1: \documentclass[aps,prd,showpacs,preprint,floatfix,nofootinbib]{revtex4}
2: \usepackage{amsmath}
3: \usepackage{amstext}
4: \usepackage{amsthm}
5: \usepackage{amssymb}
6: \usepackage{graphicx}
7: %\bibliographystyle{article}
8: %\usepackage{cite}
9:
10: \newcommand{\be}{\begin{equation}}
11: \newcommand{\ee}{\end{equation}}
12: \newcommand{\ba}{\begin{eqnarray}}
13: \newcommand{\ea}{\end{eqnarray}}
14: \newcommand{\bc}{\begin{center}}
15: \newcommand{\ec}{\end{center}}
16: \newcommand{\vs}{\vspace*{3mm}}
17: \newcommand{\dis}{\displaystyle}
18: %\newcommand{\text}{\textstyle}
19: \def\rp{r_{+}}
20: \def\re{r_{\rm EH}}
21:
22: \def\RN{Reis\-sner-Nord\-str\"{o}m }
23: %\documentstyle[aps,eqsecnum,manuscript,epsfig,floats,subeqn]{revtex}
24: \begin{document}
25:
26: \preprint{MZ-TH/06-04}
27: \begin{center}
28:
29: {\LARGE \textsc{\\}}
30: {\LARGE \textsc{\\}}
31: {\LARGE \textsc{Spacetime structure of an evaporating black hole in quantum gravity}}
32:
33: \vspace{1.4cm}
34: {\large A.~Bonanno}\\
35:
36: \vspace{0.7cm}
37: \noindent
38: \textit{INAF - Osservatorio Astrofisico di Catania,
39: Via S.Sofia 78, I-95123 Catania, Italy\\
40: INFN, Via S. Sofia 64, I-95123 Catania, Italy}\\
41:
42: \vspace{0.7cm}
43: {\large M.~Reuter}\\
44:
45: \vspace{0.7cm}
46: \noindent
47: \textit{Institute of Physics, University of Mainz\\
48: Staudingerweg 7, D-55099 Mainz, Germany}\\
49:
50: \end{center}
51: \vspace*{0.6cm}
52: \date{\today}
53:
54: \begin{abstract}
55: The impact of the leading quantum gravity effects on the dynamics of the Hawking evaporation
56: process of a black hole is investigated. Its spacetime structure is described by a renormalization group
57: improved Vaidya metric. Its event horizon, apparent horizon, and timelike limit surface are obtained
58: taking the scale dependence of Newton's constant into account. The emergence of a quantum
59: ergosphere is discussed. The final state of the evaporation process is a cold, Planck size remnant.
60: \end{abstract}
61:
62: \pacs{97.60.Lf, 11.10.Hi, 04.60.-m}
63: \maketitle
64:
65: \renewcommand{\theequation}{1.\arabic{equation}}
66: \setcounter{equation}{0}
67:
68: \section{introduction}
69: One of the very remarkable features of black hole radiance \cite{hawking} is the observation that the
70: global spacetime structure of a black hole losing mass by the evaporation process is far more complicated
71: than that of its static counterpart \cite{ybook,y1,y2}.
72: Even a Schwarzschild black hole, when it radiates, does
73: not have a single horizon that fully characterizes its structure, and
74: %in the more general case of a
75: %non-spherically symmetric radiance
76: one must distinguish at least three important horizon-like loci.
77: %The future event horizon (EH) is the null three-surface which is the locus
78: %of outgoing future-directed null geodesic rays that never manage to reach arbitrarily large distances
79: %from the hole.
80: The future event horizon (EH) is the boundary of the causal past of future null infinity,
81: %$\partial J^{-}({\cal J}^{+})$,
82: and it represents the locus of outgoing future-directed null geodesic
83: rays that never manage to reach arbitrarily large distances from the hole.
84: The apparent horizon (AH) is defined as the outermost marginally trapped surface for the
85: outgoing photons. Classically it can be null or spacelike, in presence of quantum radiance it can be timelike
86: also, when regarded as a 3-dimensional surface. The third important locus is the timelike limit surface
87: (TLS) or ``quasi-static limit" which is defined as the locus where static observers become lightlike.
88: %For black holes that possess in their exterior region an exact timelike
89: %Killing vector $\partial /\partial t$, is provided by the vanishing of the metric component
90: %$g_{tt}=g(\partial_t,\partial_t)$. For almost static black holes, with a small luminosity (or accretion) $L$, one defines
91: %the TLS as the locus where $g_{vv}\equiv g(\partial_v,\partial_v) =0$, where $\partial / \partial v$ is the timelike vector of an
92: %observer at rest at a large distance from the hole and with respect to which the luminosity $L=-dM/dv$ is
93: %measured where $M$ is the mass of the hole.
94: The TLS can be null, timelike, or spacelike \cite{ybook}.
95: For a classical Schwarzschild black hole (which does not radiate), the three surfaces EH, AH, and TLS are all
96: identical. Upon ``switching on" the Hawking evaporation this degeneracy is partially lifted. According to
97: the analysis by York \cite{ybook,y1} the AH continues to coincide with the TLS for a spherically
98: symmetric emission, but the EH is different from AH=TLS.
99:
100: In particular, if we approximate the stress-energy tensor near the horizon as a radial
101: influx of negative energy which balances the outward Hawking flux at infinity,
102: the event horizon is located {\it inside} the AH \cite{is}, the portion of spacetime between the two surfaces
103: forming the so-called ``quantum ergosphere".
104: This name stems from the analogy with the classical (stationary) Kerr black hole for which
105: EH=AH$\not =$TLS. Here the ergosphere is the space between ``the" horizon EH=AH and the TLS,
106: usually called the ``static limit". In both cases particles and light signals can escape from within the
107: ergosphere and reach infinity.
108:
109: The definition of the EH via the locus of outgoing photons
110: that can never reach large distances from the hole has the unfortunate ``teleological"
111: property of requiring knowledge of the entire future history of the
112: hole \cite{ybook, poi}. In particular when the black hole radiance is described semiclassically
113: (quantized matter in a classical geometry), the Bekenstein-Hawking temperature and the luminosity diverge
114: for $M\rightarrow 0$, as $T_{\rm BH} \propto 1/M$ and $L\propto 1/M^2$, respectively. As a result,
115: this approximation breaks down for very light holes, and in order to determine the final state
116: of the evaporation process a much more precise treatment, including backreaction and quantum gravitational
117: effects, is required.
118:
119: In York's work \cite{ybook,y1}, which is strictly within the semiclassical approximation,
120: the ``teleological" problem is
121: circumvented by relaxing the definition of the EH in the following way. Rather than demanding that the
122: photons ``never" reach infinity he demands only that they are imprisoned by the event horizon for times
123: which are very long compared to the dynamical time scale of the hole. Using this working definition
124: of the EH he is then able to determine its location to first order in the luminosity $L$. In this manner the
125: difficult question about the real final state of the evaporation is not touched upon.
126:
127: It is the purpose of the present paper to analyze the dynamical evaporation process and the corresponding
128: spacetime structure of a radiating Schwarzschild black hole. We include the leading quantum gravitational
129: corrections of the geometry which, as we shall discuss, seem to lead to a termination of
130: the evaporation process and the formation of a cold, Planck size remnant.
131: Our main tool will be the ``renormalization group improvement" of classical solutions, a technique which is very popular
132: in conventional field theory.
133:
134: In fact, recently a lot of work went into the investigation of the nonperturbative renormalization group
135: (RG) behavior of Quantum Einstein Gravity
136: \cite{mr,percadou,oliver1,frank1,oliver2,oliver3,oliver4,oliver5,oliver6,souma,percacciperini,
137: percaper2,frank2,frankf,litimgrav,max,max2,max3,brproper,resh}
138: and its possible manifestations \cite{bh1,bh2,cosmo1,cosmo2,elo,esposito,scalfact,bauer,h1,h2,wey70}.
139: In particular, in \cite{bh2}, a ``RG-improvement" of the Schwarzschild metric has
140: been performed and the properties of the corresponding ``quantum black hole" have been explored. The improvement
141: was based upon the scale dependent (``running") Newton constant $G(k)$ obtained from the exact RG
142: equation for gravity \cite{mr} describing the scale dependence of the effective average action \cite{avact,ym1,ym2,ym3,ym4}.
143: Here $k$ denotes the mass scale of the infrared cutoff which is built into the effective average action
144: $\Gamma_k [g_{\mu\nu}]$ in such a way that it generates the field equations for a metric which has been
145: averaged over a spacetime volume of linear dimension $k^{-1}$. The running of $G$ is approximately given by
146: \be\label{2.23}
147: G(k)={G_0\over 1+\omega \; G_0 \; k^2}
148: \ee
149: where $G_0$ denotes the laboratory value of Newton's constant, and $\omega$ is a constant. At large distances
150: $(k\rightarrow 0)$, $G(k)$ approaches $G_0$, and in the ultraviolet limit $(k\rightarrow \infty)$,
151: it decreases as $G(k)\propto 1/k^2$. This is the fixed point behavior responsible for the conjectured
152: nonperturbative renormalizability of Quantum Einstein Gravity \cite{mr,oliver1,oliver2,souma}.
153:
154: In the RG improvement
155: scheme of \cite{bh2} the information about the $k$-dependence of $G$ is exploited in the following way.
156: The starting point is the classical Schwarzschild metric (in Schwarzschild coordinates)
157: \be\label{due}
158: ds^2 = -f(r) dt^2 + f(r)^{-1}dr^2 + r^2 d\Omega^2
159: \ee
160: with $d\Omega^2 \equiv d\theta^2 +\sin^2 \theta d\phi^2$ and the classical lapse function
161: $f(r)= 1-2G_0 M/r \equiv f_{\rm class}(r)$. The RG improvement is effected by substituting,
162: in $f_{\rm class}(r)$, $G_0$ by the $r$-dependent Newton constant $G(r)\equiv G(k=k(r))$ which
163: obtains from $G(k)$ via an appropriate ``cutoff identification" $k=k(r)$. In flat space
164: the natural choice would be $k \propto 1/r$. In \cite{bh2} we argued that in the
165: Schwarzschild background the correct choice,
166: in leading order at least, is $k(r)= \xi/d(r)$ where $\xi$ is a constant of the
167: order of unity, and $d(r)\equiv \int_0^r dr' | f_{\rm class}(r')|^{-1/2}$ is the proper distance from a point
168: with coordinate $r$ to the center of the black hole. (We refer to \cite{bh2} for a detailed physical
169: justification of this choice.) While the integral defining $d(r)$ can be evaluated exactly, it is sufficient
170: to use the following approximation which becomes exact for both $r\rightarrow \infty$ and $r \rightarrow 0$:
171: \be\label{tre}
172: d(r)=\left ( {r^3\over r+\gamma \; G_0\; M} \right )^{1\over 2}
173: \ee
174: The resulting $G(r)\equiv G(k=\xi/d(r))$ reads
175: \be\label{quattro}
176: G(r)={G_0 \; r^3\over r^3 +\tilde{\omega}\;G_0\; [r+\gamma G_0 M]}
177: \ee
178: where $\widetilde{\omega}\equiv \omega\xi^2$. In these equations the parameter $\gamma$ has the value
179: $\gamma =9/2$ if one sets $k=\xi/d(r)$ as above. It turns out, however, that most of the
180: qualitative properties of the improved metric, in particular all those related to the structure of its horizons,
181: are fairly insensitive to the precise value of $\gamma$. In particular, $\gamma=0$ (corresponding to $k=\xi/r$)
182: and $\gamma =9/2$ where found \cite{bh2} to lead to rather similar results throughout.
183: For this reason we shall adopt the choice $\gamma=0$ in the present paper. It has the advantage that with this choice
184: many calculations can be performed analytically which require a numerical treatment otherwise.
185:
186: The metric of the RG improved Schwarzschild black hole is given by the line element (\ref{due}) with
187: \be\label{cinque}
188: f(r)=1-\frac{2 G(r) M}{r}
189: \ee
190: Let us briefly list its essential features\footnote{All formulas quoted refer to $\gamma=0$, but the
191: qualitative features are the same for $\gamma =9/2$; see \cite{bh2} for details.}.
192:
193: a) There exists a critical mass value\footnote{We define the (standard) Planck mass and length
194: in terms of the laboratory value $G_0$: $m_{\rm Pl}=\ell_{\rm Pl}^{-1}=1/\sqrt{G_0}$.}
195: \be\label{sei}
196: M_{\rm cr}=\sqrt{\widetilde{\omega}/G_0}=\sqrt{\widetilde{\omega}} \; m_{\rm Pl}
197: \ee
198: such that $f(r)$ has two simple zeros at $r_{-}$ and $r_{+}>r_{-}$ if $M>M_{\rm cr}$, one double zero
199: at $r_{+}=r_{-} = \sqrt{\widetilde{\omega}G_0}$ if $M=M_{\rm cr}$, and no zero at all if $M<M_{\rm cr}$.
200: For $M>M_{\rm cr}$ the zeros are at
201: \be\label{sette}
202: r_{\pm} = G_0 M \; [1\pm \sqrt{1-\Omega}]
203: \ee
204: with the convenient abbreviation
205: \be\label{otto}
206: \Omega \equiv \frac{M_{\rm cr}^2 }{M^2} = \widetilde{\omega} \; \Big ( \frac{m_{\rm Pl}}{M} \Big )^2
207: \ee
208: The spacetime has an outer horizon at $r_{+}$ and in inner (Cauchy) horizon at $r_{-}$ . At
209: $M_{\rm cr}$, the black hole is extremal, the two horizons coincide, and
210: the spacetime is free from any horizon if the mass is sufficiently small, $M<M_{\rm cr}$.
211:
212: b) The Bekenstein-Hawking temperature $T_{\rm BH}= \kappa /2\pi$ is given by the surface gravity
213: at the outer horizon, $\kappa = {1\over 2}f'(r_{+})$. Explicitly,
214: \be\label{nove}
215: T_{\rm BH}(M) = {1\over 4\pi G_0 M}\;{\sqrt{1-\Omega}\over 1 +\sqrt{1-\Omega}}
216: ={1\over 4\pi G_0 M_{\rm cr}} \; {\sqrt{\Omega(1-\Omega)} \over 1 +\sqrt{1-\Omega}}=
217: \frac{M_{\rm cr}}{4\pi\widetilde{\omega}}
218: {\sqrt{\Omega(1-\Omega)}\over 1+\sqrt{1-\Omega}}
219: \ee
220: This temperature vanishes for $M \searrow M_{\rm cr}$,
221: {\it i.e.} $\Omega \nearrow 1$, thus motivating the interpretation of the improved Schwarzschild
222: metric with $M=M_{\rm cr}$ as describing a ``cold" remnant of the evaporation process.
223:
224: c) The energy flux from the black hole, its luminosity $L$, can be estimated using Stefan's law.
225: It is given by $L=\sigma {\cal A}(M) T_{\rm BH}(M)^4$ where $\sigma$ is a constant and ${\cal A}\equiv
226: 4\pi r_{+}^2$ denotes the area of the outer horizon. With (\ref{sette}) and (\ref{nove})
227: we obtain
228: \be\label{dieci}
229: L(M) = {\sigma \; M_{\rm cr}^2\over (4\pi)^3 \; \widetilde{\omega}^2}\;
230: {\Omega (1-\Omega)^2\over [1+\sqrt{1-\Omega}]^2}
231: \ee
232: For a single massless field with two degrees of freedom one has $\sigma = \pi^2 /60$.
233: (We use units such that $\hbar=c=k_B=1$.)
234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
235: \renewcommand{\theequation}{2.\arabic{equation}}
236: \setcounter{equation}{0}
237: \section{The quantum-corrected Vaidya metric}
238: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
239: Our aim is to find a metric which describes the history of an evaporating Schwarzschild black hole
240: and its gravitational field. In the small luminosity limit $(L\rightarrow 0)$ this metric is supposed
241: to reduce to the static metric of the RG improved Schwarzschild spacetime.
242:
243: We begin by reexpressing the metric (\ref{due}) with the improved lapse function (\ref{cinque}) in terms
244: of ingoing Eddington-Finkelstein coordinates $(v,r,\theta,\phi)$. We trade the Schwarzschild time $t$
245: for the advanced time coordinate
246: \be\label{dueuno}
247: v=t+r^\star, \;\;\;\;\;\;\;\;\;\;\;\; r^\star \equiv \int^r dr' /f(r')
248: \ee
249: Here $r^\star$ is a generalization of the familiar ``tortoise" radial coordinate to which it
250: reduces if $G(r)=const$. For $G(r) \not = const$ the function $r^\star= r^\star(r)$ is more complicated, but its
251: explicit form will not be needed here. Eq.(\ref{dueuno}) implies
252: $dv= dt+dr/f(r)$, turning (\ref{due}) with (\ref{cinque}) into
253: \be\label{duedue}
254: ds^2=-[1-2G(r)M/r] \; dv^2 + 2 dv dr +r^2d\Omega^2
255: \ee
256: Eq.(\ref{duedue}) is exactly the Schwarzschild metric in Eddington-Finkelstein coordinates, with $G_0$
257: replaced by $G(r)$. It is thus reassuring to see that the two operations, the RG improvement
258: $G_0\rightarrow G(r)$ and the change of the coordinate system, can be performed in either order, they
259: ``commute".
260:
261: The thermodynamical properties derived in \cite{bh2} and summarized in the previous section refer to the
262: metric (\ref{duedue}). In the exterior of the hole the spacetime is static,
263: and while we can deduce a temperature and a corresponding luminosity from its periodicity in imaginary time
264: (or by computing the surface gravity at $r_{+}$ directly) the backreaction of the mass-loss due to the evaporation
265: is not described by (\ref{duedue}). From the static metric we obtained the mass dependence of the
266: luminosity, $L=L(M)$. Using this information we can compute the mass of the hole as seen by
267: a distant observer at time $v$, $M(v)$, by solving
268: the differential equation
269: \be\label{duetre}
270: -{d \over dv} M(v) = L(M(v))
271: \ee
272: In our case $L(M)$ is given by Eq.(\ref{dieci}).
273: To first order in the luminosity, the metric which incorporates the effect of the decreasing mass is obtained by
274: replacing the constant $M$ in (\ref{duedue}) with the $M(v)$ obtained from Eq.(\ref{duetre}):
275: \be\label{duequattro}
276: ds^2=-[1-2G(r)M(v)/r] \; dv^2 + 2 dv dr +r^2d\Omega^2
277: \ee
278: For $G(r)=const$, Eq.(\ref{duequattro}) is the Vaidya metric which frequently had been used to explore the influence of the Hawking
279: radiation on the geometry \cite{bar,his,his2}. It is a solution of Einstein's equation $G_{\mu\nu}=8\pi G_0 T_{\mu\nu}$
280: where $T_{\mu\nu}$ describes an inward moving null fluid. In this picture the decrease of $M$ is due to the
281: inflow of negative energy, as it is appropriate if the field whose quanta are radiated off is in the Unruh vacuum \cite{bd}.
282:
283: The metric (\ref{duequattro}) can be regarded as a RG improved Vaidya metric. It encapsulates two different mechanisms
284: whose combined effect can be studied here: the black hole radiance, and the modifications of the spacetime structure due to the
285: quantum gravity effects, the running of $G$ in particular.
286:
287: It is instructive to ask which energy-momentum tensor $T_{\mu\nu}$ would give rise to the improved
288: Vaidya metric (\ref{duequattro}) according to the classical equation ${G_{\mu}}^{\nu}=8\pi G_0 {T_{\mu}}^{\nu}$. Computing the
289: Einstein tensor of (\ref{duequattro}) one finds that its only non-zero components are
290: \begin{subequations}\label{2.5}
291: \ba
292: && {T^{v}}_v = {T^r}_r=-\frac{G'(r)M(v)}{8\pi G_0 r^2}\label{2.5a}\\[2mm]
293: && {T^r}_v = \frac{G(r) \dot{M}(v)}{8\pi G_0 r^2}\label{2.5b}\\[2mm]
294: && {T^{\theta}}_\theta = {T^\phi}_\phi= -\frac{G''(r) M(v)}{16\pi G_0 r}
295: \ea
296: \end{subequations}
297: Here the prime (dot) denotes a derivative with respect to $r(v)$. The non-zero components (\ref{2.5}) contain either $r$- or
298: $v$-derivatives but no mixed terms. The terms with $r$-derivatives of $G$, also present for $M(v)=const$, describe the
299: vacuum energy density and pressure of the improved Schwarzschild spacetime in absence of radiation effects.
300: They have been discussed in \cite{bh2} already. Allowing for $M(v)\not = const$, the new feature is a nonzero component ${T^r}_v\not = 0$
301: which, for $\dot {M}< 0$, describes the inflow of negative energy into the black hole.
302:
303: Taking advantage of the luminosity function $L(M)$,
304: Eq.(\ref{dieci}), we can solve the differential equation (\ref{duetre})
305: numerically and obtain the mass function $M=M(v)$. (We have
306: set $\sigma / (4\pi)^3 \widetilde{\omega} = 1$ in the
307: numerical calculations in order to reach the almost complete evaporation
308: for $v\approx 200$ in units of $r_{\rm cr}$.)
309: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% FIG 1
310: The result is shown in Fig.(\ref{fig1}) for various initial masses, in the domain $v>0$. In fact,
311: for definiteness we assume that the black hole is formed at $v=0$ by the implosion of a spherical
312: null shell \cite{his}. Hence $M(v)$ is given by Fig.(\ref{fig1}) together with $M=0$ for $v<0$. We observe that,
313: for any initial mass, $M(v)$ approaches the critical mass $M_{\rm cr}$ for $v\rightarrow \infty$.
314: This behavior is the most important manifestation of the quantum gravity effects: according to
315: Eq.(\ref{nove}), the temperature $T_{\rm BH}(M)$ goes to zero when $M$ approaches $M_{\rm cr}$ from
316: above. Hence the luminosity vanishes, too, the evaporation process stops, and $M(v)\approx M_{\rm cr}$ remains
317: approximately constant at very late times, $v \gg M_{\rm cr}^{-1}$.
318: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%% FIG 1
319: \begin{figure}
320: \includegraphics[width=12cm]{fig1.ps}
321: \caption{\label{fig1}
322: The ratio $M/M_{\rm cr}$ as a function of $v/r_{\rm cr}$
323: %$(4\pi)^3\sqrt{\widetilde{\omega}}/\sigma m_{\rm Pl}$, for various initial masses. }
324: for various initial masses. }
325: \end{figure}
326:
327: \begin{figure}
328: \includegraphics[width=12cm]{fig2.ps}
329: \caption{\label{fig2}
330: Time dependence of the Bekenstein-Hawking temperature during the evaporation process for the same
331: initial masses as in Fig.(\ref{fig1}).}
332: \end{figure}
333:
334: \begin{figure}
335: \includegraphics[width=12cm]{fig3.ps}
336: \caption{\label{fig3}
337: The black hole's luminosity as a function of $v/r_{\rm cr}$, for the same initial
338: masses as in Fig.(\ref{fig1}) and Fig.(\ref{fig2}).}
339: \end{figure}
340:
341: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
342: In Fig.(\ref{fig2}) and Fig.(\ref{fig3}) we plot the advanced time dependence of the temperature
343: $T_{\rm BH}(v)\equiv T_{\rm BH}(M(v))$ and the luminosity $L(v)\equiv L(M(v))$, respectively.
344: They are obtained by inserting the numerical solution of Eq.(\ref{duetre}) into (\ref{nove})
345: and (\ref{dieci}).
346:
347: Both the very early and the very late stages of the evaporation process can be described analytically.
348: A black hole with $M(v=0)\gg M_{\rm cr}$
349: starts in what we call the ``Hawking regime". It
350: is defined by the approximation $\Omega \approx 0$ which is realized
351: if the hole is very heavy ($M\gg M_{\rm cr}$)
352: or if $\widetilde{\omega} =0 $ in the semiclassical limit
353: where the quantum gravity corrections are ``switched off".
354: In the Hawking regime, (\ref{nove}) and (\ref{dieci}) reproduce the familiar results
355: \begin{subequations}\label{2.6}
356: \ba
357: && T_{\rm BH} (M)= \frac{1}{8\pi G_0 \; M} \label{2.6a}\\[2mm]
358: && L(M) = \frac{B}{G_0^2 \; M^2}, \;\;\;\;\;\;\;\;\;\;\; B\equiv\frac{\sigma}{4(4\pi)^3} \label{2.6b}
359: \ea
360: \end{subequations}
361: It is easy to solve the differential equation $-\dot{M}=L(M)$ for the luminosity (\ref{2.6b}). With the initial
362: condition $M(v=0)=M_0$ the solution reads
363: \be\label{duesette}
364: M(v)=\Big [M_0^3-3(B/G^2_0) \;v \Big ]^{1/3}.
365: \ee
366: This is the mass function during the early stages of the evaporation process, valid as long as $M(v)$
367: is well above the critical mass. If one naively extrapolates (\ref{duesette}) to small masses one finds $M(v_0)=0$,
368: implying a final ``explosion" with $T\rightarrow \infty$ and $L\rightarrow \infty$, after a finite time
369: $v_0 = G_0^2 M_0^3 / (3 B)$. As a consequence of the quantum gravity effects, this is not what really happens, however.
370:
371: The final part of the evaporation process, where the cold remnant forms, is in the ``critical regime". It is
372: described by those terms in the above expressions which are dominant for $M\searrow M_{\rm cr}$, or $\Omega \nearrow 1$. From (\ref{nove})
373: and (\ref{dieci}) we obtain in this approximation:
374: \begin{subequations}
375: \ba
376: && T_{\rm BH}(M) = \frac{1}{4\pi \widetilde{\omega}}\sqrt{M^2-M^2_{\rm cr}} \label{2.8a}\\[2mm]
377: && L(M) = \frac{\sigma G_0}{(4\pi\widetilde{\omega})^3}\bigg (M^2-M^2_{\rm cr} \bigg )^2 \label{2.8b}
378: \ea
379: \end{subequations}
380: Solving $-\dot{M} = L(M)$ with (\ref{2.8b}) one finds
381: \be\label{duenove}
382: M(v)=M_{\rm cr}+\frac{M_1 -M_{\rm cr}}{1+\alpha (M_1-M_{\rm cr})(v-v_1)}
383: \ee
384: Here $\alpha\equiv \sigma/(16\pi^3\widetilde{\omega}^2)$, and $v_1$ is a time, already in the
385: critical regime, where $M(v_1)=M_1$ is imposed.
386: For $v\rightarrow \infty$, the difference $M(v) - M_{\rm cr}$ vanishes proportional to $1/v$, as
387: a consequence of which $T_{\rm BH}(v)\rightarrow 0$ and $L(v)\rightarrow 0$.
388:
389: We mentioned already that the RG improved Vaidya metric (\ref{duequattro}) can be a correct description
390: only to first order in $L$. In fact, deriving the surface gravity and luminosity from (\ref{duequattro}) the results differ from
391: those for the improved Schwarzschild metric by terms due to the $v$-dependence of $M$. In our approximation those
392: terms are neglected as they would contain additional factors of $\dot{M} = -L$.
393: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
394: \renewcommand{\theequation}{3.\arabic{equation}}
395: \setcounter{equation}{0}
396: \section{apparent horizon and timelike limit surface}
397: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
398: Next we turn to the various horizon-like loci of the improved Vaidya metric. Regarded as $3$-surfaces, all of them are
399: histories of spherical $2$-surfaces.
400:
401: The apparent horizon is a marginally trapped surface. We determine it from the condition that one of the congruences of
402: radial null geodesics, in affine parametrization, has vanishing expansion scalar there, $\Theta=0$. The improved Vaidya
403: metric (\ref{duequattro}) has the structure
404: \be\label{treuno}
405: ds^2=[-f(r,v)dv+2dr]dv+r^2d\Omega^2
406: \ee
407: Along outgoing radial null geodesics we have $f dv = 2dr $. Hence, parametrizing them as $r=r(v)$, they obey the differential
408: equation $\dot{r}(v)=f \big (r(v),v \big)/2$ where the dot denotes a derivative with respect to $v$. We can rewrite this equation in the
409: autonomous form
410: \be\label{tredue}
411: \frac{d }{d\lambda} x^\mu(\lambda)= u^\mu (x(\lambda))
412: \ee
413: with the null vector field
414: \be\label{tretre}
415: u^{\mu}\equiv (u^v,\; u^r, \; u^\theta, \; u^\phi)=(1,\;{1\over 2} f, \;0, \; 0 )
416: \ee
417: A short calculation reveals that the geodesic equation holds in the
418: form $u^\nu D_\nu u^\mu= {\cal K}u^\mu$ with a nonzero function
419: \be\label{trequattro}
420: {\cal K} = \frac{1}{2} \partial_r f
421: \ee
422: Hence the parameter $\lambda$ in (\ref{tredue}) is not an affine one. In order for the standard discussion \cite{poi,he}
423: to be applicable we must reexpress the null geodesics in terms of an affine parameter $\lambda_\ast$. Given a solution $x(\lambda)$
424: of (\ref{tredue}) we compute the function $\lambda_\ast(\lambda)$ by integrating
425: \be\label{trecinque}
426: %\frac{d}{d \lambda} \lambda_\ast(\lambda)=\exp \int_{\lambda=\lambda(\lambda_\ast)}^{\lambda} d\lambda' {\cal K}(x(\lambda'))
427: \frac{d}{d \lambda} \lambda_\ast(\lambda)=\exp \int^{\lambda} d\lambda' \; {\cal K}(x(\lambda'))
428: \ee
429: and determine its inverse $\lambda=\lambda(\lambda_\ast)$. Then we define $x^{\mu}_\ast({\lambda_\ast})\equiv
430: x^{\mu}(\lambda(\lambda_\ast))$ which satisfies
431: \be\label{3.6}
432: \frac{d }{d\lambda_\ast} x_\ast^\mu(\lambda_\ast)= n^\mu (x_\ast(\lambda_\ast))
433: \ee
434: Here
435: \be\label{3.7}
436: n^\mu (x) = e^{-\Gamma(x)}u^{\mu}(x)
437: \ee
438: is a new vector field, with $\Gamma(x)$ satisfying
439: \be\label{3.8}
440: u^\mu \partial_\mu \Gamma ={\cal K}
441: \ee
442: Using (\ref{3.8}) one easily verifies that $n^\nu D_\nu n^\mu=0$, implying that $\lambda_\ast$ is an affine parameter \cite{poi}.
443:
444: The expansion scalar $\Theta$ which determines the location of the AH is the divergence of $n^\mu$. Using (\ref{3.8}) we find
445: \be
446: %NONUMBER
447: \Theta \equiv D_\mu n^\mu = e^{-\Gamma}[D_\mu u^\mu - {\cal K}].
448: \ee
449: For $f(r,v)=1-2G(r)M(v)/r$ we have
450: \be\label{3.9}
451: D_\mu u^\mu = \frac{1}{r} f + \frac{1}{2}\partial_r f
452: \ee
453: Together with (\ref{trequattro}) this yields the expansion scalar
454: \be\label{3.10}
455: \Theta=\frac{1}{r} e^{-\Gamma} f
456: \ee
457: This is the same result as for the classical Vaidya metric; the $r$-dependence of $G$ did not lead to extra terms.
458:
459: Eq.(\ref{3.10}) tells us that $\Theta$ vanishes if, and only if, $f$ vanishes. According to point a) of the Introduction
460: this is the case at $r=r_{+}$ and $r=r_{-}$ with $r_{\pm}$ defined by (\ref{sette}) with (\ref{otto}). (We assume that
461: the $r$-dependence of $G$ is given by (\ref{quattro}) with $\gamma=0$). Since
462: $r_{\pm}$ depends on $M$, it has become a function of the advanced time $v$ now:
463: $r_{\pm}(v)\equiv r_{\pm}(M(v))$. Defined as the outermost trapped surface, the AH is characterized by the implicit
464: equation $r=r_{+}(v)$ where, explicitly,
465: \be\label{3.11}
466: r_{+}(v)=\frac{\widetilde\omega}{M_{\rm cr}} \; \Bigg [ \frac{M(v)}{M_{\rm cr}}+ \sqrt{ \Big (\frac{M(v)}{M_{\rm cr}} \Big)^2-1} \Bigg ]
467: \ee
468:
469: The second horizon surface, the TLS, is defined as the locus where the 4-velocity of static observers
470: $u^{\alpha} \propto \delta^{\alpha}_\tau$ becomes lightlike, with $\partial / \partial \tau $
471: a vector orthogonal to the $r=const$ hypersurfaces. Since we consider
472: a spherically symmetric spacetime, this vector is precisely $\partial / \partial v$ and AH and TLS
473: coincide in this case, being $u^\alpha u^\beta g_{\alpha\beta}=g_{vv}=-f$.
474:
475: Strictly speaking, apart from the outer AH $r=r_{+}(v)$ there exists also an inner TLS=AH
476: $r=r_{-}(v)$, at which the vector field $\partial /\partial v$ switches back from spacelike to timelike.
477: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
478: \begin{figure}
479: \includegraphics[width=12cm]{fig4a.ps}
480: \caption{\label{fig6}
481: Light rays of the outgoing null congruence for $M/M_{\rm cr} = 2$.
482: The thick solid line is the EH determined numerically, the dashed lines are the outer and inner TLS=AH,
483: and the thin solid lines are generic null rays, inside and outside the EH.
484: }
485: \end{figure}
486: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
487: \renewcommand{\theequation}{4.\arabic{equation}}
488: \setcounter{equation}{0}
489: \section{The event horizon}
490: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% SEC IV
491: The radial light rays $r=r(v)$ of the outgoing null congruence are to be found by solving the differential equation
492: $\dot{r}(v)=f(r(v),v)/2$, or explicitly,
493: \be\label{4.1}
494: \frac{d r(v)}{dv} = \frac{1}{2} \Big ( 1-\frac{2 G(r(v)) M(v)}{r(v)}\Big )
495: \ee
496: Depending on their initial points, light rays can, or can not, escape to infinity for $v\rightarrow \infty$.
497: By definition, the ``separatrix" separating those two classes of solutions is the event horizon, the
498: outermost locus traced by outgoing photons that can never reach arbitrarily large distances.
499:
500: It is easy to understand why for a radiating (as opposed to an accreting) black hole the EH is {\it inside} the TLS.
501: Inside (outside) the TLS, $f$ is negative (positive), implying that the light ray's $r(v)$ decreases (increases)
502: if $r(v)<r_{\rm TLS}(v)$ $ \big ( r(v)>r_{\rm TLS}(v) \big )$.
503: Under certain conditions it can happen that due to the hole's mass
504: loss the radius of the TLS decreases faster than $r(v)$. As a result,
505: the light ray intersects the TLS, with $\dot{r}=0$ there, and then escapes from the hole with $\dot{r}>0$.
506:
507: The situation is illustrated in Fig.(\ref{fig6}). The region between $r_{\rm EH}$ and $r_{\rm TLS}$ forms the hole's
508: ``quantum ergosphere" which owes its existence entirely to the evaporation process.
509:
510: Clearly a determination of the EH's $r_{\rm EH}(v)$ requires
511: knowledge of $M(v)$ for arbitrarily late times $v$ even. Therefore the semiclassical
512: approximation is not sufficient to find the EH since it breaks down for small $M$.
513: As a way out, York \cite{ybook,y1} proposed to replace
514: the above rigorous definition of the EH by an approximate criterion which is local in $v$ and does not require the ``teleological"
515: information about $M(v)$ at late $v$. It is supposed to be valid for small $L$. In the following we analyze (\ref{4.1})
516: both using the approximate criterion and the exact definition of the EH. Since we have an explicit prediction for the final stages of the
517: evaporation process we are in a position to determine the EH exactly, and thus to assess the validity and precision of York's approximate
518: ``working definition".
519:
520: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
521: \begin{figure}
522: \includegraphics[width=10cm]{york.eps}
523: \caption{\label{yo}
524: Light rays of the outging null congruence with $\ddot{r}>0$, $\ddot{r}=0$, and $\ddot{r}<0$ at $v_{1}$. Those
525: with $\ddot{r}>0$ are most likely to eventually cross the TLS.
526: }
527: \end{figure}
528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
529:
530: Let us begin with this working definition, applied to the RG improved spacetime. While according to Eq.(\ref{4.1}), the ``velocity" $\dot{r}(v)$
531: is strictly negative inside the TLS, the acceleration $\ddot{r}(v)$ can have either sign there. York proposes to approximately
532: identify $r_{\rm EH}(v)$ with the radius where the light ray's ``acceleration" vanishes: $\ddot{r}({\rm EH})=0$.
533: In pictorial terms
534: one could think of this condition as separating the light rays curving downward ($\ddot{r}<0$) from those curving upward $(\ddot{r}>0)$. Clearly
535: the former (latter) are very unlikely (likely) to subsequently cross the TLS, even if this is not a rigorous criterion, of course.
536: (See Fig.(\ref{yo}) for a schematic sketch.)
537: At least in the Hawking regime, photons are imprisoned by the approximate horizon for times long compared to the dynamical time scale
538: of the evaporating hole \cite{ybook, y1}.
539:
540: Taking a second $v$-derivative of (\ref{4.1}) we obtain the ``acceleration"
541: \be\label{4.2}
542: {\ddot r}(v)= L(v) \frac{G(r)}{r} + G(r) M(v)\frac{\dot r}{r^2} -
543: {G}'(r) M(v) {\dot r \over r}
544: \ee
545: Here we assumed that $G(r)$ does not have a parametric dependence on $M$, which is actually true for $\gamma=0$:
546: \be\label{4.3}
547: G(r)=\frac{G_0 r^2}{r^2+{\widetilde\omega}G_0}
548: \ee
549: (For general $\gamma$ there appears an additional term $\propto dG/dM$ in (\ref{4.2}) which is irrelevant qualitatively.)
550: Eq.(\ref{4.2}) tells us that, when $\ddot r = 0$, the radius $r$, the velocity $\dot r$, and the time $v$ are related by
551: \be\label{4.4}
552: \dot{r}=-L \; G(r) \Big [ \frac{G(r) M(v)}{r} - M(v){G}'(r)\Big ]^{-1}
553: \ee
554: Since $r_{\rm EH}=r_{\rm TLS}$ if the hole would not radiate, and since higher orders in the luminosity are neglected,
555: the difference $r_{\rm TLS}-r_{\rm EH}$ which we would like to compute is of order $L$. For this reason we may replace
556: $r\equiv r_{\rm EH}$ on the RHS of (\ref{4.4}) with $r_{\rm TLS}=r_{+}(v)$, the error being of order $L^2$:
557: \be\label{4.5}
558: \dot{r}= - 2 L G(r_{+}) \Big [\frac{2 G(r_{+}) M(v)}{r_{+}}-2 M(v) G'(r_{+}) \Big ]^{-1}
559: = {-2L G(r_{+}) \over [ 1-2 M(v)G'(r_{+})] }
560: \ee
561: In the second equality of (\ref{4.5}) we used that $2G(r_{+})M=r_{+}$ which follows from $f(r_{+})=0$.
562: Eq.(\ref{4.5}) provides us with the ``velocity" $\dot{r}$ at the point where the acceleration vanishes. The corresponding
563: coordinate, $r_{\rm EH}$, is the approximate location of the EH. We obtain it by using $\dot{r}=f(r)/2$ in order to rewrite
564: the LHS of (\ref{4.5}) in the form $f(r_{\rm EH})/2$, and then inverting the function $f$. This inversion is easy to perform
565: since, again, we may expand in $r_{\rm EH}-r_{+}=O(L)$:
566: \be\label{x}\nonumber
567: f(r_{\rm EH})=f(r_{+})+(r_{\rm EH}-r_{\rm +})f'(r_{+}) + O(L^2) .
568: \ee
569: Hence, to order $L$, ${\dot r}({\rm EH})=(r_{\rm EH}-r_{+})f'(r_{+})/2$ which yields, together with (\ref{4.5}),
570: \be\label{4.6}
571: r_{\rm EH}=r_{+}- \frac{4 L G(r_{+})}{f'(r_{+})[1-2M G'(r_{+})]}
572: \ee
573: Differentiating $f(r)=1-2G(r)M/r$ and using $2G(r_{+})M=r_{+}$ one obtains the relation $f'(r_{+})=[1-2 M G'(r_{+})]/r_{+}$.
574: This leads to the following explicit formula for the $v$-dependence of $r_{\rm EH}$:
575: \be\label{4.7}
576: r_{\rm EH}(v)=r_{+}(v)\Bigg \{ 1-\frac{4 \; L(v) \; G(r_{+}(v))}{[1-2\;M(v)\;G'(r_{+}(v))]^2} \Bigg \}
577: \ee
578: Eq.(\ref{4.7}) is valid for an arbitrary $r$-dependence of $G$, still. If $G=const$ we recover
579: York's result $r_{\rm EH}=r_{+}[1-4G_0L]$ with $r_{+}=2G_0 M$.
580:
581: Let us now specialize for the function $G(r)$ motivated by quantum gravity, Eq.(\ref{4.3}). This $r$-dependence
582: of Newton's constant implies that, at the TLS,
583: $G(r_{+})=r_{+}/(2M)$ and $G'(r_{+}) ={\widetilde \omega}/(2M^2 r_{+})$.
584: These equations were simplified using the relation $r_{+}^2+\widetilde{\omega} G_0 = 2 G_0 M r_{+}$ which is equivalent to
585: $f(r_{+})=0$ with (\ref{4.3}). They lead to the following result for the position of the EH:
586: \begin{subequations}\label{4.8}
587: \ba\label{4.8a}
588: &&\re=\rp \big [1-4\;G_0 \; L(M) \; Y(\Omega) \big ]\\[2mm]
589: &&Y(\Omega)\equiv \frac{1+\sqrt{1-\Omega}}{2(1-\Omega)}
590: \ea
591: \end{subequations}
592: The radius $\rp$ is given by Eq.(\ref{sette}) or (\ref{3.11}), respectively,
593: and $M$ and $\Omega\equiv M_{\rm cr}^2/M(v)^2$ are understood to be functions of $v$, of course. The correction factor $Y(\Omega)$
594: measures the deviations from the semiclassical result; we have $Y(\Omega)\approx 1$ everywhere in the Hawking regime
595: ($\Omega \approx 0$). For $\Omega \nearrow 1$ the corrections become large; in fact, $Y(\Omega)$ diverges at $\Omega =1$.
596: Nevertheless, the product $LY$ vanishes for $\Omega\nearrow 1$, $M\searrow M_{\rm cr}$, which becomes clear when we insert
597: (\ref{dieci}) into (\ref{4.8a}):
598: \begin{subequations}\label{4.9}
599: \ba\label{4.9a}
600: \re&&=\rp \Bigg [ 1-\frac{2\sigma}{(4\pi)^3\widetilde{\omega}}\; \frac{\Omega (1-\Omega)}{[1+\sqrt{1-\Omega}]}\Bigg ]\\[2mm]
601: &&=\rp-\frac{2\sigma}{(4\pi)^3}\; \frac{1}{M}\Big (1-\frac{M_{\rm cr}^2}{M^2} \Big )
602: \ea
603: \end{subequations}
604: This is our final result for the radius of the event horizon, as given by York's approximate local criterion.
605: In the early stages of the evaporation we recover the semiclassical result. For $v\rightarrow \infty$
606: however, $M(v)\rightarrow M_{\rm cr}$, and both $r_{\rm EH}$ and $r_{+}(v)$ approach the same limiting value asymptotically:
607: $r_{\rm cr}\equiv r_{\pm}(M_{\rm cr})= \sqrt{\widetilde{\omega} G_0}=\sqrt{\widetilde{\omega}}\ell_{\rm pl}$. This behavior can
608: be seen in Fig.(\ref{yyy}).
609: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
610: \begin{figure}
611: \includegraphics[width=12cm]{fig4b.ps}
612: \caption{\label{yyy}
613: The EH determined numerically (solid thick line) and by York's approximate criterion (dashed line)
614: for $M/M_{\rm cr} = 2$. The solid thin line is the AH.
615: }
616: \end{figure}
617: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
618:
619: As we anticipated, we now see that $\re$ is indeed smaller than $\rp$ during the entire evaporation process, the EH is
620: {\it inside} the (outer) TLS, thus giving rise to a quantum ergosphere.
621:
622: It is an important observation that, while $r_{\rm EH}$ is smaller than $\rp$, it is always larger than
623: $r_{-}$, provided the radiation effects are small compared to the quantum gravity effects in an appropriate sense.
624: Comparing the radius (\ref{4.9}) to $r_{-}$ as given by (\ref{sette}) we find that
625: $\re (M) > r_{-}(M)$ provided
626: \be\label{4.10}
627: \frac{\sigma}{(4\pi)^3\widetilde{\omega}}< \frac{1}{\Omega{\sqrt{1-\Omega}}}
628: \ee
629: For $0\leq \Omega \leq 1$ the RHS of (\ref{4.10}) is bounded below by the constant $3\sqrt{3}/2$. As a result,
630: $\re$ is larger than $r_{-}$ during the entire evaporation process provided
631: \be\label{4.11}
632: \sigma < \frac{3}{2}\sqrt{3} (4\pi)^3 \; \widetilde{\omega}
633: \ee
634: If (\ref{4.11}) is satisfied, we have $\rp(v)>\re (v) >r_{-} (v)$ for any finite $v$, and the EH touches both the outer and the inner TLS only
635: in the limit $v\rightarrow \infty$ where $\rp,\re,r_{-}\rightarrow r_{\rm cr}$.
636:
637: If (\ref{4.11}) is violated our method is inapplicable, most probably, and the improved Vaidya metric is not a reliable
638: description of the spacetime structure. This metric is valid to first order in $L$ only, which means that the dimensionless luminosity
639: has to be much smaller than unity, $G_0 L \ll 1$. If so, Eq.(\ref{dieci}) implies $\sigma/[(4\pi)^3\widetilde{\omega}]\ll 1$,
640: and (\ref{4.11}) is indeed satisfied. (Note that $M_{\rm cr}^2/\widetilde{\omega}^2= G_0^{-1}/\widetilde{\omega}$.) It is
641: reassuring to see that, for pure gravity, Eq.(\ref{4.11}) is fulfilled with a very broad margin; the expansion parameter
642: $\sigma/[(4\pi)^3\widetilde{\omega}]$ assumes the tiny value $8.3 \cdot 10^{-5} / \tilde{\omega}$
643: in this case, with $\tilde\omega = O(1)$.
644:
645: Let us return to the exact definition of the EH now. We compute it for our model by numerically solving (\ref{4.1}) for a set of
646: initial conditions $r(0)$. For a certain range of $r(0)$'s the trajectories will ultimately cross the TLS and escape to infinity,
647: while the others remain at radii below
648: $r_{\rm TLS}$ for arbitrarily late advanced times. The light ray separating those two classes of solutions defines the EH,
649: a null hypersurface by construction.
650:
651: In Fig.(\ref{fig6}) we show various trajectories of either class, as well as the EH,
652: determined numerically from its {\it exact} definition. We have numerically evolved
653: several initial conditions $r(0)$ in
654: order to determine the boundary between trapped and escaping null geodesics in the limit $v\rightarrow \infty$.
655: %%In particular for $r(0)=2.692$ and $v=3.e+05$ one gets $r_{\rm EH}=0.9958$ which is a reasonable approximation for our purposes.
656: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
657: \begin{figure}
658: \includegraphics[width=6cm]{bhevp.eps}
659: \caption{\label{fig7}
660: The conformal diagram of the evaporating quantum black hole:
661: region I is a flat spacetime, and
662: region II is the evaporating BH spacetime,
663: $EH$ is the event horizon, $CH$ is the inner (Cauchy) horizon,
664: and $A$ is the apparent horizon. }
665: \end{figure}
666: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
667: In Fig.(\ref{yyy}) the true EH is compared to the prediction of York's criterion. Obviously the
668: latter provides a rather accurate approximation to the true horizon.
669:
670: The global structure of the spacetime is depicted in the conformal diagram in Fig.(\ref{fig7}).
671: Region I is a flat spacetime, while at $V=V_0$ ($V$ is the Kruskal advanced time coordinate,
672: defined as $V =-\exp (-\kappa v)$ being $\kappa$ the surface gravity of the outer horizon)
673: an imploding null shell is present (strictly speaking it must have a negative tension
674: in order to balance the flux of negative energy on its future side \cite{baris}).
675: Region II is the evaporating black hole spacetime. The AH is a timelike hypersurface which
676: ``meets'' the EH at future null infinity in the conformal diagram.
677: The null ray which is tangent to the earliest portion of the apparent horizon
678: $A$ would have been the EH if the hole were not radiating.
679: The final state of the black hole is an extremal black hole whose inner
680: and outer horizons have the same radius $(r=r_{\rm cr})$ and are located at
681: the event horizon $EH$ and the inner (Cauchy) horizon $CH$ in Fig.(\ref{fig7}).
682:
683: It is instructive to compare the areas ${\cal A}$ of the various horizons. They are defined by intersecting the EH, AH, and TLS with the
684: incoming null surfaces $v= const$. Thus ${\cal A}_{\rm TLS} (v)= 4\pi \rp(v)^2$ and ${\cal A}_{\rm EH}(v)=4\pi\re(v)^2$.
685: From Eq.(\ref{sette})
686: we obtain for ${\cal A}_{\rm TLS} \equiv {\cal A}_{\rm AH}$
687: \be\label{4.12}
688: {\cal A}_{\rm TLS} = 4\pi G_0^2 M^2 \Big [1+\sqrt{1-(M_{\rm cr}/M)^2}\Big ]^2
689: \ee
690: and the approximate result (\ref{4.9a}) for the event horizon implies
691: \be\label{4.13}
692: {\cal A}_{\rm EH}={\cal A}_{\rm TLS} \Big [1-\frac{4\sigma}{(4\pi)^3\widetilde{\omega}}
693: \frac{\Omega(1-\Omega)}{1+\sqrt{1-\Omega}} \Big ]
694: \ee
695: where a term of second order in $\sigma/(4\pi)^3\widetilde{\omega}$ has been neglected. The difference
696: $\delta {\cal A}\equiv {\cal A}_{\rm TLS}-{\cal A}_{\rm EH}$ is
697: given by
698: \be\label{4.14}
699: \delta {\cal A}=\frac{\sigma}{4\pi^2} \; \ell_{\rm Pl}^2 \; (1-\Omega) \big [1+\sqrt{1-{\Omega}} \big ]
700: \ee
701: During the early stages of the evaporation process,
702: $\delta {\cal A}\approx \sigma l_{\rm Pl}^2 /(2\pi^2) = 128 \pi B \ell_{\rm Pl}^2$ which coincides with the known result
703: \cite{ybook} for the Hawking regime, while $\delta {\cal A}$ vanishes proportional to $(M^2-M^2_{\rm cr})\rightarrow 0$
704: for $v \rightarrow \infty$.
705: It had been emphasized by York \cite{ybook,y1} that in the Hawking regime he considered, $\delta {\cal A}$ is a universal
706: ({\it i.e.} $M$ independent) quantity which depends only on $\sigma$, thus counting the degrees of freedom of the
707: field quanta which can be evaporated off. Looking at Eq.(\ref{4.14}) we see that this universality does not
708: persist beyond the semiclassical approximation.
709: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
710: \renewcommand{\theequation}{5.\arabic{equation}}
711: \setcounter{equation}{0}
712: \section{Conclusion}
713: The renormalization group improvement of black hole spacetimes
714: according to Quantum Einstein Gravity leads to concrete predictions on the final state of
715: the evaporation process.
716: Unlike previous studies based on {\it ad hoc} modifications
717: of the equation of state of matter at very high (Planckian) densities \cite{muka,dymni1,hay05},
718: or models based on loop quantum gravity \cite{ashte05},
719: the mass of the remnant can be calculated explicitly:
720: $M_{\rm cr}=\sqrt{\widetilde{\omega}} \ell_{\rm Pl}$.
721: Its precise value is determined by the value of $\widetilde{\omega}$
722: which is a measurable quantity in principle
723: \cite{donog}.
724: No naked singularity forms, at variance with the paradigm proposed in \cite{his,hay05}, so that
725: the remnant is a mini-black hole of Planckian size.
726: On the other hand, it is intriguing to note that exactly solvable semiclassical gravity-dilaton models
727: predict a final state described by an extremal configuration which is reached in an infinite amount
728: of time \cite{fabbri}. (See also \cite{grumi} and
729: \cite{jac} for analogous semiclassical models, and
730: \cite{ward1,ward2} for an approach based on special resummations of higher order graviton
731: loops. )
732:
733: It would be interesting
734: to investigate the possible astrophysical implications of a population of
735: stable Planck size mini-black holes produced in the Early Universe or by the interaction of
736: cosmic rays with the interstellar medium \cite{barrau05,barrau03}. We hope to address these issues in
737: a subsequent publication.
738:
739: \section{Acknowledgments}
740: We would like to thank Werner Israel for helpful suggestions on an earlier version of this manuscript.
741: A.B. would also like to thank the Department of Physics of Mainz University and M.R.
742: the Catania Astrophysical Observatory for the financial support and for the cordial hospitality
743: extended to them while this work was in progress.
744: We are also grateful to INFN, Sezione di Catania for financial support.
745:
746: \bibliography{bb.bbl}
747:
748: \end{document}
749:
750:
751: