vector field $X=\partial_t=(1/a)e_0$ and suppose that
871: \begin{equation}
872: \label{A1.0} a(r)={\ell\over r } +
873: o(r)\;,\quad e_1(a)= -{1\over r } + o(r)\;, \end{equation} for
874: some constant $\ell$. We then have $|X|\approx \ell|\znabla
875: X|/\sqrt 2\approx{r/ \ell}$ and $d\mu_b= r^{n-2}\sqrt{\det
876: h}\,dr\, dv^2\ldots dv^{n}$.
877: When $g$ and $b$ are Einstein, the condition \eq{c1} will be
878: satisfied if
879: \begin{equation}
880: \label{C4}
881: e^{ab}= O(r^{-\beta})\;, \quad e_a(e^{bc}) = O(r^{-\beta})\;, \quad
882: b_{ab}e^{ab}= O(r^{-\gamma})\;,
883: \end{equation}
884: with
885: \begin{equation}
886: \label{C5}
887: \beta>n/2\;, \qquad \gamma > n\;.
888: \end{equation}
889: (We note that the generalized $n+1$ dimensional Kottler metrics
890: \eq{Kottler2} satisfy \eq{C4} with $\beta=n$, and with
891: $\gamma=2n$.) An identical convergence analysis applies for all
892: ``rotational'' $b$-Killing vector fields $X^A(v^B)\partial_{A}$
893: (whenever occurring) for all the metrics \eq{A1}-\eq{A1.0}, as
894: well as all the remaining Killing vectors for the
895: $(n+1)$-dimensional anti-de Sitter metric (listed in
896: Appendix~\ref{Aai1}), showing that all the corresponding charges
897: are finite when the conditions \eq{C4}-\eq{C5} hold. Surprisingly,
898: in retrospect the analysis in the case $\Lambda\ne 0$ turns out to
899: be simpler than that for the asymptotically Minkowskian case,
900: where $\Lambda = 0$: in the latter case the requirement of
901: convergence of angular momentum or of boost integrals imposes more
902: stringent conditions on the metric than that of convergence of the
903: energy-momentum integrals.
904:
905: The conditions presented above are sufficient, but certainly not
906: necessary, for convergence of the integrals \eq{toto}: indeed, the
907: metric considered in Proposition~\ref{P2} below has a convergent
908: mass integral, but the conditions of Theorem~\ref{T0} fail to
909: hold. However, there is a potential essential ambiguity in the
910: definition of the integrals \eq{toto}, which we will describe
911: now. Proposition~\ref{P2} below then shows that \eq{C4}-\eq{C5}
912: are essentially sharp, if one requires that the integrals
913: \eq{toto} are convergent and {\em background-independent}.
914:
915: The ambiguity in the integrals \eq{toto} arises as follows: to
916: define those integrals we have fixed a model background metric $b$
917: with the corresponding coordinate system $(t,r,v^A)$ as in
918: \eq{A1}, as well as an orthonormal frame as in \eq{frame}. Once
919: this has been done, let $g$ be any metric such that the frame
920: components $g^{ab}$ of $g$ tend to $\eta^{ab}$ as $r$ tends to
921: infinity in such a way that the integrals $m(\hyp,g,b, X)$ given
922: by \eq{toto} (labeled by the background Killing vector fields $X$
923: or perhaps by a subset thereof) converge. Consider another
924: coordinate system $(\hat t,\hat r,\hat v^A)$ with the associated
925: background metric $\hat b$:
926: \begin{equation}
927: \label{A1hat}
928: \hat b= -{1\over a^2(\hat r)} d\hat t^2 + a^2(\hat r) d\hat r^2 +\hat r^2 \hat h\;, \qquad
929: \hat h=h_{AB}(\hat v^C)d\hat v^A d\hat v ^B\;,
930: \end{equation}
931: together with an associated frame $\hat e^a$,
932: \begin{eqnarray}
933: \label{hframe}
934: \hat e_0 = a(\hat r) \partial_{\hat t}\;, \qquad \hat e_1 = {1
935: \over
936: a(\hat r) } \partial _{\hat r}
937: \quad
938: \hat e_{\A} = \frac {1 }{\hat r} \hat{\ce}_{\A}\;,
939: \end{eqnarray}
940: and suppose that in the new hatted coordinates the integrals
941: defining the charges $m(\hhyp,g,\hat b, \hat X)$ converge again.
942: An obvious way of obtaining such coordinate systems is to make a
943: coordinate transformation
944: \begin{equation}
945: \label{eq:A5}
946: t\to \hat t=t+\delta t\;,\qquad r\to \hat r=r+\delta r\;,\qquad
947: v^A\to \hat v^A=v^A+\delta v^A\;,
948: \end{equation}
949: with $(\delta t, \delta r,\delta v^A)$ decaying sufficiently fast,
950: as \emph{e.g.\/} in the statement of Theorem~\ref{TP1} below.
951: (However, we do not know \emph{a priori\/} that the hatted
952: coordinates are related to the unhatted one by the simple
953: coordinate transformation \eq{eq:A5} with $(\delta t, \delta
954: r,\delta v^A)$ decaying as $r\to\infty$, or behaving in some
955: controlled way --- the behaviour of $(\delta t, \delta r,\delta
956: v^A)$ could be very wild.) The question then arises, how do the
957: $m(\hhyp,g,\hat b, \hat X)$'s relate to the $m(\hyp,g,b, X)$'s. A
958: geometric definition of mass should be coordinate-independent,
959: therefore one would like to have a simple relation between those
960: integrals.
961:
962: At this point it is worth recalling that there exist several
963: expressions for mass alternative to \eq{toto}, which might or
964: might not coincide with each other when the decay of the metric is
965: too slow. For example, we show in Appendix~\ref{AAD} that
966: \eq{toto} coincides with the Abbott-Deser \cite{AbbottDeser} mass
967: for all metrics satisfying the decay conditions \eq{C4}-\eq{C5}
968: for Killing vectors such that $|X|=O(r)$ with, say, $a(r)$ as in
969: \Eq{eq:a1}. Now, if $X=\partial_t$, for background metrics of the
970: form \eq{eq:a1}, in space-times of dimension $4$, the integral
971: defining the Abbott-Deser $m_{AD}$ can be written in a
972: particularly simple form \cite{ChruscielSimon}
973: \begin{eqnarray}
974: m_{AD}(\{t=0\},g, b, \partial_t)
975: =\lim_{R\to\infty} \frac {R^3}
976: {16\pi\ell^2}\int_{\Sigma\cap\{r=R\}} \left(r\sum_A{\frac {
977: \partial e^{AA}}{\partial r}}-2e^{\one\one}%+e^{\hA\hA}
978: \right) d^2\mu_{h}\;.
979: \label{massequation}
980: \end{eqnarray}
981: Generalizing an argument of \cite{DenisovSolovev}, we show that if
982: the decay conditions in \eq{C4} are too weak then one can obtain
983: essentially any value of $m_{AD}(\hhyp,g,\hat b, \hat X)$ by performing
984: coordinate transformations of the form \eq{eq:A5}. We do this
985: explicitly in $n=3$, the same argument applies in any dimension $n$:
986:
987: \begin{Proposition}\label{P2}
988: Let the physical metric $g$ equal the background metric $b$, and
989: let $\{r,v^A\}$ be coordinates so that $b$ takes the form
990: \eq{eq:a1a}. Consider a new set of coordinates defined as
991: \be\label{ct1} \hat{r} = r+ \frac{\ac}{r^{1/2}} \;,\quad
992: \hat{v}^A= v^{A}\;,\ \ee
993: where $\ac$ is a constant. (This leads to $\hat e^{ab}=O(r^{-3/2})$.)
994: If $\ac\ne 0$ then the mass $m_{AD}(\{t=0\},g,\hat b,
995: \partial_t) $ of $g$ with respect to the background metric $\hat b$
996: defined by the coordinates $\{\hat{r},\hat{v}^A\}$ does not
997: vanish.
998: \end{Proposition}
999:
1000: \proof First notice that this transformation satisfies
1001: $\frac{\delta r}{r} = O\left( r^{-3/2}\right)$. Then, by
1002: straightforward computations one has, assuming without loss of
1003: generality that $\ell=1$,
1004: $$
1005: \hat{e}_{1} = \left[ 1+ \frac{3\ac}{2r^{3/2}} +
1006: \frac{3\ac^2}{4r^3} + O( {r^{-7/2}})\right] e_1\;,
1007: $$
1008: $$
1009: \hat{e}_{A} = \left[ 1- \frac{\ac}{r^{3/2}} + \frac{\ac^2}{r^3}
1010: + O( {r^{-7/2}})\right] e_A\;,
1011: $$
1012: and so,
1013: $$
1014: \hat{e}^{11} = \frac{3\ac}{r^{3/2}} + \frac{15 \ac^2}{4r^3} +O(
1015: {r^{-7/2}})\;,
1016: $$
1017: $$
1018: \sum_A{e}^{AA} = -\frac{4\ac}{r^{3/2}} + \frac{6\ac^2}{r^3} + O(
1019: {r^{-9/2}})\;.
1020: $$
1021: Hence
1022: $$
1023: r\sum_A\partial_r({e}^{AA}) - 2 {e}^{11} = -\frac{51 \ac^2}{2
1024: r^3} + O( {r^{-7/2}})\;,
1025: $$
1026: and the result follows. \qed
1027:
1028:
1029: While the above shows that the Abbott-Deser mass ceases to be well
1030: defined below the threshold $o(r^{-3/2})$ in dimension $3+1$, this
1031: still leaves open the unlikely possibility that the Hamiltonian
1032: mass \eq{toto} could be well defined. In order to see that this is
1033: not the case let us, first, calculate the mass integrand for
1034: metrics of the form \bel{acom} g= -\mu
1035: \left(\frac{r^2}{\ell^2}+k\right) dt^2 +\displaystyle
1036: \frac{\nu}{\frac{r^2}{\ell^2}+k} dr^2 + \sigma r^2
1037: h_{AB}dv^Adv^B\;, \ee where $\mu$, $\nu$ and $\sigma$ are
1038: arbitrary functions. One finds
1039: \begin{eqnarray}
1040: \ourU^{tr}
1041: &=&\frac{r^{n}\sigma^{(n-1)/2}\sqrt{\det{h_{AB}}}}{16\pi
1042: \ell^2\sqrt{\mu\nu}} \left\{
1043: \frac{2\sqrt{\mu\nu}}{\sigma^{(n-1)/2}}-
1044: (\mu+\nu)\right.\nn\\
1045: &&+ \left. (n-1)\left(1+\frac{\ell^2k}{r^2}\right)\mu
1046: \left[\frac{1}{\sigma}\left(r\frac{\partial \sigma}{\partial r}
1047: -\nu\right)+ 1\right]\right\}\;.\label{acomm}
1048: \end{eqnarray}
1049: Suppose that ---in space-time dimension $n+1$ --- $g$ is the
1050: metric $b$ expressed in a hatted coordinate system $(\hr,\hv^A)$,
1051: and consider the coordinate transformation
1052: $$ \hat{r} = r+ \frac{\ac}{r^{n/2-1}} \;,\quad \hat{v}^A= v^{A}\;,
1053: $$
1054: where $\ac$ is a constant. The metric $g$, when expressed in the
1055: unhatted coordinates $(r,v^A)$, satisfies \eq{C4} with
1056: $\beta=\gamma=n/2$, and is of the form \eq{acom} so that
1057: \eq{acomm} applies. A {\sc Mathematica} calculation then shows
1058: that $g$ has a mass integral \eq{toto} with respect to the
1059: unhatted background $b$ equal to
1060: $$
1061: \frac{\text{Vol}_h(M)}{8\pi\ell^2} \left(n+{n^2\over
1062: 8}\right)\ac^2\;,$$ which is non-zero for any $\ac\ne 0$ and for
1063: any $n\in\N$. Here $\text{Vol}_h(M)$ is the $n-1$-dimensional volume of $M$
1064: ---
1065: area if $n-1=2$ --- with respect to the metric $h$.
1066:
1067: The coordinate transformation \eq{ct1} is not yet as good
1068: as one would wish, because it leads --- in space dimension three
1069: -- to coordinates in which the trace of the metric $e^{ab}b_{ab}$
1070: is $O(r^{-n/2})$, quite a bit above the threshold $r^{-n}$ set
1071: forth in \Eqs{C4}{C5}. We note that the change of coordinates
1072: \eq{ct1} accompanied by a further time redefinition (which clearly
1073: does not change the mass as given by \Eq{massequation})
1074: $$ \bar t =
1075: t(1+cr^{-3/2})\;,$$ with an appropriate choice of the constant
1076: $c$, will lead to a metric which at $\bar t = t=0$ satisfies
1077: \begin{equation}
1078: \label{C4.}
1079: e^{ij}= O(r^{-3/2})\;, \quad e_k(e^{ij}) = O(r^{-3/2})\;, \quad
1080: b_{ab}e^{ab}= O(r^{-3})\;,
1081: \end{equation}
1082: where the indices $i,j$ run from $0$ to $3$. Note that the above
1083: fall-off conditions will not hold for some of the $e^{0a}$'s, and
1084: for some $e_0$ derivatives of the $e_{ab}$'s, but this turns out
1085: irrelevant for the problem at hand: the new hypersurface $\bar
1086: t=0$ coincides with the previous one, therefore its extrinsic
1087: curvature will not change. One can check~\cite{ChHerzlich} that
1088: --- similarly to the ADM case
1089: --- conditions on the induced metric on the surface $t=0$ and on
1090: its extrinsic curvature are sufficient for a well defined notion
1091: of mass, so that the result in~\cite{ChHerzlich} complete the
1092: proof of sharpness of the condition on $\gamma$ in \eq{C5}.
1093:
1094: Let us show that the decay rates \eq{C4}-\eq{C5} guarantee
1095: non-occurrence of the behaviour exhibited in Proposition~\ref{P2}:
1096:
1097: \begin{Theorem}\label{TP1}
1098: Consider an $n+1$ dimensional space-time
1099: $(\cM, g)$, and let $b$ and $\hat{b}$ be two background metrics of
1100: the form \eq{A1} and \eq{A1hat}, with $a(r)$ as in \Eq{eq:a1}, in
1101: coordinates $\{t,r,v^{A}\}$ and $\{\hat{t},\hat{r},\hat{v}^{A}\}$
1102: with ranges covering $\{r\ge R_0\}\times\mn$ and $\{\hat r\ge \hat
1103: R_0\}\times\mn$ for some $R_0,\hat R_0\in \R$. Suppose that $b$
1104: satisfies the vacuum Einstein equations with a negative
1105: cosmological constant, that the conditions of Theorem \ref{T0}
1106: hold both for the hatted and unhatted coordinates, and that we
1107: have
1108: \begin{eqnarray}
1109: e^{ab}= o(r^{-n/2})\;, \quad e_c(e^{ab})= o(r^{-n/2})\;.
1110: \label{eq:deccond}
1111: \end{eqnarray}
1112: Let $X
1113: =X^a({t},{r},{v}^A)e_a\in \cK$ be a Killing vector field of the
1114: metric $b$ satisfying
1115: \begin{eqnarray}
1116: |X| + |\znabla X| = O(r)\;, \label{eq:deccond1}
1117: \end{eqnarray}
1118: and let $\hat{X}=X^a(\hat{t},\hat{r},\hat{v}^A)\he_a \in \hcK$ be
1119: its hatted counterpart (with the $\he_a$ components of $\hat X$
1120: given by the {\em same} functions $X^a$ of the hatted variables as
1121: the $e_a$ components of $X$ in the unhatted variables). Let $\hyp$
1122: and $\hhyp$ be the hypersurfaces given by $t=0$ and $\hat{t} =0$
1123: respectively. If the coordinate transformation satisfies
1124: $$
1125: \hat{t} = t + o(r^{-(1+n/2)})\;,\qquad e_a(\hat{t}) =
1126: \ell\,\delta_a^0 + o(r^{-(1+n/2)})\;,
1127: $$
1128: $$
1129: \hat{r} = r + o(r^{1-n/2})\;,\qquad e_a(\hat{r}) =
1130: \frac{\delta_a^1}{\ell} + o(r^{1-n/2})\;,
1131: $$
1132: \be \hat{v}^A = v^A + o(r^{-(1+n/2)})\;,\qquad e_a(\hat{v}^A) =
1133: \delta_a^A + o(r^{-(1+n/2)}) \;, \label{coordtrP}\ee then
1134: $$m(\hyp,g,b,X)=m(\hhyp,g,\hat{b},\hat{X})\;.$$
1135: \end{Theorem}
1136:
1137:
1138: \proof The idea of the proof is to compute the background metric
1139: $\hat{b}$ in a frame related to the unhatted coordinates,
1140: obtaining an expression in terms of $b$ plus correction terms.
1141: Then, we compute (\ref{Fsup2new}) for $\hat{b}$, similarly
1142: obtaining an expression in terms of (\ref{Fsup2new}) for $b$ plus
1143: additional terms. We show that these terms integrate to zero, up
1144: to terms vanishing in the limit as $r$ tends to infinity, keeping
1145: thus the mass invariant.
1146:
1147: In terms of $\delta t$, $\delta r$, and $\delta v^A$ defined as in
1148: \Eq{eq:A5}, the decay conditions \eq{coordtrP} imply
1149: \begin{equation}
1150: \label{fc-dx} \sqrt{k+r^2/\ell^2} \, \delta t =o(r^{-n/2})\;,
1151: \quad \frac{\delta r}{ \sqrt{k+r^2/\ell^2}} = o(r^{-n/2})\;, \quad
1152: r \delta v^A = o(r^{-n/2})\;.
1153: \end{equation}
1154: From \Eq{hframe} one finds the following relation between the
1155: hatted and non-hatted coframes
1156: $$
1157: \hat{\theta}^0 = \left(1+\frac{r\delta r}{{r^2+ k
1158: \ell^2}}\right)\theta^0 +
1159: {\sqrt{k+r^2/\ell^2}}d(\delta t) +
1160: o(r^{-n})\;,
1161: $$
1162: $$
1163: \hat{\theta}^1 = \left(1-\frac{r\delta r}{{r^2+ k \ell^2}}\right)
1164: \theta^1 +
1165: \frac 1{\sqrt{k+r^2/\ell^2}}d(\delta r) +
1166: o(r^{-n})\;,
1167: $$
1168: \be\label{tettr} \hat{\theta}^A = \theta^A + \frac{\delta r}{r}
1169: \theta^A + r \left\{ \frac{\partial \alpha^A{}_B}{\partial
1170: v^C}\,\delta v^C \, dv^B + \alpha^A{}_B \,d(\delta v^B) \right\} +
1171: o(r^{-n})\;. \ee
1172: where $\alpha^A$ ia a co-frame on $M$ dual to $\beta_A$, and
1173: we have introduced the notation $\alpha^A = \alpha^A{}_B dv^B$,
1174: with the index $A$ being a tetrad index, while the index $B$ is a
1175: coordinate index. All the terms denoted by $o(r^{-n})$ above have
1176: $o(r^{-n})$ coefficients when expressed in terms of the $\theta^a$
1177: frame. Actually, the term in the curly brackets in the right hand
1178: side of the last equation gives a clue to a convenient way of
1179: writing these equations, since it can be rewritten as
1180: $\lie_{\delta v^B
1181: \partial/\partial v^B }\alpha^A$, where $\lie$ denotes a Lie
1182: derivative; in order to justify such a procedure, we use the
1183: following artifact: As explained in \cite[Section~4]{CT93},
1184: embedding $\mn$ in $\R^{2(n-1)}$ and extending the metric
1185: appropriately, we can without loss of generality assume that the
1186: coordinates $v^A$ and the frames $\theta^A$ are globally defined
1187: on $\mn$. We then set $$ \zeta = \delta t \frac{\partial}{\partial
1188: t} + \delta r \frac{\partial}{\partial r} + \delta v^A
1189: \frac{\partial }{\partial v^A}\;; $$ one sees from (\ref{fc-dx})
1190: that the components of $\zeta$ in the $e_a$ frame are all of the
1191: same order $o(r^{-n/2})$, and one can check that \Eq{tettr} for
1192: the hatted tetrad reduces to $$ \hat{\theta}^a = \theta^a +
1193: \lie_{\zeta}\theta^a + o(r^{-n})\;;
1194: $$ we note that $\lie_{\zeta}\theta^a = o(r^{-n/2})$. Let us write
1195: $$\hat{e}_a = e_a + \delta e_a\;;$$ one verifies that the tetrad
1196: components of $\delta e_a$ are $o(r^{-n/2})$, and from the
1197: condition $\theta^a(e_b) = \delta^a_b$ and its hatted analogue one
1198: has
1199: \begin{eqnarray*} \theta^a(\delta e_b) &=&- \lie_{\zeta}\theta^a(e_b)
1200: + o(r^{-n})
1201: \\ & = &
1202: - \underbrace{\lie_{\zeta}\left(\theta^a(e_b)\right)}_0+
1203: \theta^a(\lie_{\zeta}e_b) + o(r^{-n})
1204: \\ & = &
1205: \theta^a(\lie_{\zeta}e_b) + o(r^{-n})
1206: \;,
1207: \end{eqnarray*}
1208: which shows that
1209: $$
1210: \hat{e}_a = e_a + \lie_{\zeta}e_a + o(r^{-n})\;,
1211: $$
1212: with $\lie_{\zeta}e_a = o(r^{-n/2})$. This looks to leading order
1213: like a change of tetrad under an infinitesimal transformation, but
1214: we emphasize that \emph{we are not} assuming that the
1215: transformation is infinitesimal. Denoting $\{x^{\mu}\} =
1216: \{t,r,v^A\}$, and using $\hat{b}^{\mu\nu} = \eta^{ab}
1217: \hat{e}_a{}^{\mu}\hat{e}_b{}^{\nu}$, and $b^{\mu\nu} = \eta^{ab}
1218: e_a{}^{\mu}e_b{}^{\nu}$, we obtain
1219: \begin{equation}
1220: \label{b-freedom} \hat{b}^{\mu\nu} = b^{\mu\nu} +
1221: \lie_{\zeta}b^{\mu\nu} + \delta b^{\mu\nu}\;,
1222: \end{equation}
1223: with $$\delta b^{ab}\equiv \delta
1224: b^{\mu\nu}\theta^a{}_{\mu}\theta^b{}_{\nu} = o(r^{-n})\;.$$ The
1225: expression \eq{b-freedom} is the first step to compute the change
1226: in the integrand of (\ref{toto}). The next step is to rewrite
1227: (\ref{Fsup2new}) in the following more convenient way $$
1228: \ourU^{\alpha\beta} = \frac{3}{8\pi} \sqrt{|\frg|}
1229: g_{\gamma\sigma} g^{\kappa[\alpha} X^{\beta}
1230: \zn_{\kappa}g^{\gamma]\sigma} + \frac{1}{8\pi} \left(\sqrt{|\frg|}
1231: g^{\kappa[\alpha}b^{\beta]\gamma} - \sqrt{|\fb|}
1232: b^{\kappa[\alpha}b^{\beta]\gamma}\right)
1233: b_{\gamma\sigma}\zn_{\kappa}X^\sigma\;, $$ with
1234: $\frg=\det(g_{\mu\nu})$, and $\fb=\det(b_{\mu\nu})$. Let
1235: $\widehat{\ourU}^{\alpha\beta}$ be defined as the expression above
1236: with $b$ and $X$ replaced with $\hat{b}$ and $\hat X$; from
1237: \Eq{b-freedom} we obtain
1238: \begin{eqnarray*}
1239: \widehat{\ourU}^{\alpha\beta} -\ourU^{\alpha\beta} &=&
1240: -\frac{\sqrt{|\fb|}}{8\pi} \left[
1241: 3\delta_{\lambda\nu\mu}^{\alpha\beta\gamma} X^{\mu}
1242: b^{\kappa\lambda} b_{\gamma\sigma} (\zn_{\kappa}\zn^{\nu}
1243: \zeta^{\sigma} + \zn_{\kappa}\zn^{\sigma}\zeta^{\nu}) \right] \\
1244: && - \frac{\sqrt{|\fb|}}{8\pi} \left[
1245: b^{\kappa[\alpha}b^{\beta]\gamma} \zn_{\mu}\zeta^{\mu} -
1246: b^{\gamma[\beta}\zn^{|\kappa|}\zeta^{\alpha]} - b^{\gamma[\beta}
1247: \zn^{\alpha]}\zeta^{\kappa} \right] \zn_{\kappa}X_{\gamma} +
1248: \delta\ourU^{\alpha\beta}\;,
1249: \end{eqnarray*}
1250: with $$\delta \ourU^{ab}\equiv\delta \ourU^{\alpha\beta}
1251: \theta^a{}_{\alpha} \theta^b{}_{\beta} = o(r^{-n})\;.$$ We have
1252: used the fact that $\lie_{\zeta} b^{\alpha\beta}= -2
1253: \zn^{(\alpha}\zeta^{\beta)}$, and that $\sqrt\frg =\sqrt\fb(1 +
1254: \zn_{\mu}\zeta^{\mu} - e_\mu{}^\mu/2) + o(r^{-n})$.
1255:
1256: The idea now is to write the right hand side above as a total
1257: divergence of a totally antisymmetric tensor density. The first
1258: term at the right hand side above can be written as
1259: \begin{eqnarray*}
1260: \lefteqn{ 3\delta_{\lambda\nu\mu}^{\alpha\beta\gamma} X ^{\mu}
1261: b^{\kappa\lambda} b_{\gamma\sigma} (\zn_{\kappa}\zn^{\nu}
1262: \zeta^{\sigma} + \zn_{\kappa}\zn^{\sigma}\zeta^{\nu}) =
1263: 3 \zn_{\gamma} \left[
1264: \delta^{\alpha\beta\gamma}_{\lambda\nu\mu} X
1265: ^{\mu}\zn^{\lambda}\zeta^{\nu}\right] } && \\
1266: &&+\zRm_{\gamma\rho}{}^{\alpha\beta} X ^{\gamma}\zeta^{\rho} + 2
1267: X ^{[\beta}\zRm^{\alpha]}{}_{\rho} \zeta^{\rho} - (\zn_{\gamma}X
1268: ^{\beta}) \zn^{[\gamma}\zeta^{\alpha]} +(\zn_{\gamma}X ^{\alpha})
1269: \zn^{[\gamma}\zeta^{\beta]} \;.
1270: \end{eqnarray*}
1271: Since $\zn_{\alpha}X _{\beta}$ is antisymmetric, one obtains
1272: \begin{eqnarray*}\hat{\ourU}^{\alpha\beta} -\ourU^{\alpha\beta} &=&
1273: -\frac{\sqrt{|b|}}{8\pi} \left\{3\zn_{\gamma} \left[ \,
1274: \delta^{\alpha\beta\gamma}_{\lambda\nu\mu} X
1275: ^{\mu}\zn^{\lambda}\zeta^{\nu}\right]
1276: +\zRm_{\gamma\rho}{}^{\alpha\beta} X ^{\gamma}\zeta^{\rho} \right.\\
1277: && \left.+ 2 X ^{[\beta}\zRm^{\alpha]}{}_{\rho} \zeta^{\rho} + 3
1278: (\zn^{[\alpha}X ^{\beta})(\zn_{\gamma}\zeta^{\gamma]})\right\}+
1279: \delta \ourU^{\alpha\beta}\;.
1280: \end{eqnarray*}
1281: Finally, with the remaining terms we construct the divergence of a
1282: totally antisymmetric tensor, as follows: $$ 3
1283: (\zn_{\gamma}\zeta^{[\gamma})(\zn^{\alpha}X ^{\beta]}) = 3
1284: \zn_{\gamma}\left[\zeta^{[\gamma}(\zn^{\alpha}X ^{\beta]})\right]
1285: - \zeta^{\gamma}\zn_{\gamma}\zn^{\alpha}X ^{\beta} - 2
1286: \zeta^{[\alpha}\zRm^{\beta]}{}_{\rho} X ^{\rho}.
1287: $$ Recalling that a Killing vector satisfies $\zn_{\alpha}\zn_{\beta}
1288: X _{\gamma} = \zRm^{\rho}{}_{\alpha\beta\gamma} X _{\rho}$, and
1289: that $\zRm_{\alpha\beta} = 2\Lambda b_{\alpha\beta}/(n-1)$, one
1290: finds $$ \widehat{\ourU}^{\alpha\beta} -\ourU^{\alpha\beta} =
1291: -\frac{3\sqrt{|\fb|}}{8\pi} \zn_{\gamma} \left[ X^{[\gamma}
1292: \zn^{\alpha}\zeta^{\beta]} + \zeta^{[\gamma}
1293: \zn^{\alpha}X^{\beta]} \right] + \delta \ourU^{\alpha\beta}\;.
1294: $$
1295: %with $\delta \ourU^{\alpha\beta} \theta^a{}_{\alpha}
1296: %\theta^b{}_{\beta} = o(r^{-n})$.
1297: (It would suffice that $\zRm_{\alpha\beta} = 2\Lambda
1298: b_{\alpha\beta}/(n-1) +O(r^{-n/2})$ for the argument to go
1299: through.) The first term on the right hand side above integrates
1300: out to zero on $(n-1)$ dimensional boundaryless compact
1301: submanifolds, and the remainder is order $o(r^{-n})$, so that
1302: \begin{equation}
1303: \label{eq:*}
1304: \int_{ r = R} \hat \ourU^{\alpha\beta}dS_{\alpha\beta} = \int_{ r
1305: = R} \ourU^{\alpha\beta}dS_{\alpha\beta} + o(1)\;,
1306: \end{equation}
1307: with $o(1)$ tending to zero as $r$ goes to infinity. We also have
1308: \begin{equation}
1309: \label{eq*}
1310: \int_{ r = R} \hat \ourU^{\alpha\beta}dS_{\alpha\beta} = \int_{
1311: \hat r
1312: = \hat R} \hat \ourU^{\alpha\beta}dS_{\alpha\beta}
1313: +2 \int_{V_{R,\hat R}} \znabla_\alpha \hat
1314: \ourU^{\alpha\beta}dS_{\beta} \;,
1315: \end{equation}
1316: where $V_{R,\hat R}$ is a set the boundary of which is the union
1317: of the coordinates sets $\{ r = R\}$ and $\{\hat r = \hat R\}$.
1318: Conditions \eq{c1} guarantee that the volume integral in \Eq{eq*}
1319: tends to zero when both $R$ and $\hat R$ tend to infinity, which
1320: together with \eq{eq:*} establishes our claims.\qed
1321:
1322:
1323:
1324:
1325:
1326:
1327:
1328: Let us finally show that the integrals \eq{toto} are {\em
1329: covariant} under isometries of the background. In what follows
1330: $\hyp$ is an arbitrary hypersurface on which the charge integrals
1331: \eq{toto} converge:
1332: \begin{Lemma}
1333: \label{L1}
1334: Let $\Phi:\cM\to\cM$ be an isometric diffeomorphism of $(\cM,b)$
1335: such that $\Phi(\hyp)=\hyp.$ Then
1336: \begin{equation}
1337: \label{eq:l1}
1338: m(\hyp,\Phi^*g,b,(\Phi_*)^{-1}X) = m(\hyp,g,b,X)\;.
1339: \end{equation}
1340: \end{Lemma}
1341:
1342:
1343: \proof Formula \eq{toto} for mass is invariant under
1344: diffeomorphisms, hence
1345: \begin{eqnarray*}
1346: m(\Phi(\hyp),g,b,\Phi_* Y) = m(\hyp, \Phi^*g,\Phi^*b,Y)\;,
1347: \end{eqnarray*}
1348: and the result follows from $\Phi(\hyp)=\hyp$, $\Phi^*b=b$.\qed
1349:
1350:
1351: \section{Asymptotic isometries - the Riemannian problem}
1352: \label{Sai}
1353:
1354: Throughout this section, \emph{in contradistinction with the
1355: remainder of this paper}, $g$ will denote the Riemannian metric
1356: induced by the space-time metric on $\hyp$. Similarly the letter
1357: $b$ will denote the associated Riemannian background
1358: metric\footnote{One could also consider background metrics of the
1359: form $b= a^2(\bar r)d\bar r^2 +q^2(\bar r) h$; however, if $q$ is
1360: sufficiently differentiable, then $b$ can be brought to the form
1361: \eq{eq:b00} in the asymptotic region by a change of coordinates $r
1362: = q(\bar r)$ provided that $dq/d\bar r$ has no zeros for large
1363: $\bar r$'s.} of the form
1364: \begin{equation}
1365: \label{eq:b00}
1366: b= a^2(r)dr^2 +r^2 h\;, \qquad h=h_{AB}(v^C)dv^A dv^B\;,
1367: \end{equation}
1368: with the indices $A$ running from $2$ to $n$. We assume that
1369: $r\in[R,\infty)$ for some $R$, while the coordinates $v^A$ are
1370: local coordinates on some compact $n-1$ dimensional manifold
1371: $\mn$. Unless explicitly stated otherwise, we use the symbol
1372: $\cO(r^\beta)$ to denote either $O(r^\beta)$ {\em throughout} this
1373: section, {\em or} $o(r^\beta)$ {\em throughout} this section. We
1374: shall mainly be interested in background metrics for which
1375: \begin{deqarr}
1376: & ra(r) = \ell + \newa(r)\;,\quad \newa = O(r^{-m_1})\;, \quad
1377: \R\ni m_1>0\;, &
1378: \label{eq:b0-1}
1379: \\ & \newa'(r)= O(r^{-1-m_1})\;. &
1380: \label{equicond} \arrlabel{eq:b0}\end{deqarr} for some constants
1381: $m_1,\ell>0$.\footnote{ A typical example is
1382: %\begin{eqnarray} \label{eq:b0-}
1383: $a(r) = \frac \ell r\left(1 + \sum_{i=m_1}^{m_2}
1384: {a_i\over r^i} +\cO(r^{-\alpha})\right)%\;.
1385: $ %\end{eqnarray}
1386: for some constants $m_1\ge 1$, $m_2\ge m_1$, $a_i$ and
1387: $\alpha$.\label{aexp} } When the vacuum Einstein equations
1388: \eq{Ee} are satisfied by the metric \eq{eq:b00} we
1389: have$^{\mathrm{\ref{footWoolgar}}}$ $a(r)=1/\sqrt{r^2/\ell^2+k}$,
1390: where $k$ is a constant, which can be written in the form
1391: \eq{eq:b0} with $m_1=2$, as well as in the form of
1392: footnote~\ref{aexp} (with $m_2$ of that footnote as large as
1393: desired). However, the hypothesis that the vacuum Einstein
1394: equations hold plays no role in this section, therefore in
1395: \eq{eq:b0} any $m_1>0$ will be allowed.
1396: % Clearly terms $a_ir^{-i}$ can be absorbed
1397: %into $\cO(r^{-\alpha})$ when $i> \alpha$, so that in
1398: %\eq{eq:b0-} one could without loss of generality assume that $a_i=0 $
1399: %for all $i>\alpha$; equivalently $m_2\le\lfloor \alpha
1400: %\rfloor$, the integer part of $\alpha$. In particular one could choose
1401: %all the $a_i$'s to be zero when $\alpha \le m_1$.
1402: %In order to check that
1403: %$a$ given by
1404: %\eq{eq:b0-} satisfies \eq{eq:b0-2,} one notes that
1405: %$$ {1 \over r^i} - {1 \over \hr^i} = {(r+\delta r)^i - r^i \over r^i
1406: %(r+\delta r)^i} = O(r^{-i-1}) \delta r\;, $$ hence
1407: %\begin{eqnarray*}
1408: %& ra(r) - \hr a(\hr) = O(r^{-m_1-1}) \delta r
1409: %\;.
1410: %\end{eqnarray*}
1411: Let us mention that \eq{equicond} is equivalent to the condition
1412: \begin{equation}\newa(\hr) - \newa(r) = O(r^{-1-m_1})(\hr -r )
1413: \label{eq:b0-2m}
1414: \end{equation}
1415: for $r$ large enough, and for
1416: $|\hr-r|\le r/2$. Indeed, under \eq{equicond} we have $$ \newa(\hr) -
1417: \newa(r) = \left(\int_0^1 \newa'(t\hr + (1-t)r)dt\right)(\hr -r)\;,$$
1418: and \Eq{eq:b0-2m} follows. The implication the other way round is
1419: straightforward using the fact that $\newa$ is smooth (recall that
1420: local smoothness of the
1421: metric is assumed throughout). Condition~\eq{eq:b0-2m} is actually
1422: the one which is needed in the arguments below.
1423:
1424: Let $\theta^i$, $i=1,\ldots,n$ be an orthonormal coframe for $b$,
1425: with $\theta^1= a(r)dr$, and let $e_i$ be the dual frame; we
1426: denote by
1427: $$\zo^{i}{}_{jk}\equiv \theta^i(\zn_{e_k}e_j)$$ the associated
1428: connection coefficients, where $\zn$ is the Levi-Civita connection
1429: of $b$. One easily finds
1430: \begin{equation}
1431: \label{eq:b3}
1432: \zo^A{}_{1B}= \frac 1 {ra(r)} \delta ^A_B= -\zo^1{}_{AB}\;.
1433: \end{equation}
1434: If we denote by \begin{equation}
1435: \label{eq:b2}\alpha^A\equiv \alpha(v^C)^A{}_{B} dv^B\end{equation}
1436: an orthonormal frame for the metric $h$, and by $\beta^A{}_{BC}$
1437: the associated Levi-Civita connection coefficients, then
1438: \begin{equation}
1439: \label{eq:b4}
1440: \zo^A{}_{BC}= \frac 1r \beta^A{}_{BC}\;.
1441: \end{equation}
1442: All connection coefficients other than those in \eq{eq:b3} or
1443: \eq{eq:b4}vanish.
1444:
1445:
1446: \begin{Lemma}
1447: \label{L2} Let $\theta^i$ be an orthonormal
1448: coframe for the metric $b$ as in \eq{eq:b00}, let
1449: $g=g_{ij}\theta^i\otimes \theta^j$, and suppose that
1450: $$g_{ij}\to_{r\to\infty} \delta^i_j\;. $$ Denote by $\sR$,
1451: respectively $\sRz$, the $g$-geodesic distance along $\hyp$,
1452: respectively the $b$-geodesic distance along $\hyp$, from the set
1453: $\{r=R\}$. There exists a function $C(R)\ge 1$ satisfying
1454: $\lim_{R\to\infty}C(R)=1$ such that \begin{equation} \label{eq:l2}
1455: C(R)^{-1} \sRz \le \sR \le C(R) \sRz\;. \end{equation}
1456: \end{Lemma}
1457:
1458:
1459: \medskip
1460:
1461: \proof For $s\in[R,r]$ let $\gamma(s)=(s,v^A)$, then
1462: \begin{eqnarray}
1463: \sR(r,v^A) & \le & \int_R^r \sqrt{g(\dot \gamma,\dot \gamma)(s,v^A)}\ds
1464: \nn\\ & = & \int_R^r \sqrt{g_{ij}\theta^i(\dot
1465: \gamma)\theta^j(\dot \gamma)(s,v^A)}\ds \nn\\ & = & \int_R^r
1466: \sqrt{g_{11}(s,v^A)} a(s) \ds\label{fsr1}
1467: \\\nn & \le & (1+o(1))\sRz(r,v^A)\;.
1468: \end{eqnarray}
1469: To obtain the reverse inequality, we note that for points
1470: $(r,v^A)$ such that $r\ge R$ it holds that
1471: $$
1472: \forall X\quad g(X,X) \ge (1+o(1)) b(X,X)\;,$$ with $o(1)$ going
1473: to zero as $R\to \infty$, hence for every curve $\gamma$ from
1474: $\{r=R\}$ to $(r,v^A)$ we have
1475: $$\int_\gamma \sqrt{g(\dot \gamma,\dot \gamma)(s)}\ds\ge
1476: (1+o(1))\int_\gamma \sqrt{b(\dot \gamma,\dot \gamma)(s)}\ds\;,
1477: $$
1478: therefore
1479: \begin{eqnarray}
1480: \sR&=&\inf_\gamma \int_\gamma \sqrt{g(\dot \gamma,\dot
1481: \gamma)(s)}\ds \nn\\ &\ge& (1+o(1)) \inf_\gamma \int_\gamma
1482: \sqrt{b(\dot \gamma,\dot
1483: \gamma)(s)}\ds
1484: \nn\\ &= & (1+o(1))\sRz\;. \label{fsr2}
1485: \end{eqnarray} \qed
1486:
1487: The proof of Lemma~\ref{L2}
1488: uses only the product structure of $b$, and does not require
1489: Equation~\eq{eq:b0} to hold. If, however, that last equation
1490: holds, then clearly
1491: $$\sRz(r,v^A)= \int_R^r a(s)ds =\ell \ln (r/R) +O(R^{-m_1})\approx \ell \ln (r/R) $$ for large
1492: $r$, and \eq{eq:l2} implies that for all $\epsilon>0$ there exists
1493: $R_\epsilon\ge R$ and a constant $\hat C(\epsilon)$ such that for
1494: all $r\ge R_\epsilon $ we have
1495: \begin{equation} \label{eq:l2aold} \hat C(\epsilon)^{-1} r^{1-\epsilon}
1496: \le \exp{(\sR/\ell)} \le \hat C(\epsilon) r^{1+\epsilon}\;.
1497: \end{equation}
1498: We will need a sharper version of this:
1499:
1500:
1501: \begin{Lemma}
1502: \label{L2.1} Under the hypotheses of Lemma~\ref{L2}, suppose further that
1503: \Eq{eq:b0} holds, and that there exists a constant $\alpha>0$ such that
1504: $$g_{ij}-\delta^i_j=\cO(r^{-\alpha})\;. $$ Then
1505: %there exists a constant $C'>0$ such that
1506: for $r\ge \max[R,1]$ we have
1507: \begin{equation} \label{eq:l2a0} \exp{(\sR/\ell)} = r/R + O(R^{-m_1}) +
1508: \cO(R^{-\alpha})\;,
1509: \end{equation}
1510: in particular
1511: \begin{equation} \label{eq:l2a1} r/C'
1512: \le \exp{(\sR/\ell)} \le C' r\;.
1513: \end{equation}
1514: \end{Lemma}
1515:
1516: \proof
1517: %Let us suppose that $\cO=O$, the proof with $\cO=0$
1518: %requires obvious modifications which are left to the reader.
1519: Here and elsewhere in this paper the letter $C$ denotes a constant
1520: which may vary from line to line; if $\cO=o$ the constants in the
1521: current proof can be chosen as small as desired by choosing $R$
1522: large enough. In \Eq{fsr1} we can estimate $\sqrt{g_{11}}$ by $1+C
1523: s^{-\alpha}$, obtaining thus
1524: \begin{eqnarray*}
1525: \sR(r,v^A) &\le &\sRz(r,v^A)+ CR^{-\alpha}
1526: \\ &=& \ell \ln (r/R) + O(R^{-m_1}) + CR^{-\alpha}\;.
1527: \end{eqnarray*}
1528: Similarly, \Eq{fsr2} is rewritten as
1529: \begin{eqnarray}
1530: \sR&=&\inf_\gamma \int_\gamma \sqrt{g(\dot \gamma,\dot
1531: \gamma)(s)}\ds \nn\\ &\ge&\inf_\gamma
1532: \int_\gamma(1-Cs^{-\alpha})\sqrt{b(\dot \gamma,\dot \gamma)(s)}\ds
1533: \;. \label{fsr3}
1534: \end{eqnarray}
1535: The last term in \Eq{fsr3} is the distance from the set $\{r=R\}$
1536: in the metric
1537: $$ (1-Cr^{-\alpha})^2\left(a^2(r)dr^2 +r^2 h\right)\;,$$
1538: which equals
1539: $$
1540: \int_R^r (1-Cs^{-\alpha})a(s)\ds= \ell \ln(r/R) + O(R^{-m_1}) -
1541: O(R^{-\alpha})\;,
1542: $$
1543: and our claims immediately follow. \qed
1544:
1545: The key result in our work is the following:
1546: \begin{Theorem}[Asymptotic symmetries] \label{T1} Let $(r,v^A)$ coordinatize
1547: $\Omega\subset \hyp$ so that $\Omega
1548: \approx\{r\in[R,\infty)\}\times\mn$, and let $(\hr,\hv^A)$ be
1549: another set of coordinates on $\Omega$ so that
1550: $\Omega\approx\{(\hv^A)\in \hM ,\hr\in [\hat R(\hv^A),\infty)\}$ for
1551: some continuous function $\hat R$. We further assume that $v^A$ and
1552: $\hv^A$ are consistently oriented, and that $(r,v^A)$ and
1553: $(\hr,\hv^A)$ are also consistently oriented. Let $b$, $\theta^i$,
1554: $e^i$, \emph{etc.}, be as at the beginning of this section, with $a$
1555: of the form
1556: \eq{eq:b0}, and let $\hb$, $\hth^i$,
1557: $\he_i$, \emph{etc.} be the hatted equivalents thereof, so
1558: that $$\hb =
1559: a^2(\hr)d\hr^2 +\hr^2 \hh= \sum_i \hth^i\otimes\hth^i\;,
1560: %\qquad
1561: %\hh=h_{AB}(\hat v^C)d\hv^A d\hv^B\;,
1562: $$
1563: for some Riemannian metric $\hh$ on $\hM $. Suppose that there
1564: exists $\alpha
1565: >0$ such that
1566: \begin{eqnarray}
1567: \label{eq:hd}
1568: & g(e_i,e_j)-\delta_i^j = \cO(r^{-\alpha})\;,\quad e_k(g(e_i,e_j))
1569: = \cO(r^{-\alpha})\;, & \\ &
1570: g(\he_i,\he_j)-\delta_i^j = \cO(\hr^{-\alpha})\;,\quad
1571: \he_k(g(\he_i,\he_j)) = \cO(\hr^{-\alpha})\;.
1572: \end{eqnarray}
1573: Then:
1574: \begin{enumerate}
1575: \item There exists a $C^{\infty}$ map
1576: $\Psi:\mn\to\hM $ satisfying
1577: $$\Psi^*\hh = e^{-2\psi} h$$
1578: for some $C^\infty$ function $\psi:\mn\to\R$, and\ptc{ would make
1579: sense to add the second derivative equations here as well, and
1580: give a coordinate counterpart of those second derivative equations
1581: }
1582: \begin{eqnarray}
1583: \label{eq:t1}& \hr = e^{\psi} r + O(r^{1-\beta})\;, \quad
1584: e_i (\hr) = e_i(e^{\psi}r) +
1585: O(r^{1-\beta})\;, & \\ & \hv^A = \psi^A(v^B)+ O(r^{-2})\;, \quad
1586: e_i(\hat{v}^A) = e_i(\psi^A(v^B)) + O(r^{-2}) \;, \label{eq:t2}
1587: &\end{eqnarray} in local coordinates with $\Psi =
1588: (\psi^A)$, with $\beta=\min{(m_1,\alpha,2)}$.
1589: \item If
1590: $\Psi$ is the identity and if $\psi=0$, then for $\alpha >1$ we
1591: further have
1592: \begin{eqnarray}
1593: \label{t1}& \hr = r + \cO(r^{1-\alpha})\;, \quad
1594: e_i (\hr) = e_i(r) +
1595: \cO(r^{1-\alpha})\;, & \\ & \hv^A = v^A+ \cO(r^{-\alpha-1})\;,
1596: \quad e_i(\hat{v}^A) = e_i(v^A) + \cO(r^{-\alpha-1}) \;,
1597: \label{t2}&
1598: \\ &e_i\left(e_j(\hat{r}-r)\right) =
1599: \cO(r^{1-\alpha})\;, \quad e_i\left(e_j(\hat{v}^A-v^A)\right) =
1600: \cO(r^{-\alpha-1}) \;. \label{t2a}&\end{eqnarray}
1601: \end{enumerate}
1602: \end{Theorem}
1603:
1604: \remarks
1605: 1. For reference we note the partial derivatives estimates
1606: \begin{eqnarray}
1607: \label{eq:t1+1}& \displaystyle\frac{\partial \hr}{\partial r} = e^{\psi} + O(r^{-\beta})\;,
1608: \quad
1609: \displaystyle\frac{\partial \hr}{\partial v^A} =
1610: \frac{\partial e^{\psi}}{\partial v^A}r+ O(r^{1-\beta})\;,
1611: & \\ & \displaystyle\frac{\partial\hv^A}{\partial r} = O(r^{-3})\;, \quad
1612: \frac{\partial\hat{v}^A}{\partial v^B} =
1613: \frac{\partial\psi^A}{\partial v^B}+ O(r^{-\beta}) \;,
1614: \label{eq:t2+1}
1615: &\end{eqnarray}
1616: with the second estimate in \eq{eq:t2+1} being somewhat stronger
1617: than its counterpart in \eq{eq:t2}.
1618: %\medskip
1619:
1620: \noindent 2. It should be noted
1621: that in point 1. above we do not assume that $\hM =M$
1622: and $\hh=h_{AB}(\hat v^C)d\hat v^Ad\hat v^B$; this fact plays a
1623: role in~\cite{ChHerzlich}. The arguments in that last reference
1624: show that $\Psi$ is a diffeomorphism, in particular $\hM$ is
1625: necessarily diffeomorphic to $M$. If $\hh=h_{AB}(\hat v^C)d\hat
1626: v^Ad\hat v^B$ and $\Psi$ is the identity, then clearly $\psi=0$
1627: follows.
1628:
1629: \medskip
1630:
1631: \noindent 3. We stress that we do not assume $M$ or $\hM $ to be
1632: parallelizable, thus \Eqs{eq:t1}{t2a} have to be understood in the
1633: sense of finite coverings of $M$ and $\hM $, with corresponding
1634: frames, on which the claimed estimates hold. However, if $M$ or
1635: $\hM$ are parallelizable, then the estimates are global.
1636: \medskip
1637:
1638: \proof Let $\ncO_p$ be a conditionally compact subset of an open
1639: domain of a local coordinate system $(v^A)$ around $p=(v_0^A)\in
1640: M$, and let, on a neighbourhood of $\overline{\ncO_p}$, $\beta_A$
1641: be a $h$-orthonormal frame dual to $\alpha^A$. We note that the
1642: connection coefficients $\beta^A{}_{BC}$ are uniformly bounded on
1643: $\ncO_p$. Consider the radial ray $$\gamma_p(r):=(r,v_0^A)\;,$$
1644: which, in hatted coordinates, can be written as
1645: \be\label{gamhat}[R,\infty)\ni r\to
1646: \gamma_p(r)=\left(\hr(r,v_0^A),\pgp(r):=\left(\hv^A(r,v_0^A)\right)\right)\in
1647: [\hat R,\infty)\times \hM \subset \hyp\;;\ee here and in what
1648: follows we identify $[R,\infty)\times \ncO_p$ with the
1649: corresponding subset of $\hyp$, similarly for sets of the form
1650: $[\hat R,\infty)\times \widehat\mcU$, where $\widehat\mcU\subset
1651: \hM $. It should be clear from \eq{gamhat} that the operation
1652: ``$\gamma_p\to\pgp$" above is a coordinate projection which
1653: consists in forgetting the $\hr$ coordinate in a coordinate system
1654: $(\hr,\hv^A)$. We have
1655: \begin{eqnarray*}
1656: \hh_{AB}\frac{\partial \hv^A}{\partial r}\frac{\partial
1657: \hv^B}{\partial r} & \le & \frac 1 {\hr^2}\left(\hr^2
1658: \hh_{AB}\frac{\partial \hv^A}{\partial r}\frac{\partial
1659: \hv^B}{\partial r} + a^2(\hr)\left(\frac{ \partial \hr}{\partial
1660: r}\right)^2\right)
1661: \\ &= &\frac 1 {\hr^2} \hat b(\frac{ \partial }{\partial
1662: r},\frac{ \partial }{\partial r}) \ \le \ C\frac 1 {\hr^2}
1663: g(\frac{
1664: \partial }{\partial r},\frac{ \partial }{\partial r})
1665: \\ &\le &C^2\frac 1 {\hr^2} b(\frac{ \partial }{\partial
1666: r},\frac{ \partial }{\partial r})\ =\ C^2\frac {a^2(r)} {\hr^2}\;.
1667: \end{eqnarray*}
1668: %Equation \eq{eq:l2a} together with its
1669: %hatted equivalent show that for any $\epsilon >0$ there exists
1670: %$R_\epsilon\ge R$ and a constant $\tilde C(\epsilon)$ such that
1671: %for all $r\ge R_\epsilon $ wWe have
1672: %\begin{equation} \label{eq:l2a1} \tilde C(\epsilon)^{-1} r^{1-\epsilon}
1673: %\le \hr \le \tilde C(\epsilon) r^{1+\epsilon}\;.
1674: %\end{equation}
1675: Let $d_{\hh}$ denote the $\hh$ distance on $\hM $, for $r_2\ge
1676: r_1$ we thus obtain
1677: \begin{eqnarray}\label{hatdist}
1678: d_{\hh}\left(\pgp(r_1),\pgp(r_2)\right) & \le & \int _{r_1}^{r_2}
1679: \sqrt{\hh_{AB}\frac{\partial \hv^A}{\partial r}\frac{\partial
1680: \hv^B}{\partial r}(s)}\; ds
1681: \\ \nn&\le &C\int _{r_1}^{r_2}\frac {a(s)} {\hr(s,v_o^A)}\; ds
1682: \\ \nn &\le& C^2\int _{r_1}^{r_2}\frac {1} {s^{2}}\; ds = C^2\left(
1683: \frac {1} {r_1}-\frac {1} {r_2}\right)\;,
1684: \end{eqnarray} and Lemma~\ref{L2.1} has been used.
1685: It then easily follows that
1686: $$\hat p:= \lim_{r\to\infty}\pgp(r)$$ exists, with
1687: \be\label{*} d_{\hh}\left(\pgp(r),\hat p\right)\le \frac {C^2}
1688: {r}\;.\ee Let $\hOp$ be a conditionally compact subset of a domain
1689: of local coordinates $\hv^A$ around $\hat p$, \Eq{*} shows that
1690: $\gamma_p$ enters and remains in $[\hat R,\infty)\times \hOp$ for
1691: $r$ large enough. In what follows only such $r$'s will be
1692: considered.
1693:
1694: Consider, now, a point $q\in\ncO_p$; we wish to show that the
1695: corresponding ray $\gamma_q$ will stay within $[\hat
1696: R,\infty)\times \cOp$ if $q$ is close enough to $p$. In order to
1697: do that, consider an $h$-geodesic segment $\gamma \subset M$
1698: parameterized by proper length such that $\gamma(0)=p$ and $
1699: \gamma(d_h(p,q))=q$.
1700: %As long as
1701: %$(r,\gamma(s))$ remains within $[\hat R,\infty)\times \cOp$ we
1702: Expressing the path $$s\to\Gamma(s):=(r,\gamma(s))\in
1703: [R,\infty)\times\ncO_p\subset \hyp$$ in terms of the barred
1704: coordinates we have
1705: \begin{eqnarray*}
1706: \hh_{AB}\frac{d\hv^A}{ds}\frac{d \hv^B}{ds} & \le & \frac 1
1707: {\hr^2}\left(\hr^2 \hh_{AB}\frac{d\hv^A}{ds}\frac{d \hv^B}{ds} +
1708: a^2(\hr)\left(\frac{ d \hr}{ds}\right)^2\right)
1709: \\ &= &\frac 1 {\hr^2} \hat b(\frac{ d\Gamma }{ds},\frac{d\Gamma}{ds})
1710: \ \le \ C\; \frac 1 {\hr^2} b(\frac{ d\Gamma
1711: }{ds},\frac{d\Gamma}{ds})
1712: \\ & = &C\; \frac {r^2} {\hr^2}\ \le\ C^2\;,
1713: \end{eqnarray*}
1714: An estimation as in \Eq{hatdist} gives
1715: \begin{eqnarray*}%\label{hatdist2}
1716: d_{\hh}\left(\pgp(r),\pgq(r)\right) & \le & C\;
1717: d_h(p,q)\;.\end{eqnarray*} Passing to a subset of $\ncO_p$ if
1718: necessary we thus obtain that for all $q\in \ncO_p$ the rays
1719: $\gamma_q$ enter and remain in $[\hat R,\infty)\times \cOp$, for
1720: $r\ge R_p$ for some $R_p$.
1721:
1722: Let, on an open neighbourhood of $\overline{\cOp}$, $\hat\alpha^A$
1723: be a $\hh$-ON frame with uniformly bounded connection coefficients
1724: $\hat\beta^A{}_{BC}$, and let $\hat\beta_A$ be a $\hh$-orthonormal
1725: frame dual to $\hat\alpha^A$. Equations \eq{eq:b3} and \eq{eq:b4}
1726: show then that all the $\zo^i{}_{kj}$'s and $\hzo^i{}_{kj}$'s ---
1727: the connection coefficients of $b$ and of $\hb$
1728: --- are uniformly bounded along the rays $\gamma_q$, $q\in\ncO_p$; the reader will note that the same
1729: will be true for the constants controlling various error terms
1730: $\cO(r^{\cdot})$ in the calculations below. The idea of the
1731: argument below is then to derive the desired estimates along the
1732: $\gamma_q$'s, $q\in\ncO_p$; covering the compact manifold $\mn$ by
1733: a finite number of coordinate patches $\ncO_{p_i}$,
1734: $i=1,\ldots,I$, will establish our claims.
1735:
1736: Let $f_i$, respectively $\hf_i$, be a $g$-orthonormal frame
1737: obtained by a Gram-Schmidt orthonormalisation procedure using
1738: $\{e_i\}_{i=1}^n$, respectively $\{\he_i\}_{i=1}^n$. The explicit
1739: form of the $f_i$'s and $\hf_i$'s in terms of the $e_i$'s and
1740: $\he_i$'s shows that
1741: \begin{eqnarray}
1742: & f_i = e_i +\delta f_i\;,\quad \delta f_i = \delta f_i{^j}
1743: e_j\;, \quad \delta f_i{^j} = \cO(r^{-\alpha})\;,\quad e_k(\delta
1744: f_i{^j}) = \cO(r^{-\alpha}) \;, & \nn\\ && \label{eq:B1a}\\
1745: & \hf_i = \he_i +\delta \hf_i\;,\quad \delta \hf_i = \delta
1746: \hf_i{^j} \he_j\;, \quad \delta \hf_i{^j} =
1747: \cO(\hr^{-\alpha})\;,\quad \he_k(\delta \hf_i{^j}) =
1748: \cO(\hr^{-\alpha}) \;. \nn \\ && \label{eq:B1b}
1749: \end{eqnarray}
1750: By construction we actually have
1751: \bel{Gseq} f_1= \left(1+
1752: \cO(r^{-\alpha})\right) e_1 = \left(1+ \cO(r^{-\alpha})
1753: +O(r^{-m_1} )\right)\frac r \ell \partial_r \;. \ee The uniform
1754: boundedness of the $\zo^i{}_{kj}$'s further shows that
1755: \be\label{eq:B1c} \zn_{e_i}\delta f_j = \cO(r^{-\alpha})\;,\ee
1756: similarly for the $\hb$-covariant derivatives of the $\delta
1757: \hf_j$'s with respect to the $\he_i$'s. Recall that
1758: \be\label{b0}\zo^i{}_{kj}=\frac 12 \left\{\theta^j([e_i,e_k]) -
1759: \theta^i([e_k,e_j])-\theta^k([e_j,e_i]) \right\}\;;\ee
1760: Equation~\eq{b0} together with its $g$-equivalent and
1761: \eq{eq:B1a}-\eq{eq:B1c} imply \be\label{eq:B1d}\omega^i{}_{jk} =
1762: \zo^i{}_{jk} + \cO(r^{-\alpha})\;,\ee similarly
1763: \be\label{eq:B1e}\home^i{}_{jk} \equiv
1764: \hat\phi^i(\nabla_{\hf_k}\hf_j)= \hzo^i{}_{jk} +
1765: \cO(\hr^{-\alpha})\;.\ee
1766: %Here $\hzo^i{}_{jk}$ denotes
1767: %$\zo^i{}_{jk}$ with $(r,v^A)$ replaced by $(\hr,\hv^A)$, wherever
1768: %occurring; w
1769: We use the symbols $\phi^i$ and $\hat\phi^i$ to denote coframes
1770: dual to $f_i$ and $\hf_i$. Now, both the $f_i$'s and $\hf_i$'s are
1771: orthonormal frames for $g$, hence there exists a field of rotation
1772: matrices $\Lambda=(\Lambda_i{}^j) \in O(n)$ such that \be
1773: \label{B4a+}\hf_i = \Lambda_i{}^j f_j\;.\ee
1774: We recall that for rotation matrices we
1775: have\footnote{We use the convention summation throughout, so that
1776: repeated indices in different positions have to be summed over. We
1777: will explicitly indicate the summation only in those equations in
1778: which we need to sum over repeated indices which are both
1779: subscripts or both superscripts.}
1780: \begin{equation}
1781: \label{eq:B4a}\sum_k \Lambda_i{}^k\Lambda_j{}^k = \delta^i_{j}\;,\ee
1782: in particular
1783: $$(\Lambda^{-1})_i{}^j = \Lambda _j {}^i\;,$$
1784: so that $\hphi^i= \Lambda _j {}^i\phi^j$, and $\phi^j = \sum_i
1785: \Lambda_j{}^i\hphi^i$. Further, from \Eq{eq:B4a} we obtain
1786: \begin{equation}
1787: \label{eq:B4} \sum_k \Lambda_i{}^k\Lambda_i{}^k=1\quad
1788: \Longrightarrow \quad \forall\; i,j\ \; |\Lambda_i{}^j| \le 1\;.
1789: \end{equation}
1790: From the definition of the connection coefficients we have
1791: \begin{eqnarray*}
1792: \hat{\omega}^\ell{}_{ji} & = & \langle\hat \phi^\ell,\nabla_{\hf_i}
1793: \hf_j\rangle
1794: \\ & = & \langle\Lambda_k{}^\ell \phi^k,\nabla_{\Lambda_i{}^m f_m}
1795: \left(\Lambda_j{}^n f_n\right)\rangle
1796: \\ & = & \Lambda_k{}^\ell\Lambda_i{}^m \langle \phi^k,{f_m}(
1797: \Lambda_j{}^n) f_n+ \Lambda_j{}^n\nabla_{f_m}
1798: f_n\rangle \;,
1799: \end{eqnarray*}
1800: leading to the well known transformation law
1801: $$\hat{\omega}^l{}_{ji} \Lambda_l{}^k = \Lambda_i{}^l
1802: f_l(\Lambda_j{}^k) + \Lambda_j{}^l \Lambda_i{}^n
1803: \omega^k{}_{ln}\;,
1804: $$
1805: which we intepret as an equation for the $\Lambda_j{}^k$'s:
1806: \begin{equation}
1807: \label{eq:B5}
1808: f_i(\Lambda_j{}^k) = (\Lambda^{-1})_i{}^l\hat{\omega}^n{}_{jl}
1809: \Lambda_n{}^k - \Lambda_j{}^l \omega^k{}_{li}\;.
1810: \end{equation}
1811: From \Eqsone{eq:b0} and \eq{eq:B1a} we obtain
1812: \begin{deqarr}
1813: \phi^m(\partial_r) &=& {a( r)} \left(\delta^m_1+
1814: \cO(r^{-\alpha})\right)\label{eq:B6-}
1815: \\ &=& \frac \ell r \left(\delta^m_1+
1816: \cO(r^{-\beta})\right) \;, \label{eq:B6}\arrlabel{B6}
1817: \end{deqarr}
1818: with $\beta=\min(\alpha,m_1)$, except if $\cO=o$ and $\alpha >m_1$
1819: in which case either $\beta$ should be taken to be any number
1820: smaller than $m_1$, or $\cO$ should be understood as $O$.
1821: Rescaling $r$ and the metric $g$ by a constant conformal factor
1822: we may without loss of
1823: generality assume that $\ell=1$; similarly for $\hr$. \Eq{eq:l2a1}
1824: together with Equations~\eq{eq:B1d}-\eq{B6} leads to
1825: \begin{deqarr}
1826: \frac{\partial \Lambda_i{}^{j}}{\partial r} &=&
1827: %\frac {a( r)}{\hr a(\hr)}
1828: {a( r)}
1829: \left(
1830: \sum_k\hzo{}^\ell{} _{ik} \Lambda _\ell{}^j \Lambda _k {} ^1 +
1831: \cO(r^{-\alpha})\right) \label{eq:B7a-} \\ &=& \frac
1832: 1 r \left(
1833: \sum_k\hzo^\ell{} _{ik} \Lambda _\ell{}^j \Lambda _k {} ^1 +
1834: \cO(r^{-\beta})\right)\;,
1835: \label{eq:B7a}\arrlabel{B7a}
1836: \end{deqarr}
1837: in particular
1838: \begin{deqarr}
1839: \frac{\partial \Lambda_1{}^{j}}{\partial r} &=& \frac
1840: {a( r)}{\hr a(\hr)} \left( \sum_A\Lambda _A{}^j \Lambda _A{} ^1+
1841: \cO(r^{-\alpha})\right) \label{eq:B8-}
1842: \\&=& \frac
1843: 1r \left( \sum_A\Lambda _A{}^j \Lambda _A{} ^1+
1844: \cO(r^{-\beta})\right)
1845: \;.\label{eq:B8}\arrlabel{B8}
1846: \end{deqarr}
1847: Now, the transpose of a rotation matrix is again a rotation
1848: matrix, therefore we also have
1849: $$\sum_k\Lambda_k{}^i \Lambda_k {}^j = \delta^i_{j}\;,$$ which gives
1850: \begin{equation}
1851: \label{eq:B8a}
1852: \sum_A\Lambda _A{}^1 \Lambda_A{} ^1 = 1 - (\Lambda_1{}^1)^2\;,
1853: \end{equation}
1854: and it follows that
1855: \begin{deqarr}
1856: \frac{\partial \Lambda_1{}^{1}}{\partial r} &= & \frac
1857: {a( r)}{\hr a(\hr)} \left(1-
1858: (\Lambda _1{}^1)^2 +
1859: \cO(r^{-\alpha}) \right)
1860: \label{eq:B9-}
1861: \\
1862: &= & \frac
1863: 1r \left(1-
1864: (\Lambda _1{}^1)^2 \right)+
1865: \cO(r^{-\beta-1})
1866: \label{eq:B9}
1867: \;.
1868: \arrlabel{B9}
1869: \end{deqarr}
1870: We have the following:
1871: \newcommand{\mysigma}{\sigma}
1872: \begin{Lemma}
1873: \label{L3} For all $\mysigma<\min(m_1,\alpha,2)$ we
1874: have
1875: \begin{eqnarray}
1876: \label{eq:B10}
1877: \Lambda _1{}^1 & = & 1+ \cO(r^{-\mysigma}) \;,
1878: \\ r \frac{\partial \hr}{\partial r}& = & \hr +
1879: \cO(r^{1-\mysigma}) \label{eq:B10a}\;.
1880: \end{eqnarray}
1881: \end{Lemma}
1882: \proof Let $\chi$ denote the $ \cO(r^{-\beta-1}) $ term in
1883: Equation~\eq{eq:B9}, set $g:=\Lambda _1{}^1$, and denote by
1884: $\phi(r,v^A) = \int_r^\infty \chi (s,v^A)\ds= \cO(r^{-\beta})$;
1885: Equation~\eq{eq:B9} shows that
1886: $$\frac{\partial (g-\phi)}{\partial r} = \frac{1-g^2}{r}\ge 0\;.$$
1887: It follows that $g-\phi$ is non-decreasing, and therefore has a
1888: limit as $r$ goes to infinity; since $\phi$ tends to zero in this
1889: limit we conclude that
1890: $$g_\infty\equiv \lim_{r\to\infty}g$$
1891: exists. Equation~\eq{eq:B9} shows that
1892: \begin{equation}
1893: \label{eq:last}
1894: \lim_{r\to\infty} r {\partial g \over \partial r} =
1895: 1-g_\infty^2\;.
1896: \end{equation}
1897: Now Equation~\eq{eq:B4} gives $|g|\le 1$, while
1898: Equation~\eq{eq:last} implies a logarithmic divergence of $g$
1899: unless $g_\infty=\pm 1$; thus $g_\infty=\pm 1$. Define $f\ge 0$ by
1900: the equation
1901: $$g= g_\infty(1-f)\;,$$
1902: then $f\to_{r\to\infty} 0$ and we have
1903: $${\partial f \over \partial r} = -{g_\infty f(2-f) \over r} -
1904: g_\infty\chi\;.$$ Suppose, first, that $g_\infty =1$; since
1905: $f\to_{r\to\infty}=0$ it follows that for every $\delta>0$ there
1906: exists $r_\delta$ such that $f\le \delta$ for $r\ge r_\delta$; for
1907: such $r$ we then obtain
1908: \begin{equation}
1909: \label{eq:B11}
1910: {\partial f \over \partial r} \le -{ f(2-\delta) \over r} -
1911: \chi\;,
1912: \end{equation}
1913: and by integration
1914: \begin{eqnarray*}
1915: r^{2-\delta} f(r,\cdot) - r_\delta^{2-\delta} f(r_\delta,\cdot) \le
1916: -\int_{r_\delta}^r \chi(s,\cdot)s^{2-\delta}\ds =
1917: \cO(r^{2-\delta-\beta}) + \cO(r_\delta^{2-\delta-\beta})\;,
1918: \end{eqnarray*}
1919: so that
1920: \begin{equation}
1921: \label{eq:B12a}
1922: f(r) = \cO(r^{\delta-2}) + \cO(r^{ -\beta})\;.
1923: \end{equation}
1924: Choosing $\delta$ appropriately we obtain \eq{eq:B10} with any
1925: $\mysigma<\min(m_1,\alpha,2)$, under the assumption that
1926: $g_\infty\ne -1$. In the case $g_\infty= -1$ similar, but simpler,
1927: manipulations lead again to Equation~\eq{eq:B12a} with $\delta=0$.
1928: From Equation~\eq{eq:B8a} and from $ \Lambda_1{}^1 = g_\infty +
1929: \cO(r^{-\mysigma})$ we obtain
1930: \begin{equation}
1931: \label{eq:B14a}
1932: \Lambda_A{}^1 = \cO(r^{-\mysigma/2})\;.
1933: \end{equation}
1934: Equation~\eq{eq:B4a} similarly implies
1935: \begin{equation}
1936: \label{eq:B14b}
1937: \Lambda_1{}^A = \cO(r^{-\mysigma/2})\;.
1938: \end{equation}
1939: Equation~\eq{eq:B7a-} gives
1940: %\ptc{equation expanded for further use}
1941: \begin{eqnarray*} %{deqarr}\label{lauf}
1942: \frac{\partial \Lambda_A{}^{1}}{\partial r}
1943: %& =& \frac 1 r \frac{ra(r)}{\hr a(\hr)}\left(
1944: % - \Lambda_1 {}^1\Lambda_A {}^1+
1945: % \cO(r^{-\alpha})\right)
1946: %\\ \label{laus}
1947: & =& \frac 1 r \left(
1948: - \Lambda_1 {}^1\Lambda_A {}^1+
1949: \cO(r^{-\mysigma})\right) \;,
1950: \end{eqnarray*}%{deqarr}
1951: and integration in $r$ together with \eq{eq:B14a} yields (assuming
1952: without loss of generality that $\sigma\ne 1$)
1953: \begin{equation}
1954: \label{eq:B15-}
1955: \Lambda_A{}^1 = O(r^{-1}) + \cO(r^{-\mysigma})\;.
1956: \end{equation}
1957: %so that for $\mysigma<\min(m_1,\alpha,1)$ it holds that
1958: %\begin{equation}
1959: % \label{eq:B15}
1960: % \Lambda_A{}^1 = \cO(r^{-\mysigma})\;.
1961: %\end{equation}
1962: \ptc{unnecessary equation and redefinition of sigma removed}
1963: Integrating in $r$ Equation~\eq{eq:B8} and using
1964: \eq{eq:B14b}-\eq{eq:B15-} we similarly obtain\ptc{term added since
1965: sigma's redefinition removed}
1966: \begin{equation}
1967: \label{eq:B14}
1968: \Lambda_1{}^A = O(r^{-1})+\cO(r^{-\mysigma})\;.
1969: \end{equation}
1970: From the definition of the $\hf_i$'s and from \eq{eq:b0} (with
1971: $\ell=1$) we have
1972: $$\hf_1(\hr) = \hr + \cO(\hr^{1-\beta})\;, \quad
1973: \hf_A(\hr)=\cO(\hr^{1-\alpha})\;,$$ hence\ptc{$f_A(\hr)$ equation
1974: commented out since not needed; still in the file}
1975: \begin{eqnarray*}
1976: f_1(\hr) & =& \Lambda_1{}^1 \hf_1(\hr) + \Lambda_1{}^A \hf_A(\hr)
1977: \\ & = & g_\infty \hr + \cO(r^{1-\mysigma})\;,
1978: %\\ f_A(\hr) & =& \Lambda_A{}^1 \hf_1(\hr) + \Lambda_A{}^B \hf_B(\hr)
1979: %\\ & = & \cO(r^{1-\mysigma})\;.
1980: \end{eqnarray*}
1981: Inverting \Eq{Gseq} we have
1982: %the relation $f_i=e_i+\delta
1983: %f_i{}^je_j$ and using Equation~\eq{eq:B1a} shows that $e_i = f_i +
1984: %\sum_j \cO(r^{-\alpha}) f_j$, which implies
1985: \begin{eqnarray*}
1986: \left(1+ O(r^{-m_1})\right) r {\partial
1987: \hr \over \partial r}
1988: & = & e_1(\hr)
1989: \\ & = &\left(1+\cO(r^{-\alpha})\right) f_1(\hr)
1990: % + \sum_j \cO(r^{-\mysigma}) f_j(\hr)
1991: \\ & = & g_\infty \hr + \cO(r^{1-\mysigma})\;.
1992: \end{eqnarray*}
1993: We have finally obtained
1994: \begin{equation}
1995: \label{eq:B16}
1996: r {\partial \hr \over \partial r } = g_\infty\hr
1997: +\cO(r^{1-\mysigma}) \;,
1998: \end{equation}
1999: which is compatible with the fact that the coordinate systems
2000: $(r,v^A)$ and $(\hr,\hv^A)$, as well as $(v^A)$ and $(\hv^A)$,
2001: carry the same orientations if and only if $g_\infty=1$, and the
2002: lemma is established. \qed
2003:
2004: Returning to the proof of Theorem~\ref{T1}, it is useful to
2005: introduce new coordinates $x$ and $\hx$ defined as
2006: $$r=e^x\;, \qquad \hr = e^{\hx}\;.$$
2007: \newcommand{\mysigmax}{\mysigma x}
2008: In terms of those variables Equation~\eq{eq:B10a} can be rewritten
2009: as \be\label{hxx}{\partial \hx \over \partial x } = 1 + \phi\;,
2010: \qquad\phi =\cO(e^{-\mysigmax})\;,\ee for an appropriately defined
2011: function $\phi$: More precisely, if we write \be\label{fnotation}
2012: e_i=e_i{}^jf_j\;,\quad f_i= f_i{}^je_j\;, \ee with the obvious
2013: hatted equivalents, then \be\label{phiequation} \phi:= \frac
2014: {ra(r)}{\hr a(\hr)} \sum_{k,j} e_1{}^j \Lambda^j{}_k\hat
2015: f_k{}^1-1\;.\ee
2016: Integration of \Eq{hxx} gives
2017: \begin{eqnarray}
2018: \hx(x,v^A)& = & x-\hx(x_0,v^A) + \int_{x_0} ^x \phi(s,v^A)\ds
2019: \nonumber \\ & = &
2020: x-\hx(x_0,v^A) + \int_{x_0} ^\infty \phi(s,v^A)\ds- \int_{x} ^\infty \phi(s,v^A)\ds
2021: \nonumber \\ & =: & x+\psi(v^A) + \cO(e^{-\mysigmax})\;,
2022: \label{eq:B16a}
2023: \end{eqnarray}
2024: which establishes the existence of a continuous function
2025: $\psi:\mn\to\R$ such that
2026: \begin{equation}
2027: \label{eq:B17}
2028: \hr(r,v^A) = e^{\psi(v^A)} r + \cO(r^{1-\mysigma})
2029: \end{equation}
2030: (continuity of $\psi$ follows from the Lebesgue (dominated)
2031: theorem on continuity of integrals with parameters; the Lebesgue
2032: theorem is used in a similar way without explicit reference at
2033: several places below).
2034:
2035: Let us write
2036: \Eq{eq:B9-} as
2037: \begin{eqnarray}
2038: \frac{\partial (\Lambda_1{}^{1}-1)}{\partial r} &= &
2039: \chi_1(\Lambda_1{}^{1}-1) +\chi_2 \;,
2040: \label{eq:B9-a}
2041: \end{eqnarray}
2042: where
2043: $$ \chi_1:= -(1+\Lambda_1{}^{1})\frac
2044: {a( r)}{\hr a(\hr)} =-{2\over r}+\cO(r^{-1-\sigma})\;,
2045: \qquad \chi_2 = \cO(r^{-\alpha-1})\;;$$ we obtain
2046: \begin{eqnarray}
2047: { \Lambda_1{}^{1}}{(r,v^A)} &= & 1 + \left(f_1{}^{1}(v^A)
2048: + \int_{r_0}^r
2049: s^2\exp\left\{\int_s^{\infty}\left(\chi_1(v,v^A)+{2\over
2050: v}\right)dv\right\}\chi_2(s,v^A)ds\right)\times
2051: \nonumber \\
2052: & & r^{-2}\exp\left\{-\int_{r}^\infty\left(\chi_1(v,v^A)+{2\over
2053: v}\right)dv\right\} \;,
2054: \label{eq:B9-b}
2055: \end{eqnarray}
2056: with some continuous function $f_1{}^{1}:M\to \R$. In particular
2057: \begin{eqnarray}
2058: \Lambda_1{}^1 &= & 1 + O(r^{-2})+\cO(r^{-\alpha})+ \delta
2059: _\alpha^2\cO(r^{-\alpha}\ln r)%+O(r^{-\sigma_1})\;,
2060: \;. \label{eq:B7a--}
2061: \end{eqnarray}
2062: %where $\sigma_1$ is any number smaller than
2063: %$\min{(m_1,\alpha,2)}+2$.
2064: From \Eq{eq:B7a-} we obtain
2065: \begin{eqnarray}
2066: \frac{\partial \Lambda_A{}^{1}}{\partial r} &=& \frac
2067: {a( r)}{\hr a(\hr)} \left(-\Lambda _1{}^1 \Lambda _A {} ^1 +
2068: O(r^{-2})+
2069: \cO(r^{-2\sigma})+\cO(r^{-\alpha})\right)\;, \label{eq:B7a-a}
2070: \end{eqnarray}
2071: which integrated in a manner similar to that for \Eq{eq:B9-a}
2072: shows that there exists a continuous function $f_A{}^1: M\to \R$
2073: such that\ptc{terms added since $\alpha \le 1$ allowed now; there
2074: was an error here before anyway, so \eq{deltaeq} added }
2075: \begin{deqarr}
2076: \Lambda_A{}^1(r,v^B) &= &
2077: \frac{f_A{}^1(v^B)}{r}+\cO(r^{-\alpha})+\delta_1^\alpha \cO(
2078: r^{-\alpha}\ln r) + O(r^{-1-\delta}) \nn \\ &&
2079: \label{eq:B8--1}\\
2080: & = & O(r^{-1})+\delta_1^\alpha \cO(r^{-\alpha}\ln
2081: r)+\cO(r^{-\alpha})\; , \label{eq:B8--}
2082: \end{deqarr}
2083: with any $\delta$ satisfying\ptc{maybe sigma $\le$ and not $<$?}
2084: \bel{deltaeq} \delta<\min(1,2\sigma-1)\;.\ee It is easily seen now
2085: that the $r^{-2}\ln r$ terms which could potentially be present in
2086: \Eq{eq:B7a--} cannot occur: clearly they are only relevant for
2087: $\alpha = 2$; in that last case it immediately follows from
2088: \Eqsone{eq:B8a} and \eq{eq:B8--1} that no such terms are allowed.
2089: It follows that
2090: \begin{eqnarray}
2091: \Lambda_1{}^1 &= & 1 + O(r^{-2})+\cO(r^{-\alpha})\;. \label{B7a--}
2092: \end{eqnarray}
2093: %Inserting \Eq{eq:B8--}
2094: %into \Eq{eq:B8-} and integrating in $r$ one further finds that
2095: %\begin{eqnarray}
2096: %\Lambda_1{}^A (r,v^B)&= & O(r^{-1})\;.
2097: %\label{eq:B8--an}
2098: %\end{eqnarray}
2099: Repeating the argument after \Eq{eq:B14} one is led to
2100: \begin{deqarr}
2101: {\partial \hr \over \partial r} & = & {\hr \over r} %e^{\psi(v^A)}
2102: +
2103: \cO(r^{-\alpha}) +O(r^{-m_1}) + O(r^{-2})\;, \label{eq:B18first}\\
2104: \hr &=& e^{\psi(v^A)} r + \cO(r^{1-\alpha})+O(r^{1-m_1}) +
2105: O(r^{-1})\;;
2106: \label{eq:B18}\arrlabel{eq:B18m}
2107: \end{deqarr}
2108: \eq{eq:B18} has been obtained from \eq{eq:B18first} by calculating
2109: $\partial \hx / \partial x$ and integrating the resulting
2110: equation, compare \Eqs{hxx}{eq:B16a}. Equations~\eq{eq:B7a-} and
2111: \eq{eq:B18} yield\ptc{log term added cause of small alphas}
2112: \begin{eqnarray*}
2113: {\partial \Lambda_A{}^B \over \partial r} &= & a(r) \left(
2114: \sum_k\hzome^\ell{}_{Ak} \Lambda_\ell{}^B\Lambda_k{}^1
2115: +\cO(r^{-\alpha})\right)
2116: \\ &= & a(r) \left(
2117: \sum_C\hzome^\ell{}_{AC} \Lambda_\ell{}^B\Lambda_C{}^1
2118: +\cO(r^{-\alpha})\right)
2119: \\ & = & \cO(r^{-\alpha-1})
2120: +\delta_1^\alpha \cO(r^{-\alpha-1}\ln r) +O(r^{-3})\;,
2121: \end{eqnarray*}
2122: and by integration one finds that there exists a continuous matrix
2123: valued function $R=(R_A{}^B):\mn\to O(n-1)$ such that
2124: \begin{equation}
2125: \label{eq:B19}
2126: \Lambda_A{}^B(r,v^C)= R_A{}^B(v^C)+ \cO(r^{-\alpha})+
2127: \delta_1^\alpha \cO(r^{-\alpha}\ln^2 r) +O(r^{-2})\;.
2128: \end{equation}
2129: Repeating the argument which led to \Eq{eq:B14} and using
2130: \eq{eq:B19} one finds now that there exists a continuous function
2131: $f_1{}^A: M\to \R$ such that\ptc{there is a problem with the delta
2132: terms which might need clarifying; there was a wrong log term in
2133: the previous equation}
2134: \begin{eqnarray}
2135: \Lambda_1{}^A (r,v^B)&= &
2136: \frac{f_1{}^A(v^B)}{r}\left(1+O(r^{-m_1})\right) \nn \\&&
2137: +\cO(r^{-\alpha})+ \delta _\alpha^1\cO(r^{-\alpha}\ln r)
2138: %\delta_\alpha^2\cO(r^{-\alpha}\ln r)
2139: + \cO(r^{-2\sigma})\;, \label{eq:B8--ao}
2140: \end{eqnarray}
2141: compare \Eq{eq:B7a-a}; without loss of generality we have assumed
2142: that $\sigma \ne 1$. The hatted equivalent of \eq{eq:b2},
2143: $$\hat\alpha^A\equiv \hat \alpha(\hv^C)^A{}_{B} d\hv^B\;,$$ gives
2144: \begin{eqnarray*}
2145: d\hv^A & = & \hbeta^A{}_B \hat \alpha ^B
2146: =\frac 1 \hr \hbeta^A{}_B\hat \theta ^B
2147: \\ & = & \frac 1 \hr \hbeta^A{}_B
2148: \left((1 +\cO(r^{-\alpha}))\Lambda_C{}^B \theta^C + (1
2149: +\cO(r^{-\alpha}))\Lambda_1{}^B \theta^1\right)\;,
2150: \end{eqnarray*}
2151: where $\beta^A{}_B$ denotes the matrix inverse to $\alpha^A{}_B$,
2152: while the symbol $\hbeta^A{}_B$ is used to denote the the matrix
2153: inverse to $\hat \alpha^A{}_B$. This implies\begin{eqnarray}
2154: \label{eq:B20}
2155: {\partial \hv ^A \over \partial r } &= &O(r^{-3})\;,
2156: \\ \label{eq:B21}
2157: {\partial \hv ^A \over \partial v^C } &= &\frac r {\hr}
2158: \hbeta^A{}_B \Lambda_D{}^B \alpha ^D{_C} +\cO(r^{-\alpha}) \;.
2159: \end{eqnarray}
2160: Integrating \eq{eq:B20} in $r$ one obtains that the limits
2161: $$\psi^A\equiv \lim_{r\to\infty}\hv ^A$$
2162: exist and are continuous functions, with \bel{eq:B21+}
2163: \hv^A-\psi^A = O(r^{-2})\;.\ee Moreover, it follows from
2164: \eq{eq:B21} that the limits as ${r}$ tends to infinity of $
2165: {\partial \hv^A /
2166: \partial v^B}$ exist and are continuous.\ptc{Argument rearranged,
2167: since the previous one did not work for small alpha}
2168: Passing
2169: to the limit $r\to\infty$ in Equation~\eq{eq:B21} one obtains
2170: \bel{B22++}\Psi^*\hat \alpha^A= e^{-\psi} R^A{}_B \alpha ^B\;,\ee
2171: hence
2172: \begin{eqnarray*}
2173: \Psi^*\hat h &= &\Psi^*\sum_A\hat\alpha^A\otimes\hat\alpha^A
2174: \\ & = & e^{-2\psi} \sum_AR^A{}_B R^A{}_C \alpha ^B\otimes\alpha ^C
2175: \\ & = & e^{-2\psi} \sum_A\alpha^A\otimes\alpha^A = e^{-2\psi} h\;,
2176: \end{eqnarray*}
2177: where we have used the fact that $R=(R_A{}^B)$ is a rotation
2178: matrix. It follows that the map $\Psi=(\psi^A)$ is a conformal
2179: local diffeomorphism from $(M,h)$ to $(\hM,\hh)$. We can thus use
2180: a deep result of Lelong-Ferrand~\cite{Lelong-Ferrand} to conclude
2181: that $\Psi$ is smooth, in particular so is $\psi$.
2182: Equation~\eq{B22++} shows then smoothness of $R^A{}_B$. Further
2183: \begin{eqnarray}
2184: \frac 1 r \beta^B{}_A {\partial \hr \over \partial v^B}
2185: & = & e_A(\hr) \nonumber \\ & = & (1
2186: +\cO(r^{-\alpha}))\sum_B\Lambda_B{}^A \hf_B(\hr)
2187: + (1 +\cO(r^{-\alpha}))\Lambda_1{}^A \hf_1(\hr)
2188: \nonumber\\ & = & O(1)+\cO(r^{1-\alpha})\;, \label{B22}
2189: \end{eqnarray}
2190: hence \bel{B22+} {\partial \hx \over \partial v^A} = \frac 1 \hr
2191: {\partial \hr \over \partial v^A} = O(1)+\cO(r^{1-\alpha})\;. \ee
2192: This, together with \Eqs{eq:B20}{eq:B21+}, establishes point 1 for
2193: $0<\alpha\le 1$.
2194:
2195: If $\alpha>1$, a closer inspection of the $O(1)$ terms in
2196: \Eq{B22+}, making use of \Eq{eq:B8--ao}, shows that the limits
2197: $\lim_{r\to\infty} {\partial \hx / \partial v^A}$ exist, and are
2198: continuous functions of the $v^A$'s. Now, in the current range of
2199: $\alpha$'s it is easy to show that the function $\psi$ in
2200: \eq{eq:B16a} is continuously differentiable without invoking the
2201: Lelong-Ferrand theorem, as follows: Let $\phi$ be the function
2202: appearing at the right-hand-side of \Eq{hxx}, from
2203: \eq{phiequation} and from what has been said with a little work
2204: one finds
2205: $$\frac{\partial \phi}{\partial v^A} = \cO(e^{(1-\alpha)x}) + O(e^{-x})\;; $$
2206: the differentiability of $\psi$ follows now from its definition
2207: \eq{eq:B16a} and from Lebesgue's dominated theorem on
2208: differentiability of integrals with parameters. The last estimate
2209: together with \Eq{eq:B16a} also show that
2210: \be\label{hxvder}\frac{\partial \hx}{\partial v^A} =
2211: \frac{\partial\psi}{\partial v^A}+\cO(e^{(1-\alpha)x}) +
2212: O(e^{-x})\;.\ee \ptc{end of new stuff; a spurious piece of
2213: previous argument commented out} and point 1 is established.
2214:
2215: % Next, Equation \eq{eq:B5}
2216: %shows that the $\partial \Lambda_A{}^B/
2217: %\partial v^C $'s are uniformly bounded, hence the
2218: %$\Lambda_A{}^B$'s are uniformly Lipschitzian. This implies that
2219: %the $R_A{}^B$'s are Lipschitzian functions, as pointwise limits of
2220: %Lipschitzian functions. Equation~\eq{eq:B21} implies that the
2221: %limits $\lim_{r\to\infty} {\partial \hv^A/
2222: %\partial v^B}$ are Lipschitzian functions, and one concludes that
2223: %$\Psi=(\psi^A)$ is a $C^{1,1}$ map from $\mn$ to itself. Passing
2224: %to the limit $r\to\infty$ in \Eq{eq:B5} multiplied by $r$ shows
2225: %that the limits
2226: %$$
2227: %\lim_{r\to\infty} \partial_A \Lambda_B{}^C(r,v^C)$$ exist, and are
2228: %continuous functions of $v^C$. It follows that the $R_A{}^B$'s are
2229: %continuously differentiable. Passing, next, to the limit
2230: %$r\to\infty$ in \Eq{eq:B21} we obtain --- from uniform convergence
2231: %as $r$ tends to infinity of the $v^A(r,\cdot)$ and their
2232: %derivatives to continuous limiting functions\comment
2233: %--- that the limits
2234: %$$\lim_{r\to\infty}{\partial \hv ^A \over \partial v^C }$$ are
2235: %continuously differentiable functions of the $v^B$'s, hence $\Psi$
2236: %is $C^2$. Straightforward algebra establishes the remaining claims
2237: %in point 1.
2238:
2239: To establish point 2, suppose that $\Psi$ is the identity and that
2240: $\psi=0$. The calculation in \Eq{B22} shows that
2241: $$0 =\lim_{r\to\infty} {\partial \hx \over
2242: \partial v^A} =
2243: \lim_{r\to\infty} {1\over \hr}{\partial \hr \over \partial v^A} =
2244: \lim_{r\to\infty} r\alpha^B{}_A e_B(\hr)
2245: =\lim_{r\to\infty} \alpha^B{}_A\Lambda_B{}^1\;,
2246: $$
2247: hence the function $f_A{}^1$ appearing in \Eq{eq:B8--} vanishes.
2248: The identity
2249: \begin{equation}
2250: \Lambda_1{}^A\Lambda_1{}^1+\sum_B\Lambda_B{}^A\Lambda_B{}^1 =0
2251: \label{B23-}\end{equation} shows that the function $f_1{}^A$ from
2252: \Eq{eq:B8--ao} vanishes as well. If $1<\alpha<2$ we thus obtain
2253: \begin{equation}
2254: \Lambda_i{}^j =\delta_i^j + \cO(r^{-\alpha})\;. \label{B23}
2255: \end{equation}
2256: \ptc{From now on changes} For $\alpha\ge2$ a closer inspection of
2257: \Eq{eq:B8-} is needed: \begin{eqnarray} \nn \frac{\partial
2258: \Lambda_1{}^A}{\partial r} &=& \frac 1 r \frac {ra(r)}{\hr a
2259: (\hr)} \left( \sum_C\Lambda_C{}^A\Lambda_C{}^1 +
2260: \cO(r^{-\alpha})\right)
2261: \\
2262: &=& \frac 1 r \frac {ra(r)}{\hr a (\hr)} \left(- \Lambda_1{}^1
2263: \Lambda_1{}^A + \cO(r^{-\alpha})\right) \;,
2264: \label{B8-++}\end{eqnarray} where we have used \Eq{B23-}. \ptc{an
2265: elegant but unneeded argument commented out}
2266: % For $\epsilon\ge 0$
2267: %let us set
2268: %$$f_\epsilon:=r^{-2} \sqrt{\sum_A\Lambda_1{}^A\Lambda_1{}^A +
2269: %\epsilon^2}\;.$$ It follows from \Eq{B8-++} that for $\epsilon>0$
2270: %we have
2271: %\begin{eqnarray*} &\displaystyle
2272: %\frac{\partial f_\epsilon}{\partial r} \le \chi_5 f_\epsilon +
2273: %\chi_6\;, &
2274: %\\
2275: %& \chi_5 = O(r^-5) + O(r^{-m_1-3})\;, \quad \chi_6 =
2276: %\cO(r^{-\alpha-3})\;.
2277: %\end{eqnarray*}
2278: %which implies
2279: %$$f_\epsilon \le \exp\left\{-\int_r^\infty \chi_5(v) dv\right\} \left(\epsilon -
2280: %\int_r^\infty \exp\left\{\int_s^\infty \chi_5(v) dv\right\}
2281: %\chi_6(s)\; ds\right)\;.
2282: %$$
2283: %Passing with $\epsilon $ to zero we then obtain
2284: %$$|\Lambda_1{}^A| \le r^2 f_0 = \cO(r^{-\alpha})\;.
2285: %$$
2286: Integrating this equation in a way somewhat similar to
2287: \Eq{eq:B9-a} shows that
2288: \begin{eqnarray}
2289: \Lambda_1{}^A (r,v^B)= \cO(r^{-\alpha})+ \delta
2290: _\alpha^2\cO(r^{-\alpha}\ln r)\;. \label{eq:B8--ao+}
2291: \end{eqnarray}
2292: If $\alpha=2$, suppose for the moment that there is no $\ln r$
2293: term in \eq{eq:B8--ao+}; it then follows from \Eqsone{B7a--} and
2294: \eq{B23-} that \Eq{B23} holds. On the other hand, for $\alpha>2$
2295: \Eq{eq:B8a} forces the function $f_1{}^1$ from \Eq{eq:B9-b} to
2296: vanish, which in turn implies that \Eq{B23} holds again. The
2297: formula inverse to \Eq{B4a+} reads \be \label{B4b+}f_j =
2298: \sum_i\Lambda_i{}^j \hf_i\;,\ee in particular
2299: $$f_1(\hr) = \Lambda_1{}^1 \hf_1(\hr)+\sum_A\Lambda_A{}^1 \hf_A(\hr)\;,$$
2300: which implies
2301: \begin{equation}{\partial \hr \over \partial r } = {a(r) \over a(\hr) }
2302: \left(1+\cO(r^{ -\alpha})\right)+\cO(r^{ -\alpha})\;. \label{B24}
2303: \end{equation}
2304: \Eq{eq:b0-2m} together with the identities
2305: \begin{deqarr}
2306: \arrlabel{B24+}
2307: & \displaystyle
2308: {ra(r) \over \hr a(\hr)} = { ra(r) - \hr a(\hr)\over \hr a(\hr)}
2309: +1 = 1 + O(r^{-m_1-1}) \delta r \;, & \\ & \displaystyle {a(r)
2310: \over a(\hr)} = { ra(r) \over \hr a(\hr)} \times {\hr \over r} =
2311: { ra(r) \over \hr a(\hr)} \left(1+{\delta r\over r}\right) = 1 +
2312: \left({1\over r} +O(r^{-m_1-1})\right) \delta r \;.\end{deqarr}
2313: shows that \Eq{B24} can be rewritten as
2314: \begin{equation}
2315: {\partial \delta r \over \partial r} = \chi_3
2316: \delta r + \chi_4 \;,
2317: \label{B25}
2318: \end{equation}
2319: with
2320: $$\chi_3 = {1\over r} + O (r^{-m_1-1})\;, \qquad \chi_4 =
2321: \cO(r^{-\alpha})\;.
2322: $$
2323: Hence, for $r_2>r$ we have
2324: \begin{eqnarray*}
2325: {\delta r(r_2) \over r_2} &=& \left( {\delta r (r)\over r} +
2326: \int_r^{r_2} \exp\left\{-\int_t^\infty \left(\chi_3(s)-{1\over
2327: s}\right)ds\right\} {\chi_4(t)\over t}dt\right) \times\\ &&
2328: \exp\left\{\int_r^\infty \left(\chi_3(s)-{1\over
2329: s}\right)ds\right\} \;.
2330: \end{eqnarray*}
2331: Passing with $r_2$ to infinity and using the fact that $\delta r =
2332: o(r)$ shows that
2333: \begin{eqnarray}
2334: \delta r &=& -r \int_r^{\infty}
2335: \exp\left\{-\int_t^\infty \left(\chi_3(s)-{1\over
2336: s}\right)ds\right\} {\chi_4(t)\over t}dt \nonumber \\ &=&
2337: \cO(r^{1-\alpha})\;. \label{B27-}\end{eqnarray} From \Eq{B24} we
2338: obtain
2339: \begin{equation}{\partial \delta r \over \partial r } = \cO(r^{ -\alpha})\;.
2340: \label{B27}
2341: \end{equation}
2342: \Eq{B4b+} implies
2343: \begin{eqnarray}
2344: e_j &=&\left(\delta_j^k +\cO(r^{ -\alpha})\right) f_k \nonumber \\
2345: &=& \sum_i\left(\delta_j^k +\cO(r^{ -\alpha})\right) \Lambda_i{}^k
2346: \left(\delta_i^k +\cO(r^{ -\alpha})\right)
2347: \he_k
2348: \nonumber \\ &=& \he_j +\sum_i \cO(r^{ -\alpha}) \he_i \;.
2349: \label{B4c+}\end{eqnarray} It follows that
2350: \begin{deqarr}
2351: e_j(\hr) &=& \delta_j^1\hr +\cO(r^{1 -\alpha})
2352: \nonumber\\ &=&
2353: e_j(r) +\cO(r^{1 -\alpha})\;,
2354: \\
2355: e_j(\hv^A) &=& \he_j(\hv^A)
2356: +\cO(r^{-1 -\alpha})
2357: \;. \label{B4d+}
2358: \end{deqarr}
2359: In particular
2360: \begin{eqnarray}
2361: &\displaystyle{\partial \hv^A \over \partial r} = a(r)
2362: e_1(\hv^A) = \cO(r^{-2-\alpha})
2363: %& \nonumber \\ &
2364: \quad \Longrightarrow \quad
2365: %&\nonumber \\ &
2366: \hv^A - v^A = \cO(r^{-1-\alpha})\;.
2367: \label{B4e+}\end{eqnarray} \Eqsone{B27-} and \eq{B4e+} allow us to
2368: rewrite \Eq{B4d+} as
2369: \begin{eqnarray}
2370: e_j(\hv^A) &=& e_j(v^A)
2371: +\cO(r^{-1 -\alpha})
2372: \;. \label{B4f+}\end{eqnarray} \Eqs{eq:t1}{eq:t2} are thus
2373: established. A straightforward analysis of the equations
2374: \begin{eqnarray*}
2375: \label{B4g+}e_k\left(f_j(\hr)\right) &=&
2376: e_k\left(\sum_i\Lambda_i{}^j \hf_i(\hr)\right)
2377: \;,\\
2378: e_k\left(f_j(\hv^A)\right) &=& e_k\left(\sum_i\Lambda_i{}^j
2379: \hf_i(\hv^A)\right) \;,
2380: \end{eqnarray*}
2381: leads to \Eq{t2a}, and the theorem is established for $\alpha\ne
2382: 2$, or for $\alpha=2$ provided that no log terms are present in
2383: \eq{eq:B8--ao+}.\reword
2384:
2385: Let us return to the case $\alpha=2$; then \eq{B23} holds with any
2386: $\alpha<2$ and therefore the calculations that follow remain valid
2387: with any $\alpha<2$. Further \eq{B23} holds with $i=j=1$ and
2388: $\alpha=2$ by \Eq{B7a--}. \Eqsone{eq:B5} and \eq{B24+} give then
2389: $$\frac{\partial(r\Lambda_A{}^1)}{\partial r} = \cO(r^{-2}) \quad
2390: \Longrightarrow \quad \Lambda_A{}^1= \cO(r^{-2})\;,
2391: $$
2392: so that no $\log$ terms can occur in $\Lambda_A{}^1$. \Eq{B23-}
2393: implies then
2394: $$ \Lambda_1{}^A= \cO(r^{-2})\;,
2395: $$
2396: and \eq{eq:B5} establishes \eq{B23} with $\alpha=2$,\reword and
2397: the theorem follows. \qed
2398:
2399: In the next section we shall need the following:
2400: \begin{Corollary}
2401: \label{C1} Let $\Psi(r,v^A)=(\hr(r,v^A),\hv^B(r,v^A))$
2402: be an isometry of the background metric $b$:
2403: $$\Psi^*(a^2(r)dr^2 +r^2 h) =a^2(\hr)d\hr^2 +\hr^2 \hh\;, \qquad \hh=h_{AB}(\hv^C)d\hv^A
2404: d\hv^B\;.$$ If there exists $\alpha >1$ such that the physical
2405: metric $g$ satisfies
2406: \begin{eqnarray}
2407: \label{eq:hd1}
2408: & g(e_i,e_j)-\delta_i^j = \cO(r^{-\alpha})\;,\quad e_k(g(e_i,e_j))
2409: = \cO(r^{-\alpha})\;, &
2410: \end{eqnarray}
2411: where $e_i$, $i=1,\ldots,n$ is the usual ON frame for the metric
2412: $b$ as in \eq{frame}, then
2413: \begin{eqnarray}
2414: \label{eq:hd2}
2415: &
2416: g(\he_i,\he_j)-\delta_i^j = \cO(\hr^{-\alpha})\;,\quad
2417: \he_k(g(\he_i,\he_j)) = \cO(\hr^{-\alpha})\;,
2418: \end{eqnarray}
2419: where $\he_i$ is the corresponding hatted frame.
2420: \end{Corollary}
2421:
2422: \proof Applying point 1. of Theorem~\ref{T1} to $g=b$ we obtain
2423: that
2424: \begin{eqnarray}
2425: \label{eq:t1+}& \hr = e^{\psi} r + O(r^{1-\beta})\;,
2426: \end{eqnarray}
2427: with $\beta=\min{(m_1,2)}$. Since $\Psi$ is an isometry we have
2428: $\he_i = \Lambda_i{}^je^j$ for some rotation matrix
2429: $\Lambda_i{}^j$, which gives
2430: $$ \he^{ij}= (g-b)(\htheta^i,\htheta^j)
2431: =\Lambda_k{}^i\Lambda_\ell{}^j(g-b)(\theta^k,\theta^\ell)=
2432: \cO(r^{-\alpha})=\cO(\hr^{-\alpha})\;.$$ Further,
2433: $$e_r({\he^{ij}})=e_r\left( \Lambda_k{}^i\Lambda_\ell{}^je^{kl}\right)\;,$$
2434: and --- since $e_i( \Lambda_k{}^j)=O(1)$ by \eq{eq:B5} --- the
2435: result easily follows. \qed
2436: \section{Global charges}\label{Sgc}
2437: \newcommand{\Isot}{\text{Iso}^\uparrow(\hyp,b)}
2438: \newcommand{\Iso}{\text{Iso}(\hyp,b)}
2439: \newcommand{\Isoo}{\text{Iso}_0(\hyp,b)}
2440: %\newcommand{\Isoto}{\text{Iso}^\uparrow_0(\hyp,b)}
2441: \newcommand{\cC}{\mycal C}
2442: Let us give here a general prescription how to assign global
2443: geometric invariants to hypersurfaces $\hyp$ in the class of
2444: space-times with metrics asymptotic to backgrounds \eq{A1}:
2445: consider such a background metric $b$ and consider a hypersurface
2446: $\hyp$ given by the equation $\{t=0\}$ in the coordinate system of
2447: \eq{A1}. Let $\cK$ denote the set of all Killing vector fields of
2448: $b$; the hypersurface $\hyp$ singles out two subsets of $\cK$: a)
2449: the set $\cKS$ of those Killing vector fields of $b$ which are
2450: normal to $\hyp$, and the set $\cKt$ of all $b$-Killing vector
2451: fields which are tangent to $\hyp$. Consider any metric $g$ for
2452: which the fall-off hypotheses of Theorem~\ref{T0} are met, with
2453: $X\in\cKS$, or with $X\in\cKt$, or perhaps with all $X\in\cK$. (In
2454: that theorem we have assumed that $b$ satisfies Equation~\eq{Ee},
2455: but it would be sufficient that \eq{Ee} holds only up to terms
2456: which decay sufficiently fast when $r$ tends to infinity, the same
2457: for $g$.) Let $\Isot$ be the group of all time-orientation
2458: preserving isometries of $b$ which leave $\hyp$
2459: invariant.\footnote{Some further invariants can sometimes be
2460: obtained by considering the connected component of the identity of
2461: $\Isot$, but this seems to require a case by case analysis, so
2462: that no general discussion will be given here.} We shall suppose
2463: further that the following condition holds:
2464: \begin{eqnarray}\nonumber
2465: & \text{\emph{for every orientation-preserving
2466: conformal isometry $\Psi$ of $(M,h)$ there exists }} & \\
2467: % \label{H}
2468: \nonumber &\text{\emph{$R_*>0$ and a $b$-isometric map $\Phi:
2469: [R_*,\infty)\times M \to [R,\infty)\times M$, such that}} & \\
2470: \label{H} & \lim_{r\to\infty} \Phi(r,\cdot) = \Psi(\cdot)\;.&
2471: \end{eqnarray} It follows from
2472: \cite[Vol.~II,Theorem~18.10.4]{Berger} and from what is said in
2473: Appendix \ref{Aai1} that this condition holds for the
2474: $(n+1)$-dimensional anti-de Sitter metrics, $n\ge 3$; the case
2475: $n=2$ is handled by the discussion of toroidal $\mn$'s below.
2476: Further, the above condition obviously holds for those metrics for
2477: which every conformal isometry of $(M,h)$ is an isometry, as is
2478: the case for the $(M,h)$'s considered in Appendix~\ref{Aai2}: the
2479: desired extension $\Phi$ is
2480: $$\Phi(r,v^A)=(r,\Psi(v^A))\;.$$
2481: In fact, it is shown in \cite{ChHerzlich} that condition~\eq{H}
2482: always holds when $a(r)=1/\sqrt{r^2+k}$ regardless of the metric
2483: $h$.
2484:
2485: Let $\cC$ denote the collection of positively oriented coordinate
2486: systems $c=\left(\cO,(r,v^A)\right)$, where $\cO$ is the domain of
2487: definition of the collection of functions $(r,v^A)$, with the
2488: associated background metrics and orthonormal tetrads, for which
2489: Equations~\eq{c1}, \eq{eq:deccond} and \eq{eq:deccond1} hold. For
2490: each such coordinate system $c$ we can calculate the set of
2491: integrals \eq{toto}, where $X$ runs over $\cKS$, or over $\cKt$,
2492: or over $\cK$, whichever appropriate. By Theorem~\ref{T1} every
2493: two coordinate system $c_1,c_2$ in $\cC$ differ by a coordinate
2494: transformation, say $\Upsilon$, the $M$-part of which asymptotes
2495: to an orientation preserving conformal isometry $\Psi:M\to M$. By
2496: the hypothesis \eq{H} $\Psi$ can be extended to an isometry $\Phi$
2497: of $b$ which leaves $\hyp$ invariant. Writing $\Upsilon$ as
2498: $$\Upsilon = (\Upsilon\circ \Phi^{-1})\circ\Phi$$ we can decompose
2499: $\Upsilon$ into an isometry of $b$ and a map
2500: $\Upsilon\circ\Phi^{-1} $ which asymptotes to the identity. By
2501: Corollary~\ref{C1} the metric $\Phi^*g$ satisfies the hypotheses
2502: of Theorem~\ref{T1}, in the new coordinates $c_3$ as made precise
2503: by that Corollary, so that the conclusions of Theorem~\ref{T1}
2504: apply to $\Upsilon\circ \Phi^{-1}$. Let $b$ be the background
2505: associated with the first coordinate system $c_1$, and let $\hb$
2506: be that associated with $c_2$; since $\Phi$ is an isometry, the
2507: background metric associated with $c_3$ coincides with that
2508: associated with $c_1$. Now, by Theorem~\ref{TP1}, the integrals
2509: \eq{toto} are invariant under the change of background which is
2510: associated to $\Upsilon\circ\Phi^{-1}$:
2511: \begin{equation}
2512: \label{eq:t1*}
2513: m(\hyp,g,b, X)= m(\hyp,g,\hb, \hat X)\;,\end{equation}
2514: where the $\hb$-Killing vector $\hat X$ is associated to the
2515: $b$-Killing vector field $X$ as described in the statement of
2516: Theorem~\ref{TP1}. On the other hand, Lemma~\ref{L1} shows that
2517: the
2518: isometry $\Phi$ reshuffles the integrals associated with
2519: different Killing vectors,
2520: \begin{equation}
2521: \label{eq:t1**}m(\hyp,\Phi^*g,b,(\Phi_*)^{-1}X) =
2522: m(\hyp,g,b,X)\;,\end{equation}
2523: according to the action of $\Isot$ by push-forward on $\cK$, or
2524: $\cKS$, or $\cKt$. (We note that since $\Phi$ is an $b$-isometry
2525: preserving $\hyp$, $\Phi_*$ preserves the field of $b$-unit
2526: normals to $\hyp$, hence the space $\cKS$ of those Killing vector
2527: fields which are normal to $\hyp$. Similarly $\Phi$ preserves the
2528: space $\cKt$ of Killing vector fields tangent to $\hyp$.)
2529: Equations~\eq{eq:t1*}-\eq{eq:t1**} show that any invariant of the
2530: action of $\Isot$ on $\cK$, or on $\cKS$, or on $\cKt$, gives a
2531: geometric invariant which can be associated to $\hyp$,
2532: independently of the choice the coordinate
2533: systems in $\cC$. % which define the backgrounds $b$.
2534:
2535: When $b$ is the $(n+1)$-dimensional anti-de Sitter metric, the
2536: relevant invariants based on Killing vector fields in $\cKS$ have
2537: already been discussed in detail in the introduction,
2538: Section~\ref{Si}. Consider the remaining Killing vector fields
2539: $L_{(\mu)(\nu)}\in\cKt$, as given by \Eq{Lor2}. Equation~\eq{remK}
2540: shows that under the action of $\Isot=O^\uparrow(1,n)$, the
2541: orthochronous $(n+1)$-dimensional Lorentz group, the integrals
2542: $$Q_{(\mu)(\nu)}\equiv m(\hyp,g,b,L_{(\mu)(\nu)})$$ transform as
2543: the components of a two-covariant antisymmetric tensor. One then
2544: obtains a
2545: geometric invariant of $\hyp$ by calculating
2546: \begin{equation}
2547: \label{eq:ch}Q\equiv Q_{(\mu)(\nu)}Q_{(\alpha)(\beta)}\eta^{(\mu)(\alpha)}\eta^{(\nu)(\beta)}\end{equation}
2548: (for conventions see Appendix~\ref{Aai1}). In dimension $3+1$
2549: another independent geometric invariant is obtained from
2550: \begin{equation}
2551: \label{eq:ch1}
2552: Q^*\equiv
2553: Q_{(\mu)(\nu)}Q_{(\alpha)(\beta)}\epsilon^{(\mu)(\alpha)(\nu)(\beta)}\;.
2554: \end{equation}
2555: In higher dimensions further invariants are obtained by
2556: calculating $\text{tr}(P^{2k})$, $2\le 2k\le (n+1)$, where
2557: $P^{(\alpha)}{}_{(\beta)}= \eta^{(\alpha)(\mu)}Q_{(\mu)(\beta)}$.
2558: (In this notation $Q$ given by \Eq{eq:ch} equals
2559: $\text{tr}(P^{2})$.) When $n+1$ is even one also has obvious
2560: generalisations of \eq{eq:ch1}.
2561:
2562: Consider, next, a (compact) strictly negatively curved $(\mn,h)$,
2563: as considered in Appendix~\eq{Aai2}. In that case all Killing
2564: vector fields are in $\cKS$, the action of $\Isot$ on $\cKS$ is
2565: trivial, and all the geometric invariants of $\hyp$ given by
2566: \eq{toto} are provided by the mass integrals considered in the
2567: Introduction.
2568:
2569: Let, finally, $(\mn,h)$ be a flat $(n-1)$-dimensional torus
2570: $\mathbb{T}^n$, $n\ge 2$; as discussed in Appendix~\eq{Aai2}, all
2571: conformal isometries of $(\mn,h)$ are isometries and the action
2572: of $\Isot$ on $\cK$ is trivial. It follows that in addition to the
2573: mass we have $n-1$ independent invariants
2574: $$m_A(\hyp,g)\equiv m(\hyp,g,b, \partial_A)$$
2575: associated with the Killing vectors $\partial_A$ of $(\mn,h)$;
2576: here
2577: the $\partial_A$'s have been chosen so that they are
2578: tangent to the $S^1$ factors of $\T^N=S^1\times\cdots\times S^1$,
2579: and normalized to have unit length; such vector fields can loosely
2580: be thought of as generating ``rotations'' of the torus
2581: $\mathbb{T}^n$ into itself, giving the $m_A$'s an angular-momentum
2582: type character.
2583: %Here the $h$-Killing vector filed
2584: %$X=X^A\partial_A$ on $\mn$ has been extended to
2585: %a space-time Killing vector field in the obvious way.
2586:
2587:
2588: \appendix
2589: \section{The phase space and the Hamiltonians}
2590: \label{AHam} {In \cite{ChAIHP} the starting
2591: point of the analysis is the Hilbert Lagrangian for vacuum Einstein
2592: gravity, $$\mathcal{L}= \sqrt{- \det g_{\mu\nu}}~\frac{g^{\alpha
2593: \beta} R_{\alpha \beta}}{16\pi}\;. $$ With our signature
2594: $(-+\cdots+)$ one needs to repeat the analysis in \cite{ChAIHP}
2595: with
2596: $\mathcal{L}$ replaced by $$\frac{\sqrt{- \det
2597: g_{\mu\nu}}}{16\pi}\left(g^{\alpha \beta} R_{\alpha \beta}
2598: -2\Lambda\right)\;,$$ and without making the assumption $n+1=4$ done
2599: there; we follow the presentation in \cite{CJK}:
2600: Consider the Ricci tensor, \be \label{JK1} R_{\mu\nu} =
2601: \partial_\alpha \left[
2602: {\Gamma}^{\alpha}_{\mu\nu} - {\delta}^{\alpha}_{(\mu}
2603: {\Gamma}^{\kappa}_{\nu ) \kappa} \right] - \left[
2604: {\Gamma}^{\alpha}_{\sigma\mu} {\Gamma}^{\sigma}_{\alpha\nu}-
2605: {\Gamma}^{\alpha}_{\mu\nu} {\Gamma}^{\sigma}_{\alpha\sigma} \right]
2606: \;, \ee where the $\Gamma$'s are the Christoffel symbols of $g$.
2607: Contracting $R_{\mu\nu}$ with the contravariant density of metric,
2608: \be \label{JK1.1}
2609: \Kp^{\mu\nu} := %\textstyle
2610: \frac 1{16 \pi} \sqrt{-\det g } \ g^{\mu\nu} \;, \ee one obtains
2611: the following expression for the Hilbert \Lagrangian density:
2612: \begin{eqnarray}
2613: {\tilde L} & = & \frac 1{16 \pi} \sqrt{-\det g } R = \Kp^{\mu\nu}
2614: R_{\mu\nu} \nonumber \\ \label{JK3} & = &
2615: \partial_\alpha \left[
2616: \Kp^{\mu\nu} \left( {\Gamma}^{\alpha}_{\mu\nu} -
2617: {\delta}^{\alpha}_{(\mu} {\Gamma}^{\kappa}_{\nu ) \kappa} \right)
2618: \right] + \Kp^{\mu\nu} \left[ {\Gamma}^{\alpha}_{\sigma\mu}
2619: {\Gamma}^{\sigma}_{\alpha\nu}-
2620: {\Gamma}^{\alpha}_{\mu\nu} {\Gamma}^{\sigma}_{\alpha\sigma} \right] \;.
2621: \end{eqnarray}
2622: Here we have used the metricity condition of $\Gamma$, which is
2623: equivalent to the following identity: \be\label{JK2}
2624: \Kp^{\mu\nu}_{\ \
2625: , \alpha} := \partial_\alpha \Kp^{\mu\nu} = \Kp^{\mu\nu}
2626: {\Gamma}^{\sigma}_{\alpha\sigma} - \Kp^{\mu\sigma}
2627: {\Gamma}^{\nu}_{\sigma\alpha} - \Kp^{\nu\sigma}
2628: {\Gamma}^{\mu}_{\sigma\alpha} \;. \ee Suppose now, that
2629: ${\Bgamma}^{\alpha}_{\sigma\mu}$ is another symmetric connection
2630: in $M$, which will be used as a background (or reference)
2631: connection. Denote by $\zR_{\mu\nu}$ its Ricci tensor. From the
2632: metricity condition \eq{JK2} we similarly obtain
2633: \begin{eqnarray}
2634: \Kp^{\mu\nu} \zR_{\mu\nu} \nonumber & = &
2635: \partial_\alpha \left[
2636: \Kp^{\mu\nu} \left( {\Bgamma}^{\alpha}_{\mu\nu} -
2637: {\delta}^{\alpha}_{(\mu} {\Bgamma}^{\kappa}_{\nu ) \kappa} \right)
2638: \right] - \Kp^{\mu\nu} \left[ {\Bgamma}^{\alpha}_{\sigma\mu}
2639: {\Bgamma}^{\sigma}_{\alpha\nu}-
2640: {\Bgamma}^{\alpha}_{\mu\nu} {\Bgamma}^{\sigma}_{\alpha\sigma} \right]
2641: \nonumber \\
2642: & & +
2643: \Kp^{\mu\nu} \left[ {\Gamma}^{\alpha}_{\sigma\mu}
2644: {\Bgamma}^{\sigma}_{\alpha\nu} + {\Bgamma}^{\alpha}_{\sigma\mu}
2645: {\Gamma}^{\sigma}_{\alpha\nu} - {\Gamma}^{\alpha}_{\mu\nu}
2646: {\Bgamma}^{\sigma}_{\alpha\sigma} - {\Bgamma}^{\alpha}_{\mu\nu}
2647: {\Gamma}^{\sigma}_{\alpha\sigma} \right]
2648: \;. \label{JK4}
2649: \end{eqnarray}
2650: It is useful to introduce the tensor field \be \KA^\alpha_{\mu\nu}
2651: :=\left( {\Bgamma}^{\alpha}_{\mu\nu} -
2652: {\delta}^{\alpha}_{(\mu} {\Bgamma}^{\kappa}_{\nu ) \kappa} \right) -\left({\Gamma}^{\alpha}_{\mu\nu} - {\delta}^{\alpha}_{(\mu}
2653: {\Gamma}^{\kappa}_{\nu ) \kappa}\right) \label{Pend} \;. \ee Once
2654: the reference connection $\Bgamma^\alpha_{\mu\nu}$ is given, the
2655: tensor $\KA^\alpha_{\mu\nu}$ encodes the entire information about
2656: the connection $\Gamma^\alpha_{\mu\nu}$: \[
2657: {\Gamma}^{\alpha}_{\mu\nu} = {\Bgamma}^{\alpha}_{\mu\nu} -
2658: \KA^{\alpha}_{\mu\nu} + \frac 2{n} {\delta}^{\alpha}_{(\mu}
2659: \KA^{\kappa}_{\nu ) \kappa} \] (recall that the space-time
2660: dimension is $n+1$). Subtracting Equation~\eq{JK4} from \eq{JK3},
2661: and using the definition of $\KA^\alpha_{\mu\nu} $, we arrive at
2662: the equation
2663: \[\Kp^{\mu\nu}
2664: R_{\mu\nu} -\frac{\sqrt{- \det
2665: g_{\mu\nu}}}{8\pi}\Lambda= -\partial_\alpha \left( \Kp^{\mu\nu}
2666: \KA^\alpha_{\mu\nu} \right) + L \;,\] where
2667: \begin{eqnarray*} L &:=& \Kp^{\mu\nu}
2668: \left[ \left(
2669: {\Gamma}^{\alpha}_{\sigma\mu} - {\Bgamma}^{\alpha}_{\sigma\mu}
2670: \right) \left( {\Gamma}^{\sigma}_{\alpha\nu}-
2671: {\Bgamma}^{\sigma}_{\alpha\nu} \right) - \left(
2672: {\Gamma}^{\alpha}_{\mu\nu} - {\Bgamma}^{\alpha}_{\mu\nu} \right)
2673: \left( {\Gamma}^{\sigma}_{\alpha\sigma} -
2674: {\Bgamma}^{\sigma}_{\alpha\sigma} \right) + \zR_{\mu\nu} \right]
2675: \\ && -\frac{\sqrt{- \det
2676: g_{\mu\nu}}}{8\pi}\Lambda\;.\end{eqnarray*} This result may be
2677: used as follows: the quantity $L$ differs by a total divergence
2678: from the gravitational Lagrangian, and hence the associated
2679: variational principle leads to the same equations of motion.
2680: Further, the metricity condition \eq{JK2} enables us to rewrite
2681: $L$ in terms of the first derivatives of $\Kp^{\mu\nu}$: indeed,
2682: replacing in \eq{JK2} the partial derivatives $
2683: \Kp^{\mu\nu}_{\ \;, \alpha} $ by the covariant derivatives
2684: $\Kp^{\mu\nu}_{\ \ ; \alpha} $, calculated with respect to the
2685: background connection $\Bgamma$,\be\label{JK2.1} \Kp^{\mu\nu}_{\ \
2686: ; \alpha} := \Kp^{\mu\nu} \left(
2687: {\Gamma}^{\sigma}_{\alpha\sigma} - {\Bgamma}^{\sigma}_{\alpha\sigma}
2688: \right) - \Kp^{\mu\sigma} \left( {\Gamma}^{\nu}_{\sigma\alpha} -
2689: {\Bgamma}^{\nu}_{\sigma\alpha} \right) - \Kp^{\nu\sigma} \left(
2690: {\Gamma}^{\mu}_{\sigma\alpha} - {\Bgamma}^{\mu}_{\sigma\alpha}
2691: \right) \;, \ee we may calculate $\KA^{\alpha}_{\mu\nu}$ in terms
2692: of the latter derivatives. The final result is:
2693: \begin{eqnarray}\label{Pend2}
2694: \KA^{\lambda}_{\mu\nu} & = & \frac 12
2695: \Kp_{\mu\alpha} \Kp^{\lambda\alpha}_{\ \ ; \nu}
2696: + \frac 12
2697: \Kp_{\nu\alpha} \Kp^{\lambda\alpha}_{\ \ ; \mu}
2698: - \frac 12
2699: \Kp^{\lambda\alpha} \Kp_{\sigma\mu} \Kp_{\rho\nu}
2700: \Kp^{\sigma\rho}_{\ \ ; \alpha}
2701: \nonumber \\ & &
2702: +\frac 1{2(n-1)}
2703: \Kp^{\lambda\alpha} \Kp_{\mu\nu} \Kp_{\sigma\rho}
2704: \Kp^{\sigma\rho}_{\ \ ; \alpha} \;,
2705: \end{eqnarray}
2706: where by $\Kp_{\mu\nu}$ we denote the matrix inverse to
2707: $\Kp^{\mu\nu}$; we assume that $n\ge 2$. Further,
2708: \[ {\Gamma}^{\alpha}_{\mu\nu}-
2709: {\Bgamma}^{\alpha}_{\mu\nu} = - \KA^{\alpha}_{\mu\nu} + \frac 1{n-1}
2710: \Kp_{\sigma\rho}
2711: \Kp^{\sigma\rho}_{\ \ ; (\mu}{\delta}^{\alpha}_{\nu)} \;. \]
2712: We have \be\label{JK5alt} \frac{\partial L}{\partial
2713: \Kp^{\mu\nu}_{\ \;, \lambda}} = \frac{\partial L}{\partial
2714: \Kp^{\mu\nu}_{\ \ ; \lambda}} = \frac{\partial L}{\partial
2715: \Gamma^{\alpha}_{\beta\gamma}}\frac{\partial
2716: \Gamma^{\alpha}_{\beta\gamma}}{\partial \Kp^{\mu\nu}_{\ \ ;
2717: \lambda}} = \KA^{\lambda}_{\mu\nu}\;, \ee with the last equality
2718: being obtained by tedious but otherwise straightforward algebra.
2719: It follows that the tensor $ \KA^{\lambda}_{\mu\nu}$ is the
2720: momentum canonically conjugate to the contravariant tensor density
2721: $\Kp^{\mu\nu}$; prescribing this last object is of course
2722: equivalent to prescribing the metric. Alternatively, one can
2723: calculate
2724: \begin{eqnarray}
2725: L & = & \frac 12 \Kp_{\mu\alpha} \Kp^{\mu\nu}_{\ \ ; \lambda}
2726: \Kp^{\lambda\alpha}_{\ \ ; \nu} - \frac 14 \Kp^{\lambda\alpha}
2727: \Kp_{\sigma\mu} \Kp_{\rho\nu} \Kp^{\mu\nu}_{\ \ ; \lambda}
2728: \Kp^{\sigma\rho}_{\ \ ; \alpha} \nonumber \\ & & + \frac 18
2729: \Kp^{\lambda\alpha} \Kp_{\mu\nu} \Kp^{\mu\nu}_{\ \ ; \lambda}
2730: \Kp_{\sigma\rho} \Kp^{\sigma\rho}_{\ \ ;
2731: \alpha}+\Kp^{\mu\nu}\zR_{\mu\nu} -\frac{\sqrt{- \det
2732: g_{\mu\nu}}}{8\pi}\Lambda\;.\label{defL}
2733: \end{eqnarray}
2734: and check directly that \be\label{JK5} \frac{\partial
2735: L}{\partial \Kp^{\mu\nu}_{\ \ ; \lambda}} =
2736: \KA^{\lambda}_{\mu\nu}\;. \ee
2737:
2738: Given a symmetric background connection
2739: $B$ on $M$, we take $L$ given by Equation~\eq{defL} as the
2740: \Lagrangian for the theory. The canonical momentum $
2741: \KA^{\lambda}_{\mu\nu}$ is defined by Equation~\eq{Pend2} or,
2742: equivalently, by \Eq{JK5}. If ${\hyp}$ is any piecewise smooth
2743: hypersurface in $M$, we define {\em the space-time phase bundle
2744: over $\hyp$ } as the collection of the $
2745: (\KA^{\lambda}_{\mu\nu},\Kp^{\alpha\beta}) $'s over $\hyp$. If $
2746: (\delta_a\KA^{\lambda}_{\mu\nu},\delta_a \Kp^{\alpha\beta}) $,
2747: $a=1,2$, are two sections over $\hyp$ of the bundle of vertical
2748: vectors tangent to the space-time phase bundle, following
2749: \cite{KijowskiGRG} we set \be\label{N.3g}
2750: \Omega_{\hyp}((\delta_1\KA^{\lambda}_{\mu\nu},\delta_1
2751: \Kp^{\alpha\beta}),(\delta_2\KA^{\lambda}_{\mu\nu},\delta_2
2752: \Kp^{\alpha\beta}))=\int_{\hyp}
2753: (\delta_1\KA^{\mu}_{\alpha\beta}\delta_2
2754: \Kp^{\alpha\beta}-\delta_2\KA^{\mu}_{\alpha\beta}\delta_1
2755: \Kp^{\alpha\beta})\rd S_\mu \, , \ee with the fields
2756: $(\delta_1\KA^{\lambda}_{\mu\nu},\delta_1 \Kp^{\alpha\beta})$ and
2757: $(\delta_2\KA^{\lambda}_{\mu\nu},\delta_2 \Kp^{\alpha\beta})$ such
2758: that the integrals converge. Here $\rdS_{\mu}$ is defined as
2759: \begin{equation}
2760: \label{eq:dSdef}
2761: \frac{\partial}{\partial x^\mu}\lrcorner \rd x^0 \wedge\cdots
2762: \wedge\rd x^{n} \;,
2763: \end{equation}
2764: where $\lrcorner$ denotes contraction.
2765: This can be loosely thought of as being the ``symplectic'' form on
2766: the gravitational phase space; however we will avoid this
2767: terminology since the definition of a symplectic form involves
2768: non-degeneracy conditions, which are quite subtle in an infinite
2769: dimensional context, and which we do not want to address.
2770:
2771: To be more specific, let $\hyp$ be a hypersurface which is the
2772: union of a compact set with an asymptotic region $\hypext\approx
2773: [R_0,\infty)\times M$ parameterized by $(r,v^A)$ as in the body of
2774: this paper. Consider a background metric $b$ of the form
2775: \eq{eq:a1} defined on $\hypext$, with its associated tetrad $e_a$;
2776: we define the phase space $\cP_b$ as the space of those
2777: smooth\footnote{The condition of smoothness of the relevant fields
2778: is certainly not needed, and should be relaxed if an attempt is
2779: made to obtain a full symplectic description of the situation at
2780: hand.} sections $ (\KA^{\lambda}_{\mu\nu},\Kp^{\alpha\beta}) $
2781: along $\hyp$ of the space-time phase bundle which satisfy the
2782: following conditions:
2783: \begin{enumerate}
2784: \item[${\mycal C} 1$.]
2785: First, we only allow those sections of the space-time phase bundle
2786: which arise from solutions of the vacuum Einstein equations with
2787: cosmological constant $\Lambda$ --- in particular the general
2788: relativistic constraint equations with cosmological constant
2789: $\Lambda$ have to be satisfied by the fields $
2790: (\KA^{\lambda}_{\mu\nu},\Kp^{\alpha\beta}) $.
2791: \item[${\mycal C} 2$.] Next, the $e_a$-tetrad
2792: components of $g$ are required to be bounded on $\hypext$.
2793: Moreover, we impose the integral condition
2794: \begin{equation}\int_\hypext r\left(\sum_{a,b,c}
2795: |\znabla_ae^{bc}|^2 + \sum_{d,e} |e^{de}|^2\right) \; d\mu_b <
2796: \infty\;, \label{P1}\end{equation} where $e^{ab}$ are the
2797: $e_a$-tetrad components of $g-b$. In \eq{P1} $d\mu_b$ is, as
2798: before, the measure arising from the metric induced on $\hyp$ by
2799: the background metric $b$; in local coordinates such that
2800: $\hypext=\{t=0\}$ we have $d\mu_b=\sqrt{\det b_{ij}}drd^{n-1}v$,
2801: with the indices $i,j$ running from $1$ to $n$.
2802: \item[${\mycal C} 3$.] Further,
2803: the fall-off conditions
2804: \begin{eqnarray}
2805: e^{ab}= o(r^{-n/2})\;, \quad e_c(e^{ab})= o(r^{-n/2})\;.
2806: \label{P1-}
2807: \end{eqnarray}
2808: are assumed to hold on $\hypext$.
2809: \item[${\mycal C} 4$.]Finally, we shall assume that the
2810: following ``volume normalization condition'' is satisfied:
2811: \begin{equation}\int_\hypext r|b_{cd}e^{cd}| \; d\mu_b <
2812: \infty\;. \label{P1+}\end{equation}
2813: \end{enumerate}
2814: (Recall that when $M$ is not parallelizable, then
2815: conditions \eq{P1}, \eq{P1-}, {\em etc.}, should be understood as
2816: the requirement of { existence of a covering of $M$ by a finite
2817: number of open sets ${\mycal O}_i$ together
2818: with frames defined on $[R_0,\infty)\times {\mycal O}_i$
2819: satisfying the relevant conditions.})
2820:
2821: Whenever we consider variations
2822: $(\delta\KA^{\lambda}_{\mu\nu},\delta\Kp^{\alpha\beta}) $ of the
2823: fields in $\cP_b$, we will require that those variations satisfy
2824: the same decay conditions as the fields in $\cP_b$.
2825:
2826: From now on we shall assume that $B^\alpha_{\beta\gamma}$ is the
2827: Levi-Civita connection of the background metric $b$. A condition
2828: equivalent to \eq{P1}, and slightly more convenient to work with,
2829: is
2830: \begin{equation}\int_\hypext r \left(\sum_{a,b,c}
2831: |e_a(e^{bc})|^2 + \sum_{d,e} |e^{de}|^2\right) \; d\mu_b <
2832: \infty\;. \label{P1equiv}\end{equation}
2833: This follows immediately from \Eqsone{eq:b3} and \eq{eq:b4}, which
2834: show that the $\znabla$-connection coefficients are bounded in the
2835: frame $e_a$. It follows that the fall-off conditions
2836: \eq{C4}-\eq{C5} will ensure that ${\mycal C} 2$--${\mycal C} 4$
2837: hold.
2838:
2839: Let us show that
2840: \Eq{P1} guarantees that the integral defining $\Omega_\hyp$
2841: converges. In order to see that, consider the identity:
2842: \begin{eqnarray*}\znabla_{\alpha}e^{\mu\nu} & = &
2843: \znabla_{\alpha}g^{\mu\nu} =
2844: (\znabla_{\alpha}-\nabla_{\alpha})g^{\mu\nu}
2845: \\ & = & -g^{\mu\sigma}(\Gamma^\nu_{\sigma\alpha} -
2846: B^\nu_{\sigma\alpha})-g^{\nu\sigma}(\Gamma^\mu_{\sigma\alpha} -
2847: B^\mu_{\sigma\alpha}) \;.
2848: \end{eqnarray*}
2849: The usual cyclic permutations calculation allows one to express
2850: $\Gamma^\alpha_{\beta\gamma} - B^\alpha_{\beta\gamma}$ as a linear
2851: combination of the $\znabla_{\alpha}e^{\mu\nu} $'s. It then
2852: follows from \Eq{Pend} that the tetrad coefficients $p^a_{bc}$ of
2853: $p^\alpha_{\beta\gamma}$ are, on $\hypext$, linear combinations
2854: with bounded coefficients of the $\znabla_ae^{bc}$'s. In local
2855: coordinates on $\hypext$ we have
2856: $$\sqrt{|\det b_{\mu\nu}|}\sim r \sqrt{\det b_{ij}}\;,$$
2857: hence
2858: $$\int_{\hypext} |\delta p^t_{ab}|\;|\delta
2859: \Kp^{ab}| dr d^{n-1}v \le C \sum_{a,b,c,d,e}\int_{\hypext} r
2860: |\znabla_c\delta e^{de}|\;|\delta e^{ab}| d\mu_b <\infty\;.$$ Here
2861: the coordinate $x^0\equiv t$ has been chosen so that
2862: $\hypext=\{t=0\}$. Thus, $\Omega_\hyp$ is well defined on $\cP_b$,
2863: as desired.
2864:
2865: Recall, now, that $\Omega_\hyp$ coincides up to boundary terms
2866: with the more familiar ``ADM symplectic form'' \cite{Kij1,Kij2}:
2867: one sets
2868: \begin{eqnarray} \label{Pkl}
2869: P^{kl} & := & \sqrt{\det g_{mn}} \ (\gthreeup^{ij}K_{ij}
2870: \gthreeup^{kl} - K^{kl} )\;,
2871: \end{eqnarray}
2872: where $K_{kl}$ is the extrinsic curvature of $\hyp$,
2873: \begin{eqnarray} \label{Kl} K_{kl} & := & - \frac
2874: 1{\sqrt{|g^{tt}|}} {\Gamma}^t_{kl}
2875: %= - \frac 1{\sqrt{|g^{tt}|}} A^t_{kl}
2876: \;,
2877: \end{eqnarray}
2878: with $\gthreeup^{kl}$ --- the three-dimensional inverse of the
2879: induced metric $g_{kl}$ on $\hyp$; the indices on $K^{kl}$ have
2880: been raised using $\gthreeup^{kl}$. If we further choose the
2881: coordinate $x^3$ in such a way that
2882: $\partial\hypext=\{t=0,x^3=1\}$, then the ``symplectic'' form
2883: (\ref{N.3g}) can be rewritten as~\cite{Kij1,Kij2}
2884: \begin{eqnarray}\nonumber
2885: \lefteqn{\Omega_{\hyp}((\delta_1 p^{\lambda}_{\mu\nu},\delta_1
2886: \Kp^{\alpha\beta}),(\delta_2 p^{\lambda}_{\mu\nu},\delta_2
2887: \Kp^{\alpha\beta})) = \frac 1{16 \pi} \int_\hyp \left( \delta_1
2888: g_{kl} \delta_2
2889: P^{kl} -\delta_2 g_{kl} \delta_1 P^{kl} \right) d^{n}x} \\ &
2890: \phantom{xxxxxxxxx} + \displaystyle\frac 1{16 \pi} \int_{\partial
2891: \hyp} \left(\delta_1 N^3 \delta_2 \frac{\sqrt{\det g_{kl}}}{N} -
2892: \delta_2 N^3 \delta_1 \frac{\sqrt{\det g_{kl}}}{N} \right)
2893: d^{n-1}v\;, \label{nN.3.1jj}\end{eqnarray} where
2894: $$N=1/\sqrt{-g^{tt}}\;,\quad N_k=g_{tk}\;,\quad
2895: N^3=(g^{3k}-\frac{g^{t3}g^{tk}}{g^{tt}})N_k\;.$$ Let us show that
2896: $\Omega_\hyp$ actually coincides with the ADM ``symplectic form'' on
2897: $\cP_b$. It clearly follows from \eq{P1} (with
2898: $e^{ab}$ replaced by $\delta e^{ab}$) that the volume integral there
2899: converges as before; it remains to show that the boundary integral
2900: vanishes.
2901: We have
2902: \begin{eqnarray*}
2903: \delta \left( \frac {\sqrt{\det g_{ij}}} {N} \right) & = & \delta
2904: \left(
2905: \sqrt{|\det g_{\mu\nu}|}\right)
2906: \\
2907: & = & \delta \left( \sqrt{\frac {\det g_{\mu\nu}}{\det
2908: b_{\mu\nu}}}\right) \sqrt{|{\det b_{\mu\nu}} |}
2909: \\
2910: &=& o(r^{-n/2})O(r^{n-1})= o(r^{n/2-1})\;.
2911: \end{eqnarray*}
2912: One easily checks the identity
2913: $$ N^3= \frac {f_0(r)}{\sqrt{|g^{tt}|}} \;,$$
2914: where $f_0$ is the future directed $g$-unit-normal to $\hyp$. We
2915: have
2916: \begin{eqnarray*}
2917: g^{tt}\equiv g(dt,dt) &=& (\eta^{ab}+e^{ab})e_a(t)e_b(t)
2918: \\ &=& (\eta^{00}+e^{00})\left(e_0(t)\right)^2
2919: \\
2920: &=& (\eta^{00}+e^{00})|b^{tt}|\;,
2921: \end{eqnarray*}
2922: which gives
2923: $$\frac 1 {\sqrt{|g^{tt}|}}=O(r)\;,$$
2924: $$ \delta\left(
2925: \frac 1 {\sqrt{|g^{tt}|}} \right) = \delta\left(
2926: \sqrt{\frac {b^{tt}}{g^{tt}}}\right)\frac 1{\sqrt{|b^{tt}|}} =
2927: o(r^{-n/2+1})\;.$$ Further, $$ f_0(r) = f_0{}^be_b(r) =
2928: f_0{}^1e_1(r)= o(r^{-n/2+1})\;,$$ $$\delta f_0(r) =\delta
2929: f_0{}^1e_1(r)= o(r^{-n/2+1})\;,$$ where $e_a$ is a $b$-orthonormal
2930: frame as in \Eq{frame}, and the vanishing of the boundary term in
2931: \eq{nN.3.1jj} readily follows.
2932:
2933: According to \cite{KijowskiTulczyjew} (see also
2934: \cite{CJK,KijowskiGRG}) the Hamiltonian associated with a one
2935: parameter family of maps of the phase space into itself which
2936: arise from the flow of a vector field $X$ on the space-time
2937: equals \be\label{N.5g} H(X,{\hyp}\przecinekJped )=\int_{\hyp}
2938: (\KA^{\mu}_{\alpha\beta}\lie_X \Kp^{\alpha\beta} -X^\mu L)\rd
2939: S_\mu \, \ee {\em provided that all the integrals involved are
2940: well defined, and that the boundary integral in the variational
2941: formula
2942: \begin{eqnarray}\nonumber
2943: -\delta H
2944: & = & \int_{\hyp}\left(
2945: \cLX\KA^{\lambda}{_{\mu\nu}}\delta\Kp^{\mu\nu}-\cLX\Kp^{\mu\nu}
2946: \delta\KA^{\lambda}{_{\mu\nu}}\right) \rd S_\lambda
2947: \phantom{xxxxxxxxxxx} \\ & & +
2948: \int_{\partial\hyp} X^{[\mu}{ \KA^{\nu ]}{_{\alpha\beta}} } \delta
2949: \Kp^{\alpha\beta} \rdS_{\mu\nu} \;,
2950: \label{Hvar1g}
2951: \end{eqnarray}
2952: vanishes.} In the case when $\Bgamma$ is the metric connection of
2953: a given background metric $\bmetric_{\mu\nu}$, and when $X$ is a
2954: Killing vector field of $\bmetric_{\mu\nu}$, the identification
2955: \begin{eqnarray}
2956: m(\hyp,g,b, X)&= &H(X,{\hyp}\przecinekJped )\;,
2957: \label{toto2}\end{eqnarray} together with the calculations in
2958: \cite{ChAIHP} leads to \Eqs{toto}{Freud2.0}. More precisely, let
2959: $E^{\nu\lambda}$ be given by the formula \cite{ChAIHP}
2960: \begin{eqnarray}
2961: E^{\nu\lambda}&= & \displaystyle{\frac{2|\det
2962: \bmetric_{\mu\nu}|}{ 16\pi\sqrt{|\det g_{\rho\sigma}|}}}
2963: g_{\beta\gamma}(e^2 g^{\gamma[\nu}g^{\lambda]\kappa})_{;\kappa}
2964: X^\beta \nonumber \\ && + \frac 1{8\pi} \sqrt{|\det
2965: g_{\rho\sigma}|}~g^{\alpha[\nu}\delta^{\lambda]}_\beta
2966: {X^\beta}_{;\alpha}
2967: \;.\label{Freud2.01} %\nonumber
2968: \\
2969: \label{mas2.1}
2970: e&=& {\sqrt{|\det
2971: g_{\rho\sigma}|}}/{\sqrt{|\det\bmetric_{\mu\nu}|}}\; .
2972: \end{eqnarray}
2973: It can be checked that all the formulae of
2974: \cite[Appendix~B]{ChAIHP} are dimension independent, and lead to
2975: the identity \be\label{N.5g1} E^\lambda :=
2976: \KA^{\lambda}_{\alpha\beta}\lie_X \Kp^{\alpha\beta} -X^\lambda L =
2977: E^{\nu\lambda}{}_{;\nu} +T^\lambda{}_\kappa X^\kappa\;, \ee where
2978: the matter energy-momentum tensor has been defined in \eq{N.5g2}.
2979: Now, when $b$ is the anti de Sitter metric, the integral of
2980: $E^\lambda dS_\lambda$ over large ``balls'' $B_R:=\{r\le R\}$
2981: within $\hyp$ would diverge if we tried to pass with the radius of
2982: those balls to infinity because we have $$ E^\lambda\Big|_{g=b}=
2983: -(\zRs-2\Lambda)X^\lambda/16\pi\;,$$ with $\zRs$
2984: --- the Ricci scalar of the background metric $b$, and
2985: $\zRs-2\Lambda=4 \Lambda/(n-1)$ in an $(n+1)$-dimensional
2986: space-time. We therefore add to $E^\lambda$ a $g$-independent term
2987: which will cancel this divergence: indeed, such terms can be
2988: freely added to the Hamiltonian because they do not affect the
2989: variational formula that defines a Hamiltonian. {}From an energy
2990: point of view such an addition corresponds to a choice of the zero
2991: point of the energy. We thus set $$ \ourU^{\alpha\beta}:=
2992: E^{\alpha\beta}-E^{\alpha\beta}\Big|_{g=b} \;.$$ From the
2993: definition of $E^\lambda$ and from \Eq{N.5g1} one easily finds
2994: \begin{eqnarray}
2995: \nonumber
2996: 16\pi\znabla_\beta \ourU^{\alpha\beta} &= & \left( \sqrt{|\det g|}
2997: g^{ab} - \sqrt{|\det b|} b^{ab}\right) \zRm_{ab} X^\beta
2998: \\ & & +2 \Lambda \left(\sqrt{|\det b| }- \sqrt{|\det g| }\right)X^\beta
2999: \nonumber \\ & & +
3000: \left(\mathring{T}^\lambda{}_\kappa-T^\lambda{}_\kappa\right)X^\kappa
3001: \nonumber \\ & & + \sqrt{|\det b|} \;\left( Q^\alpha{}_{\beta}
3002: X^\beta + Q^\alpha{}_{\beta \gamma}
3003: \znabla^\beta X^\gamma\right)\;, \label{C3}\end{eqnarray} where
3004: $Q^\alpha{}_{\beta}$ is a quadratic form in $e_a(e^{bc})$, and
3005: $Q^\alpha{}_{\beta \gamma}$ is bilinear in $e_a(e^{bc})$ and
3006: $e^{ab}$, both with bounded coefficients. Further,
3007: $\mathring{T}^\lambda{}_\kappa$ is defined as in \Eq{N.5g2} with
3008: $g$ replaced by $b$.
3009:
3010: From now on we assume that both $g$ and $b$ are Einstein, and we
3011: only consider vector fields $X$ which are $b$-Killing vector
3012: fields and satisfy \be \label{A1.}|X|+ |\zn X| \le C r \ee
3013: for some constant $C$; this holds for
3014: all the backgrounds considered in Appendix~\ref{Aiso}, in
3015: particular for the generalized Kottler metrics \eq{Kottler2}.
3016: Theorem~\ref{T0} then shows that the integral defining $H$
3017: converges for fields in $\cP_b$.
3018:
3019: Suppose, further, that the $b$-Killing vector field $X$ has the
3020: property that {\em the associated variations of the fields are
3021: compatible with the boundary conditions} imposed on fields in
3022: $\cP_b$. This means in particular that we must have
3023: \begin{equation}\int_\hyp r\sum_{a,b,c}
3024: | \lie_X \left(\znabla_a e^{bc}\right)|^2 \;
3025: d\mu_b < \infty\;.
3026: \label{eqP2}\end{equation}
3027: Clearly the volume
3028: integral in the variational formula \eq{Hvar1g} converges under
3029: \eq{eqP2} together with the remaining conditions set forth above.
3030: Further, the boundary integral there vanishes under \eq{P1-}, so
3031: that \Eq{N.5g} does indeed provide the required Hamiltonian on
3032: $\cP_b$.
3033:
3034: For Killing vectors satisfying \eq{A1.} \Eq{eqP2} will hold if
3035: \begin{equation}\int_\hyp r^3\sum_{a,b,c,d}
3036: | \left(\znabla_a \znabla_d e^{bc}\right)|^2 \; d\mu_b < \infty\;,
3037: \label{eqP2.1}\end{equation} but we emphasize that the weaker
3038: condition \eq{eqP2} suffices.
3039:
3040: \section{Isometries and Killing vectors of the background}\label{Aiso}
3041:
3042:
3043: \subsection{$(n+1)$-dimensional anti-de Sitter metrics}
3044: \label{Aai1}
3045:
3046: \newcommand{\Y}{{\mycal Y}}
3047: %{\R^{n,2}}%{{\mycal Y}}
3048:
3049: For $n\ge 2$ consider the $(n+1)$-dimensional anti-de Sitter
3050: space-time $(\cM,b)$, thus $b$ is given by \eq{eq:a1a}
3051: %\begin{equation}
3052: %\label{app-AdS}
3053: %b = -\left( \frac{r^2}{\ell^2} +1\right) dt^2 + \left( \frac{r^2}{\ell^2}
3054: %+1\right)^{-1} dr^2 + r^2 h\;,
3055: %\end{equation}
3056: with $h$ --- the unit round metric on the $(n-1)$-dimensional
3057: sphere ${}^{(n-1)}S$. As elsewhere we set $\hyp = \{t=0\}.$ When
3058: $\mn$ is the two-dimensional sphere, the Killing vectors of $b$
3059: are given in \cite{HT}. For higher dimensional spheres the
3060: $b$-Killing vector fields are easily found by thinking of $b$ as
3061: the metric induced on the covering space of the hyperboloid
3062: \begin{equation}
3063: \label{hyp}
3064: \eta_{(a)(b)} y^{(a)}y^{(b)} =-\ell^2
3065: \end{equation}
3066: in the $(n+2)$-dimensional manifold $\Y$ with the
3067: metric\footnote{$\Y$ can be identified with the universal covering
3068: of the space obtained by removing the set $y^{(0)}=y^{(n+1)}=0$
3069: from $\R^{n+2}$; $\Y$ then inherits the local coordinates
3070: $y^{(a)}$ used in \Eq{locy}. However, in order to understand the
3071: geometry of $\cM$ in a neighbourhood of $\hyp$ it is sufficient
3072: --- and most convenient --- to think of $\Y$ as of $\R^{n+2}$.}
3073: \be\label{locy} \eta_{(a)(b)}dy^{(a)}dy^{(b)} = -(dy^{(0)})^2 +
3074: \sum_{(i)=(1)}^{(n)}(dy^{(i)})^2 -(dy^{(n+1)})^2\;. \ee Throughout
3075: this section the indices $(a),(b)$, \emph{etc.,\/} run from $ (0)$
3076: to $(n+1)$. The hyperboloid can be locally parameterized by
3077: coordinates $t$, $x^i$ implicitly defined by the
3078: equations\footnote{The spherical coordinates associated to the
3079: ``cartesian'' coordinates $x^i$ give the form (\ref{eq:a1a}) of the
3080: metric $b$.}
3081: \begin{equation}
3082: \label{transf1} y^{(0)} = \ell\cos (t/\ell)\sqrt{1+r^2/\ell^2} \;,
3083: \end{equation}
3084: \begin{equation}
3085: \label{transf2} y^{(n+1)} = \ell \sin (t/\ell)\sqrt{1+r^2/\ell^2}
3086: \;,
3087: \end{equation}
3088: \begin{equation}
3089: \label{transf3} y^{(i)} =x^i\;,
3090: \end{equation}
3091: with $r^2 = \sum_{i=1}^n (x^i)^2$, where $x^i=r n^i$, and
3092: $n^i\in{}^{(n-1)}S$ can eventually be expressed in terms of
3093: coordinates on the sphere ${}^{(n-1)}S$. For example, for $n=3$ we
3094: can use $x^1=r \sin(\theta)\cos(\varphi)$, $x^2 =r \sin(\theta)
3095: \sin(\varphi)$, $x^3= r \cos(\theta)$, with $\theta,\varphi$ ---
3096: the usual spherical coordinates. It is also convenient to
3097: represent the hypersurface $\hyp \subset \cM$ given by $\hyp
3098: =\{t=0\}$ as $\{\eta_{(a)(b)} y^{(a)}y^{(b)} =-\ell^2\}\cap
3099: \{y^{(n+1)} =0, y^{(0)}
3100: >0\} \subset \Y$. We set
3101: \begin{equation}
3102: \label{Lor2}
3103: L_{(a)(b)} = y_{(a)} \frac{\partial~~}{\partial y^{(b)}} -y_{(b)}
3104: \frac{\partial~~}{\partial y^{(a)}}\;,
3105: \end{equation}
3106: where $y_{(a)}=\eta_{(a)(b)} y^{(b)}$. The $L_{(a)(b)}$'s are
3107: Killing vector fields of $(\Y,\eta_{(a)(b)})$. Further they are
3108: tangent to the hyperboloid $\{\eta_{(a)(b)} y^{(a)}y^{(b)}
3109: =-\ell^2\}$ and hence define, by restriction, Killing vector
3110: fields of the hyperboloid with the induced metric. In fact they
3111: span the space of all the Killing vectors of $b$ because there is
3112: the right maximal number of them. From the coordinate
3113: transformation (\ref{transf1})-(\ref{transf3}) one can compute the
3114: corresponding Killing vectors of anti de Sitter space-time in the
3115: coordinates $\{t,x^i\}$, obtaining
3116: $$
3117: L_{(n+1)(0)} = \ell\frac{\partial }{\partial t}\;,
3118: $$
3119: $$
3120: L_{(i)(n+1)} = \frac{x^i}{\sqrt{1+r^2/\ell^2}} \cos(t/\ell)
3121: \frac{\partial }{\partial t} + \ell\sqrt{1+r^2/\ell^2}
3122: \sin(t/\ell)
3123: %(\delta_{i}^j+n^in^j)
3124: \frac{\partial}{\partial x^i}\;,
3125: $$
3126: $$
3127: L_{(i)(0)} = -\frac{x^i}{\sqrt{1+r^2/\ell^2}} \sin(t/\ell)
3128: \frac{\partial }{\partial t} +\ell\sqrt{1+r^2/\ell^2} \cos(t/\ell)
3129: %(\delta_{i}^j+n^in^j)
3130: \frac{\partial}{\partial x^i}\;,
3131: $$
3132: $$
3133: L_{(i)(j)} = x^i \frac{\partial}{\partial x^j} - x^j
3134: \frac{\partial}{\partial x^i}\;.
3135: $$
3136: Let $\cKS$ be the set of Killing vector fields of $b$ which are
3137: orthogonal to $\hyp$; from the expressions above it is not
3138: difficult to check that a vector basis for $\cKS$ is given by
3139: $\frg_{(\mu)} = L_{(n+1)(\mu)}|_{t=0}$, where $(\mu)$ runs from
3140: $(0)$ to $ (n)$.
3141:
3142: \begin{Proposition}
3143: \label{app-lemma1} Let $\Phi:\cM \to \cM$ be an isometry of $b$
3144: such that $\Phi(\hyp)=\hyp$. Then there exists a Lorentz
3145: transformation matrix $\Lambda^{(\nu)}{}_{(\mu)}$ such that the
3146: basis vectors of $\cKS$ satisfy
3147: $$\Phi_{*}\frg_{(\mu)} = \Lambda^{(\nu)}{}_{(\mu)} \frg_{(\nu)}\;.$$
3148: \end{Proposition}
3149: \remark We note that the property $\Phi_*(\cKS) =\cKS$ follows
3150: from the fact that $\Phi$ preserves $\hyp$, which implies that
3151: $\Phi$ maps the field of unit normals to $\hyp$ into itself.
3152:
3153: \medskip
3154:
3155: \proof As is well known, for every isometry $\Phi:\cM \to \cM$ of
3156: $b$ there exists a diffeomorphism $\hat{\Phi}:\Y \to \Y$, isometry
3157: of $\eta_{(a)(b)}$, such that $\Phi$ is the restriction of
3158: $\hat{\Phi}$ to the hyperboloid \eq{hyp}. In coordinates we have
3159: \begin{equation}
3160: \label{Lor0}\hat{\Phi}^{(a)}(y) =
3161: \Lambda^{(a)}{}_{(b)}y^{(b)}\;,\end{equation}
3162: where $\Lambda^{(a)}{}_{(b)}$ is a matrix satisfying
3163: \begin{equation}
3164: \label{Lor}
3165: \Lambda^{(c)}{}_{(a)} \Lambda^{(d)}{}_{(b)} \eta_{(c)(d)} =
3166: \eta_{(a)(b)}\;.
3167: \end{equation}
3168: The hypersurface $\hyp \subset \Y$ is given by
3169: $\eta_{(a)(b)}y^{(a)}y^{(b)} =-1$ and $y^{(n+1)} =0$ together with
3170: the condition $y^{(0)} >0$, so that the condition
3171: $\hat{\Phi}(\hyp) =\hyp$ implies
3172: $$
3173: \Lambda^{(a)}{}_{(b)} = \left[
3174: \begin{array}{cc}
3175: \Lambda^{(\mu)}{}_{(\nu)} & 0 \\
3176: 0 & \pm 1
3177: \end{array}
3178: \right]\;,
3179: $$
3180: where we split the indices as $(a)=(\mu), (n+1)$.
3181: Equation~\eq{Lor} shows that $\Lambda^{(\mu)}{}_{(\nu)}$ is a
3182: $n+1$-dimensional Lorentz transformation,
3183: $(\Lambda^{(\mu)}{}_{(\nu)})\in O(1,n)$. Equations~\eq{Lor0} and
3184: \eq{Lor2} imply that under push-forward by $\hat{\Phi}$ the
3185: Killing vectors of $\eta_{(a)(b)}$ transform as
3186: $$
3187: \hat{\Phi}_{*}L_{(a)(b)} = \Lambda^{(c)}{}_{(a)}
3188: \Lambda^{(d)}{}_{(b)} L_{(c)(d)}
3189: $$
3190: in particular, the basis vectors of $\cKS$ transform as
3191: \begin{eqnarray*}
3192: \hat{\Phi}_{*}\frg_{(\mu)} &=&
3193: \hat{\Phi}_{*}L_{(n+1)(\mu)}\Big |_{\hat{\Phi}(y^{(n+1)}) =0} \\
3194: &=& \Lambda^{(c)}{}_{(n+1)} \Lambda^{(d)}{}_{(\mu)}
3195: L_{(c)(d)}\Big |_{y^{(n+1)} =0} \\
3196: %&=& \Lambda^{(d)}{}_{(\mu)} L_{(n+1)(d)}\Big |_{y^{(n+1)} =0} \\
3197: %&=& \Lambda^{(\nu)}{}_{(\mu)} L_{(n+1)(\nu)}\Big |_{y^{(n+1)} =0} \\
3198: %\hat{\Phi}_{*}\frg_{(\mu)}
3199: &=& \pm \Lambda^{(\nu)}{}_{(\mu)} \frg_{(\nu)}.
3200: \end{eqnarray*}
3201: Replacing $\Lambda^{(\nu)}{}_{(\mu)}$ by
3202: $-\Lambda^{(\nu)}{}_{(\mu)}$ if necessary, the result follows.
3203: \qed
3204:
3205: Let $\cKt$ be the space of $b$- Killing vectors spanned by the
3206: $L_{(\mu)(\nu)}$'s, thus $\cKt$ contains all the $L_{(a)(b)}$'s
3207: which are not in $\cKS$. An identical calculation as in the proof
3208: above shows that under isometries of $b$ preserving $\hyp$ we have
3209: \begin{eqnarray} \label{remK}
3210: \hat{\Phi}_{*}L_{(\mu)(\nu)} &=&
3211: \Lambda^{(\sigma)}{}_{\mu} \Lambda^{(\rho)}{}_{(\nu)}
3212: L_{(\sigma)(\rho)}\;.
3213: \end{eqnarray}
3214: It follows %that $\cKt$ is mapped into itself by such isometries, and
3215: that the resulting representation of the Lorentz group on $\cKt$
3216: is equivalent to a representation on two-contravariant
3217: anti-symmetric tensors.
3218:
3219: %$\Phi(\cK)=\cK$, where $\cK$ the set of killing vectors of $b$
3220: % which are tangent to the hypersurface $\hyp$.
3221:
3222:
3223:
3224:
3225:
3226: \subsection{$h$'s with a non-positive Ricci tensor} \label{Aai2}
3227: We consider metrics \eq{A1}, as in Section~\ref{S2}. In what
3228: follows we shall only consider $(M,h)$'s with a non-positive Ricci
3229: curvature, with $n\ge 3$, the case $n=2$ being covered by the
3230: previous section. We shall further assume that the scalar
3231: curvature $R_h$ of $h$ (the Ricci scalar) is a constant. We note
3232: that the vector fields
3233: \begin{equation}
3234: \label{A5}
3235: X=X^0 n= {\lambda\over a}n = \lambda \partial_t \;,\qquad \lambda \in \R\;,
3236: \end{equation}
3237: where $n=e_0$ is the field of future pointing unit normals to the
3238: hypersurfaces $\{t=\mathit{const}\}$, are Killing vector fields
3239: for the metric $b$ whatever $a=a(r)$. The non-vanishing connection
3240: coefficients, $$\zo^{a}{}_{bc}\equiv \theta^a(\zn_{e_c}e_b)\;,$$
3241: with respect to this frame are
3242: \begin{equation}
3243: \label{A2}
3244: \zo^A{}_{1B}= \frac 1 {ra(r)} \delta ^A_B= -\zo^1{}_{AB}\;,\quad
3245: \zo^A{}_{BC}= \frac 1r \beta^A{}_{BC}\;, \quad \zo_{100}=-\zo_{010}
3246: = -{a'(r)\over a^2(r)}\;,
3247: \end{equation}
3248: where the $\beta^A{}_{BC}$ are the
3249: Levi-Civita connection coefficients of $h$ with respect to the frame
3250: $\alpha^A$.
3251: The $AB$ components of the Killing equations read
3252: \begin{equation}
3253: \label{eq:A1}
3254: \caD_A X_B + \caD_B X_A + {1\over a(r)}h_{AB}X^1 =0\;,
3255: \end{equation}
3256: where $\caD$ is the covariant derivative operator associated with
3257: the metric $h$, which shows that $X^B\partial_B$ is a conformal
3258: Killing vector field on $M$. Uniqueness of solutions of the
3259: volume-normalized Yamabe equation in the case under consideration
3260: implies that conformal Killing vector fields of $(\mn,h)$ are
3261: necessarily Killing vectors, hence
3262: $$X^1\equiv 0\;.$$
3263: The $00$, $01$ and $0A$ components of the Killing equations read
3264: \begin{eqnarray}
3265: \label{eq:A2}
3266: e_0(X_0)& = & 0\;,
3267: \\ \label{eq:A3}
3268: e_1(X_0)+ {a'\over a^2} X_0& = & 0\;,
3269: \\ \label{eq:A4}
3270: e_A(X_0)+e_0(X_A)& = & 0\;.
3271: \end{eqnarray}
3272: Equation~\eq{eq:A2} shows that $X_0$ is $t$-independent.
3273:
3274: Suppose, first, that the Ricci tensor of $h$ is strictly negative.
3275: It is well known\footnote{\label{Kf}The Killing equations imply
3276: $X_B\caD_C\caD^C X^B =
3277: -R_{AB} X^AX^B$, where $R_{AB}$ is the Ricci tensor of $h$;
3278: integration of this equation over $\mn$ shows that $X^A$ is
3279: covariantly constant when $R_{AB}$ is non-positive, and vanishes
3280: when $R_{AB}$ is strictly negative.} that in this case $(\mn,h)$
3281: has no non-trivial Killing vector fields so that $X^A\equiv 0$,
3282: and Equation~\eq{eq:A4} shows that $X_0$ is $v^A$-independent.
3283: Integrating \eq{eq:A3} yields then the one parameter family of
3284: Killing vector fields \eq{A5}, which shows that the algebra of all
3285: Killing vector fields of $b$ is one-dimensional.
3286:
3287: Suppose, next, that $(M,h)$ is an $(n-1)$-dimensional flat torus
3288: $(\T^{n-1},\delta)$. Then the $X^A$'s are covariantly
3289: constant$^{\mathrm{\ref{Kf}}}$ vector fields on $\T^{n-1}$, which
3290: shows that $X^A=X^A(t,r)$ in coordinates $v^A$ in which the metric
3291: $\delta$ has constant entries. Integrating \eq{eq:A4} over
3292: $\T^{n-1}$ gives
3293: $$0= \int_{\T^{n-1}} e_A(X_0) = -e_0(X_A)\;\mathrm{Vol}(\T^{n-1})\;,$$
3294: hence $X_A=X_A(r)$. Equation~\eq{eq:A4} implies then that $X_0$ is
3295: $v^A$-independent, so that $X_0=X_0(r)$, from \eq{eq:A3} we
3296: recover \eq{A5}, and $\cKS$ is again one-dimensional, as claimed.
3297: We note that the $1A$ component of the Killing equation implies
3298: that the $X^A$'s are in fact $r$-independent, which gives a
3299: complete description of the set of Killing vector fields occurring
3300: in this case.
3301:
3302: The above arguments extend to all manifolds with constant scalar
3303: curvature and non-positive Ricci curvature, as follows: suppose
3304: that $(\mn,h)$ has non-trivial Killing vector fields. The $1A$
3305: component of the Killing equations gives
3306: $$e_1(X_A) - {1\over ra} X_A = 0 \quad \Longrightarrow \quad X^A =
3307: X^A(t,v^A)\;.$$
3308: Integration of Equation~\eq{eq:A3} gives
3309: \begin{equation}
3310: \label{A6}
3311: X^0 = {\lambda(v^A)\over a(r)}\;,
3312: \end{equation}
3313: for some function $\lambda$ on $\mn$. Equation~\eq{A5} inserted
3314: into \eq{eq:A4} gives
3315: $$e_A(\lambda)=-a^2(r)\partial_t X^A\;,$$
3316: which is compatible with \Eq{A5} only if $\partial_t X^A=0$, hence
3317: $e_A(\lambda)=0$. Summarizing, we have proved
3318:
3319: \begin{Proposition}
3320: \label{PA1} If $(\mn,h)$ has non-positive Ricci curvature and
3321: constant scalar curvature, then all Killing vector fields of the
3322: metric $b$ given by \Eq{A1} are of the form
3323: $$X={\lambda \over a} n + X^A(v^B)\partial _A\;, \qquad\lambda \in
3324: \R\;,$$ where $X^A(v^B)\partial _A$ is a Killing vector field of
3325: the metric $h$.
3326: \end{Proposition}
3327:
3328: \section{Equality of the Hamiltonian mass with the Abbott-Deser one}
3329: \label{AAD}
3330:
3331: In this appendix we consider a subset $\R\times \Sigma_{\ext}$ of
3332: a four dimensional space-time $(\cM,g)$ defined by
3333: a coordinate system $\{x^{\alpha}\}$; we identify
3334: $\Sigma_{\ext}$ with the set $ \{x^{\alpha} : x^0 =0\}$. The space
3335: coordinates $(x^i)$ on $\Sigma_{\ext}$ will be written as $(r,v^A)$,
3336: with the range of $r$ being $[R_0,\infty)$, and with the $v^A$ being
3337: local coordinates on some compact two dimensional manifold. Assume that
3338: there exists a frame $\{ e_a\}_{a=0}^3$ defined on
3339: $\Sigma_{\ext}$, which defines a background metric
3340: $b^{\alpha\beta} = \eta^{ab} e_a^{\alpha}e_b^{\beta}$, where
3341: $\eta^{ab} = \mbox{diag}(-1,1,1,1)$. In other words, the tetrad
3342: $\{e_a\}$ is orthonormal with respect to $b_{\alpha\beta}$. Assume
3343: that in this frame the space-time metric $g$ has the form
3344: $$g_{ab} = \eta_{ab} + e_{ab}\;,$$
3345: where $$e_{ab} = o(1/r^{\alpha})\:,\qquad e_a(e_{bc}) =
3346: o(1/r^{\alpha}) %\;,\qquad \alpha >3/2
3347: \;,$$ for some $\alpha>0$. The Abbott-Deser mass $M_{AB}$
3348: associated with $X$ is defined as \cite{AbbottDeser}
3349: \begin{equation}
3350: \label{m-ab} M_{AB} = \frac{1}{2} \lim_{R \to \infty}
3351: \int_{\partial \Sigma_R} V^{\alpha\beta} dS_{\alpha\beta}\;,
3352: \end{equation}
3353: where
3354: \begin{equation}
3355: \label{m-ab1} V^{\alpha\beta}(h)= \frac{1}{8\pi}\, b \left(
3356: K^{\alpha\beta\sigma\kappa}{}_{;\kappa} X_{\sigma} -
3357: K^{\alpha\kappa\sigma\beta} X_{\sigma;\kappa} \right) \;, \ee with
3358: $$K^{\alpha\beta\sigma\kappa} =
3359: b^{\alpha[\kappa}H^{\sigma]\beta} +
3360: H^{\alpha[\kappa}b^{\sigma]\beta}\;,\quad H^{\alpha\beta} =
3361: e^{\alpha\beta} -\frac{1}{2} e_{\gamma}{}^{\gamma}
3362: b^{\alpha\beta}\;,$$
3363: $$ b= \sqrt{|\det b_{\mu\nu}|}\;.$$
3364: Note that
3365: $\tK^{\alpha\beta\sigma\kappa}$ has the same symmetries as the
3366: Riemann tensor. Let $\ourU^{\alpha\beta}$ be the ``Hamiltonian
3367: superpotential'' defined by \eq{Fsup2new}; assume that
3368: \begin{equation}
3369: \label{m-ab2} (|X| + |\nabla X|)b = O(r^\beta)\;, \ee
3370: we then
3371: claim that
3372: $$\ourU^{\alpha\beta} = \myvareps
3373: \,\tV^{\alpha\beta} + \oepsb\;.$$ In order to establish this,
3374: recall that
3375: \begin{eqnarray*}
3376: \det(g_{\alpha\beta}) &=&
3377: \det\left(b_{\alpha\gamma}\left[\delta^{\gamma}{}_{\beta} +
3378: \myvareps
3379: b^{\gamma\sigma}\th_{\sigma\beta} + \oeps\right]\right) \\
3380: &=& \det(b_{\alpha\gamma}) \det(\delta^{\gamma}{}_{\beta} +
3381: \myvareps e^{\gamma}{}_{\beta})\;,
3382: \end{eqnarray*}
3383: where $e^{\gamma}{}_{\beta} = b^{\gamma\sigma}\th_{\sigma\beta} +
3384: \oeps$. A well known identity gives
3385: $$
3386: \det(g_{\alpha\beta}) = \det(b_{\alpha\beta}) \left(1 + \myvareps
3387: \, \th_{\gamma}{}^{\gamma} + \oeps \right)\;.
3388: $$
3389: Let us write $\ourU^{\alpha\beta} =\ourU^{\alpha\beta}{}_{\gamma}
3390: X^{\gamma} + \widehat{\ourU}^{\alpha\beta}$, where
3391: $$
3392: \ourU^{\alpha\beta}{}_{\gamma} := \frac{1}{8\pi} \frac{b}{e}
3393: \left(e^2 g^{\sigma[\alpha} g^{\beta]\kappa}\right)_{;\kappa}
3394: g_{\gamma\sigma}\;,
3395: $$
3396: $$
3397: \widehat{\ourU}^{\alpha\beta} := \frac{1}{8\pi} \left( \sqrt{|\det
3398: g_{\rho\sigma}|} \; g^{\kappa[\alpha}\delta^{\beta]}{}_{\gamma} -
3399: b \; b^{\kappa[\alpha} \delta^{\beta]}{}_{\gamma} \right)
3400: X^{\gamma}{}_{;\kappa} \;.
3401: $$
3402: We have
3403: $$
3404: e^2 = 1+ \myvareps \,\th_{\alpha}{}^{\alpha} + \oeps \;,$$ so that
3405: the first term above can be written as
3406: \begin{eqnarray*}
3407: {\ourU}^{\alpha\beta}{}_{\gamma} &=& \frac{1}{8\pi} b \left[ (1+
3408: \myvareps \th_{\alpha}{}^{\alpha} + \oeps) \left(b^{\sigma[\alpha}
3409: b^{\beta]\kappa} - \myvareps \,
3410: b^{\sigma[\alpha}\th^{\beta]\kappa} - \myvareps
3411: \,\th^{\sigma[\alpha} b^{\beta]\kappa}
3412: + \oeps\right)\right]_{;\kappa} b_{\gamma\sigma}\\
3413: &=& \frac{b}{8\pi} \,\myvareps \, b_{\gamma\sigma} \left(
3414: \th_{\rho}{}^{\rho} \,b^{\sigma[\alpha} b^{\beta]\kappa} -
3415: b^{\sigma[\alpha} \th^{\beta]\kappa} - \th^{\sigma[\alpha}
3416: b^{\beta]\kappa} \right)_{;\kappa}
3417: + \oepsb \\
3418: &=&- \frac{b}{8\pi}\, \myvareps \,b_{\gamma\sigma} \left(
3419: b^{\sigma[\alpha} \tH^{\beta]\kappa} + \tH^{\sigma[\alpha}
3420: b^{\beta]\kappa} \right)_{;\kappa} + \oepsb
3421: \end{eqnarray*}
3422: so that
3423: $$
3424: {\ourU}^{\alpha\beta}{}_{\gamma} = \frac{b}{8\pi} \, \myvareps\,
3425: b_{\gamma\sigma} \tK^{\alpha\beta\sigma\kappa}{}_{;\kappa} +
3426: \oepsb\;.
3427: $$
3428: Similarly,
3429: \begin{eqnarray*}
3430: \widehat{\ourU}^{\alpha\beta} &:=& \frac{1}{8\pi} \left(
3431: \sqrt{|\det g_{\rho\sigma}|} \;
3432: g^{\kappa[\alpha}\delta^{\beta]}{}_{\gamma} - b \;
3433: b^{\kappa[\alpha} \delta^{\beta]}{}_{\gamma} \right)
3434: X^{\gamma}{}_{;\kappa}\\
3435: &=& \frac{1}{8\pi} \left( \sqrt{|\det g_{\rho\sigma}|} \;
3436: g^{\kappa[\alpha}b^{\beta]\gamma} - b \; b^{\kappa[\alpha}
3437: b^{\beta]\gamma} \right)
3438: X_{\gamma;\kappa} \\
3439: &=& \frac{b}{8\pi} \left[ \left(1+ \frac{1}{2}\, \myvareps \th +
3440: \oeps\right) \left(b^{\kappa[\alpha}b^{\beta]\gamma} - \myvareps
3441: \, \th^{\kappa[\alpha}b^{\beta]\gamma} + \oeps\right)
3442: - b^{\kappa[\alpha}b^{\beta]\gamma} \right] X_{\gamma;\kappa}\\
3443: &=& - \frac{b}{8\pi} \,\myvareps
3444: \tH^{\kappa[\alpha}b^{\beta]\gamma} X_{\gamma;\kappa} + \oepsb\;.
3445: \end{eqnarray*}
3446: Now, $X_{\gamma}$ is a Killing vector of $b_{\mu\nu}$, therefore
3447: \begin{eqnarray*}
3448: \tH^{\kappa[\beta}b^{\alpha]\gamma} X_{\gamma;\kappa}
3449: &=& \tH^{\kappa[\beta}b^{\alpha]\gamma}X_{[\gamma;\kappa]}\\
3450: &=& \frac{1}{2} \left(\tH^{\kappa[\beta}b^{\alpha]\gamma}
3451: - \tH^{\gamma[\beta}b^{\alpha]\kappa} \right) X_{\gamma;\kappa}\\
3452: &=& \frac{1}{2} \tK^{\kappa\gamma\alpha\beta}X_{\gamma;\kappa}\;,
3453: \end{eqnarray*}
3454: so, we have obtained
3455: $$
3456: \widehat{\ourU}^{\alpha\beta} = \frac{b}{16\pi} \, \myvareps
3457: \tK^{\kappa\gamma\alpha\beta}X_{\gamma;\kappa} + \oepsb\;.
3458: $$
3459: The two terms together give
3460: \begin{eqnarray*}
3461: \ourU^{\alpha\beta}&=& \frac{b}{8\pi}\, \myvareps \left(
3462: \tK^{\alpha\beta\sigma\kappa}{}_{;\kappa} X_{\sigma} +
3463: \frac{1}{2}\,\tK^{\kappa\gamma\alpha\beta}X_{\gamma;\kappa}\right)
3464: + \oepsb\\
3465: &=& \frac{b}{8\pi}\, \myvareps \left(
3466: \tK^{\alpha\beta\sigma\kappa}{}_{;\kappa} X_{\sigma} -
3467: \frac{1}{2}\,\tK^{\alpha\beta\gamma\kappa}X_{\gamma;\kappa}\right)
3468: + \oepsb\;.
3469: \end{eqnarray*}
3470: As $\tK^{\alpha\beta\gamma\kappa}$ has the same symmetries as the
3471: Riemann tensor, we have $\tK^{\alpha[\beta\gamma\kappa]} =0$,
3472: which implies that $\frac{1}{2} \, \tK^{\alpha\beta\gamma\kappa} =
3473: \tK^{\alpha[\kappa\gamma]\beta}$, and
3474: \begin{eqnarray*}
3475: \ourU^{\alpha\beta}&=& \frac{b}{8\pi}\, \myvareps \left(
3476: \tK^{\alpha\beta\sigma\kappa}{}_{;\kappa} X_{\sigma} -
3477: \tK^{\alpha[\kappa\gamma]\beta}X_{\gamma;\kappa}\right)
3478: + \oepsb\\
3479: &=& \myvareps \, \tV^{\alpha\beta} + \oepsb\;.
3480: \end{eqnarray*}
3481: If \be\label{abcond} \beta - 2 \alpha \le 0\;,\ee we obtain
3482: equality of the Abbott-Deser mass with the Hamiltonian one; recall
3483: that $\beta=n$ for the anti-de Sitter type metrics considered in
3484: the body of the paper, and \Eq{abcond} reproduces the condition
3485: $\alpha \ge n/2$, identical to that which arises in the proof of
3486: coordinate-invariance of the mass integral.
3487:
3488: Summarizing, we have proved:
3489:
3490: \begin{Proposition}
3491: Suppose that \Eqsone{m-ab1}, \eq{m-ab2} and \eq{abcond} hold. Then
3492: the Hamiltonian mass coincides with the Abbott-Deser one; in
3493: particular, either they are both undefined, or both diverge, or
3494: both converge to the same values.
3495: \end{Proposition}
3496:
3497: \textbf{Acknowledgements:} PTC acknowledges useful discussions
3498: with S.~Ilias, A.~Polombo and A.El Soufi. We appreciated the
3499: friendly hospitality of the Max Planck Institute for Gravitation
3500: in Golm during final stages of work on this paper.
3501:
3502: %%% Local Variables:
3503: %%% mode: latex
3504: %%% TeX-master: "a"
3505: %%% End:
3506:
3507: %\cite{Aros,WaldZoupas,YorkBrown}
3508: \addcontentsline{toc}{1}{\bigskip \noindent {\bf References \hfill}}
3509:
3510: \bibliographystyle{amsplain}
3511: %\bibliographystyle{/usr/share/texmf/tex/revtex/prsty}
3512: \bibliography{%ptjjk,
3513: ../../references/newbiblio,%
3514: ../../references/reffile,%
3515: ../../references/bibl,%
3516: ../../references/Energy,%
3517: ../../references/hip_bib,%
3518: ../../references/netbiblio}
3519: %\texttt{\input{READMEl}}
3520: \end{document}
3521:
3522: %%% Local Variables:
3523: %%% mode: latex
3524: %%% TeX-master: t
3525: %%% End:
3526: