1: \chapter{Grey-body Factors in $(4+n)$ Dimensions}
2: \chaptermark{Grey-body Factors}
3: \label{greybody}
4: \let\om=\omega
5: \let\si=\sigma
6: \let\Ga=\Gamma
7:
8: \section{Introduction}
9: \label{gbfintro}
10:
11: The idea of grey-body factors, both for four-dimensional and extra-dimensional black holes, was introduced in Chapter~\ref{bhintro}. Extra-dimensional black holes are of interest because, if scenarios like that of ADD \cite{Arkani-Hamed:1998rs,Arkani-Hamed:1998nn,Antoniadis:1998ig} are realized in nature, they could possibly be produced at high-energy colliders. To understand the particle emission probabilities and spectra for these black holes it is necessary to calculate the extra-dimensional grey-body factors.
12:
13: For ease of reference, equations~(\ref{3flux}) and (\ref{3power}) describing the flux and power spectra are reproduced below:
14: %%%%%%%%
15: %\begin{align}
16: %\label{4flux}
17: % \frac{dN^{(s)}(\om)}{dt}& = \sum_{\ell} \sigma^{(s)}_{\ell}(\om)\,
18: %\frac{1}{\exp\left(\om/T_\text{H}\right) \mp 1}
19: %\,\frac{d^{n+3}p}{(2\pi)^{n+3}}\,;\\
20: %\frac{dE^{(s)}(\om)}{dt}& = \sum_{\ell} \sigma^{(s)}_{\ell}(\om)\,
21: %\frac{\om}{\exp\left(\om/T_\text{H}\right) \mp 1}\,\frac{d^{n+3}p}{(2\pi)^{n+3}}\,.
22: %\label{4power}
23: %\end{align}
24: \begin{equation}
25: \label{4flux}
26: \frac{dN^{(s)}(\om)}{dt} = \sum_{\ell} \sigma^{(s)}_{\ell}(\om)\,
27: \frac{1}{\exp\left(\om/T_\text{H}\right) \mp 1}
28: \,\frac{d^{n+3}p}{(2\pi)^{n+3}}\,;
29: \end{equation}
30: %%%%%
31: and
32: %%%%%
33: \begin{equation}
34: \frac{dE^{(s)}(\om)}{dt} = \sum_{\ell} \sigma^{(s)}_{\ell}(\om)\,
35: \frac{\om}{\exp\left(\om/T_\text{H}\right) \mp 1}\,\frac{d^{n+3}p}{(2\pi)^{n+3}}\,.
36: \label{4power}
37: \end{equation}
38: %%%%%%%%
39: $\sigma^{(s)}_{\ell} (\om) $ is the grey-body factor\footnote{The quantity $\sigma^{(s)}_{\ell} (\om) $ is alternatively called the absorption cross section. It is also common in the literature to refer to the absorption probability $|{\cal A}^{(s)}_\ell |^2$, related to $\sigma^{(s)}_{\ell} (\om) $ through eq.~(\ref{greydef}), as the grey-body factor. The notation $\Gamma_\ell^{(s)}$ is sometimes used for $|{\cal A}^{(s)}_\ell|^2$.} which can be determined by solving the
40: equation of motion of a particular degree of freedom in the gravitational
41: background and computing the corresponding absorption coefficient
42: ${\cal A}^{(s)}_\ell$. The absorption coefficient is related to the grey-body factor by \cite{Gubser:1997yh}
43: %%%%%%%%%%%%%
44: \begin{equation}
45: \si^{(s)}_\ell(\om) = \frac{2^{n}\pi^{(n+1)/2}\,\Ga[(n+1)/2]}{ n!\, \om^{n+2}}\,
46: \frac{(2\ell+n+1)\,(\ell+n)!}{\ell !}\, |{\cal A}^{(s)}_\ell|^2.
47: \label{greydef}
48: \end{equation}
49: %%%%%%%%%%%
50:
51: As already mentioned, the grey-body factors modify the spectra of emitted particles from that of a perfect thermal black body \emph{even} in four dimensions \cite{Hawking:1975sw}. For a 4-dimensional Schwarzschild black hole, geometric arguments show that, in the high-energy ($\omega\gg T_\text{H}$) regime,
52: $\Sigma_{\ell}\,|{\cal A}_\ell^{(s)}|^2 \propto(\omega r_\text{h})^2$. Therefore
53: the grey-body factor at high energies is independent of $\omega$ and the
54: spectrum is exactly like that of a black body for every particle species
55: \cite{misner,Sanchez:1978si,Sanchez:1978vz,Page:1976df}. The low-energy behaviour, on the other hand, is
56: strongly spin-dependent. A common feature for fields with spin $s=0$,
57: 1/2 and 1 is that the grey-body factors reduce
58: the low-energy emission rate significantly below the geometrical optics value
59: \cite{Page:1976df,MacGibbon:1990zk}. The result is that both the power and flux spectra peak at higher energies than those for a black body at the same temperature. As $s$ increases, the low-energy suppression increases and so the emission probability decreases. These particle emission probabilities have been available in the literature for over twenty years \cite{Page:1976df,Sanchez:1978si,Sanchez:1978vz}.
60:
61: These characteristics of the four-dimensional grey-body factors demonstrate why it is important to calculate the equivalent extra-dimensional factors. Studying their dependence on the dimensionality of space-time is important if detected Hawking radiation emitted by small black holes is to be used to try to determine the value of $n$. Both the energy spectra and the relative emissivities of scalars, fermions and gauge bosons will be potentially useful in this task. The main motivation for this work was to obtain grey-body factors which could be used in the Monte Carlo event generator described in Chapter~\ref{generator}, thereby allowing the possible determination of $n$ to be studied in detail.
62:
63: The procedure for calculating grey-body factors needs to be generalized to include emission from small higher-dimensional black holes with $r_\text{h} < R$; there is an added complication in this case because black holes can emit radiation either on the brane or in the bulk. The emission of particle modes on the brane is the most phenomenologically interesting effect since it involves Standard Model particles. However gravitons will certainly be emitted into the bulk (since it is the propagation of gravity in the bulk which accounts for its apparent weakness and hence solves the hierarchy problem) and it is possible that a $(4+n)$-dimensional black hole will also emit bulk scalar modes. These are potentially important because in some extra dimension models there are additional scalars propagating in the bulk; for example, supersymmetric models inspired by string theory involve graviton supermultiplets which can include bulk scalars (see e.g.\ \cite{Scherk:1979aj,Scherk:1979rh}).\footnote{Scalars propagating in the bulk would affect many of the constraints described in \secref{edcon}, although in model-dependent ways since the couplings to SM particles would not be universal (unlike those of the graviton).} Numerical results for bulk and brane scalar modes are necessary to allow comparison of the total bulk and brane emissivities. This would allow the usual heuristic argument for brane dominance (because there are many more SM brane degrees of freedom than there are
64: degrees of freedom in the bulk---see \secref{basic}) to be confirmed or disproved.
65:
66: There have already been some analytic studies for extra-dimensional black holes \cite{Kanti:2002nr,Kanti:2002ge,Ida:2002ez} which have suggested that the grey-body factors can have a strong dependence on the number of extra dimensions. However it is known that results from the power series expansions in these papers are only accurate in a very limited low-energy regime (especially in the case of scalar fields) and even the full analytic results are not necessarily reliable above the intermediate-energy regime.
67:
68: The work described in this chapter focuses on the emission of energy from uncharged $(4+n)$-dimensional black holes; initially only Schwarzschild-like black holes are considered although there are also some results for rotating holes. Section~\ref{branegbf} gives some details of a master equation describing the motion of scalars, fermions and gauge bosons in the 4-dimensional induced background---the basic steps of Kanti's calculation \cite{Harris:2003eg} are given in Appendix~\ref{appb}. This is particularly useful in that it helps to resolve ambiguities present in similar equations which have previously appeared in the literature. The brane grey-body factors and emission rates are also defined in \secref{branegbf}, and analytic and numerical methods for computing these quantities are discussed. Exact numerical results for emission on the brane (for scalars, fermions and gauge bosons) are presented in \secref{numres}. This allows a comparison with the earlier analytic studies of the Schwarzschild phase as well as a discussion of the relative emissivities and their dependence on $n$. A calculation of the total power and flux emitted (and hence the lifetime and number of emitted particles expected for extra-dimensional black holes) is also possible in this section. The emission of bulk scalar modes is studied in \secref{embulk}, and the results for the relative bulk-to-brane emissivities are also presented in this section. The analysis provides, for the first time, exact results for the relative bulk and brane emissivities in all energy regimes and for various values of $n$. Finally section~\ref{rotbh} includes a discussion of the emission of scalars from rotating black holes; this allows some comments to be made on super-radiance for extra-dimensional black holes. The conclusions of the work in this chapter are summarized in \secref{gbfconc}.
69:
70: \section{Grey-body factors for emission on the brane}
71: \label{branegbf}
72:
73: This section and \secref{numres} focus on the emission of brane-localized modes, leaving the study of bulk emission and of relative bulk-to-brane emissivity until \secref{embulk}.
74:
75: The brane-localized modes propagate in a 4-dimensional black hole background
76: which is the projection of the higher-dimensional one, given in eq.~(\ref{metric-D}), onto the brane. The induced metric tensor follows by fixing
77: the values of the extra angular co-ordinates ($\theta_i=\pi/2$ for $i \geq 2$)
78: and is found to have the form
79: %%%%
80: \begin{equation}
81: ds^2=-h(r)\,dt^2+h(r)^{-1}dr^2+r^2\,(d\theta^2 + \sin^2\theta\,d \varphi^2)\,.
82: \label{non-rot}
83: \end{equation}
84:
85: The grey-body factors are determined from the amplitudes of in-going and
86: out-going waves at infinity so the essential requirement is to solve the
87: equation of motion for a particle propagating in the above background.
88: For this purpose, a generalized {\it master equation} was derived by Kanti (see Appendix~\ref{appb}) for a particle with arbitrary spin $s$; this equation is similar to the 4-dimensional one derived by Teukolsky \cite{Teukolsky:1973ha}.
89: For $s=$1/2 and 1, the equation of motion was
90: derived using the Newman-Penrose method \cite{Newman:1962qr,chandra}, while for $s=0$
91: the corresponding equation follows by the evaluation of the double covariant
92: derivative $g^{\mu\nu} D_\mu D_\nu$ acting on the scalar field. The derived
93: master equation is separable in each case and, by using the factorization
94: %%%%%%%%%%%%%%
95: \begin{equation}
96: \label{sepsoln}
97: \Psi_s=e^{-i\omega t}\,e^{im\varphi}\,R_s(r)\,{}_sS^m_{\ell}(\theta)\,,
98: \end{equation}
99: %%%%%%%%%%%%
100: we obtain the radial equation
101: %%%%%%%%%%%
102: \begin{equation}
103: \label{radial}
104: \Delta^{-s} \frac{d}{dr}\left(\Delta^{s+1}\,\frac{d R_s}{dr}\right)+
105: \left(\frac{\omega^2 r^2}{h}+2i\omega s r-\frac{is\omega r^2 h'}{h}+
106: s(\Delta''-2)-{}_s\lambda_{\ell} \right)R_s (r)=0\,,
107: \end{equation}
108: where $\Delta=hr^2$. The corresponding angular equation has the form
109: %%%%%%%%%%%
110: \begin{equation}
111: \frac{1}{\sin\theta}\,\frac{d}{d\theta}\left(\sin\theta\,\,
112: \frac{d\,{}_sS_{\ell}^m}{d\theta}\right)+
113: \left(-\frac{2ms\cot\theta}{\sin\theta}-\frac{m^2}{\sin^2\theta}+
114: s-s^2\cot^2\theta+{}_s\lambda_{\ell}\right){}_sS^m_{\ell}(\theta)=0\,,
115: \label{angular}
116: \end{equation}
117: %%%%%%%%%%%%%%%%
118: where $e^{im\varphi}\,{}_sS^m_{\ell}(\theta)={}_sY^m_{\ell}(\Omega_2)$ are known as the spin-weighted spherical harmonics and ${}_s\lambda_{\ell}$ is a separation constant which is found to have the value ${}_s\lambda_{\ell}=\ell(\ell+1)-s(s+1)$ \cite{Goldberg:1967uu}.
119:
120: For $s=0$, equation~(\ref{radial}) reduces as expected to eq. (41) of
121: ref. \cite{Kanti:2002nr} which was used for the analytical study of the emission of
122: brane-localized scalar modes from a spherically-symmetric higher-dimensional
123: black hole. Under the redefinition $R_s=\Delta^{-s} P_s$, eq.~(\ref{radial})
124: assumes a form similar to eq. (11) of ref. \cite{Kanti:2002ge} which was
125: used for the study of brane-localized fermion and gauge boson emission.
126: The two equations differ due to an extra term in the expression of the latter
127: one, which although vanishing for $s=$1/2 and 1 (leading to the
128: correct results for fermion and gauge boson fields)
129: gives a non-vanishing contribution
130: for all other values of $s$. Therefore, the generalized equation derived by
131: Cvetic and Larsen \cite{Cvetic:1998ap} cannot be considered as a master equation
132: valid for all types of fields.\footnote{A similar equation was derived in
133: ref. \cite{Ida:2002ez} but due to a typographical error the multiplicative factor $s$ in front of the $\Delta''$-term was missing, leading to an apparently different equation for general $s$.} Hence the derivation of a consistent master equation was imperative, before addressing the question of the grey-body factors in the brane background. This task was performed by Kanti and led to eqs.~(\ref{radial}) and (\ref{angular}) above.
134:
135: For the derivation of grey-body factors associated with the emission of
136: fields from the projected black hole, we need to know the asymptotic solutions
137: of eq.~(\ref{radial}) both as $r\rightarrow r_\text{h}$ and as $r\rightarrow \infty$.
138: In the former case, the solution is of the form
139: %%%%%%%%%
140: \begin{equation}
141: \label{near}
142: R_s^{(\text{h})}=A_\text{in}^{(\text{h})}\,\Delta^{-s}\,e^{-i\omega r^{*}}+
143: A_{out }^{(\text{h})}\,e^{i\omega r^{*}},
144: \end{equation}
145: %%%%%%%%%%
146: where
147: %%%%%%%%%%%%
148: \begin{equation}
149: \label{rstarr}
150: \frac{dr^*}{dr}=\frac{1}{h(r)}\,.
151: \end{equation}
152: %%%%%%%%%
153: We impose the boundary condition that there is no out-going solution near the
154: horizon of the black hole, and therefore set $A_\text{out}^{(\text{h})}=0$.
155: The solution at infinity is of the form
156: %%%%%%%%%%%
157: \begin{equation}
158: \label{far}
159: R_s^{(\infty)}=A_\text{in}^{(\infty)}\,\frac{e^{-i\omega r}}{r}+
160: A_\text{out}^{(\infty)}\,\frac{e^{i\omega r}}{r^{2s+1}}\,,
161: \end{equation}
162: and comprises both in-going and out-going modes.
163: \enlargethispage{-2\baselineskip}
164:
165: The grey-body factor $\sigma^{(s)}_{\ell}(\om)$ for the emission of brane-localized
166: modes is related to the energy absorption coefficient ${\cal A}_\ell$
167: through the simplified relation
168: %%%%%%%%
169: \begin{equation}
170: \hat \si^{(s)}_\ell(\om) =\frac{\pi}{\om^2}\,(2 \ell +1)\,
171: |\hat {\cal A}^{(s)}_\ell|^2,
172: \label{brane-loc}
173: \end{equation}
174: %%%%%%%%%%%%
175: where from now on a `hat' will denote quantities associated with
176: the emission of brane-localized modes. The above relation follows from
177: eq.~(\ref{greydef}) by setting $n=0$ since the emission of brane-localized
178: modes is a 4-dimensional process. The absorption coefficient itself is defined as
179: %%%%%%%%%%%
180: \begin{equation}
181: |\hat {\cal A}^{(s)}_\ell|^2=1-\frac{{\cal F}_\text{out}^{(\infty)}}
182: {{\cal F}_\text{in}^{(\infty)}}=
183: \frac{{\cal F}_\text{in}^{(\text{h})}}{{\cal F}_\text{in}^{(\infty)}}\,,
184: \label{absorption}
185: \end{equation}
186: in terms of the out-going and in-going energy fluxes evaluated either at infinity or at the horizon.
187: The two definitions are related by simple energy conservation and lead to
188: the same results. The choice of which is used for the determination of the absorption coefficient may depend on numerical issues.
189:
190: The flux and power spectra (equations~(\ref{4flux}) and (\ref{4power}) respectively) of the Hawking radiation emitted on the brane can be computed, for massless particles, from the
191: 4-dimensional expressions:
192: %%%%%%%%%%%%%
193: \begin{align}
194: \frac{d \hat N^{(s)}(\om)}{dt}&= \sum_{\ell} (2\ell +1) |\hat {\cal A}^{(s)}_\ell|^2\,
195: \frac{1}{\exp\left(\om/T_\text{H}\right) \mp 1}\,\frac{d\om}{2\pi}\,;
196: \label{fdecay-brane}\\
197: \frac{d \hat E^{(s)}(\om)}{dt}&= \sum_{\ell} (2\ell +1) |\hat {\cal A}^{(s)}_\ell|^2\,
198: \frac{\om}{\exp\left(\om/T_\text{H}\right) \mp 1}\,\frac{d\om}{2\pi}\,.
199: \label{pdecay-brane}
200: \end{align}
201: %%%%%%%%%%%%%
202: Note, however, that the absorption coefficient $\hat {\cal A}_\ell^{(s)}$ still depends on the number of extra dimensions since the projected metric tensor of eq.~(\ref{non-rot}) carries a signature of the dimensionality of the bulk space-time through the expression of the metric function $h(r)$.
203:
204: \subsection{Analytic calculation}
205:
206: The grey-body factors have been determined analytically for the 4-dimensional
207: case in refs. \cite{Page:1976df,Sanchez:1978si,Sanchez:1978vz} both for a rotating and non-rotating black
208: hole. In the $(4+n)$-dimensional case, refs. \cite{Kanti:2002nr,Kanti:2002ge} have provided
209: analytic expressions for grey-body factors for the emission of scalars,
210: fermions and gauge bosons from a higher-dimensional Schwarzschild-like
211: black hole. Both bulk and brane emission were considered for an arbitrary number of extra dimensions $n$. Reference \cite{Ida:2002ez} presented
212: analytic results for brane-localized emission from a Kerr-like
213: black hole, but only in the particular case of $n=1$. All the above results were derived
214: in the low-energy approximation where the procedure used is as follows:
215:
216: \begin{itemize}
217: \item{Find an analytic solution in the near-horizon regime and expand it as
218: in-going and out-going waves so that the $A^{(\text{h})}$ coefficients can be extracted;}
219: \item{Apply the boundary condition on the horizon so that the wave is purely
220: out-going;}
221: \item{Find an analytic solution in the far-field regime and again expand it as
222: in-going and out-going waves;}
223: \item{Match the two solutions in the intermediate regime;}
224: \item{Extract $|{\cal A}_\ell|^2$ and expand it in powers of $\omega r_\text{h}$.}
225: \end{itemize}
226:
227: The solution obtained by following the above approximate method is a power
228: series in $\omega r_\text{h}$, which is only valid for low energies and expected to
229: significantly deviate from the exact result as the energy of the emitted
230: particle increases (this was pointed out in \cite{Page:1976df,MacGibbon:1990zk} for the
231: 4-dimensional case). In \cite{Kanti:2002ge}, the full analytic result for the absorption
232: coefficient (before the final expansion was made) was used for the evaluation
233: of the emission rates for all particle species.\footnote{This full analytic approach was pursued in \cite{Kanti:2002ge} as a result of some of the numerical results presented in this thesis.} The range of validity of these
234: results, although improved compared to the power series, was still limited since the assumption of small
235: $\omega r_\text{h}$ was {\it still} made during the matching of the two solutions
236: in the intermediate regime. As the equations of motion are too complicated to solve analytically for all values of $\om r_\text{h}$, it becomes clear that only an exact numerical analysis can give the full exact results for grey-body factors and emission rates.
237:
238: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%5
239:
240: \subsection{Numerical calculation}
241: \label{numcal}
242:
243: There are various numerical issues which arise while trying to do the full
244: calculation of the absorption coefficient and the complexity of these problems
245: depends strongly on the spin $s$ of the emitted particle. The usual numerical
246: procedure starts by applying the relevant horizon boundary condition (a
247: vanishing out-going wave) to $R_s (r)$. Then, the solution is integrated out
248: to `infinity' and the asymptotic coefficient $A_\text{in}^{(\infty)}$ is extracted
249: in terms of which ${\cal A}_\ell^{(s)}$ can be calculated.
250:
251: By looking at the asymptotic solution at infinity in eq.~(\ref{far}) we can
252: see that, for the scalar case, the in-going and out-going waves are of
253: comparable magnitude. Therefore it is relatively easy to extract the
254: coefficients at infinity and thus determine the grey-body factor. However for
255: fields with non-vanishing spin, i.e.\ fermions, vector bosons and
256: gravitons, this is not an easy task. First of all, different components carry
257: a different
258: part of the emitted field: the upper component $\Psi_{+s}$ consists mainly of
259: the in-going wave with the out-going one being greatly suppressed, while for
260: the lower component $\Psi_{-s}$ the situation is reversed (for a field with
261: spin $s \neq 0$, only the upper and lower components are radiative).
262: Distinguishing between the two parts of the solution (in-going
263: and out-going) is not an easy task no matter which component is used, and it
264: becomes more difficult as the magnitude of the spin increases. In addition,
265: the choice of either positive or negative $s$ to extract the grey-body factor (i.e.\ using either the upper or the lower component) affects
266: the numerical issues.
267:
268: If $s$ is negative, then the horizon boundary condition is easy to apply
269: because components of the in-going solution on the horizon will be
270: exponentially suppressed, as can be seen from eq.~(\ref{near}). However eq.~(\ref{far}) shows that negative $s$ also means the out-going solution at infinity is enhanced by $r^{-2s}$
271: with respect to the in-going one; this makes accurate
272: determination of $A_\text{in}^{(\infty)}$ very difficult. Hence the negative $s$
273: approach is not used in this work.
274:
275: On the other hand if $s$ is positive, it is easy to accurately extract
276: $A_\text{in}^{(\infty)}$ because the out-going solution at infinity is suppressed by $r^{-2s}$. However, close to the horizon the out-going solution is
277: exponentially smaller than the in-going one.
278: Therefore the solution for $R_s(r)$ can be easily contaminated by
279: components of the out-going solution. This problem becomes worse for larger
280: $s$ and first becomes significant for $s=1$.
281: \enlargethispage{-\baselineskip}
282:
283: For $n=0$, various methods (see e.g.\ \cite{Press:1974}) have been used to
284: solve the numerical problems which arise in the gauge boson case. The
285: approach of Bardeen in \cite{Press:1974} is also applicable for $n=1$; some details of this method are briefly outlined below.
286:
287: The method starts from a mathematically equivalent form of the radial equation in Kerr in-going co-ordinates:\footnote{This was derived, for general $n$, in \cite{Kantiprivate} and for $n=0$ reduces to the result found in \cite{Press:1974}.}
288: %%%%%%%%%%%
289: \begin{equation}
290: \Delta^{-s}\frac{d}{dr} \left(\Delta^{s+1}\frac{dR_s}{dr}\right)-2 i r^2\omega \frac{dR_s}{dr}-\left[2(2s+1)i\omega r-s\Delta''+{}_s\Lambda_{\ell}\right]R_s=0\,,
291: \end{equation}
292: %%%%%%%%%%%
293: where ${}_s\Lambda_{\ell}=\ell(\ell+1)-s(s-1)$. The asymptotic solutions are now different from those of eqs.~(\ref{near}) and (\ref{far}); they are given by
294: %%%%%%%%
295: \begin{equation}
296: R_s^{(\text{h})}=B_\text{in}^{(\text{h})}+B_\text{out}^{(\text{h})}\Delta^{-s} e^{2i\omega r^{*}},
297: \end{equation}
298: %%%%%%%%%%
299: and
300: %%%%%%%%%%
301: \begin{equation}
302: R_s^{(\infty)}=B_\text{in}^{(\infty)}\frac{1}{r^{2s+1}}+B_\text{out}^{(\infty)} \frac{e^{2i\omega r}}{r}\,.
303: \end{equation}
304: %%%%%%%%%%%%
305:
306: We can follow the method in Appendix A of \cite{Press:1974} and apply the co-ordinate transformation
307: %%%%%%%%%
308: \begin{equation}
309: x=\frac{r}{r_\text{h}}-1\,.
310: \end{equation}
311: %%%%%%%%%
312: This transformation is easily applicable for either $n=0$ or $n=1$ since in both cases $\Delta''=2$. In the case $s=-1$ (which is mathematically equivalent to the $s=1$ discussed earlier) applying this transformation for $n=0$ gives
313: %%%%%%%%%%
314: \begin{equation}
315: \label{xradial}
316: x(x+1)\frac{d^2R_{-1}}{dx^2}-2i\omega r_\text{h}\left(x^2+2x+1\right)\frac{dR_{-1}}{dx}-\left[-2i\omega r_\text{h} (x+1)+\lambda\right]R_{-1}=0\,,
317: \end{equation}
318: %%%%%%%%%%
319: where $\lambda=\ell(\ell+1)$. The asymptotic solutions can now be written in terms of $x$, with more care being taken over sub-leading terms in the in-going solution at the horizon:
320: %%%%%%%%%%%
321: \begin{equation}
322: R_{-1}^{(\text{h})}=B_\text{in}^{(\text{h})} (1+a_1x+a_2x^2)+B_\text{out}^{(\text{h})} r_\text{h}^2 e^{2i\omega r^*}x(1+b_1x+\ldots)\,.
323: \end{equation}
324: %%%%%%%%%%%%
325:
326: It is easy to obtain the coefficients $a_1$ and $a_2$ by substituting the in-going solution into the radial equation and comparing terms in different powers of $x$. This leads to
327: %%%%%%%%%%
328: \begin{equation}
329: a_1=1+i\frac{\lambda}{2\omega r_\text{h}}\,,
330: \label{a1n0}
331: \end{equation}
332: %%%%%%%%%%
333: and
334: %%%%%%%%%%%
335: \begin{equation}
336: a_2=\frac{1}{2(1-2i\omega r_\text{h})}\left[(\lambda+2i\omega r_\text{h})a_1-2i\omega r_\text{h}\right].
337: \label{a2n0}
338: \end{equation}
339: %%%%%%%%%%%%
340: Instead of solving ${\mathcal L}R=0$, the equation ${\mathcal L}Y=f$ is solved, where $Y=R-(1+a_1x)$. This means that
341: %%%%%%%%%%%
342: \begin{equation}
343: f=2(1-2i\omega r_\text{h})a_2x\,.
344: \label{fn0}
345: \end{equation}
346: %%%%%%%%%%%%%
347: Close to the horizon the leading-order terms in $Y$ and $Y'$ are $a_2x^2$ and $2a_2x$---these are used to apply the horizon boundary condition. Bardeen found that the equation in this form can now be stably integrated from $x=0$ to $x=\infty$. The $B_\text{in}$ and $B_\text{out}$ coefficients extracted can be used to obtain the grey-body factors using similar formulae to those discussed later for $A_\text{in}$ and $A_\text{out}$.
348:
349: For $n=1$, equations~(\ref{xradial}) and (\ref{a1n0})--(\ref{fn0}) are slightly modified and become
350: %%%%%%%%%%
351: \begin{gather}
352: x(x+2)\frac{d^2R_{-1}}{dx^2}-2i\omega r_\text{h}\left(x^2+2x+1\right)\frac{dR_{-1}}{dx}-\left[-2i\omega r_\text{h} (x+1)+\lambda\right]R_{-1}=0\,;\\
353: a_1=1+i\frac{\lambda}{2\omega r_\text{h}}\,;
354: \label{a1n1}\\
355: a_2=\frac{1}{4(1-i\omega r_\text{h})}\left[(\lambda+2i\omega r_\text{h})a_1-2i\omega r_\text{h}\right];
356: \label{a2n1}\\
357: f=4(1-i\omega r_\text{h})a_2x\,.
358: \label{fn1}
359: \end{gather}
360: %%%%%%%%%%%%%
361:
362: Unfortunately no analogous method has been found for $n \geq 2$ because there is then no simple equivalent of eq.~(\ref{xradial}). Instead, an
363: alternative transformation of the radial equation~(\ref{radial}) was used.
364: Writing $y=r/r_\text{h}$ and $R_1=y\,F(y)\,e^{-i\omega r^*}$, the wave equation becomes
365: %%%%%%%%%%%%
366: \begin{equation}
367: (hy^2)\,\frac{d^2F}{dy^2}+2y\,(h-i\omega r_\text{h} y)\,\frac{dF}{dy}-\ell(\ell+1)F=0\,.
368: \end{equation}
369: %%%%%%%%%%%%
370: Since on the horizon $y=1$ and $h=0$, the boundary conditions become $F(1)=1$
371: and
372: %%%%%%%%%%%%
373: \begin{equation}
374: \left. \frac{dF}{dy} \right|_{y=1}=\frac{i \ell(\ell+1)}{2 \omega r_\text{h}}\,.
375: \end{equation}
376: %%%%%%%%%%%%%%
377: Using both this method and the Bardeen method for the $n=1$ case provided a useful cross-check on the spin-1 numerical results.
378:
379: For fermions no special transformation was necessary and therefore the radial
380: equation~(\ref{radial}) was used. However the application of the
381: boundary condition at the horizon is made slightly easier by the
382: transformation $P_s=\Delta^s R_s$ so that the asymptotic solution at
383: the horizon becomes
384: %%%%%%%%%%%
385: \begin{equation}
386: P_s^{(\text{h})}=A_\text{in}^{(\text{h})}\,e^{-i\omega r^{*}}+
387: A_{out }^{(\text{h})}\,\Delta^s \,e^{i\omega r^{*}}.
388: \label{nh-new}
389: \end{equation}
390: %%%%%%%%%%%%
391: Since we require $A_\text{out}^{(\text{h})}=0$, a suitable boundary condition to apply as $r \rightarrow r_\text{h}$ is that
392: %%%%%%%%%%
393: \begin{equation}
394: P_s=1\,,
395: \end{equation}
396: %%%%%%%%%%%%%%
397: while using eq.~(\ref{rstarr}) we also obtain
398: %%%%%%%%%%%%%%%
399: \begin{equation}
400: \frac{dP_s}{dr}=-i\omega\, \frac{dr^*}{dr}=-\frac{i\omega}{h(r)}\,.
401: \end{equation}
402: %%%%%%%%%%%%
403: The above boundary conditions ensure that $|A_\text{in}^{(\text{h})}|^2=1$.
404: The asymptotic form for $P_s$ at infinity now looks like
405: %%%%%%%%%%%%
406: \begin{equation}
407: P_s^{(\infty)}=A_\text{in}^{(\infty)}\,\frac{e^{-i\omega r}}{r^{1-2s}}+
408: A_\text{out}^{(\infty)}\, \frac{e^{i\omega r}}{r}\,
409: \label{ff-new}
410: \end{equation}
411: %%%%%%%%%%%
412: since $\Delta \rightarrow r^2$ as $r \rightarrow \infty$.
413:
414: The numerical work described in this chapter was almost entirely performed using the \texttt{NDSolve} package in \texttt{Mathematica}. However in some cases Fortran programs calling {\small NAG} routines (specifically \texttt{D02EJF} and its associated routines) were used as a check on the numerical robustness of the results obtained in \texttt{Mathematica}. The {\small NAG} routines implement the Backward Differentiation Formulae (BDF) whilst in \texttt{Mathematica} either the BDF method or the implicit Adams method is used.
415:
416: There are various considerations which must be taken into account in order to
417: obtain results to the desired accuracy (at least three significant figures).
418: Firstly, although the horizon boundary condition cannot be applied exactly at
419: $r_\text{h}$ (due to singularities in the boundary condition and the differential
420: equation) the error introduced by applying the condition at $r=r_\text{n}$ (where $\Delta r_\text{h}=r_\text{n}-r_\text{h} \ll r_\text{h}$) must be small. This can be investigated by studying changes in the
421: grey-body factors for order-of-magnitude changes in $\Delta r_\text{h}$. Similarly it
422: must be verified that the value $r_{\infty}$ used as an approximation for `infinity'
423: does not introduce errors which will affect the accuracy of the result. This is illustrated in Figure~\ref{farstab} in which $\hat{\sigma}^{(0)}_\text{abs}=\sum_{\ell}\hat{\sigma}^{(0)}_{\ell}$ is plotted in units of $\pi r_\text{h}^2$. The oscillations in the numerical result die away as the asymptotic form given in eq.~(\ref{far}) is obtained. A value of $r_{\infty}=10000$ was used in the work that follows to ensure the grey-body factor was accurate to at least three significant figures.
424:
425: \begin{figure}
426: \unitlength1cm
427: \psfrag{x}[][][1.4]{$r_{\infty}$}
428: \psfrag{y}[][][1.4]{$\hat{\sigma}^{(0)}_\text{abs}/\pi r_\text{h}^2$}
429: \begin{center}
430: \begin{minipage}[t]{3.35in}
431: \scalebox{0.65}{\rotatebox{0}{\includegraphics{nfars0n0.eps}}}
432: \end{minipage}
433: \hfill
434: \begin{minipage}[t]{2.75in}
435: \scalebox{0.65}{\rotatebox{0}{\includegraphics{nfars0n0mag.eps}}}
436: \end{minipage}
437: \capbox{Stability of grey-body factor for changes in `infinity'}{Stability of the $n=0$, $s=0$, $\omega r_\text{h}=0.5$ grey-body factor as $r_{\infty}$ is changed. The right plot is a magnified version of part of the left plot to illustrate the accuracy obtained.\label{farstab}}
438: \end{center}
439: \end{figure}
440:
441: Care must also be taken that the numerical integration procedure is sufficiently accurate that significant integration errors are avoided even at large values of $r$. Plots like those in Figure~\ref{farstab} are also useful for investigating this because they can be used to confirm that there is no significant gradient on a straight line drawn through the centre of the oscillations.
442:
443: Finally, for each energy being considered, enough angular momentum
444: modes must be included in the summation so that only modes which do not
445: contribute significantly are neglected. For any value of $\ell$ the absorption coefficient will approach unity at high enough energies and the $(2\ell+1)$ factor in eq.~(\ref{brane-loc}) means that higher angular momentum modes will start to dominate the grey-body factor. It was found that for the highest values of $\omega r_\text{h}$ considered in this work it was necessary to include contributions from in excess of ten angular momentum modes.
446:
447: \section{Numerical results for brane emission}
448: \label{numres}
449:
450: In this section, results are presented for grey-body factors and emission rates for brane-localized scalar, fermion and gauge boson
451: fields, as obtained by numerically solving the corresponding equations of
452: motions. The relationship between the absorption coefficient ${\cal A}_\ell^{(s)}$
453: and the $A$ coefficients in equations~(\ref{nh-new}) and (\ref{ff-new}) is different in each case, so each will be considered separately.
454:
455: \subsection{Spin 0 fields}
456: \label{numscalar}
457:
458: The numerical integration of eq.~(\ref{radial}) for $s=0$ yields the solution
459: for the radial function $R_0(r)$ which smoothly interpolates between the
460: asymptotic solutions of eqs.~(\ref{nh-new}) and (\ref{ff-new}) in the near-horizon
461: and far-field regimes respectively. The absorption coefficient is easily
462: defined in terms of the in-going and out-going energy fluxes at infinity,
463: or equivalently by the corresponding wave amplitudes ($A_\text{in}^{(\infty)}$ and $A_\text{out}^{(\infty)}$) in the same asymptotic regime.
464: Therefore eq.~(\ref{absorption}) may be written in the form \cite{Kanti:2002nr}
465: %%%%%%%%%%%
466: \begin{equation}
467: |\hat {\cal A}^{(0)}_\ell|^2=1-|\hat {\cal R}^{(0)}_\ell|^2=
468: 1-\Biggl|\frac{A_\text{out}^{(\infty)}}{A_\text{in}^{(\infty)}}\Biggr|^2,
469: \label{scalars}
470: \end{equation}
471: %%%%%%%%%%
472: where $\hat {\cal R}_\ell$ is the corresponding reflection coefficient.
473:
474: \begin{figure}
475: \begin{center}
476: \psfrag{x}[][][1.75]{$\omega r_\text{h}$}
477: \psfrag{y}[][][1.75]{$\hat{\sigma}^{(0)}_\text{abs}(\om)/\pi r_\text{h}^2$}
478: %\psfrag{n=0}[][][1.75]{$n=0$}\psfrag{n=1}[][][1.75]{$n=1$}\psfrag{n=2}[][][1.75]{$n=2$}\psfrag{n=6}[][][1.75]{$n=6$}
479: %\psfrag{0}[][][1.75]{0}\psfrag{0.2}[][][1.75]{0.2}\psfrag{0.4}[][][1.75]{0.4}\psfrag{0.6}[][][1.75]{0.6}\psfrag{0.8}[][][1.75]{0.8}\psfrag{1}[][][1.75]{1}\psfrag{1.2}[][][1.75]{1.2}\psfrag{2}[][][1.75]{2}\psfrag{4}[][][1.75]{4}\psfrag{6}[][][1.75]{6}\psfrag{8}[][][1.75]{8}\psfrag{10}[][][1.75]{10}
480: \scalebox{0.5}
481: {\rotatebox{-90}{\includegraphics[width=\textwidth]{ngrey0.ps}}}
482: \capbox{Grey-body factors for scalars on the brane}{Grey-body factors for scalar emission on the brane from a $(4+n)$D
483: black hole.\label{grey0}}
484: \end{center}
485: \end{figure}
486:
487: The plot presented in Figure~\ref{grey0} shows, for several values of $n$,
488: the grey-body factor for the emission of scalar fields on the brane (for completeness, the plots include values of $n$ ruled out
489: on astrophysical grounds, i.e.\ $n=1$ and 2).\footnote{Throughout the numerical analyses the horizon radius $r_\text{h}$ is an
490: arbitrary input parameter which remains fixed.} The grey-body
491: factor is obtained by using eq.~(\ref{brane-loc}) and summing over the angular momentum number $\ell$.
492:
493: For $n=0$ and
494: $\omega r_\text{h} \rightarrow 0$, the grey-body factor assumes a non-zero value
495: which is equal to $4\pi r_\text{h}^2$---that is, the grey-body factor for scalar
496: fields with a very low energy is given exactly by the area of the black hole
497: horizon. As the energy increases, the factor soon starts oscillating
498: around the geometrical optics
499: limit $\hat{\sigma}_\text{g} = 27 \pi r_\text{h}^2/4$ which corresponds to the spectrum of a
500: black body with an absorbing area of radius $r_\text{c}=3 \sqrt{3}\,r_\text{h}/2$
501: \cite{Sanchez:1978si,Sanchez:1978vz,misner}. The $n=0$ result agrees exactly with Page's result presented in Figure 1 of \cite{MacGibbon:1990zk} (this is also found to be the case for the fermion and gauge boson grey-body factors shown later).
502: \enlargethispage{-\baselineskip}
503:
504: In the extra-dimensional case, the grey-body factor has an asymptotic low-energy value which is the same for all values of $n$,
505: and at high energies it again starts oscillating around a limiting value. This is always lower than the 4-dimensional geometrical optics limit because the effective radius $r_\text{c}$ depends on the dimensionality of the bulk space-time through the metric tensor of the projected space-time in which the particle moves. For arbitrary $n$, it adopts the value \cite{Emparan:2000rs}
506: %%%%%%%%%%%
507: \begin{equation}
508: r_\text{c}=\biggl(\frac{n+3}{2}\biggr)^{1/n+1}\,\sqrt{\frac{n+3}{n+1}}\,\,r_\text{h}\,.
509: \label{effective}
510: \end{equation}
511: %%%%%%%%%%%%
512: The above quantity is a strictly decreasing function of $n$ which causes the
513: asymptotic grey-body factor, $\hat{\sigma}_\text{g}=\pi r_\text{c}^2$\,, to become more and more
514: suppressed as the number of extra dimensions projected onto the
515: brane increases. The values of $\hat{\sigma}_\text{g}$ are tabulated in Table~\ref{brgolim} for different values of $n$.\footnote{For the larger values of $n$ it was found that these asymptotic values are not approached until relatively high energies, typically $\omega r_\text{h} \sim n$.}
516:
517: %%%%%%%%%%%%%%%%%%%
518: \begin{table}
519: \def\arraystretch{1.1}
520: \begin{center}
521: \begin{tabular}{|c|c|c|c|c|c|c|c|c|}
522: \hline
523: $n$ & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 \\
524: \hline
525: $\hat{\sigma}_\text{g}/\pi r_\text{h}^2$ & 6.75 & 4 & 3.07 & 2.60 & 2.31 & 2.12 & 1.98 & 1.87\\
526: \hline
527: \end{tabular}
528: \capbox{High-energy limits of grey-body factors for brane emission}{High-energy limits of grey-body factors for brane emission, given in units of $\pi r_\text{h}^2$.\label{brgolim}}
529: \end{center}
530: \def\arraystretch{1.0}
531: \end{table}
532: %%%%%%%%%%%%%%%%%%
533:
534: The power series expression of the grey-body factor determined in \cite{Kanti:2002nr} matches the exact solution only in a very limited low-energy
535: regime. In the limit $\omega r_\text{h} \rightarrow 0$, the asymptotic
536: value $4\pi r_\text{h}^2$ is recovered as expected; however, as the energy increases
537: the exact solution rapidly deviates from the behaviour dictated by the
538: dominant term in the $\omega r_\text{h}$ expansion. This was confirmed in \cite{Kanti:2002ge} where full analytic results for the grey-body
539: factors were plotted, although still in the low-energy approximation. The behaviour depicted in Figure 3 of ref. \cite{Kanti:2002ge} is much closer to the exact one, shown here in Figure~\ref{grey0}, and
540: successfully reproduces some of the qualitative features including the
541: suppression of the grey-body factor as the dimensionality of the bulk space-time increases. Nevertheless, as previously stated, even this result rapidly breaks down as the energy increases; it is particularly unreliable for the $n=0$ case and fails to reproduce the high-energy oscillations for any value of $n$. This means that the exact numerical solution obtained in this work is the only reliable source of information concerning the form of the extra-dimensional grey-body factor throughout the energy regime.
542:
543: \begin{figure}
544: \begin{center}
545: \psfrag{x}[t][b][1.75]{$\omega r_\text{h}$}
546: \psfrag{y}[][][1.75]{$r_\text{h} d^2 \hat{E}^{(0)}/dtd\om$}
547: %\psfrag{n=0}[][][1.75]{$n=0$}\psfrag{n=1}[][][1.75]{$n=1$}\psfrag{n=2}[][][1.75]{$n=2$}\psfrag{n=4}[][][1.75]{$n=4$}\psfrag{n=6}[][][1.75]{$n=6$}
548: \scalebox{0.5}{\rotatebox{0}{\includegraphics[width=23cm, height=15cm]{nscalars-dec.eps}}}
549: \capbox{Power emission for scalars on the brane}{Energy emission rates for scalar fields on the brane from
550: a $(4+n)$D black hole.\label{sc-dec}}
551: \end{center}
552: \end{figure}
553:
554: The numerical solution for the grey-body factor allows the computation of the energy emission rate for scalar fields on the brane. Using eq.~(\ref{pdecay-brane}) it is found that the suppression of the grey-body factor with $n$ does not lead to the suppression of the emission rate itself. The behaviour of the differential energy emission rate is given in Figure~\ref{sc-dec}. As $n$ increases, the increase in the temperature of the black hole, and therefore in its emitting power, overcomes the decrease in the grey-body factor and leads to a substantial enhancement of the energy emission rate. Figure~\ref{sc-dec} clearly shows that the enhancement of the peak of the emission curve can be up to several orders of magnitude compared to the 4-dimensional case. In addition the peak in the spectrum is shifted to higher values of the energy parameter $\omega r_\text{h}$ since for fixed $r_\text{h}$ an increase in $n$ corresponds to an increase in the temperature of the radiating body.
555:
556: In order to quantify the enhancement of the emission rate as the number of extra dimensions projected onto the brane increases, the total flux and power emissivities were computed, for a range of values of $n$, by integrating eqs.~(\ref{4flux}) and (\ref{4power}) with respect to $\om$. The results obtained are displayed in Table~\ref{sc-tab}. The relevant emissivities for different values of $n$ have been normalized in terms of those for $n=0$. From the entries of this table, the order-of-magnitude enhancement of both the flux and power radiated by the black hole as $n$ increases is again clear.
557: %%%%%%%%%%%%%%%%%%%
558: \begin{table}
559: \begin{center}
560: \begin{tabular}{|l|c|c|}
561: \hline
562: $n$& Flux & Power\\
563: \hline
564: 0 & 1 & 1\\
565: 1 & 4.75 & 8.94\\
566: 2 & 13.0 & 36.0\\
567: 3 & 27.4 & 99.8\\
568: 4 & 49.3 & 222\\
569: 5 & 79.9 & 429\\
570: 6 & 121 & 749\\
571: 7 & 172 & 1220\\
572: \hline
573: \end{tabular}
574: \capbox{Emissivities for scalar fields on the brane}{Flux and power emissivities for scalar fields on the brane.\label{sc-tab}}
575: \end{center}
576: \end{table}
577:
578: %%%%%%%%%%%%%%%%%%
579:
580:
581:
582: \subsection{Spin 1/2 fields}
583:
584: Unlike the case of scalar fields, the study of the emission of fields with
585: non-vanishing spin involves, in principle, fields with more than
586: one component. Equation~(\ref{radial}) depends on the helicity number $s$ which can be either $+$1/2 or $-$1/2 and this leads to radial equations for the two different components of the field. As mentioned in \secref{numcal}, the upper ($+$1/2) and lower ($-$1/2) components carry mainly the in-going and out-going parts of the field respectively. Although knowledge of both components is necessary to construct the complete solution for the emitted field, the determination of either is sufficient to compute the absorption coefficient
587: $\hat {\cal A}_j^{(s)}$ ($j$ is the total angular momentum number). For
588: example, if the in-going wave is known in the case of the emission of
589: fields with spin $s=$1/2, eq.~(\ref{absorption})
590: may be directly written as \cite{Kanti:2002ge,Cvetic:1998ap}
591: %%%%%%%%%%%
592: \begin{equation}
593: |\hat {\cal A}^{(1/2)}_j|^2=\Biggl|\frac{A_\text{in}^{(\text{h})}}
594: {A_\text{in}^{(\infty)}}\Biggr|^2.
595: \label{fermions}
596: \end{equation}
597: %%%%%%%%%%
598: The above follows by defining the incoming flux of a fermionic field as the
599: radial component of the conserved current,
600: $J^\mu=\sqrt{2}\,\sigma^\mu_{AB}\,\Psi^A\,\bar \Psi^B$, integrated over a
601: two-dimensional sphere and evaluated at both the horizon and infinity.
602:
603: The grey-body factor is again related to the absorption probability through eq.~(\ref{brane-loc}) with $\ell$ now being replaced by the total angular momentum $j$. Numerically solving the radial equation~(\ref{radial}) and computing $\hat {\cal A}^{(1/2)}_j$ gives the behaviour of the grey-body factor, in terms of the energy parameter $\omega r_\text{h}$ and the number of extra dimensions $n$. This is shown in Figure~\ref{grey05} for four different values of $n$.
604:
605: As in the scalar case, at low energies the grey-body factor assumes a n-zero asymptotic value; this depends on the dimensionality of
606: space-time and increases with $n$. The enhancement of
607: $\hat{\sigma}^{(1/2)}_\text{abs}(\omega)$ with $n$ in the low-energy regime continues up to intermediate values of $\omega r_\text{h}$ where the situation is reversed: as $n$ becomes larger, the high-energy grey-body factor becomes more and more suppressed as was found in the scalar case. The full analytic results
608: derived in \cite{Kanti:2002ge} provide a reasonable description of the low-energy behaviour except in the $n=0$ case; however, as expected, they fail to
609: give accurate information for the high-energy regime. As for scalar fields, the high-energy grey-body factors for fermions are shown to oscillate around asymptotic values determined by the effective radius of eq.~(\ref{effective}).
610:
611: \begin{figure}
612: \begin{center}
613: \psfrag{x}[][][1.75]{$\omega r_\text{h}$}
614: \psfrag{y}[][][1.75]{$\hat{\sigma}^{(1/2)}_\text{abs}(\om)/\pi r_\text{h}^2$}
615: %\psfrag{n=0}[][][1.75]{$n=0$}\psfrag{n=1}[][][1.75]{$n=1$}\psfrag{n=2}[][][1.75]{$n=2$}\psfrag{n=6}[][][1.75]{$n=6$}
616: %\psfrag{0}[][][1.75]{0}\psfrag{0.2}[][][1.75]{0.2}\psfrag{0.4}[][][1.75]{0.4}\psfrag{0.6}[][][1.75]{0.6}\psfrag{0.8}[][][1.75]{0.8}\psfrag{1}[][][1.75]{1}\psfrag{1.2}[][][1.75]{1.2}\psfrag{2}[][][1.75]{2}\psfrag{4}[][][1.75]{4}\psfrag{6}[][][1.75]{6}\psfrag{8}[][][1.75]{8}\psfrag{10}[][][1.75]{10}
617: \scalebox{0.5}{\rotatebox{-90}{\includegraphics[width=\textwidth]{ngrey05.ps}}}
618: \capbox{Grey-body factors for fermions on the brane}{Grey-body factors for fermion emission on the brane from a $(4+n)$D
619: black hole.\label{grey05}}
620: \end{center}
621: \end{figure}
622:
623: \begin{figure}
624: \begin{center}
625: \psfrag{x}[t][b][1.75]{$\omega r_\text{h}$}
626: \psfrag{y}[][][1.75]{$r_\text{h} d^2 \hat{E}^{(1/2)}/dtd\om$}
627: %\psfrag{n=0}[][][1.75]{$n=0$}\psfrag{n=1}[][][1.75]{$n=1$}\psfrag{n=2}[][][1.75]{$n=2$}\psfrag{n=4}[][][1.75]{$n=4$}\psfrag{n=6}[][][1.75]{$n=6$}
628: \scalebox{0.5}{\rotatebox{0}{\includegraphics[width=23cm, height=15cm]
629: {nfermions-dec.eps}}}
630: \capbox{Power emission for fermions on the brane}{Energy emission rates for fermions on the brane from a $(4+n)$D black hole.\label{fer-dec}}
631: \end{center}
632: \end{figure}
633: %
634: The energy emission rate for fermion fields on the brane is shown in Figure~\ref{fer-dec} for various values of $n$.
635: As $n$ increases, the power emission is found to be significantly enhanced at both low and high energies. The emission curves show the same features as for the emission of scalar fields, i.e.\ the peak increases in height by several orders of magnitude and shifts towards higher energies (due to the increasing black hole temperature). Quantitative results regarding the enhancement with $n$ of both the flux and power fermion emission spectra were obtained by integrating the data in Figure~\ref{fer-dec}; they are shown in Table~\ref{fer-tab}. Once again, the enhancement is substantial, and in fact is even more significant than for scalar emission.
636: %%%%%%%%%%%%%%%%%%%
637: \begin{table}[b]
638: \begin{center}
639: \begin{tabular}{|l|c|c|}
640: \hline
641: $n$& Flux & Power\\
642: \hline
643: 0 & 1 & 1\\
644: 1 & 9.05 & 14.2\\
645: 2 & 27.6 & 59.5\\
646: 3 & 58.2 & 162\\
647: 4 & 103 & 352\\
648: 5 & 163 & 664\\
649: 6 & 240 & 1140\\
650: 7 & 335 & 1830\\
651: \hline
652: \end{tabular}
653: \capbox{Emissivities for fermions on the brane}{Flux and power emissivities for fermions on the brane.\label{fer-tab}}
654: \end{center}
655: \end{table}
656: %%%%%%%%%%%%%%%%%%
657:
658: \subsection{Spin 1 fields}
659:
660: In the case of the emission of gauge boson fields, the incoming flux can be
661: computed from the energy-momentum tensor $T^{\mu\nu}=2 \sigma^{\mu}_{AA'} \sigma^{\nu}_{BB'} \Psi^{AB}\,\bar\Psi^{A'B'}$; the time-radial component is
662: integrated over a two-dimensional sphere and then evaluated at the horizon
663: and at infinity. By making use of the solution for the in-going wave, the
664: following expression for the absorption probability of eq.~(\ref{absorption}),
665: is obtained \cite{Kanti:2002ge,Cvetic:1998ap}:
666: %%%%%%%%%%%
667: \begin{equation}
668: |\hat {\cal A}^{(1)}_j|^2=\frac{1}{r_\text{h}^2}\,
669: \Biggl|\frac{A_\text{in}^{(\text{h})}}{A_\text{in}^{(\infty)}}\Biggr|^2.
670: \label{bosons}
671: \end{equation}
672: %%%%%%%%%%
673:
674: \begin{figure}
675: \begin{center}
676: \psfrag{x}[][][1.75]{$\omega r_\text{h}$}
677: \psfrag{y}[][][1.75]{$\hat{\sigma}^{(1)}_\text{abs}(\om)/\pi r_\text{h}^2$}
678: %\psfrag{n=0}[][][1.75]{$n=0$}\psfrag{n=1}[][][1.75]{$n=1$}\psfrag{n=2}[][][1.75]{$n=2$}\psfrag{n=6}[][][1.75]{$n=6$}
679: %\psfrag{0}[][][1.75]{0}\psfrag{0.2}[][][1.75]{0.2}\psfrag{0.4}[][][1.75]{0.4}\psfrag{0.6}[][][1.75]{0.6}\psfrag{0.8}[][][1.75]{0.8}\psfrag{1}[][][1.75]{1}\psfrag{1.2}[][][1.75]{1.2}\psfrag{2}[][][1.75]{2}\psfrag{4}[][][1.75]{4}\psfrag{6}[][][1.75]{6}\psfrag{8}[][][1.75]{8}\psfrag{10}[][][1.75]{10}
680: \scalebox{0.5}{\rotatebox{-90}{\includegraphics{ngrey1.ps}}}
681: \capbox{Grey-body factors for gauge bosons on the brane}{Grey-body factors for gauge boson emission on the brane from a $(4+n)$D black hole.\label{grey1}}
682: \end{center}
683: \end{figure}
684:
685:
686: \begin{figure}
687: \begin{center}
688: \psfrag{x}[t][b][1.75]{$\omega r_\text{h}$}
689: \psfrag{y}[][][1.75]{$r_\text{h} d^2 \hat{E}^{(1)}/dtd\om$}
690: %\psfrag{n=0}[][][1.75]{$n=0$}\psfrag{n=1}[][][1.75]{$n=1$}\psfrag{n=2}[][][1.75]{$n=2$}\psfrag{n=4}[][][1.75]{$n=4$}\psfrag{n=6}[][][1.75]{$n=6$}
691: \scalebox{0.5}{\rotatebox{0}{\includegraphics[width=23cm, height=15cm]
692: {ngauge-dec.eps}}}
693: \capbox{Power emission for gauge fields on the brane}{Energy emission rates for gauge fields on the brane from a $(4+n)$D
694: black hole.\label{gauge-dec}}
695: \end{center}
696: \end{figure}
697:
698: The exact results for the grey-body factors and energy emission rates for
699: gauge boson fields are given in Figures~\ref{grey1} and \ref{gauge-dec}
700: respectively. A distinct feature of the grey-body factor for gauge fields,
701: previously observed in the 4-dimensional case, is that it vanishes when
702: $\omega r_\text{h} \rightarrow 0$. The same behaviour is observed for all values of $n$. This leads to the suppression of the energy emission rate, in the low-energy regime, compared to those for scalar and fermion fields. Up to intermediate energies the grey-body factors exhibit enhancement with increasing $n$ as in the case of fermion fields. An asymptotic behaviour, similar to the previous cases, is observed in the high-energy regime with each grey-body factor assuming, after oscillation, the geometrical optics value (which decreases with increasing $n$). This result establishes the existence of a universal behaviour of all types of particles emitted by black holes at high energies. This behaviour is independent of the particle spin but dependent on the number of extra dimensions projected onto the brane.
703: \enlargethispage{-\baselineskip}
704:
705: Again the total flux and power emissivities for gauge fields on the brane can be obtained for different values of $n$. The exact results obtained by numerically integrating with respect to energy are given in Table~\ref{gauge-tab}.
706: As anticipated, the pattern of enhancement with $n$ is also observed
707: for the emission of gauge bosons. It is worth noting that
708: the relative enhancement observed in this case is the largest among all particle types---this is predominantly due to the large suppression of gauge boson emission in four dimensions.
709:
710: %%%%%%%%%%%%%%%%%%%
711: \begin{table}
712: \begin{center}
713: \begin{tabular}{|l|c|c|}
714: \hline
715: $n$& Flux & Power\\
716: \hline
717: 0 & 1 & 1\\
718: 1 & 19.2 & 27.1\\
719: 2 & 80.6 & 144\\
720: 3 & 204 & 441\\
721: 4 & 403 & 1020\\
722: 5 & 689 & 2000\\
723: 6 & 1070 & 3530\\
724: 7 & 1560 & 5740\\
725: \hline
726: \end{tabular}
727: \capbox{Emissivities for gauge fields on the brane}{Flux and power emissivities for gauge fields on the brane.\label{gauge-tab}}
728: \end{center}
729: \end{table}
730: %%%%%%%%%%%%%%%%%%
731:
732:
733:
734:
735: \subsection{Relative emissivities for different species}
736:
737: It is interesting to investigate how the relative numbers of scalars,
738: fermions and gauge bosons emitted on the brane change
739: as the number of extra dimensions projected onto the brane varies. In other
740: words, this means finding out what type of particles the black hole prefers
741: to emit, for different values of $n$, and what fraction of the total energy
742: each particular type of particle carries away during emission.
743:
744: \begin{figure}
745: \unitlength1cm
746: \begin{center}
747: \psfrag{x}[][][1.25]{$\omega r_\text{h}$}
748: \psfrag{y}[][][1.25]{$r_\text{h} d^2 \hat{E}^{(0)}/dtd\om$}
749: \begin{minipage}[t]{3.05in}
750: \scalebox{0.75}{\rotatebox{0}
751: {\includegraphics{nrelativen0.eps}}}
752: \end{minipage}
753: \hfill
754: \begin{minipage}[t]{3.05in}
755: \scalebox{0.75}{\rotatebox{0}
756: {\includegraphics{nrelativen6.eps}}}
757: \end{minipage}
758: \capbox{Power emission on the brane for $n=0$ and $n=6$}{Energy emission rates for the emission of scalars, fermions and gauge bosons on the brane with {\bf(a)} $n=0$ and {\bf(b)} $n=6$.\label{relan6}}
759: \end{center}
760: \end{figure}
761:
762: Comparing the energy emission rates for different types of particles and for
763: fixed $n$, gives the qualitative behaviour. Figures~\ref{relan6}(a) and \ref{relan6}(b) show (with a linear scale) the a power spectra for $n=0$ and $n=6$ respectively; these two figures demonstrate very clearly the effects already discussed, namely the orders-of-magnitude enhancement of the emission rates and the displacement of the peak to higher energies, as $n$ increases. Figure~\ref{relan6}(a) reveals that, in the absence of any extra dimensions, most of the energy of the black hole emitted on the brane is in the form of scalar particles;
764: the next most important are the fermion fields, and less significant are the
765: gauge bosons. As $n$ increases, the emission rates for all species are enhanced but at different rates. Figure~\ref{relan6}(b) clearly shows that, for a
766: large number of extra dimensions, the most effective `channel' during
767: the emission of brane-localized modes is that of gauge bosons;
768: the scalar and fermion fields follow second and third respectively.
769: The changes in the flux spectra are similar as $n$ increases.
770:
771: \begin{table}[b]
772: \begin{minipage}[t]{3.05in}
773: \begin{center}
774: \begin{tabular}{|c|c|c|c|}
775: \hline {\rule[-3mm]{0mm}{8mm} }
776: $\!\!n$ & $s=0$ & $s=\frac{1}{2}$ & $s=1$ \\
777: \hline
778: 0&1&0.37&0.11\\
779: 1&1&0.70&0.45\\
780: 2&1&0.77&0.69\\
781: 3&1&0.78&0.83\\
782: 4&1&0.76&0.91\\
783: 5&1&0.74&0.96\\
784: 6&1&0.73&0.99\\
785: 7&1&0.71&1.01\\
786: \hline
787: Black body&1&0.75&1\\
788: \hline
789: \end{tabular}
790: \capbox{Flux emission ratios}{Flux emission ratios.\label{fratios}}
791: \end{center}
792: \end{minipage}
793: \hfill
794: \begin{minipage}[t]{3.05in}
795: \begin{center}
796: \begin{tabular}{|c|c|c|c|}
797: \hline {\rule[-3mm]{0mm}{8mm} }
798: $\!\!\!n$ & $s=0$ & $s=\frac{1}{2}$ & $s=1$ \\
799: \hline
800: 0&1&0.55&0.23\\
801: 1&1&0.87&0.69\\
802: 2&1&0.91&0.91\\
803: 3&1&0.89&1.00\\
804: 4&1&0.87&1.04\\
805: 5&1&0.85&1.06\\
806: 6&1&0.84&1.06\\
807: 7&1&0.82&1.07\\
808: \hline
809: Black body&1&0.87&1\\
810: \hline
811: \end{tabular}
812: \capbox{Power emission ratios}{Power emission ratios.\label{pratios}}
813: \end{center}
814: \end{minipage}
815: \end{table}
816:
817: The above behaviour can be more helpfully quantified by computing the relative emissivities for scalars, fermions and gauge bosons emitted on the brane by integrating the flux and power emission spectra. The relative emissivities obtained in this way are shown in Tables~\ref{fratios} and \ref{pratios} (they are normalized to the scalar values). The ratios for $n\geq 1$ are available for
818: the first time in the literature as a result of this numerical work, while
819: the $n=0$ results would appear to be the most accurate ones available. From \cite{MacGibbon:1990zk}, flux and power ratios are found to be 1\,:\,0.36\,:\,0.11 and 1\,:\,0.56\,:\,0.23 respectively, in good agreement with those ratios shown in Tables~\ref{fratios} and \ref{pratios}. In \cite{Giddings:2001bu} the power ratio for $n=0$ is given as 40\,:\,19\,:\,7.9 i.e.\ 1\,:\,0.48\,:\,0.20 which is in less good agreement. More careful examination shows that the relative power emitted by the $s=1/2$ and $s=1$ degrees of freedom agrees exactly with Table~\ref{pratios} and so the disagreement seems to come from the scalar value. This is unsurprising since the scalar value quoted in \cite{Giddings:2001bu} was only estimated from a plot in \cite{Sanchez:1978si}.
820: \enlargethispage{-\baselineskip}
821:
822: The entries in these tables reflect the qualitative behaviour
823: illustrated above for some extreme values of the number of extra dimensions.
824: For $n=0$, the scalar fields are the type of particle which are
825: most commonly produced and the ones which carry away most of the energy of
826: the black hole emitted on the brane; the fermion and gauge fields carry
827: approximately 1/2 and 1/4, respectively, of the energy emitted in scalar fields, and their fluxes are only 1/3 and 1/10 of the scalar
828: flux. For intermediate values of $n$, the fermion and
829: gauge boson emissivities are considerably enhanced compared to
830: the scalar one and have become of approximately the same magnitude---e.g.\
831: for $n=2$ the amount of energy spent by the black hole in the emission
832: of fermions and gauge bosons is exactly the same, although the net number
833: of gauge bosons is still subdominant. For large values of $n$, the situation
834: is reversed: the gauge bosons dominate both flux and power spectra, with
835: the emission of fermions being the least effective channel both in
836: terms of number of particles produced and energy emitted. The reader is reminded that the above results refer to the emission of individual scalar,
837: fermionic or bosonic degrees of freedom and not to the elementary particles. However it is easy to combine the ratios in Table~\ref{pratios} with the relevant numbers of degrees of freedom (see Table~\ref{pprobs}) to obtain relative emissivities for different particle types.
838:
839: These numerical results confirm that the relative emissivities of different particles produced by small, higher-dimensional black holes depend on the number of extra dimensions projected onto the brane. Therefore, if Hawking radiation from such objects is detected, this could provide a way of determining the number of extra dimensions existing in nature.
840:
841: \subsection{Total flux and power emitted}
842: \label{totfandp}
843:
844: Whilst it is often only necessary to know the relative emissivities of the fields of different spins, sometimes (for example, when estimating the black hole lifetime) the absolute values are required. For the case of brane emission from a non-rotating black hole, the flux emitted in a field of spin $s$ can be written as
845: %%%%%%
846: \begin{equation}
847: F^{(s)}=\int_0^{\infty} \frac{\Gamma^{(s)} (\omega r_\text{h})}{\exp\left(\omega/T_\text{H}\right)\mp1}\frac{d\omega}{2\pi}=\int_0^{\infty} \frac{\Gamma^{(s)} (\omega r_\text{h})}{\exp\left(\frac{4\pi\omega r_\text{h}}{n+1}\right)\mp1}\frac{d\omega}{2\pi}\,.
848: \label{totflux}
849: \end{equation}
850: %%%%%%
851: This expression is obtained by integrating equation~(\ref{fdecay-brane}) using the definition
852: %%%%%%
853: \begin{equation}
854: \Gamma^{(s)}(\omega r_\text{h})=\sum_{\ell} (2\ell +1) |\hat {\cal A}^{(s)}_\ell|^2.
855: \end{equation}
856: %%%%%%
857: Since the integrand in eq.~(\ref{totflux}) is entirely a function of $\omega r_\text{h}$, the substitution $x=\omega r_\text{h}$ can be used to express $F^{(s)}$ as
858: %%%%%%
859: \begin{equation}
860: F^{(s)}=\frac{1}{r_\text{h}}\int_0^{\infty}\frac{\Gamma^{(s)}(x)}{\exp\left(\frac{4\pi x}{n+1}\right)\mp1}\frac{dx}{2\pi}\,,
861: \end{equation}
862: %%%%%%%
863: where the only dependence on the radius (or equivalently the mass) of the black hole is contained in the factor outside the integral.
864:
865: A similar procedure, starting from equation~(\ref{4power}), shows the power emitted in a field of spin $s$ to be
866: \begin{equation}
867: P^{(s)}=\frac{1}{r_\text{h}^2}\int_0^{\infty}\frac{x\Gamma^{(s)}(x)}{\exp\left(\frac{4\pi x}{n+1}\right)\mp1}\frac{dx}{2\pi}\,.
868: \end{equation}
869:
870: \enlargethispage{-\baselineskip}
871: These results of these integrations are given in Tables~\ref{fvalues} and \ref{pvalues}. The values in these tables can be combined with the total numbers of degrees of freedom for each spin. These are given in the last row
872: of Table~\ref{pprobs} and are 4, 90 and 24 for $s=0$, 1/2 and 1 respectively.
873: Therefore the total flux $F$ and power $P$ emitted by black holes of different numbers of dimensions can be calculated---the results are shown in Table~\ref{totfp}.
874:
875: \begin{table}
876: \def\arraystretch{1.1}
877: \begin{center}
878: \begin{tabular}{|l|c|c|c|}
879: \hline
880: $n$& $r_\text{h} F^{(0)}$ & $r_\text{h} F^{(1/2)}$ & $r_\text{h} F^{(1)}$ \\
881: \hline
882: 0 & 0.00133 & 0.000486 & 0.000148\\
883: 1 & 0.00631 & 0.00439 & 0.00283\\
884: 2 & 0.0173 & 0.0134 & 0.0119\\
885: 3 & 0.0364 & 0.0283 & 0.0301\\
886: 4 & 0.0655 & 0.0499 & 0.0596\\
887: 5 & 0.106 & 0.0789 & 0.102\\
888: 6 & 0.160 & 0.116 & 0.159\\
889: 7 & 0.229 & 0.163 & 0.231\\
890: \hline
891: \end{tabular}
892: \capbox{Flux for different spin fields}{Total flux emitted in fields of different spins.\label{fvalues}}
893: \end{center}
894: \def\arraystretch{1.0}
895: \end{table}
896:
897: \begin{table}
898: \def\arraystretch{1.1}
899: \begin{center}
900: \begin{tabular}{|l|c|c|c|}
901: \hline
902: $n$& $r_\text{h}^2 P^{(0)}$ & $r_\text{h}^2 P^{(1/2)}$ & $r_\text{h}^2 P^{(1)}$ \\
903: \hline
904: 0 & 0.000298 & 0.000164 & 6.72$\times 10^{-5}$\\
905: 1 & 0.00266 & 0.00232 & 0.00182\\
906: 2 & 0.0107 & 0.00973 & 0.00971\\
907: 3 & 0.0297 & 0.0265 & 0.0296\\
908: 4 & 0.0661 & 0.0575 & 0.0686\\
909: 5 & 0.128 & 0.109 & 0.135\\
910: 6 & 0.223 & 0.187 & 0.237\\
911: 7 & 0.362 & 0.299 & 0.386\\
912: \hline
913: \end{tabular}
914: \capbox{Power for different spin fields}{Total power emitted in fields of different spins.\label{pvalues}}
915: \end{center}
916: \def\arraystretch{1.0}
917: \end{table}
918:
919: \begin{table}
920: \def\arraystretch{1.1}
921: \begin{center}
922: \begin{tabular}{|l|c|c|}
923: \hline
924: $n$& $r_\text{h} F$ & $r_\text{h}^2 P$\\
925: \hline
926: 0 & 0.0526 & 0.0175\\
927: 1 & 0.489 & 0.263\\
928: 2 & 1.56 & 1.15\\
929: 3 & 3.41 & 3.21\\
930: 4 & 6.18 & 7.09\\
931: 5 & 9.98 & 13.5\\
932: 6 & 14.9 & 23.4\\
933: 7 & 21.1 & 37.6\\
934: \hline
935: \end{tabular}
936: \capbox{Total flux and power emitted}{Total flux and power emitted from black holes of different numbers of dimensions.\label{totfp}}
937: \end{center}
938: \def\arraystretch{1.0}
939: \end{table}
940:
941: It is possible to use these values for the total flux emitted from extra-dimensional black holes to comment on the usual assumption that black hole decays can
942: be considered as quasi-stationary. The simplest way of trying to verify this is to compare the typical time between emissions (given by $F^{-1}$) with $r_\text{h}$ (the time for light to cross the black hole radius in the natural units being used). To be sure that the quasi-stationary approach is valid, we would require $F^{-1}\gg r_\text{h}$ or equivalently $r_\text{h} F \ll 1$. From Table~\ref{totfp}, this can be seen to be true only for the cases $n=0$ and $n=1$.
943:
944: \enlargethispage{-\baselineskip}
945: Another requirement for the validity of a semi-classical description of black hole production and decay is that the lifetime $\tau \gg 1/M_\text{BH}$ so that the black hole is a well-defined resonance. This assumption can also be tested here by using the values of $P$ in Table~\ref{totfp}. The relationship
946: \begin{equation}
947: P=-\frac{dM_\text{BH}}{dt}=\frac{p}{r_\text{h}^2}\,,
948: \end{equation}
949: can be integrated as follows:
950: \begin{equation}
951: p\int_0^\tau dt = - \int_{M_\text{BH}}^0 r_\text{h}^2 dM_\text{BH} = \frac{1}{\pi M_{\text{P}(4+n)}^2}\left(\frac{8 \Gamma\left(\frac{n+3}{2}\right)}{n+2}\right)^{\frac{2}{n+1}} \int_0^{M_\text{BH}} \left(\frac{M_\text{BH}}{M_{\text{P}(4+n)}}\right)^{\frac{2}{n+1}} dM_\text{BH}\,.
952: \end{equation}
953: In the above, the final Planck phase and any complications caused by kinematically forbidden emissions have been ignored. Proceeding with the integration gives
954: \begin{equation}
955: \tau M_\text{BH} = \frac{1}{\pi p}\left(\frac{8 \Gamma\left(\frac{n+3}{2}\right)}{n+2}\right)^{\frac{2}{n+1}} \frac{n+1}{n+3} \left(\frac{M_\text{BH}}{M_{\text{P}(4+n)}}\right)^{\frac{2(n+2)}{n+1}}.
956: \end{equation}
957: Values of $\tau M_\text{BH}$ for different values of $n$ are shown in Table~\ref{taum} both in the case $M_\text{BH}=5M_{\text{P}(4+n)}$ and in the case $M_\text{BH}=10M_{\text{P}(4+n)}$. It should be noted that switching to convention `a' for the definition of $M_{\text{P}(4+n)}$ the high-$n$ values in the table would be significantly altered (for example, the $n=7$ values would be multiplied by a factor of 13.2). The `long'-lifetime requirement is clearly different to the quasi-stationary issue discussed above since it depends on the initial mass of the black hole. However given the limits on $M_{\text{P}(4+n)}$ and the energy available at the LHC, it would again seem that this requirement may not be satisfied for the higher values of $n$.
958:
959: \begin{table}
960: \def\arraystretch{1.1}
961: \begin{center}
962: \begin{tabular}{|c|c|c|}
963: \hline
964: & \multicolumn{2}{c|}{$\tau M_\text{BH}$} \\
965: \cline{2-3}
966: $n$& $M_\text{BH}=5M_{\text{P}(4+n)}$ & $M_\text{BH}=10M_{\text{P}(4+n)}$\\
967: \hline
968: 0 & 47500 & 761000\\
969: 1 & 202 & 1610\\
970: 2 & 23.3 & 148\\
971: 3 & 6.60 & 37.4\\
972: 4 & 2.77 & 14.6\\
973: 5 & 1.43 & 7.23\\
974: 6 & 0.846 & 4.12\\
975: 7 & 0.544 & 2.59\\
976: \hline
977: \end{tabular}
978: \capbox{Values of $\tau M_\text{BH}$}{Values of $\tau M_\text{BH}$ for different values of $n$.\label{taum}}
979: \end{center}
980: \def\arraystretch{1.0}
981: \end{table}
982:
983: \enlargethispage{\baselineskip}
984: The results presented in Table~\ref{taum} differ significantly from equivalent results which can be obtained from an expression given in \cite{Dimopoulos:2001en}. Partly this is due to inclusion of the grey-body factors here, but this does not account for order-of-magnitude differences. It would appear that the lifetime is incorrect in \cite{Dimopoulos:2001en} because of the failure to take all the degrees of freedom of the emitted particles into account.
985:
986: The numerical results presented in Table~\ref{totfp} can also be used to calculate the average particle multiplicity, again completely ignoring the complications of the Planck phase and kinematic constraints on the decay. Defining $r_\text{h} F=f$, the average particle multiplicity can be expressed as
987: %%%%%
988: \begin{equation}
989: \langle N\rangle\,=\,\frac{f}{p}\int_0^{M_\text{BH}} r_\text{h} \, dM_\text{BH} = \frac{1}{\sqrt{\pi} M_{\text{P}(4+n)}}\left(\frac{8 \Gamma\left(\frac{n+3}{2}\right)}{n+2}\right)^{\frac{1}{n+1}}\!\!\int_0^{M_\text{BH}} \left(\frac{M_\text{BH}}{M_{\text{P}(4+n)}}\right)^{\frac{1}{n+1}} dM_\text{BH}\,.
990: \end{equation}
991: %%%%%
992: Performing the integration gives
993: %%%%%
994: \begin{equation}
995: \langle N\rangle\,=\,\frac{f}{p}\frac{1}{\sqrt{\pi}}\left(\frac{8 \Gamma\left(\frac{n+3}{2}\right)}{n+2}\right)^{\frac{1}{n+1}} \frac{n+1}{n+2} \left(\frac{M_\text{BH}}{M_{\text{P}(4+n)}}\right)^{\frac{n+2}{n+1}}=\frac{f}{p}\frac{n+1}{4\pi}S_\text{BH}\,,
996: \label{avn}
997: \end{equation}
998: %%%%%
999: where the final identification has been made by referring back to equation~(\ref{bhent}). Table~\ref{avntab} shows the average multiplicities obtained for different values of $n$ along with the (initial) entropy $S_\text{BH}$. The average multiplicity values given improve on those in \cite{Dimopoulos:2001hw} (where $\langle N\rangle\sim M_\text{BH}/2\,T_\text{H}$ was assumed based on the constant temperature approximation) and \cite{Cavaglia:2003hg} (where a more careful approach was taken but still without the full numerical grey-body factors).
1000:
1001: \begin{table}
1002: \def\arraystretch{1.1}
1003: \begin{center}
1004: \begin{tabular}{|c|c|c|c|c|}
1005: \hline
1006: & \multicolumn{2}{c|}{$\langle N\rangle$} & \multicolumn{2}{c|}{$S_\text{BH}$}\\
1007: \cline{2-5}
1008: $n$& $M_\text{BH}=5M_{\text{P}(4+n)}$ & $M_\text{BH}=10M_{\text{P}(4+n)}$ & $M_\text{BH}=5M_{\text{P}(4+n)}$ & $M_\text{BH}=10M_{\text{P}(4+n)}$\\
1009: \hline
1010: 0 & 75.3 & 301 & 314 & 1260 \\
1011: 1 & 12.8 & 36.1 & 43.1 & 122\\
1012: 2 & 6.82 & 17.2 & 21.0 & 52.9\\
1013: 3 & 4.78 & 11.4 & 14.2 & 33.7\\
1014: 4 & 3.80 & 8.75 & 11.0 & 25.2\\
1015: 5 & 3.22 & 7.23 & 9.13 & 20.5\\
1016: 6 & 2.80 & 6.20 & 7.92 & 17.5\\
1017: 7 & 2.52 & 5.50 & 7.06 & 15.4\\
1018: \hline
1019: \end{tabular}
1020: \capbox{Values of $\langle N\rangle$ and $S_\text{BH}$}{Values of $\langle N\rangle$ and $S_\text{BH}$ (calculated using eqs.~(\ref{avn}) and (\ref{bhent}) respectively) for different values of $n$.\label{avntab}}
1021: \end{center}
1022: \def\arraystretch{1.0}
1023: \end{table}
1024:
1025: The average multiplicities in the first column clearly do not satisfy $\langle N\rangle\gg 1$ for $n>3$ which means that a significant amount of the decay will be affected by kinematic and Planck phase considerations. However using the alternative definition of $M_{\text{P}(4+n)}$ these results would again be modified (for $n=7$ they would be multiplied by a factor of 3.64). There is, of course, no fundamental difference between these conventions---setting $M_{\text{P}(4+n)}=1$~TeV in convention `a' simply corresponds to a lower value of the Planck mass in convention `d' and hence for a fixed black hole mass the Hawking temperature is lower and the average multiplicity higher.
1026:
1027: The tabulated entropies allow comparison with the requirement that $1/\sqrt{S_\text{BH}}$ should be much less than unity if statistical fluctuations of the number of micro-canonical degrees of freedom are to be small (as is necessary for a semi-classical description of the black hole decay to be valid).
1028:
1029: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1030:
1031:
1032: \section{Emission in the bulk}
1033: \label{embulk}
1034:
1035: An important question regarding the emission of particles by
1036: higher-dimensional black holes is how much of this energy is radiated
1037: onto the brane and how much is lost in the bulk. In the former case,
1038: the emitted particles are zero-mode gravitons and scalar fields as
1039: well as Standard Model fermions and gauge bosons, while in the latter
1040: case all emitted energy is in the form of massive Kaluza-Klein gravitons
1041: and, as discussed in \secref{gbfintro}, possibly also scalar fields. In \cite{Emparan:2000rs}, it was shown that the whole
1042: tower of KK excitations of a given particle carries approximately the
1043: same amount of energy as a massless particle emitted on the
1044: brane. Combining this result with the observation that many more types of particles
1045: live on the brane than in the bulk, it was concluded that most of the
1046: energy of the black hole goes into brane modes. The results obtained
1047: in \cite{Emparan:2000rs} were only approximate since the dependence of the grey-body factor on the energy of the emitted particle was ignored and the
1048: geometric expression for the area of the horizon (valid in the high-energy regime) was used instead.
1049:
1050: In order to provide an accurate answer to the question of how much energy
1051: is emitted into the bulk compared to on the brane, it is necessary to take into account the dependence of the grey-body factor on both energy and number of extra dimensions. There is a similar discussion in \cite{Cavaglia:2003hg} but it is incomplete because full numerical results for the grey-body factor in all energy regimes were not available. In this section the emission of scalar modes in the bulk is thoroughly investigated. Exact numerical results are produced for the grey-body factors and energy emission rates in terms of the energy and number of extra dimensions. This allows the questions mentioned above to be addressed by calculating the total bulk-to-brane relative emissivities for different values of $n$.
1052:
1053: \subsection{Grey-body factors and emission rates}
1054:
1055: The analysis in this section is relevant for gravitons and possibly scalar fields, and requires knowledge of the solutions of the
1056: corresponding equations of motion in the bulk. Only the case of scalar fields is considered here since the bulk equation of motion is known---the emission of bulk scalar modes was previously studied analytically, in the low-energy regime, in \cite{Kanti:2002nr}.\footnote{After this work was completed, some plots of grey-body factors for bulk scalars were discovered in Figure 1 of \cite{Frolov:2002as}. It is unclear how these were calculated and only the $\ell=0$ contributions for $n=$1--3 are shown; however, they provided a useful cross-check on some of the results presented here.}
1057:
1058: A scalar field propagating in the background of a higher-dimensional,
1059: non-rotating (Schwarzschild-like) black hole, with line-element given by eq.~(\ref{metric-D}), satisfies the following equation of motion \cite{Kanti:2002nr}
1060: %%%%%%%%%%%%
1061: \begin{equation}
1062: \frac{h(r)}{r^{n+2}}\,\frac{d \,}{dr}\,
1063: \biggl[\,h(r)\,r^{n+2}\,\frac{d R}{dr}\,\biggr] +
1064: \biggl[\,\om^2 - \frac{h(r)}{r^2}\,\ell\,(\ell+n+1)\,\biggr] R =0 \, .
1065: \label{scalareqn}
1066: \end{equation}
1067: %%%%%%%%%%%
1068: This radial equation was obtained by using a separable solution similar to equation~(\ref{sepsoln}) but with ${}_sS^m_{\ell}(\theta)$ replaced by $S^m_{\ell}(\theta_1,\theta_2,\ldots,\theta_n)$. The functions $e^{im\varphi}\,S^m_{\ell}(\theta_1,\theta_2,\ldots,\theta_n)$ can be written as $\tilde{Y}^m_{\ell}(\Omega_{2+n})$\,; these are generalizations of the usual spherical harmonic functions to the case of $(3+n)$ spatial dimensions \cite{muller} and account for the $\ell(\ell+n+1)$ term in eq.~(\ref{scalareqn}).
1069:
1070: As for the emission of particles on the brane, the determination of the grey-body factor for bulk emission requires the above equation to be solved over the whole radial domain. The exact solution for the radial function must interpolate between the near-horizon and far-field asymptotic solutions, given by
1071: %%%%%%%%%
1072: \begin{equation}
1073: \label{near-bulk}
1074: R^{(\text{h})}=A_\text{in}^{(\text{h})}\,e^{-i\omega r^{*}}+ A_{out }^{(\text{h})}\,e^{i\omega r^{*}},
1075: \end{equation}
1076: %%%%%%%%%%
1077: and
1078: %%%%%%%%%%%
1079: \begin{equation}
1080: \label{far-bulk}
1081: R^{(\infty)}=A_\text{in}^{(\infty)}\,\frac{e^{-i\omega r}}{\sqrt{r^{n+2}}}+
1082: A_\text{out}^{(\infty)}\,\frac{e^{i\omega r}}{\sqrt{r^{n+2}}}\,,
1083: \end{equation}
1084: %%%%%%%%%%%%%
1085: respectively. Again the boundary condition that no out-going solution should exist near the horizon of the black hole is imposed which means that $A_\text{out}^{(\text{h})}=0$. The solution at infinity comprises, as usual, both in-going and out-going modes.
1086:
1087: The absorption probability $|\tilde {\cal A}_\ell|^2$ may
1088: then be calculated either by using eq.~(\ref{scalars}) or from the ratio $|A_\text{in}^{(\text{h})}/A_\text{in}^{(\infty)}|^2$ (this latter approach is found to be more numerically stable for low energies and/or larger values of $n$). A tilde will now denote bulk quantities, as opposed to brane quantities which carry a hat. The grey-body factor $\tilde \si_\ell(\om)$ may be determined by using eq.~(\ref{greydef}). The dimensionality of the grey-body factor changes as the number of extra dimensions varies; therefore, in order to be able to compare its values for different $n$, it is normalized to the horizon area of a $(4+n)$-dimensional black hole. The discussion of the normalization procedure and the subsequent comments on the geometrical optics limit are due to Kanti \cite{Kantiprivate,Harris:2003eg}.
1089:
1090: First equation~(\ref{greydef}) is re-written in the form
1091: %%%%%%%%%%%%
1092: \begin{equation}
1093: \tilde \si_\ell(\om) = \frac{2^{n}}{\pi}\,
1094: \Ga\Bigl[\frac{n+3}{2}\Bigr]^2\,
1095: \frac{\tilde A_\text{h}}{(\om r_\text{h})^{n+2}}\, \tilde N_\ell\,
1096: |\tilde {\cal A}_\ell|^2,
1097: \label{greyb}
1098: \end{equation}
1099: %%%%%%%%%%%%
1100: where $\tilde N_\ell$ is the multiplicity of states corresponding to the
1101: same partial wave $\ell$. For a $(4+n)$-dimensional space-time it is given by
1102: %%%%%%%%%
1103: \begin{equation}
1104: \tilde N_\ell= \frac{(2\ell+n+1)\,(\ell+n)!}{\ell! \,(n+1)!}\,,
1105: \label{bulk-mult}
1106: \end{equation}
1107: %%%%%%%%%
1108: and the horizon area in the bulk is defined as
1109: %%%%%%%%%
1110: \begin{equation}
1111: \begin{split}
1112: \tilde A_\text{h} =&\,
1113: r_\text{h}^{n+2}\,\int_0^{2 \pi} \,d \varphi \,\prod_{k=1}^{n+1}\,
1114: \int_0^\pi\,\sin^k\theta_{n+1}\,
1115: \,d(\sin\theta_{n+1}) \\[3mm]
1116: =&\, r_\text{h}^{n+2}\,(2\pi)\,\prod_{k=1}^{n+1}\,\sqrt{\pi}\,\,\frac{\Ga[(k+1)/2]}
1117: {\Ga[(k+2)/2]}\\[3mm]
1118: =&
1119: \,r_\text{h}^{n+2}\,(2\pi)\,\pi^{(n+1)/2}\,\Ga\Bigl[\frac{n+3}{2}\Bigr]^{-1}.
1120: \end{split}
1121: \end{equation}
1122: %%%%%%%%%%
1123:
1124: Equation~(\ref{greyb}) allows the computation of the grey-body factor's low-energy limit once the corresponding expression for the absorption
1125: coefficient is determined. Analytic results for $\tilde {\cal A}_\ell$
1126: were derived in \cite{Kanti:2002nr} by solving eq.~(\ref{scalareqn}) in the two
1127: asymptotic regimes (near-horizon and far-field) and matching them in an intermediate zone. It was found that the low-energy expression of the $\ell=0$ absorption coefficient has the form
1128: %%%%%%%%%%%%%
1129: \begin{equation}
1130: |\tilde {\cal A}_0|^2 = \biggl(\frac{\omega r_\text{h}}{2}\biggl)^{n+2}
1131: \,\frac{4 \pi}{\Gamma[(n+3)/2]^2} + \ldots \,,
1132: \end{equation}
1133: %%%%%%%%%%%%%%%
1134: where the ellipsis denotes higher-order terms in the $\omega r_\text{h}$ power series. These terms, as well as the corresponding expressions for higher partial waves, vanish quickly in the limit $\omega r_\text{h} \rightarrow 0$, leaving the above term as the dominant one. Substituting
1135: into eq.~(\ref{greyb}) it is clear that, in the low-energy regime, the grey-body factor is given by the area $\tilde A_\text{h}$ of the black hole horizon. This behaviour is similar to the 4-dimensional case except that the area of the horizon now changes with $n$.
1136:
1137: In the high-energy regime, it is anticipated that an equivalent of the geometrical optics limit discussed in \secref{numscalar} will be recovered. In four dimensions, the low- and high-energy asymptotic limits are $4\pi r_\text{h}^2$ and $\pi r_\text{c}^2$ respectively. This has led to the na\"{\i}ve generalization that, in an arbitrary number of dimensions, the high-energy expression for the grey-body factor will be approximately $\Omega_{n+2}\,r_\text{c}^{n+2}/4$, where $\Omega_{n+2}$ is defined as in eq.~(\ref{omegapdef}). It will shortly be shown that this is in
1138: fact an over-estimate of the high-energy limit.
1139:
1140: As in four dimensions it is assumed, for high energy particles, that the grey-body factor
1141: becomes equal to the area of an absorptive body of radius $r_\text{c}$, projected on a plane parallel to the orbit of the moving particle
1142: \cite{misner}. According to ref. \cite{Emparan:2000rs}, the value of the effective radius $r_\text{c}$ remains the same for both bulk and brane particles and
1143: is given by eq.~(\ref{effective}). The area of the absorptive body depends
1144: strongly on the dimensionality of space-time and its calculation requires
1145: one of the azimuthal angles to be set to $\pi/2$. A careful calculation
1146: reveals that the `projected' area is given by
1147: %%%%%%%%%%%%
1148: \begin{equation}
1149: \tilde A_\text{p} = \frac{2 \pi}{(n+2)}\,\frac{\pi^{n/2}}{\Gamma[(n+2)/2]}\,
1150: r_\text{c}^{n+2} = \frac{1}{n+2}\,\Omega_{n+1}\,r_\text{c}^{n+2}\,.
1151: \end{equation}
1152: %%%%%%%%%%%%
1153:
1154: The above relation reduces to the usual 4-dimensional result
1155: ($\tilde A_\text{p} = \pi r_\text{c}^2$) for $n=0$ but, compared to the \,na\"{\i}ve $\Omega_{n+2}\,r_\text{c}^{n+2}/4$, leads to values reduced by 50\% for higher values of $n$. Assuming that the grey-body factor at high
1156: energies becomes equal to the absorptive area $\tilde A_\text{p}$, it can be explicitly written as
1157: %%%%%%%%%%%%%%
1158: \begin{equation}
1159: \begin{split}
1160: \tilde \si_\text{g} = & \,\frac{1}{n+2}\,\frac{\Omega_{n+1}}{\Omega_{n+2}}\,
1161: \biggl(\frac{r_\text{c}}{r_\text{h}}\biggr)^{n+2}\,\tilde A_\text{h} \\[3mm]
1162: = &\,\frac{1}{\sqrt{\pi}\,(n+2)}\,\frac{\Gamma[(n+3)/2]}{\Gamma[(n+2)/2]}\,
1163: \biggl(\frac{n+3}{2}\biggr)^{(n+2)/(n+1)}\,
1164: \biggl(\frac{n+3}{n+1}\biggr)^{(n+2)/2}\,\tilde A_\text{h}\,.
1165: \label{high}
1166: \end{split}
1167: \end{equation}
1168: %%%%%%%%%%%%
1169: In the above the same normalization is used as in the low-energy regime---that is, the normalization is in terms of $\tilde{A}_\text{h}$, the area of the $(4+n)$-dimensional horizon. The values predicted by eq.~(\ref{high}) are tabulated in Table~\ref{golim}, along with the more na\"{\i}ve $\Omega_{n+2}\,r_\text{c}^{n+2}/4$ prediction.
1170:
1171: %%%%%%%%%%%%%%%%%%%
1172: \begin{table}
1173: \def\arraystretch{1.1}
1174: \begin{center}
1175: \begin{tabular}{|l|c|c|}
1176: \hline
1177: $n$ & $\phantom{\Bigl.}\tilde \si_\text{g}/\tilde{A}_\text{h}\phantom{\Bigr.}$ & $\Omega_{n+2}\,r_\text{c}^{n+2}/4\tilde{A}_\text{h}$\\
1178: \hline
1179: 0 & 1.69 & 1.69\\
1180: 1 & 1.70 & 2\\
1181: 2 & 1.77 & 2.36\\
1182: 3 & 1.85 & 2.72\\
1183: 4 & 1.93 & 3.09\\
1184: 5 & 2.01 & 3.45\\
1185: 6 & 2.08 & 3.81\\
1186: 7 & 2.16 & 4.17\\
1187: \hline
1188: \end{tabular}
1189: \capbox{High-energy limits of grey-body factors for bulk emission}{High-energy limits of grey-body factors for bulk emission, given in units of the $(4+n)$-dimensional area $\tilde{A}_\text{h}$.\label{golim}}
1190: \end{center}
1191: \def\arraystretch{1.0}
1192: \end{table}
1193: %%%%%%%%%%%%%%%%%%
1194:
1195: Turning now to the numerical analysis, the grey-body factors can be found by using eq.~(\ref{greyb}) and the exact numerical results for the absorption
1196: coefficients. Their behaviour is shown in Figure~\ref{grey0-bulk}. As
1197: it was anticipated after the above discussion, the normalized grey-body
1198: factors, in the low-energy regime, tend to unity for all values of $n$
1199: as each one adopts the value of the black hole horizon area
1200: to which it has been normalized. As for the emission of
1201: scalar fields on the brane, the grey-body factors are suppressed with increasing $n$ in the low-energy regime, start oscillating at intermediate energies and then tend to their asymptotic high-energy limits. A simple numerical analysis shows that the na\"{\i}ve expression
1202: $\Omega_{n+2}\,r_\text{c}^{n+2}/4$ fails to describe the high-energy asymptotic
1203: limits for all values of $n$ larger than zero. In contrast, eq.~(\ref{high}) gives asymptotic values which are verified by the numerical results. As in the brane emission case, it is found that relatively large values of $\omega r_\text{h}$ are required before the asymptotic values are approached for the higher values of $n$ (this can be seen in Figure~\ref{grey0-bulk} in which the $n=4$ and 6 curves clearly do not attain the asymptotic values given in Table~\ref{golim}).
1204:
1205: \begin{figure}
1206: \begin{center}
1207: \psfrag{x}[t][b][1.75]{$\omega r_\text{h}$}
1208: \psfrag{y}[][][1.75]{$\tilde{\sigma}^{(0)}_\text{abs}(\om)/\tilde{A}_\text{h}$}
1209: %\psfrag{n=0}[][][1.75]{$n=0$}\psfrag{n=1}[][][1.75]{$n=1$}\psfrag{n=2}[][][1.75]{$n=2$}\psfrag{n=4}[][][1.75]{$n=4$}\psfrag{n=6}[][][1.75]{$n=6$}
1210: \scalebox{0.5}{\rotatebox{0}{\includegraphics[width=22cm, height=15.4cm]
1211: {ngrey-bulk.eps}}}
1212: \capbox{Grey-body factors for scalars in the bulk}{Grey-body factors for scalar emission in the bulk from a $(4+n)$D
1213: black hole.\label{grey0-bulk}}
1214: \end{center}
1215: \end{figure}
1216:
1217: \begin{figure}
1218: \begin{center}
1219: \psfrag{x}[t][b][1.75]{$\omega r_\text{h}$}
1220: \psfrag{y}[][][1.75]{$r_\text{h} d^2\tilde{E}^{(0)}/dtd\om$}
1221: %\psfrag{n=0}[][][1.75]{$n=0$}\psfrag{n=1}[][][1.75]{$n=1$}\psfrag{n=2}[][][1.75]{$n=2$}\psfrag{n=3}[][][1.75]{$n=3$}
1222: \scalebox{0.5}{\rotatebox{0}{\includegraphics[width=23cm, height=15cm]
1223: {nbulkn.eps}}}
1224: \capbox{Power emission for scalars in the bulk}{Energy emission rates for scalar fields in the bulk from a $(4+n)$D
1225: black hole.\label{rate-bulk}}
1226: \end{center}
1227: \end{figure}
1228:
1229: In general, the suppression of the grey-body factor for bulk emission at low
1230: energies is milder than the one for brane emission. However this does not
1231: lead to higher emission rates for bulk modes compared to those for brane
1232: modes: the integration over the phase-space in eq.~(\ref{4power})
1233: involves powers of $\om r_\text{h}$ which cause an increasingly
1234: suppressive effect in the low-energy regime as $n$ increases. Nevertheless,
1235: the increase in the temperature of the black hole (still given by $T_\text{H}=(n+1)/4 \pi r_\text{h}$) eventually overcomes the decrease in the grey-body factor and causes the enhancement of the emission rate with $n$ at
1236: high energies (as well as shifting the spectrum peak to higher energies). The behaviour of the differential energy emission rates as a function of the energy parameter $\om r_\text{h}$ is shown in Figure~\ref{rate-bulk} for some indicative values of $n$. Comparing these numerical results with the analytic results of ref. \cite{Kanti:2002nr} shows that the earlier results were reasonably successful in describing the low-energy behaviour of both the grey-body factors and the energy emission rates for scalar fields in the bulk.
1237:
1238: \subsection{Bulk-to-brane relative emissivities}
1239:
1240: \enlargethispage{-3\baselineskip}
1241: The aim of this section is to perform an analysis which provides an answer
1242: to the question of the relative bulk-to-brane emissivity. This requires the differential energy emission rates in the bulk and on the brane to be evaluated; the two quantities are then compared for different numbers of extra dimensions.
1243:
1244: Equation~(\ref{4power}) for energy emission in the bulk may alternatively be
1245: written, in terms of the absorption coefficient, as
1246: %%%%%%%%%%%%%
1247: \begin{equation}
1248: \frac{d \tilde E(\om)}{dt} =
1249: \sum_{\ell} \tilde N_\ell\, |\tilde {\cal A}_\ell|^2\,
1250: \frac{\om}{\exp\left(\om/T_\text{H}\right) - 1}\,\,\frac{d \om}{2\pi}\,.
1251: \label{alter-bulk}
1252: \end{equation}
1253: %%%%%%%%%%%%%
1254: The above can be compared with the corresponding expression for the emission
1255: of brane-localized modes given, as in eq.~(\ref{pdecay-brane}), by
1256: %%%%%%%%%%%%%
1257: \begin{eqnarray}
1258: \frac{d \hat E (\om)}{dt} = \sum_{\ell} \hat N_\ell\, |\hat {\cal A}_\ell|^2\,
1259: \frac{\om}{\exp\left(\om/T_\text{H}\right) - 1}\,\,\frac{d \om}{2\pi}\,,
1260: \label{emission-br}
1261: \end{eqnarray}
1262: %%%%%%%%%%%%%
1263: where $\hat N_\ell =2\ell +1$.
1264: Since the temperature is the same for both bulk and brane modes, the bulk-to-brane ratio of the two energy emission rates will be simply given by the expression
1265: \begin{figure}
1266: \begin{center}
1267: \psfrag{x}[t][b][1.75]{$\omega r_\text{h}$}
1268: \scalebox{0.5}{\rotatebox{0}{\includegraphics[width=23cm, height=15cm]
1269: {nratio-ex.eps}}}
1270: \capbox{Bulk-to-brane relative emission rates for scalars}{Bulk-to-brane energy emission rates for scalar fields from a
1271: $(4+n)$D black hole.\label{bb-ratio}}
1272: \end{center}
1273: \end{figure}
1274: %%%%%%%%%%%%%
1275: \begin{equation}
1276: \frac{d \tilde E/dt}{d \hat E/dt} = \frac{\sum_{\ell} \tilde N_\ell\,
1277: |\tilde {\cal A}_\ell|^2}{\sum_{\ell} \hat N_\ell\,
1278: |\hat {\cal A}_\ell|^2}\,,
1279: \label{ratio}
1280: \end{equation}
1281: %%%%%%%%%%%%%
1282: \enlargethispage{-\baselineskip}and will depend on the scaling of the multiplicities of states and the
1283: absorption coefficients with $n$.\footnote{The absorption coefficients are related to the grey-body factors through eq.~(\ref{greydef}). This includes a multiplicative coefficient which depends on both $\omega r_\text{h}$ and $n$, and so the coefficients might have a completely different behaviour from the grey-body factors themselves.}
1284:
1285: As is clear from eq.~(\ref{bulk-mult}), the
1286: multiplicity of bulk modes $\tilde N_\ell$ increases quickly for increasing
1287: $n$, while $\hat N_\ell$ remains the same. However, it turns out that the
1288: enhancement with $n$ of the absorption probability $|\hat {\cal A}_\ell|^2$ for
1289: brane emission is considerably greater than
1290: the one for bulk emission. This leads to the dominance of the emission of
1291: brane-localized modes over bulk modes, particularly for intermediate values of $n$. The behaviour of this ratio is shown in Figure~\ref{bb-ratio}. This figure shows that, in the low-energy regime, the ratio for large $n$ is suppressed by many orders of magnitude, compared to the value of unity for $n=0$. In the high-energy regime on the other hand, the suppression becomes smaller and the ratio seems to approach unity. A more careful examination reveals that bulk modes in fact dominate over brane modes in a limited high-energy regime which becomes broader as $n$ increases.\footnote{For the very high energies for which this is true, it is not clear that the expressions for higher-dimensional Hawking radiation will remain valid.}
1292:
1293: A definite conclusion regarding the relative amount of energy emitted in the two `channels'---bulk and brane---can only be drawn if
1294: the corresponding total energy emissivities are computed. By integrating
1295: the areas under the bulk and brane energy emission rate curves, the relative energy emission rates are determined. The results obtained, for values of $n$ from 1 to 7, are given in Table~\ref{bb-ratios}.
1296:
1297: %%%%%%%%%%%%%%%%%%%
1298: \begin{table}
1299: \def\arraystretch{1.1}
1300: \begin{center}
1301: \begin{tabular}{|c|c|c|c|c|c|c|c|c|}
1302: \hline
1303: $n$ & 0 & 1 & 2 & 3 & 4 & 5 & 6 & 7 \\
1304: \hline
1305: Bulk/Brane & 1 & 0.40 & 0.24 & 0.22 & 0.24 & 0.33 & 0.52 & 0.93\\
1306: \hline
1307: \end{tabular}
1308: \capbox{Relative bulk-to-brane emission rates for scalars}{Relative bulk-to-brane energy emission rates for scalar fields.
1309: \label{bb-ratios}}
1310: \end{center}
1311: \def\arraystretch{1.0}
1312: \end{table}
1313: %%%%%%%%%%%%%%%%%%
1314:
1315:
1316: From the entries of the Table~\ref{bb-ratios}, it becomes clear that the emission of
1317: brane-localized scalar modes is dominant, in terms of the energy emitted, for all the values of $n$ considered. As $n$ increases, the
1318: ratio of bulk-to-brane emission gradually becomes smaller, and is
1319: particularly suppressed for intermediate values, i.e.\ $n=2$--5; in these cases, the total energy emitted in a bulk mode varies between approximately 1/3 and 1/4 of that emitted in
1320: a brane mode. As $n$ increases further, the high-energy dominance of the bulk
1321: modes mentioned above gives a boost to the value of the bulk-to-brane
1322: ratio---nevertheless, the energy ratio never exceeds unity.
1323:
1324: The above analysis provides exact results for the energy emission rates of brane and bulk scalar modes and gives considerable support to
1325: earlier, more heuristic, arguments \cite{Emparan:2000rs}, according to which a
1326: $(4+n)$-dimensional black hole emits mainly brane modes. A complete confirmation would mean performing a similar analysis for the emission of gravitons, but there are still theoretical and numerical issues which prevent this.
1327:
1328: \section{Rotating black holes}
1329: \label{rotbh}
1330:
1331: We have already seen that a black hole created in a hadron collider is expected to have some angular momentum $J$ about an axis perpendicular to the plane of parton collision.
1332: Using the same na\"{\i}ve argument as in \secref{bhprod}, we assume $J=bM_\text{BH}/2$ (where $b$ is the impact parameter of the colliding partons) and that $b<2r_\text{h}$ if a black hole is to form. The maximum possible value of the rotation parameter defined in eq.~(\ref{astardef}) is then found to be
1333: %%%%%%%%%
1334: \begin{equation}
1335: \label{astarmax}
1336: a_*^\text{max}=\frac{n+2}{2}\,.
1337: \end{equation}
1338: %%%%%%%%%
1339: The equivalent of eqs.~(\ref{fdecay-brane}) and (\ref{pdecay-brane}) are now
1340: %%%%%%%
1341: \begin{equation}
1342: \frac{d \hat N^{(s)}(\om)}{dt} = \sum_{\ell,m} |\hat {\cal A}^{(s)}_{\ell,m}|^2\,
1343: \frac{1}{\exp\left[(\om-m\Omega)/T_\text{H}\right] \mp 1}\,\frac{d\om}{2\pi}\,,
1344: \label{rflux}
1345: \end{equation}
1346: %%%%%%%%%%%%%%
1347: and
1348: %%%%%%%
1349: \begin{equation}
1350: \frac{d \hat N^{(s)}(\om)}{dt} = \sum_{\ell,m} |\hat {\cal A}^{(s)}_{\ell,m}|^2\,
1351: \frac{\om}{\exp\left[(\om-m\Omega)/T_\text{H}\right] \mp 1}\,\frac{d\om}{2\pi}\,,
1352: \label{rpower}
1353: \end{equation}
1354: %%%%%%%%%
1355: where the Hawking temperature is now given by
1356: %%%%%%%%%
1357: \begin{equation}
1358: T_\text{H}=\frac{(n+1)+(n-1)a_*^2}{4\pi(1+a_*^2)r_\text{h}}\,,
1359: \end{equation}
1360: %%%%%%%%%%
1361: and $\Omega$ is defined by
1362: %%%%%%%%%%
1363: \begin{equation}
1364: \Omega=\frac{a_*}{(1+a_*^2)r_\text{h}}\,.
1365: \end{equation}
1366: %%%%%%%%%%
1367:
1368: For the rotating case, the metric takes on a more complicated form than previously. In general $(4+n)$-dimensional objects are described by $(n+3)/2$ angular momentum parameters; however, it is assumed that here there is only one non-zero parameter (about an axis in the brane). This is reasonable because the partons which collide to produce the black hole are themselves on the brane. Hence the metric reduces to
1369: %%%%%%%%%%
1370: \begin{equation}
1371: \begin{split}
1372: ds^2=\left(1-\frac{\mu}{\Sigma r^{n-1}}\right)dt^2&+\frac{2 a\mu\sin^2\theta}{\Sigma r^{n-1}}dtd\varphi-\frac{\Sigma}{\Delta}dr^2 \\[3mm]
1373: &-\Sigma d\theta^2-\left(r^2+a^2+\frac{a^2\mu\sin^2\theta}{\Sigma r^{n-1}}\right)\sin^2\theta d\varphi^2,
1374: \end{split}
1375: \end{equation}
1376: %%%%%%%%
1377: where
1378: %%%%%%%%%
1379: \begin{equation}
1380: \Delta=r^2+a^2-\frac{\mu}{r^{n-1}} \quad\mbox{ and } \quad\Sigma=r^2+a^2\cos^2\theta,
1381: \end{equation}
1382: %%%%%%%%%%
1383: with $a=a_*r_\text{h}$ and $\mu=r_\text{S}^{n+1}$ ($r_\text{S}$ is the Schwarzschild radius of a non-rotating black hole of the same mass).
1384:
1385: Again the field equation is separable using a solution of the same form as in the non-rotating case, but now the radial equation is
1386: %%%%%%%%%
1387: \begin{equation}
1388: \Delta^{-s} \frac{d}{dr}\left(\Delta^{s+1}\frac{R_s}{dr}\right)+\left(\frac{K^2-isK\Delta'}{\Delta}+4is\omega r+s\Delta''-{}_s\Lambda^m_{\ell} \right)R_s=0\,,
1389: \end{equation}
1390: %%%%%%%%
1391: where
1392: %%%%%%%%%
1393: \begin{equation}
1394: K=(r^2+a^2)\omega-am \quad\mbox{ and } \quad{}_s\Lambda^m_{\ell}={}_sE^m_{\ell}+2s+a^2\omega^2-2am\omega\,.
1395: \end{equation}
1396: %%%%%%%%%%
1397: The angular equation is now
1398: %%%%%%%%%%
1399: \begin{multline}
1400: \label{spinang}
1401: \frac{1}{\sin\theta} \frac{d}{d\theta}\left(\sin\theta\frac{d \:{}_sT^m_{\ell}(\theta)}{d\theta}\right)+ \\[3mm]
1402: \left(-\frac{2ms\cot\theta}{\sin\theta}-\frac{m^2}{\sin^2\theta}+a^2\omega^2\cos^2\theta-2a\omega s\cos\theta+s-s^2\cot^2\theta+{}_sE^m_{\ell}\right){}_sT^m_{\ell}(\theta)=0\,,
1403: \end{multline}
1404: %%%%%%%%%
1405: where $e^{im\varphi}{}_sT^m_{\ell}(\theta)$ are known as spin-weighted spheroidal harmonics.
1406:
1407: The black hole horizon is given by solving $\Delta(r)=0$. Unlike the 4D Kerr black hole for which where are inner and outer solutions for $r_\text{h}$, for $n\geq1$ there is only one solution of this equation (as can easily be seen graphically---Figure~\ref{deltacomp}). In the $n=0$ case the maximum possible value of $a_*$ is 1 otherwise there are no solutions of $\Delta=0$ (this is forbidden as it would mean there is no horizon and hence there is a naked ring singularity at $r=0$). For $n>1$ there is no fundamental upper bound on $a_*$ but only the bound given in eq.~(\ref{astarmax}) which was argued geometrically. For general $n$, the horizon radius $r_\text{h}$ is found to be
1408: %%%%%%%%%
1409: \begin{equation}
1410: r_\text{h}=\frac{r_\text{S}}{(1+a_*^2)^{\frac{1}{n+1}}}\,.
1411: \end{equation}
1412: %%%%%%%%%
1413: \begin{figure}
1414: \unitlength1cm
1415: \begin{center}
1416: \begin{minipage}[t]{2.0in}
1417: \scalebox{0.55}{\rotatebox{0}{\includegraphics{nd0.eps}}}
1418: \end{minipage}
1419: \hfill
1420: \begin{minipage}[t]{2.0in}
1421: \scalebox{0.55}{\rotatebox{0}{\includegraphics{nd1.eps}}}
1422: \end{minipage}
1423: \hfill
1424: \begin{minipage}[t]{2.0in}
1425: \scalebox{0.55}{\rotatebox{0}{\includegraphics{nd4.eps}}}
1426: \end{minipage}
1427: \capbox{Number of solutions of $\Delta(r)=0$ for different $n$}{Number of solutions of $\Delta(r)=0$ for $n=0,1 \mbox{ and } 4$. On all three plots, the two functions shown are $f(r)=r^2+a^2$ and $g(r)=\mu r^{1-n}$ for particular values of $a$ and $\mu$.\label{deltacomp}}
1428: \end{center}
1429: \end{figure}
1430:
1431: The solution of the radial equation to obtain the grey-body factor is carried out in exactly the same same way as for the non-rotating black hole. There is a slight difference in the asymptotic solution close to the horizon in that the $\omega$ in eq.~(\ref{nh-new}) is replaced by $k$ where
1432: %%%%%%%%%%
1433: \begin{equation}
1434: k=w-\frac{ma}{r_\text{h}^2+a^2}\,,
1435: \end{equation}
1436: %%%%%%%%%
1437: and that the equivalent of eq.~(\ref{rstarr}) is now
1438: %%%%%%%%%%%%
1439: \begin{equation}
1440: \frac{dr^*}{dr}=\frac{r^2+a^2}{\Delta(r)}\,.
1441: \end{equation}
1442:
1443: \subsection{Calculation of the eigenvalue}
1444:
1445: The only other complication is to determine ${}_sE^m_{\ell}$ from the angular equation. The $\theta$ wave functions ${}_sT^m_{\ell}$ are no longer like those of the spin-weighted spherical harmonics and so ${}_sE^m_{\ell} \neq {}_s\lambda_{\ell}$\,. The eigenvalues of the spin-weighted spheroidal harmonics are functions of $a\omega$ and can be obtained by a continuation method \cite{Wasserstrom:1972}. This is a generalization of perturbation theory which can be applied for arbitrarily large changes in the initial Hamiltonian for which the eigenvalues are known. Making the $a\omega$-dependence of both $T$ and $E$ clear, eq.~(\ref{spinang}) can be written as
1446: %%%%%%%%%%%%
1447: \begin{equation}
1448: \label{opang}
1449: ({\mathcal H}_0+{\mathcal H}_1)\, {}_sT^m_{\ell}(\theta,a\omega) =-{}_sE^m_{\ell} (a\omega)\, {}_sT^m_{\ell}(\theta,a\omega)\,,
1450: \end{equation}
1451: %%%%%%%%%%%%
1452: where
1453: %%%%%%%%%
1454: \begin{equation}
1455: {\mathcal H}_0=\frac{1}{\sin\theta}\frac{d}{d\theta}\left(\sin\theta\frac{d}{d\theta}\right)+\left(-\frac{2ms\cot\theta}{\sin\theta}-\frac{m^2}{\sin^2\theta}+s-s^2\cot^2\theta\right),
1456: \end{equation}
1457: %%%%%%%%%
1458: and
1459: %%%%%%%%%%
1460: \begin{equation}
1461: {\mathcal H}_1=(a^2\omega^2\cos^2\theta-2a\omega s\cos\theta)\,.
1462: \end{equation}
1463: %%%%%%%%%%
1464: From these equations it is evident that
1465: %%%%%%%%%%%%
1466: \begin{equation}
1467: {}_sT^m_{\ell}(\theta,0)= {}_sS^m_{\ell}(\theta)\,,
1468: \label{init1}
1469: \end{equation}
1470: %%%%%%%%%
1471: and that
1472: %%%%%%%%%
1473: \begin{equation}
1474: {}_sE^m_{\ell}(0)={}_s\lambda_{\ell}=\ell(\ell+1)-s(s+1)\,.
1475: \label{init2}
1476: \end{equation}
1477: %%%%%%%%%%%%
1478:
1479: The continuation method was outlined in \cite{Press:1973} although this paper unfortunately seems to contain a number of errors. First consider the case in which $a\omega$ is small and normal perturbation theory can be used. This means we have
1480: %%%%%%%%%%%%%%
1481: \begin{equation}
1482: \label{pte}
1483: {}_sE^m_{\ell}(a\omega)={}_s\lambda_{\ell}-\langle s\ell m|{\mathcal H}_1|s\ell m\rangle+\ldots\,,
1484: \end{equation}
1485: %%%%%%%%%%%%%
1486: and
1487: %%%%%%%%%%%%
1488: \begin{equation}
1489: \label{pts}
1490: {}_sT^m_{\ell}(a\omega)={}_sS^m_{\ell}-\sum_{\ell'\neq l} \frac{\langle s\ell'm|{\mathcal H}_1|s\ell m\rangle}{\ell(\ell+1)-\ell'(\ell'+1)} {}_sS^m_{\ell'}+\ldots\,,
1491: \end{equation}
1492: %%%%%%%%%%%%
1493: where
1494: %%%%%%%%%%%%
1495: \begin{equation}
1496: \label{prod}
1497: \langle s\ell'm|{\mathcal H}_1|s\ell m\rangle \equiv \int d\Omega\,({}_sS_{\ell'}^m)^* {}_sS_{\ell}^m \,{\mathcal H}_1\,.
1498: \end{equation}
1499: %%%%%%%%%%%%%
1500: Care must be taken about the signs of the terms in eqs.~(\ref{pte}) and (\ref{pts}) due to the sign of the eigenvalue ${}_sE^m_{\ell}$ in eq.~(\ref{opang}). This seems to account for some differing results in \cite{Press:1973}.
1501:
1502: The product in eq.~(\ref{prod}) can be split into terms for which the following standard formulae are useful:
1503: %%%%%%%%%%%
1504: \begin{equation}
1505: \langle s\ell'm|\cos^2\theta|s\ell m\rangle=\frac{1}{3}\delta_{\ell\ell'}+\frac{2}{3}\left(\frac{2\ell +1}{2\ell'+1}\right)^{\frac{1}{2}}\langle\ell2m0|\ell'm\rangle\langle\ell2(\!-\!s)0|\ell'(\!-\!s)\rangle,
1506: \end{equation}
1507: %%%%%%%%%%
1508: and
1509: %%%%%%%%%
1510: \begin{equation}
1511: \langle s\ell'm|\cos\theta|s\ell m\rangle=\left(\frac{2\ell +1}{2\ell'+1}\right)^{\frac{1}{2}}\langle\ell1m0|\ell'm\rangle\langle\ell1(\!-\!s)0|\ell'(\!-\!s)\rangle,
1512: \end{equation}
1513: %%%%%%%%%%
1514: where $\langle\ell_1\ell_2m_1m_2|LM\rangle$ are Clebsch-Gordan coefficients. Using the above we obtain
1515: %%%%%%%%%%
1516: \begin{equation}
1517: {}_sE^m_{\ell}= \left\{ \begin{array}{ll}
1518: {}_s\lambda_{\ell}-2a\omega \frac{s^2m}{\ell(\ell+1)}+{\mathcal O}[(a\omega)^2] & \mbox{if $s \neq 0$}\,;\\
1519: \\
1520: {}_s\lambda_{\ell}+a^2\omega^2\left[\frac{2m^2+1-2\ell(\ell+1)}{(2\ell-1)(2\ell-3)}\right]+{\mathcal O}[(a\omega)^4] & \mbox{if $s=0$}\,.\end{array} \right.
1521: \end{equation}
1522: %%%%%%%%%%
1523: Again this differs (in the $s=0$ case) from the result given in \cite{Press:1973}. However it agrees with \cite{Seidel:1989ue} where several discrepancies in the literature are helpfully clarified.
1524:
1525: To employ the continuation method we write the ${}_sT^m_{\ell}(\theta,a\omega)$ functions in the basis of the $\theta$-parts of the spin-weighted spherical harmonics:
1526: %%%%%%%%
1527: \begin{equation}
1528: {}_sT^m_{\ell}(\theta,a\omega)=\sum_{\ell'}{}_sB_{\ell\ell'}^m(a\omega)S_{s\ell'}^m(\theta)\,.
1529: \end{equation}
1530: %%%%%%%%%%
1531: By differentiating eq.~(\ref{opang}) and applying the same techniques as in perturbation theory it is possible to obtain the results
1532: %%%%%%%%%%
1533: \begin{equation}
1534: \label{conte}
1535: \frac{d \, {}_sE_{\ell}^m}{d(a\omega)}=-\sum_{\alpha,\beta}{}_sB_{\ell\alpha}^m \,{}_sB_{\ell\beta}^m\langle\alpha|\beta\rangle,
1536: \end{equation}
1537: %%%%%%%%%%
1538: and
1539: %%%%%%%%%%
1540: \begin{equation}
1541: \label{contb}
1542: \frac{d \, {}_sB_{\ell\ell'}}{d(a\omega)}=-\sum_{\alpha,\beta,\gamma \neq l}\frac{{}_sB_{\gamma\alpha} \, {}_sB_{\ell\beta}}{{}_sE^m_{\ell}-{}_sE_{\gamma}^m}\langle\alpha|\beta\rangle{}_sB_{\gamma \ell'},
1543: \end{equation}
1544: %%%%%%%%%%
1545: where $\langle\alpha|\beta\rangle \equiv \langle s\alpha m|d{\mathcal H}_1/d(a\omega)|s\beta m\rangle$. The initial conditions are obtained from eqs. (\ref{init1}) and (\ref{init2}) which imply that ${}_sB_{\ell\ell'}^m(0)=\delta_{\ell\ell'}$. By integrating eqs.~(\ref{conte}) and (\ref{contb}) it is possible to obtain the eigenvalues of the spin-weighted spheroidal functions for any $s,\ell$ and $m$ and for arbitrarily large values of $a\omega$. This integration was performed by using the \texttt{NDSolve} package in \texttt{Mathematica} to implement a Runge-Kutta method. The computational effort involved increases significantly for larger values of $\ell$ as it is necessary to include a larger number of terms in the summations, increasing the number of differential equations to be solved.
1546:
1547: Although only the $s=0$ eigenvalues were required here, as a check the continuation method was also used to calculate the $s=1$ eigenvalues as a function of $a\omega$. This allowed comparison with the polynomial approximations given in \cite{Press:1974}. Excellent agreement was found---see Figure~\ref{eigen} for an example.
1548:
1549: \begin{figure}
1550: \begin{center}
1551: \psfrag{x}[][][1.5]{$a\omega$}
1552: \psfrag{y}[][][1.5]{${}_sE^m_{\ell}(a\omega)$}
1553: \scalebox{0.7}{\rotatebox{0}{
1554: \includegraphics[width=\textwidth]{neigen.eps}}}
1555: \capbox{Comparison of eigenvalue with previous result}{Comparison of the numerical result obtained in this work for the $s=1,\ell=3,m=-2$ eigenvalue (solid line) with the polynomial approximation in \cite{Press:1974} (dashed line). The lines are exactly on top of each other (the maximum difference between them is $\sim 10^{-5}$).\label{eigen}}
1556: \end{center}
1557: \end{figure}
1558:
1559: \subsection{Calculation of the grey-body factor}
1560:
1561: In order to calculate the grey-body factors in the rotating case, it is necessary to know the equivalent of eqs.~(\ref{scalars}), (\ref{fermions}) and (\ref{bosons}). Unfortunately, for fermions and bosons these expressions are not available in the literature. However for the scalar case, since eq.~(\ref{scalars}) only involves coefficients at infinity, exactly the same formula holds even when the black hole is rotating. This means that for scalars it has been possible to numerically calculate the grey-body factors for rotating black holes.
1562:
1563: Since the denominator in eqs.~(\ref{rflux}) and (\ref{rpower}) has $m$-dependence, it is less useful to plot an equivalent of the $\hat{\sigma}$ used in the non-rotating case,\footnote{The absorption cross section $\hat{\sigma}_\text{abs}$ summed over all angular momentum modes \emph{can} still be calculated; the high-energy asymptotic value appears to increase with $a_*$, apparently in contradiction with \cite{Page:1976df} which states that roughly the same value is expected as in the non-rotating case.} so only the power spectra are shown. They are presented as a function of $\omega r_\text{h}$ for different values of $a_*$ and hence cannot be used to directly compare the spectra from two black holes of the same mass (but different $J$) since this is itself a function of $M_\text{BH}$.
1564:
1565: \begin{figure}
1566: \begin{center}
1567: \psfrag{x}[][][1.2]{$\omega r_\text{h}$}
1568: \psfrag{y}[][][1.2]{$r_\text{h} d^2 \hat{E}^{(0)}/dtd\om$}
1569: \scalebox{0.8}{\rotatebox{0}{\includegraphics[width=\textwidth]{nefiopfig2.eps}}}
1570: \capbox{Power spectra for scalar emission from rotating black holes ($n=1$)}{Power spectra for scalar emission from rotating black holes ($n=1$).\label{s0a}}
1571: \end{center}
1572: \end{figure}
1573:
1574: Figure~\ref{s0a} can be directly compared to Figure 2 in \cite{Ida:2002ez} and shows markedly different behaviour for values of $\omega r_\text{h}$ larger than $\sim 0.2$. Almost all of this effect is due to the low-energy expansion used by those authors to obtain the grey-body factors. Their approximation that ${}_sE^m_{\ell}={}_s\lambda_{\ell}=\ell(\ell+1)-s(s+1)$ is much less significant except at the very highest values of $\omega r_\text{h}$.
1575:
1576: These numerical results for the grey-body factors also allowed the statement in \cite{Giddings:2001bu} that emission is dominated by modes with $\ell=m$ to be tested. It was found that a plot like Figure~\ref{s0a} looks the same at the $\sim$~90\% level if only the the $\ell=m$ modes are included in the sum of eq.~(\ref{rpower})
1577:
1578: \subsection{Super-radiance}
1579:
1580: The numerical results obtained allow the possibility of super-radiance (i.e.\ negative grey-body factors) to be investigated in an extra-dimensional context. Previously an analytic approach \cite{Frolov:2002xf} was used to confirm super-radiance for bulk scalars incident on 5D black holes, but there do not seem to be any other results in the literature for higher-dimensional black holes. To study this phenomenon it is useful to consider angular momentum modes individually and, for comparison with the 4D scalar results presented in \cite{Press:1972}, the quantity on the horizontal axis of the plots is
1581: %%%%%
1582: \begin{equation}
1583: \frac{\omega}{m \Omega}=\frac{(1+a_*^2)(\omega r_\text{h})}{m a_*}\,,
1584: \end{equation}
1585: %%%%%
1586: rather than the $\omega r_\text{h}$ used in previous plots. The vertical axis is the absorption probability (in fact $-|\hat {\cal A}_{\ell,m}|^2$) expressed as a percentage so that it gives the percentage energy amplification of an incident wave.
1587: %%%%%%%%
1588: \begin{figure}
1589: \begin{center}
1590: \psfrag{x}[][][1.5]{$\omega/m\Omega$}
1591: \psfrag{a}[][][1.5][18]{\textsf{l=m=1}}
1592: \psfrag{b}[][][1.5][30]{\textsf{l=m=2}}
1593: \psfrag{c}[][][1.5][37]{\textsf{l=m=3}}
1594: \psfrag{d}[][][1.5][40]{\textsf{l=m=4}}
1595: \psfrag{e}[][][1.5][45]{\textsf{l=m=5}}
1596: \psfrag{f}[][][1.5]{\textsf{l=2, m=1}}
1597: \scalebox{0.5}{\rotatebox{0}{\includegraphics[width=\textwidth]{super1.eps}}}
1598: \capbox{Super-radiant scattering by 4D black holes with $a_*=1$}{Super-radiant scattering of scalars by maximally rotating ($a_*=1$) 4-dimensional black holes.\label{super1}}
1599: \end{center}
1600: \end{figure}
1601: %%%%%%%%
1602: Figure~\ref{super1} shows excellent agreement with the results in Figure 1 of \cite{Press:1972} which are for $n=0$ and $a_*=1$ (equivalent to $a=M_\text{BH}$ in the 4D case). Figure~\ref{super2} is for $n=2$ and 6 but still with the same value of the rotation parameter. It is clear that the peak amplification is more significant than previously and becomes comparable to that for gauge bosons in the 4D case. It is still found that the $\ell=m=1$ mode provides the greatest amplification. However Figure~\ref{super3} shows that for larger values of $a_*$ the maximum amplification can occur in modes with larger values of $\ell$. This trend continues as $n$ and $a_*$ increase further: for $n=6$, equation~(\ref{astarmax}) shows that the geometric maximal rotation is $a_*^\text{max}=4.0$ and for this value of the rotation parameter the maximum energy amplification is found to be just under 9\% (and occurs in the $\ell=m=7$ mode).
1603:
1604: \begin{figure}
1605: \unitlength1cm
1606: \begin{center}
1607: \psfrag{x}[][][0.8]{$\omega/m\Omega$}
1608: \psfrag{a}[][][0.8][18]{\textsf{l=m=1}}
1609: \psfrag{b}[][][0.8][28]{\textsf{l=m=2}}
1610: \psfrag{c}[][][0.8][35]{\textsf{l=m=3}}
1611: \psfrag{d}[][][0.8][40]{\textsf{l=m=4}}
1612: \psfrag{e}[][][0.8][43]{\textsf{l=m=5}}
1613: \psfrag{f}[][][0.8]{\textsf{l=2, m=1}}
1614: \psfrag{g}[][][0.8][20]{\textsf{l=m=1}}
1615: \psfrag{h}[][][0.8][30]{\textsf{l=m=2}}
1616: \psfrag{i}[][][0.8][37]{\textsf{l=m=3}}
1617: \psfrag{j}[][][0.8][42]{\textsf{l=m=4}}
1618: \psfrag{k}[][][0.8][45]{\textsf{l=m=5}}
1619: \psfrag{l}[][][0.8]{\textsf{l=2, m=1}}
1620: \begin{minipage}[t]{3.05in}
1621: %\scalebox{0.55}
1622: {\rotatebox{0}{\includegraphics{super2.eps}}}
1623: \end{minipage}
1624: \hfill
1625: \begin{minipage}[t]{3.05in}
1626: %\scalebox{0.55}
1627: {\rotatebox{0}{\includegraphics{super3.eps}}}
1628: \end{minipage}
1629: \capbox{Super-radiant scattering by $(4+n)$D black holes with $a_*=1$}{Super-radiant scattering of scalars by $(4+n)$-dimensional black holes with $a_*=1$: {\bf(a)} $n=2$ and {\bf(b)} $n=6$.\label{super2}}
1630: \end{center}
1631: \end{figure}
1632:
1633: \begin{figure}
1634: \unitlength1cm
1635: \begin{center}
1636: \psfrag{x}[][][1]{$\omega/m\Omega$}
1637: \psfrag{a}[r][][1]{\textsf{l=m=1}}
1638: \psfrag{b}[r][][1]{\textsf{l=m=2}}
1639: \psfrag{c}[][][1][60]{\textsf{l=m=3}}
1640: \psfrag{d}[][][1][60]{\textsf{l=m=4}}
1641: \psfrag{e}[l][][1]{\textsf{l=m=5}}
1642: \psfrag{f}[r][][1]{\textsf{l=m=1}}
1643: \psfrag{g}[r][][1]{\textsf{l=m=2}}
1644: \psfrag{h}[][][1]{\textsf{l=m=3}}
1645: \psfrag{i}[][][1][64]{\textsf{l=m=4}}
1646: \psfrag{j}[l][][1]{\textsf{l=m=5}}
1647: \begin{minipage}[t]{3.05in}
1648: \scalebox{0.8}
1649: {\rotatebox{0}{\includegraphics{super4.eps}}}
1650: \end{minipage}
1651: \hfill
1652: \begin{minipage}[t]{3.05in}
1653: \scalebox{0.8}
1654: {\rotatebox{0}{\includegraphics{super5.eps}}}
1655: \end{minipage}
1656: \capbox{Super-radiant scattering by $(4+n)$D black holes with $a_*>1$}{Super-radiant scattering of scalars by $(4+n)$-dimensional black holes: {\bf(a)} $n=2$, $a_*=1.5$ and {\bf(b)} $n=6$, $a_*=2$.\label{super3}}
1657: \end{center}
1658: \end{figure}
1659:
1660: As in the 4-dimensional case, the region in which the grey-body factors are negative exactly corresponds to the region in which the thermal factor in the denominator of equations like (\ref{rflux}) and (\ref{rpower}) is negative; hence the quantum emission rate always remains positive. For all angular momentum modes the grey-body factor is positive for $\omega > m \Omega$ and the change from negative to positive at $\omega = m \Omega$ occurs such that the flux and energy spectra are smooth positive functions. Therefore although super-radiance can in principle be used to extract energy from rotating black holes, its significance when considering the Hawking radiation spectra is limited.
1661:
1662: \section{Conclusions}
1663: \label{gbfconc}
1664:
1665: The revival of the idea of extra space-like dimensions
1666: in nature has led to the formulation of theories which allow gravity to become strong at a significantly lower energy scale. This has opened the way for the proposal of the creation of miniature higher-dimensional black holes during collisions of energetic particles in the earth's atmosphere or at ground-based particle colliders. The Hawking radiation emitted by such black holes is described by the relevant grey-body factors; a wide variety of these factors have been calculated numerically in the work described in this chapter and, at the same time, some open questions from previous analyses in the literature have been addressed.
1667:
1668: The exact numerical results obtained in \secref{numres} for the grey-body factors and emission rates of scalars, fermions and gauge bosons confirm that previous analytic results \cite{Kanti:2002nr,Kanti:2002ge} were only valid in limited low-energy regimes. The grey-body factors depend on both the dimensionality of space-time and the spin. They adopt the same high-energy asymptotic values (which decrease as $n$ increases) for all particle species, confirming the geometrical optics limit of \cite{Emparan:2000rs}. In the low-energy limit the grey-body factors are found to be enhanced as $n$ increases for $s=1$ and 1/2 and suppressed for $s=0$.
1669:
1670: Using these grey-body factors it was possible to accurately calculate the flux and power emission spectra for the black holes. As the number of extra dimensions projected on the brane increases there is a substantial (orders of magnitude) enhancement in the energy emission rate, mainly as a result of the increase in black hole temperature with $n$. However the effect of including the grey-body factors is that the increase in the rate depends on the spin of the particle studied. The computed relative emissivities show that scalar fields, which are the dominant degree of freedom emitted by $n=0$ black holes, are just outnumbered by gauge bosons for large values of $n$, with the fermions becoming the least effective emission channel. Therefore the emission spectra and the relative particle emissivities may possibly both lead to the determination of the number of extra dimensions.
1671:
1672: The emission spectra for modes of different spins were combined with the total number of degrees of freedom for all the particle types which could be emitted from an extra-dimensional black hole. This made it possible to calculate the total flux and power emission and hence also the expected time for, and average multiplicity of, the black hole decay. These calculations suggest that some of the assumptions usually made when modelling black hole decay by sequential emission of Hawking radiation may not be valid for the larger values of $n$.
1673: \enlargethispage{-\baselineskip}
1674:
1675: In section \ref{embulk}, the details of the emission of bulk scalar modes were investigated with the aim being to provide an accurate estimate for the amount of energy lost into the bulk. Comparing the total bulk and brane emission rates for scalar modes, integrated over the whole energy regime, the conclusion is that more of the black hole energy is emitted in brane modes than bulk modes, for all values of $n$ considered. The total emissivity in the bulk is less than 1/4 of that on the brane for $n=2$--4, while it becomes substantial for the extreme value of $n=7$ without, however, exceeding the brane value. The accurate results presented in this chapter provide firm support to the heuristic arguments made in ref. \cite{Emparan:2000rs}.
1676:
1677: No results are presented in this work concerning the emission of gravitons either on the brane or in the bulk. The derivation of a consistent equation, which can describe the motion of gravitons in the induced black hole background on the brane, is still under investigation. Nevertheless it is expected that, as for the emission of fields with spin $s=0$, 1/2 and 1 the graviton emission becomes enhanced as the dimensionality of space-time increases, but remains subdominant compared to the other species at least for small values of $n$.
1678:
1679: The work on rotating black holes in \secref{rotbh} suggests that, at least for the emission of scalars, the power spectra can be significantly modified compared to those in the Schwarzschild-like case. The spectra are dominated by the angular momentum modes with $\ell=m$ and as the rotation parameter $a_*$ increases, the peak in the spectrum shifts to much larger values of $\omega r_\text{h}$. The phenomenon of super-radiance was also observed for higher-dimensional rotating black holes. The peak energy amplification was found to be significantly larger than in the four-dimensional case and occurred in higher angular momentum modes for larger values of the rotation parameter.
1680: \enlargethispage{\baselineskip}
1681:
1682: A final comment is appropriate here concerning the validity of the results obtained in this work. As pointed out in the text, the horizon of the black hole is an input parameter of the analysis. The results are applicable for all values of $r_\text{h}$ which are smaller than the size of the extra dimensions $R$, no matter how small or large $R$ is (as long as it remains considerably larger than $\ell_\text{P}$ to avoid quantum corrections). The analysis therefore remains valid for all theories postulating the existence of flat extra dimensions and a fundamental scale of gravity even a few orders of magnitude lower than $M_\text{P}$. If the emission of Hawking radiation from these black holes is successfully detected, either at next-generation colliders or in the more distant future, the distinctive features discussed here may help in the determination of the number of extra dimensions existing in nature. Chapter~\ref{generator} describes a black hole event generator which incorporates these grey-body effects and would be a useful tool in such an analysis.