hep-th0507096/v3.tex
1: %Revised Version:Last modified by N.I. by August 13th.
2: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3: %Preamble
4: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
5: \documentclass[12pt]{article} \usepackage{epsfig,amsfonts,amssymb}
6: %%%% uncomment line below to see all the equation names
7: % \usepackage{showkeys}
8: %\usepackage[active]{srcltx}
9:  
10: \def\Tr{\hbox{Tr}} \newcommand{\be}{\begin{equation}}
11:   \newcommand{\ee}{\end{equation}} \newcommand{\bea}{\begin{eqnarray}}
12:   \newcommand{\eea}{\end{eqnarray}}
13: \newcommand{\beas}{\begin{eqnarray*}}
14:   \newcommand{\eeas}{\end{eqnarray*}} \def\kl{{\frac{2 \pi l}{\beta}}}
15: \def\km{{\frac{2 \pi m}{\beta}}} \def\kn{{\frac{2 \pi n}{\beta}}}
16: \def\kr{{\frac{2 \pi r}{\beta}}} \def\ks{{\frac{2 \pi s}{\beta}}}
17: \def\b{{\beta}} \font\cmsss=cmss8 \def\C{{\hbox{\cmsss C}}}
18: \font\cmss=cmss10 \def\bigC{{\hbox{\cmss C}}}
19: \def\scriptlap{{\kern1pt\vbox{\hrule height 0.8pt\hbox{\vrule width
20:         0.8pt \hskip2pt\vbox{\vskip 4pt}\hskip 2pt\vrule width
21:         0.4pt}\hrule height 0.4pt} \kern1pt}} \def\ba{{\bar{a}}}
22: \def\bb{{\bar{b}}} \def\bc{{\bar{c}}} \def\bphi{{\Phi}}
23: \def\kkappa{{\beta}} \def\Bigggl{\mathopen\Biggg}
24: \def\Bigggr{\mathclose\Biggg} \def\Biggg#1{{\hbox{$\left#1\vbox to
25:         25pt{}\right.\n@space$}}} \def\n@space{\nulldelimiterspace=0pt
26:   \m@th} \def\m@th{\mathsurround = 0pt}
27: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
28: % Kevin's short cuts
29: %%%%%%%%%%%%%%%%%%%%%%%
30: \newcommand{\ta}{\tilde{\alpha}}
31: \newcommand{\taod}{\frac{-\ta}{\alpha-\ta}} \newcommand{\PP}{Q_2}
32: \newcommand{\Q}{Q_1} \newcommand{\aod}{\frac{\alpha}{\alpha-\ta}}
33: \newcommand{\nosum}{\qquad\mbox{(no summation)}}
34: \newcommand{\ha}{\frac{1}{2}}
35: \newcommand{\textha}{{\textstyle{\frac{1}{2}}}}
36: 
37: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
38: 
39: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
40: % Document starts here
41: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
42: \begin{document}
43: 
44: \voffset -0.7 true cm \hoffset 1.1 true cm \topmargin 0.0in
45: \evensidemargin 0.0in \oddsidemargin 0.0in \textheight 8.6in
46: \textwidth 7.25in \parskip 10 pt
47: 
48: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
49: % title page
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: \begin{titlepage}
52:   \begin{flushright}
53:     {\small TIFR/TH/05-25} \\
54:     {\small hep-th/0507096}
55:   \end{flushright}
56: 
57:   \begin{center}
58: 
59:     \vspace{26mm}
60: 
61:     {\LARGE \bf Non-Supersymmetric Attractors}
62:     
63:  
64:  
65:     \vspace{10mm}
66: 
67:     Kevin Goldstein, Norihiro Iizuka, Rudra P. Jena and Sandip P.
68:     Trivedi
69: 
70:     \vspace{5mm}
71: 
72:     {\small \sl Tata Institute of Fundamental Research} \\
73:     {\small \sl Homi Bhabha Road, Mumbai, 400 005, INDIA} \\
74:     {\small \tt kevin, iizuka, rpjena, sandip@theory.tifr.res.in}
75:     \vspace{10mm}
76: 
77:   \end{center}
78: 
79:   \vskip 0.3 cm \centerline{\bf Abstract} \vspace{5mm}
80:   \noindent
81:   We consider theories with gravity, gauge fields and scalars in
82:   four-dimensional asymptotically flat space-time.  By studying the
83:   equations of motion directly we show that the attractor mechanism
84:   can work for non-supersymmetric extremal black holes.  Two
85:   conditions are sufficient for this, they are conveniently stated in
86:   terms of an effective potential involving the scalars and the
87:   charges carried by the black hole.  Our analysis applies to black
88:   holes in theories with ${\cal N}\leqslant 1$ supersymmetry, as well as
89:   non-supersymmetric black holes in theories with ${\cal N}=2$
90:   supersymmetry. Similar results are also obtained for extremal black
91:   holes in asymptotically Anti-de Sitter space and in higher
92:   dimensions.
93:  
94: \end{titlepage}
95: \setcounter{tocdepth}{2}
96: \tableofcontents
97: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
98: \section{Introduction}
99: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
100: 
101: Black holes in ${\cal N}=2$ supersymmetric theories are known to
102: exhibit a fascinating phenomenon called the the attractor mechanism.
103: There is a family of black hole solutions in these theories which are
104: spherically symmetric, extremal black holes, with double-zero horizons
105: \footnote{By a double-zero horizon we mean a horizon for which the
106:   surface gravity vanishes because the $g_{00}$ component of the
107:   metric has a double-zero (in appropriate coordinates), as in an
108:   extremal Reissner Nordstrom black hole.}.  In these solutions
109: several moduli fields are drawn to fixed values at the horizon of the
110: black hole regardless of the values they take at asymptotic infinity.
111: The fixed values are determined entirely by the charges carried by the
112: black hole.  This phenomenon was first discussed by
113: \cite{Ferrara:1995ih} and has been studied quite extensively since
114: then
115: \cite{Cvetic:1995bj,Strominger:1996kf,Ferrara:1996dd,Ferrara:1996um,Cvetic:1996zq,Ferrara:1997tw,Gibbons:1996af,Denefa,Denef:2001xn}.
116: It has gained considerable attention recently due to the conjecture of
117: \cite{Ooguri:2004zv} and related developments
118: \cite{LopesCardoso:1998wt,Dabholkar:2004yr,Ooguri:2005vr,Dijkgraaf:2005bp}.
119: 
120: So far the attractor phenomenon has been studied almost exclusively in
121: the context of BPS black holes in the ${\cal N}=2$ theories.  The aim
122: of this paper is to examine if it is more general and can happen for
123: non-supersymmetric black holes as well. These black holes might be
124: solutions in theories which have no supersymmetry or might be
125: non-supersymmetric solutions in ${\cal N} \geqslant 1$ supersymmetric
126: theories.
127: 
128: There are two motivations for this investigation.  First, a
129: non-supersymmetric attractor mechanism might help in the study of
130: non-supersymmetric black holes, especially their entropy. Second,
131: given interesting parallels between flux compactifications and the
132: attractor mechanisms, a non-supersymmetric attractor phenomenon might
133: lead to useful lessons for non-supersymmetric flux compactifications.
134: For example, it could help in finding dual descriptions of such
135: compactifications. This might help to single out vacua with a small
136: cosmological constant. Or it might suggest ways to weight vacua with
137: small cosmological constants preferentially while summing over all of
138: them \footnote{For a recent attempt along these lines where
139:   supersymmetric compactifications have been considered, see
140:   \cite{Ooguri:2005vr,Dijkgraaf:2005bp}.}.  These lessons would be
141: helpful in light of the vast number of vacua that have been recently
142: uncovered in string theory \cite{KKLT}.
143: 
144: 
145: An intuitive argument for the attractor mechanism is as follows. One
146: expects that the total number of microstates corresponding to an
147: extremal black hole is determined by the quantised charges it carries,
148: and therefore does not vary continuously. If the counting of
149: microstates agrees with the Bekenstein-Hawking entropy, that is the
150: horizon area, it too should be determined by the charges alone.  This
151: suggests that the moduli fields which determine the horizon area take
152: fixed values at the horizon, and these fixed values depend only on the
153: charges, independent of the asymptotic values for the moduli.  While
154: this argument is only suggestive what is notable for the present
155: discussion is that it does not rely on supersymmetry. This provides
156: further motivation to search for a non-supersymmetric version of the
157: attractor mechanism.
158: 
159: 
160: 
161: The theories we consider in this paper consist of gravity, gauge
162: fields and scalar fields.  The scalars determine the gauge couplings
163: and there by couple to the gauge fields.  It is important that the
164: scalars do not have a potential of their own that gives them in
165: particular a mass.  Such a potential would mean that the scalars are
166: no longer moduli.
167:  
168: We first study black holes in asymptotically flat four dimensions.
169: Our main result is to show that the attractor mechanism works quite
170: generally in such theories provided two conditions are met. These
171: conditions are succinctly stated in terms of an ``effective
172: potential'' $V_{eff}$ for the scalar fields, $\phi_i$.  The effective
173: potential is proportional to the energy density in the electromagnetic
174: field and arises after solving for the gauge fields in terms of the
175: charges carried by the black hole, as we explain in more detail below.
176: The two conditions that need to be met are the following. First, as a
177: function of the moduli fields $V_{eff}$ must have a critical point,
178: $\partial_iV_{eff}(\phi_{i0})=0$. And second, the matrix of second
179: derivatives of the effective potential at the critical point,
180: $\partial_{ij}V_{eff}(\phi_{i0})$, must be have only positive
181: eigenvalues. The resulting attractor values for the moduli are the
182: critical values, $\phi_{i0}$. And the entropy of the black hole is
183: proportional to $V(\phi_{i0})$, and is thus independent of the
184: asymptotic values for the moduli. It is worth noting that the two
185: conditions stated above are met by BPS black hole attractors in an
186: ${\cal N}=2$ theory.
187: 
188: 
189: 
190: The analysis for BPS attractors simplifies greatly due to the use of
191: the first order equations of motion.  In the non-supersymmetric
192: context one has to work with the second order equations directly and
193: this complicates the analysis.  We find evidence for the attractor
194: mechanism in three different ways.  First, we analyse the equations
195: using perturbation theory.  The starting point is a black hole
196: solution, where the asymptotic values for the moduli equal their
197: critical values.  This gives rise to an extremal Reissner Nordstrom
198: black hole. By varying the asymptotic values a little at infinity one
199: can now study the resulting equations in perturbation theory. Even
200: though the equations are second order, in perturbation theory they are
201: linear, and this makes them tractable.  The analysis can be carried
202: out quite generally for any effective potential for the scalars and
203: shows that the two conditions stated above are sufficient for the
204: attractor phenomenon to hold.
205:  
206: Second, we carry out numerical analysis.  This requires a specific
207: form of the effective potential, but allows us to go beyond the
208: perturbative regime.  The numerical analysis corroborates the
209: perturbation theory results mentioned above.  In simple cases we have
210: explored so far, we have found evidence for a only single basin of
211: attraction, although multiple basins must exist in general as is
212: already known from the SUSY cases.
213: 
214: Finally, in some special cases, we solve the equations of motion
215: exactly by mapping them a solvable Toda system. This allows us to
216: study the black hole solutions in these special cases in some depth.
217: Once again, in all the cases we have studied, we can establish the
218: attractor phenomenon.
219: 
220: It is straightforward to generalise these results to other settings.
221: We find that the attractor phenomenon continues to hold in Anti-de
222: Sitter space (AdS) and also in higher dimensions, as long as the two
223: conditions mentioned above are valid for a suitable defined effective
224: potential.  There is also possibly an attractor mechanism in de Sitter
225: space (dS), but in the simplest of situations analysed here some
226: additional caveats have to be introduced to deal with infrared
227: divergences in the far past (or future) of dS space.
228: 
229: 
230: 
231: This paper is structured as follows. Black holes in asymptotically
232: flat four dimensional space are analysed first, in sections 2,3,4.
233: The discussion is extended to asymptotically flat space-times of
234: higher dimension in section 5.  Asymptotically AdS space is discussed
235: next in section 6.
236: 
237: As was mentioned above our analysis in the asymptotically flat and AdS
238: cases is based on theories which have no potential for the scalars so
239: that their values can vary at infinity.  Some comments on this are
240: contained in Section 7.  With ${\cal N}\geqslant 1$ SUSY such theories can
241: arise, with the required couplings between scalars and gauge fields,
242: and are at least technically natural. In the absence of supersymmetry
243: there is no natural way to arrange this and our study is more in the
244: nature of a mathematical investigation.  We follow in section 8, with
245: some comments on the attractor phenomenon in dS.  Finally, in section
246: 9 we show that non-extremal black holes do not have an attractor
247: mechanism.  Thus, the double-zero nature of the horizon is essential
248: to draw the moduli to fixed values.
249: 
250: 
251: Several important intermediate steps in the analysis are discussed in
252: appendices \ref{sec:appa}-\ref{sec:ads}.
253: 
254: Some important questions are left for the future. First, we have not
255: analysed the stability of these black hole solutions.  It is unlikely
256: that there are any instabilities at least in the S-wave sector. We do
257: not attempt a general analysis of small fluctuations here.  Second, in this paper  we
258: have not analysed string theory situations where such
259: non-supersymmetric black holes can arise \cite{TT3}.  This could include both
260: critical and non-critical string theory. In case of ${\cal
261:   N}=1$ supersymmetry it would be interesting to explore if there is
262: partial restoration of supersymmetry at the horizon. Given the
263: rotational invariance of the solutions one can see that no
264: supersymmetry is preserved in-between asymptotic infinity and the
265: horizon in this case. 
266: 
267: Let us also briefly comment on some of the literature of especial
268: relevance. The importance of the effective potential, $V_{eff}$, for
269: ${\cal N}=2$ black holes was emphasised in \cite{Denefa}. Some
270: comments pertaining to the non-supersymmetric case can be found for
271: example in \cite{Ferrara:1997tw}. A similar  analysis using an effective  one dimensional theory,
272: and the Gauss-Bonett term, was carried out in \cite{Sen:2005kj}. Finally, while the thrust of the
273: analysis is different, our results are quite closely related to those
274: in \cite{Sen:2005wa} which appeared while this paper was in
275: preparation (see also \cite{Kraus:2005vz} for the 3-dimensional case).
276: In \cite{Sen:2005wa} the entropy (including higher derivative
277: corrections) is obtained from the gauge field Lagrangian after
278: carrying out a Legendre transformation with respect to the electric
279: parameters. This is similar to our result which is based on
280: $V_{eff}$. As was mentioned above, $V_{eff}$, is proportional to the
281: electro-magnetic energy density i.e., the Hamiltonian density of the
282: electro-magnetic fields, and is derived from the Lagrangian by doing a
283: canonical transformation with respect to the gauge fields.
284: For an action   with only  two-derivative terms, our  results
285: and those  in \cite{Sen:2005wa} agree \cite{IJ}. 
286: 
287: 
288: 
289:   
290: 
291: 
292:  
293: 
294: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
295: \section{Attractor in Four-Dimensional Asymptotically Flat Space}
296: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
297: 
298: 
299: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
300: \subsection{Equations of Motion }
301: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
302: 
303: In this section we consider gravity in four dimensions with $U(1)$
304: gauge fields and scalars. The scalars are coupled to gauge fields with
305: dilaton-like couplings.  It is important for the discussion below that
306: the scalars do not have a potential so that there is a moduli space
307: obtained by varying their values.
308: 
309: 
310: The action we start with has the form,
311: \begin{equation}
312:   S=\frac{1}{\kappa^{2}}\int d^{4}x\sqrt{-G}(R-2(\partial\phi_i)^{2}-
313:   f_{ab}(\phi_i)F^a_{\mu \nu} F^{b \ \mu \nu})
314: \end{equation}
315: Here the index $i$ denotes the different scalars and $a,b$ the
316: different gauge fields and $F^a_{\mu \nu}$ stands for the field
317: strength of the gauge field. $f_{ab}(\phi_i)$ determines the gauge
318: couplings, we can take it to be symmetric in $a,b$ without loss of
319: generality.
320: 
321: The Lagrangian is
322: \begin{equation}
323:   {\mathcal{L}}=(R-2(\partial\phi_i)^{2}-f_{ab}(\phi_i) F^a_{\mu \nu} F^{b \ \mu \nu})
324: \end{equation}
325: Varying the metric gives \footnote{In our notation $G_{\mu\nu}$ refers to the components of the metric.},
326: \begin{equation}
327:   R_{\mu\nu}-2\partial_{\mu}\phi_i\partial_{\nu}\phi_i
328:   =2 f_{ab}(\phi_i) F^a_{\phantom{a}\mu\lambda} F^{b\phantom{\nu}\lambda}_{\phantom{b}\nu}
329:   +\frac{1}{2} G_{\mu\nu}\mathcal{L}
330:   \label{eq:metric1}
331: \end{equation}
332: The trace of the above equation implies
333: \begin{equation}
334:   R-2(\partial\phi_i)^{2}=0\label{trace}
335: \end{equation}
336: The equations of motion corresponding to the metric, dilaton and the
337: gauge fields are then given by,
338: \begin{eqnarray}
339:   R_{\mu\nu}-2\partial_{\mu}\phi_i\partial_{\nu}\phi_i 
340:   & = &   f_{ab}(\phi_i) \left(2F^a_{\phantom{a}\mu\lambda}F^{b\phantom{\nu}\lambda}_{\phantom{b}\nu}-
341:     {\textstyle \frac{1}{2}}G_{\mu\nu}F^a_{\phantom{a}\kappa\lambda} F^{b \kappa\lambda} \right) \label{motion} \\
342:   \frac{1}{\sqrt{-G}}\partial_{\mu}(\sqrt{-G}\partial^{\mu}\phi_i) 
343:   & = & {1 \over 4} \partial_i(f_{ab}) F^a _{\phantom{a}\mu\nu} F^{b \mu\nu} \label{dilatonmotion} \\
344:   \partial_{\mu}(\sqrt{-G}f_{ab}(\phi_i) F^{b \mu\nu}) & = & 0.\nonumber 
345: \end{eqnarray}
346: The Bianchi identity for the gauge field is,
347: \begin{equation}
348: \label{bi}
349: \partial_\mu F_{\nu \rho} + \partial_\nu F_{\rho \mu} + \partial_\rho F_{\mu \nu}=0.
350: \end{equation}
351: We now assume all quantities to be function of $r$.  To begin, let us
352: also consider the case where the gauge fields have only magnetic
353: charge, generalisations to both electrically and magnetically charged
354: cases will be discussed shortly.  The metric and gauge fields can then
355: be written as,
356: \begin{eqnarray}
357:   ds^{2} & = & -a(r)^{2}dt^{2}+a(r)^{-2}dr^{2}+b(r)^{2}d\Omega^{2}\label{metric2}\\
358:   F^a & = & Q_m^{a}sin\theta d\theta\wedge d\phi \label{fstrength}\end{eqnarray}
359: Using the equations of motion we then get, 
360: \begin{eqnarray}
361:   R_{tt} & = & \frac{a^{2}}{b^{4}} V_{eff}(\phi_i) \label{energyd} \\
362:   R_{\theta\theta} & = & \frac{1}{b^{2}} V_{eff}(\phi_i) \label{stressd} 
363: \end{eqnarray}
364: where,
365: \begin{equation}
366:   \label{defpot}
367:   V_{eff}(\phi_i) \equiv f_{ab}(\phi_i) Q_m^a Q_m^b.
368: \end{equation}
369: This function, $V_{eff}$, will play an important role in the
370: subsequent discussion.  We see from eq.(\ref{energyd}) that up to an
371: overall factor it is the energy density in the electromagnetic field.
372: Note that $V_{eff}(\phi_i)$ is actually a function of both the scalars
373: and the charges carried by the black hole.
374: 
375: The relation, $R_{tt}=\frac{a^{2}}{b^{2}}R_{\theta\theta}$, after
376: substituting the metric ansatz implies that,
377: \begin{equation}
378:   \label{eq1}
379:   (a^{2}(r)b^{2}(r))^{''}=2.
380: \end{equation}
381: 
382: The $R_{rr}-{G^{tt}\over G^{rr}} R_{tt}$ component of the Einstein
383: equation gives
384: \begin{equation}
385:   \frac{b^{''}}{b}=-(\partial_{r}\phi)^{2}. \label{eq2}
386: \end{equation}
387: Also the $R_{rr}$ component itself yields a first order ``energy''
388: constraint,
389: \begin{equation}
390:   -1+a^{2}b^{'2}+\frac{a^{2'}b^{2'}}{2}=\frac{-1}{b^{2}}(V_{eff}(\phi_i))+a^{2}b^{2}
391:   (\phi')^{2}\label{constraint}
392: \end{equation}
393: 
394: Finally, the equation of motion for the scalar $\phi_i$ takes the
395: form,
396: \begin{equation}
397:   \label{eomdil}
398:   \partial_{r}(a^{2}b^{2}\partial_{r}\phi_i)=\frac{\partial_iV_{eff} }{2b^{2}}.
399: \end{equation}
400: We see that $V_{eff}(\phi_i)$ plays the role of an ``effective
401: potential '' for the scalar fields.
402: 
403: Let us now comment on the case of both electric and magnetic charges.
404: In this case one should also include ``axion'' type couplings and the
405: action takes the form,
406: \begin{equation}
407:   S=\frac{1}{\kappa^{2}}\int d^{4}x\sqrt{-G}(R-2(\partial\phi_i)^{2}-
408:   f_{ab}(\phi_i)F^a_{\phantom{a}\mu \nu} F^{b \, \mu \nu} 
409:   -{\textstyle{1 \over 2}} {\tilde f}_{ab}(\phi_i) F^a_{\phantom{a}\mu \nu} 
410:   F^b_{\phantom{b}\rho \sigma} \epsilon^{\mu \nu \rho \sigma} ).
411:   \label{actiongen}
412: \end{equation}
413: We note that ${\tilde f}_{ab}(\phi_i)$ is a function independent of
414: $f_{ab}(\phi_i)$, it can also be taken to be symmetric in $a,b$
415: without loss of generality.
416: 
417: The equation of motion for the metric which follows from  this action
418: is unchanged from eq.(\ref{motion}). While the equations of motion for the dilaton and the gauge field now
419:  take the form,
420: \begin{equation}
421: \label{genedg}
422: \frac{1}{\sqrt{-G}}\partial_{\mu}(\sqrt{-G}\partial^{\mu}\phi_i)
423:    =  \textstyle{1 \over 4} \partial_i(f_{ab}) F^a _{\phantom{a}\mu\nu} F^{b\, \mu\nu} +\textstyle{1\over 8} 
424:    \partial_i({\tilde f}_{ab}) F^a_{\phantom{a}\mu\nu} F^b_{\phantom{b}\rho \sigma} \epsilon^{\mu\nu\rho\sigma} \label{gendilatonmotion}
425: \end{equation}
426: \begin{equation}
427:   \partial_{\mu}\left(\sqrt{-G}\left(f_{ab}(\phi_i) F^{b\, \mu\nu} 
428: + \textstyle{1\over 2} {\tilde f}_{ab}F^b_{\phantom{b}\rho\sigma}
429: \epsilon^{\mu\nu\rho\sigma}\right) \right)  =  0.\label{geneomgf}
430: \end{equation}
431:  
432: With both electric and magnetic charges the gauge fields take the
433: form,
434: \begin{equation}
435:   \label{fstrenghtgen}
436:   F^a=f^{ab}(\phi_i)(Q_{eb}-{\tilde f}_{bc}Q^c_m) {1\over b^2} dt\wedge dr + Q_m^a sin \theta  d\theta \wedge d\phi, 
437: \end{equation} 
438: where $Q_m^a, Q_{ea}$ are constants that determine the magnetic and
439: electric charges carried by the gauge field $F^a$, and $f^{ab}$ is the
440: inverse of $f_{ab}$ \footnote{We assume that $f_{ab}$ is invertible.
441:   Since it is symmetric it is always diagonalisable. Zero eigenvalues
442:   correspond to gauge fields with vanishing kinetic energy terms,
443:   these can be omitted from the Lagrangian.}.  
444: It is easy to see that
445: this solves the Bianchi identity eq.(\ref{bi}), and the equation of motion for the
446: gauge fields eq.(\ref{geneomgf}).
447: 
448: A little straightforward algebra shows that the Einstein equations for
449: the metric and the equations of motion for the scalars take the same
450: form as before, eq.(\ref{eq1}, \ref{eq2}, \ref{constraint},
451: \ref{eomdil}), with $V_{eff}$ now being given by,
452: \begin{equation}
453:   \label{defpotgen}
454:   V_{eff}(\phi_i)=f^{ab}(Q_{ea}-{\tilde f}_{ac}Q^c_m)(Q_{eb}- {\tilde f}_{bd}Q^d_m)+f_{ab}Q^a_mQ^b_m.
455: \end{equation}
456: As was already noted in the special case of only magnetic charges,
457: $V_{eff}$ is proportional to the energy density in the electromagnetic
458: field and therefore has an immediate physical significance.  It is
459: invariant under duality transformations which transform the electric
460: and magnetic fields to one-another.
461: 
462: Our discussion below will use (\ref{eq1}, \ref{eq2}, \ref{constraint},
463: \ref{eomdil}) and will apply to the general case of a black hole
464: carrying both electric and magnetic charges.
465: 
466: It is also worth mentioning that the equations of motion,
467: eq.(\ref{eq1}, \ref{eq2}, \ref{eomdil}) above can be derived from a
468: one-dimensional action,
469: \begin{eqnarray}
470:   S=\frac{2}{\kappa^{2}}\int dr\left((a^{2}b)^{'} b^{'}
471:     -a^{2}b^{2}(\phi')^{2}-\frac{ V_{eff}(\phi_i)}{b^{2}}\right).\label{action}
472: \end{eqnarray}
473: The constraint, eq.(\ref{constraint}) must be imposed in addition.
474: 
475: One final comment before we proceed. The eq.(\ref{actiongen}) can be
476: further generalised to include non-trivial kinetic energy terms for
477: the scalars of the form,
478: \begin{equation}
479:   \label{compkinetic}
480:   \int d^4x \sqrt{-G}\left(- g_{ij}(\phi_k) \partial\phi^i \partial\phi^j\right).
481: \end{equation}
482: The resulting equations are easily determined from the discussion
483: above by now contracting the scalar derivative terms with the metric
484: $g_{ij}$.  The two conditions we obtain in the next section for the
485: existence of an attractor are not altered due to these more general
486: kinetic energy terms.
487: 
488: 
489: \subsection{Conditions for an Attractor \label{sec:cond:attr}}
490: We can now state the two conditions which are sufficient for the
491: existence of an attractor.  First, the charges should be such that the
492: resulting effective potential, $V_{eff}$, given by
493: eq.(\ref{defpotgen}), has a critical point.  We denote the critical
494: values for the scalars as $\phi_i=\phi_{i0}$.  So that,
495: \begin{equation}
496:   \label{critical}
497:   \partial_iV_{eff}(\phi_{i0})=0
498: \end{equation}
499: Second, the matrix of second derivatives of the potential at the
500: critical point,
501: \begin{equation}
502:   \label{massmatrix}
503:   M_{ij}={1\over 2} \partial_i\partial_jV_{eff}(\phi_{i0})
504: \end{equation}
505: should have positive eigenvalues. Schematically we write,
506: \begin{equation}
507:   \label{positive}
508:   M_{ij}>0
509: \end{equation}
510: Once these two conditions hold, we show below that the attractor
511: phenomenon results.  The attractor values for the scalars are
512: \footnote{Scalars which do not enter in $V_{eff}$ are not fixed by the
513:   requirement eq.(\ref{critical}). The entropy of the extremal black
514:   hole is also independent of these scalars.}  $\phi_i=\phi_{i0}$.
515: 
516: The resulting horizon radius is given by,
517: \begin{equation}
518:   \label{RH}
519:   b_H^2=V_{eff}(\phi_{i0})
520: \end{equation}
521: and the entropy is
522:  \be
523: \label{BH}
524: S_{BH}={1 \over 4} A = \pi b_H^2. 
525: \ee
526: 
527: 
528: There is one special solution which plays an important role in the
529: discussion below.  From eq.(\ref{eomdil}) we see that one can
530: consistently set $\phi_i=\phi_{i0}$ for all values of $r$. The
531: resulting solution is an extremal Reissner Nordstrom (ERN) Black hole.
532: It has a double-zero horizon. In this solution $\partial_r\phi_{i}=0$,
533: and $a,b$ are
534: \begin{eqnarray}
535:   a_0(r)=\left(1-\frac{r_{H}}{r}\right),\,\,\
536:   b_0(r)=r \label{ern} \end{eqnarray}
537: where $r_{H}$ is the horizon radius.
538: We see that $a_0^2, (a_0^2)'$ vanish at the horizon while $b_0, b_0'$ are finite there. 
539: From eq.(\ref{constraint})  it follows then that the horizon radius $b_H$ is indeed given by
540: \begin{equation}
541:   \label{horizon}
542:   r_H^2=b_H^2= V_{eff}(\phi_{i0}),
543: \end{equation}
544: and the black hole entropy is eq.(\ref{BH}).
545:  
546: If the scalar fields take values at asymptotic infinity which are
547: small deviations from their attractor values we show below that a
548: double-zero horizon black hole solution continues to exist. In this
549: solution the scalars take the attractor values at the horizon, and
550: $a^2, (a^2)'$ vanish while $b, b'$ continue to be finite there.  From
551: eq.(\ref{constraint}) it then follows that for this whole family of
552: solutions the entropy is given by eq.(\ref{BH}) and in particular is
553: independent of the asymptotic values of the scalars.
554: 
555: For simple potentials $V_{eff}$ we find only one critical point. In
556: more complicated cases there can be multiple critical points which are
557: attractors, each of these has a basin of attraction.
558: 
559: One comment is worth making before moving on.  A simple example of a
560: system which exhibits the attractor behaviour consists of one scalar
561: field $\phi$ coupled to two gauge fields with field strengths, $F^a,
562: a= 1, 2$. The scalar couples to the gauge fields with dilaton-like
563: couplings,
564: \begin{equation}
565:   \label{se}
566:   f_{ab}(\phi)=e^{\alpha_a \phi} \delta_{ab}.
567: \end{equation}
568: If only magnetic charges are turned on,
569: \begin{equation}
570:   \label{seffpot}
571:   V_{eff}=e^{\alpha_1 \phi} (Q_1)^2 + e^{\alpha_2 \phi} (Q_2)^2.
572: \end{equation}
573: (We have suppressed the subscript ``$m$'' on the charges).  For a
574: critical point to exist $\alpha_1$ and $\alpha_2$ must have opposite
575: sign.  The resulting critical value of $\phi$ is given by,
576: \begin{equation}
577:   e^{\phi_{0}}=\left(-\frac{\alpha_2 (Q_{2})^{2}}{\alpha_1 (Q_{1})^{2}}\right)^{\frac{1}{\alpha_1-\alpha_2}}
578:   \label{attractorexample}
579: \end{equation}
580: The second derivative, eq.(\ref{massmatrix}) now is given by
581: \begin{equation}
582:   \label{sdexample}
583:   {\partial^2V_{eff}\over \partial \phi^2 }=-2 \alpha^1 \alpha^2
584: \end{equation}
585: and is positive if $\alpha_1,\alpha_2$ have opposite sign.
586: 
587: This example will be useful for studying the behaviour of perturbation
588: theory to higher orders and in the subsequent numerical analysis.
589: 
590: As we will discuss further in section 7, a Lagrangian with dilaton-like couplings of the type in eq.(\ref{se}),
591: and additional axionic terms ( which can be consistently set to zero if only magnetic charges are turned on),
592:  can always  be embedded in a theory with ${\cal N}=1$ supersymmetry.
593: But for generic values of $\alpha$ we do not expect to be able to embed it in an ${\cal N}=2$ theory.
594: The resulting extremal black hole, for generic $\alpha$, will also then not  be a  BPS state. 
595: 
596: 
597: \subsection{Comparison with the ${\cal N}=2$ Case}
598: It is useful to compare the discussion above with the special case of
599: a BPS black hole in an ${\cal N}=2$ theory.  The role of the effective
600: potential, $V_{eff}$ for this case was emphasised by Denef,
601: \cite{Denefa}.  It can be expressed in terms of a superpotential $W$
602: and a Kahler potential $K$ as follows:
603: \begin{equation}
604:   \label{comparison}
605:   V_{eff}=e^K[K^{i \bar j}D_i W (D_j W)^* + |W|^2],
606: \end{equation}
607: where $D_iW\equiv \partial_iW+\partial_iK W$.  The attractor equations
608: take the form,
609: \begin{equation}
610:   \label{attractorsusy}
611:   D_iW=0
612: \end{equation}
613: And the resulting entropy is given by
614: \begin{equation}
615:   \label{susyentropy}
616:   S_{BH}=\pi |W|^2 e^K.
617: \end{equation}
618: with the superpotential evaluated at the attractor values.
619: 
620: It is easy to see that if eq.(\ref{attractorsusy}) is met then the
621: potential is also at a critical point, $\partial_i V_{eff}=0$. A
622: little more work also shows that all eigenvalues of the second
623: derivative matrix, eq.(\ref{massmatrix}) are also positive in this
624: case. Thus the BPS attractor meets the two conditions mentioned above.
625: We also note that from eq.(\ref{comparison}) the value of $V_{eff}$ at
626: the attractor point is $V_{eff}=e^K|W|^2$.  The resulting black hole
627: entropy eq.(\ref{RH}, \ref{BH}) then agrees with
628: eq.(\ref{susyentropy}).
629: 
630: We now turn to a more detailed analysis of the attractor conditions
631: below.
632: 
633: \subsection{Perturbative Analysis}
634: \subsubsection{A Summary}
635: 
636: The essential idea in the perturbative analysis is to start with the
637: extremal RN black hole solution described above, obtained by setting
638: the asymptotic values of the scalars equal to their critical values,
639: and then examine what happens when the scalars take values at
640: asymptotic infinity which are somewhat different from their attractor
641: values, $\phi_{i}=\phi_{i0}$.
642: 
643: We first study the scalar field equations to first order in the
644: perturbation, in the ERN geometry without including backreaction.  Let
645: $\phi_i$ be a eigenmode of the second derivative matrix
646: eq.(\ref{massmatrix}) \footnote{More generally if the kinetic energy
647:   terms are more complicated, eq.(\ref{compkinetic}), these eigenmodes
648:   are obtained as follows. First, one uses the metric at the attractor
649:   point, $g_{ij}(\phi_{i0})$, and calculates the kinetic energy terms.
650:   Then by diagonalising and rescaling one obtains a basis of
651:   canonically normalised scalars.  The second derivatives of $V_{eff}$
652:   are calculated in this basis and gives rise to a symmetric matrix,
653:   eq.(\ref{massmatrix}). This is then diagonalised by an orthogonal
654:   transformation that keeps the kinetic energy terms in canonical
655:   form. The resulting eigenmodes are the ones of relevance here.}.
656: Then denoting, $\delta \phi_i \equiv \phi_i -\phi_{i0}$, neglecting
657: the gravitational backreaction, and working to first order in $\delta
658: \phi_i$, we find that eq.(\ref{eomdil}) takes the form, 
659: \be
660: \label{nhpert}
661: \partial_r\left((r-r_H)^2\partial_r (\delta \phi_i)\right) = {\beta_i^2 \delta \phi_i
662:   \over r^2 }, 
663: \ee 
664: where $\beta_i^2 $ is the relevant eigenvalue of
665: ${1 \over 2} \partial_i\partial_j V(\phi_{i0})$.  In the vicinity of
666: the horizon, we can replace the factor $1/r^2$ on the r.h.s by a
667: constant and as we will see below, eq.(\ref{nhpert}), has one solution
668: that is well behaved and vanishes at the horizon provided $\beta_i^2
669: \geqslant 0$.  Asymptotically, as $r\rightarrow \infty$, the effects of the
670: gauge fields die away and eq.(\ref{nhpert}) reduces to that of a free
671: field in flat space. This has two expected solutions, $\delta \phi_i
672: \sim constant$, and $\delta \phi_i \sim 1/r$, both of which are well
673: behaved.  It is also easy to see that the second order differential
674: equation is regular at all points in between the horizon and infinity.
675: So once we choose the non-singular solution in the vicinity of the
676: horizon it can be continued to infinity without blowing up.
677: 
678: Next, we include the gravitational backreaction. The first order
679: perturbations in the scalars source a second order change in the
680: metric.  The resulting equations for metric perturbations are regular
681: between the horizon and infinity and the analysis near the horizon and
682: at infinity shows that a double-zero horizon black hole solution
683: continues to exist which is asymptotically flat after including the
684: perturbations.
685: 
686: In short the two conditions, eq.(\ref{critical}), eq.(\ref{positive}),
687: are enough to establish the attractor phenomenon to first non-trivial
688: order in perturbation theory.  
689: 
690: In  4-dimensions, for an  effective potential which can be expanded 
691: in a power series about its  minimum, one can in principle solve for the perturbations 
692: analytically to all orders in perturbation theory.  We illustrate this below for the simple case 
693: of dilaton-like couplings, eq.(\ref{se}),
694:  where the coefficients that appear in the perturbation theory can be 
695: determined easily. 
696: One finds that the
697: attractor mechanism works to all orders without conditions other than
698: eq.(\ref{critical}), eq.(\ref{positive}) \footnote{For some specific values of the exponent $\gamma_i$,
699: eq.(\ref{order1}), though, we find that   there can be an obstruction  which prevents
700:  the solution from being extended to 
701: all orders.}.
702: 
703: When we turn
704: to other cases later in the paper, higher dimensional or AdS space
705: etc., we will sometimes not have explicit solutions, but an analysis
706: along the above lines in the near horizon and asymptotic regions and
707: showing regularity in-between will suffice to show that a smoothly
708: interpolating solution exists which connects the asymptotically flat
709: region to the attractor geometry at horizon.
710: 
711: 
712: To conclude, the key feature that leads to the attractor is the fact
713: that both solutions to the linearised equation for $\delta \phi$ are
714: well behaved as $r \rightarrow \infty$, and one solution near the
715: horizon is well behaved and vanishes.  If one of these features fails
716: the attractor mechanism typically does not work.  For example, adding
717: a mass term for the scalars results in one of the two solutions at
718: infinity diverging.  Now it is typically not possible to match the
719: well behaved solution near the horizon to the well behaved one at
720: infinity and this makes it impossible to turn on the dilaton
721: perturbation in a non-singular fashion.
722: % We will come back to this point in the context of the no-hair
723: % theorems in section 7.
724: 
725:  
726:  
727: We turn to a more detailed description of perturbation theory below.
728: 
729: \subsubsection{First Order Solution}
730: 
731: We start with first order perturbation theory.  We can write,
732: \begin{equation}
733:   \label{defepsilon}
734:   \delta \phi_i \equiv  \phi_i-\phi_{i0}=\epsilon\phi_{i1}, 
735: \end{equation}
736: where $\epsilon$ is the small parameter we use to organise the
737: perturbation theory.  The scalars $\phi_i$ are chosen to be
738: eigenvectors of the second derivative matrix, eq.(\ref{massmatrix}).
739: 
740: From, eq.(\ref{eq1}), eq.(\ref{eq2}), eq.(\ref{constraint}), we see
741: that there are no first order corrections to the metric components,
742: $a,b$. These receive a correction starting at second order in
743: $\epsilon$.  The first order correction to the scalars $\phi_i$
744: satisfies the equation,
745: \begin{eqnarray}
746: \label{eodf}
747:   \partial_{r}(a_0^{2}b_0^{2}\partial_{r}\phi_{i1})
748:   =\frac{\kkappa_i^{2}}{b_0^{2}}\phi_{i1}.
749: \end{eqnarray}
750: where, $\kkappa_i^2$ is the eigenvalue for the matrix
751: eq.(\ref{massmatrix}) corresponding to the mode $\phi_i$.
752: % \begin{eqnarray}
753: %  \kkappa^{2}=\textha  \alpha_i^{2}e^{\alpha_i\phi_{0}}Q_{i}^{2}.
754: %  \label{defkappa}
755: %\end{eqnarray}
756: Substituting for $a_0,b_0,$ from eq.(\ref{ern}) we find,
757: \begin{equation}
758:   \phi_{i1}=c_{1i}\left(\frac{r-r_H}{r}\right)^{\frac{1}{2}(\pm \sqrt{1+4\kkappa_i^{2}/r_H^2}-1)}
759: \label{order1}
760: \end{equation}
761: We are interested in a solution which does not blow up at the horizon,
762: $r=r_H$. This gives,
763: \begin{equation}
764:   \phi_{i1}=c_{1i}\left(\frac{r-r_H}{r}\right)^{\gamma_i}\label{soldil1},
765: \end{equation}
766: where
767: \begin{eqnarray}
768:   \label{deft}
769:   \gamma_i&=&\textstyle\frac{1}{2}\left(\sqrt{1+\frac{4 \kkappa_i^{2}}{r_H^2}}-1\right).
770:   % &=& \textstyle\ha\left(\sqrt{1+2\alpha(-\ta)}-1\right).
771: \end{eqnarray}
772: 
773: Asymptotically, as $r \rightarrow \infty$, $\phi_{i1} \rightarrow
774: c_{1i}$, so the value of the scalars vary at infinity as $c_{1i}$ is
775: changed.  However, since $\gamma_{i}>0$, we see from
776: eq.(\ref{soldil1}) that $\phi_{i1}$ vanishes at the horizon and the
777: value of the dilaton is fixed at $\phi_{i0}$ regardless of its value
778: at infinity. This shows that the attractor mechanism works to first
779: order in perturbation theory.
780: 
781: It is worth commenting that the attractor behaviour arises because the   solution
782: to eq.(\ref{eodf}) which is non-singular at $r=r_H$, also vanishes there. 
783: To examine this further we write eq.(\ref{eodf})  in standard form,
784: \cite{MF}, 
785: \begin{equation}
786: \label{eq:standard:ode}
787: {d^2y \over dx^2} + P(x) y + Q(x) y=0,
788: \end{equation}
789: with $x=r-r_H$, $y=\phi_{i1}$.  The vanishing non-singular solution
790: arises because eq.(\ref{eodf}) has a single and double pole
791: respectively for $P(x)$ and $Q(x)$, as $x\rightarrow 0$. This results
792: in (\ref{eq:standard:ode}) having a scaling symmetry as $x\rightarrow
793: 0$ and the solution goes like $x^{\gamma_i}$ near the horizon.  The
794: residues at these poles are such that the resulting indical equation
795: has one solution with exponent $\gamma_i>0$.  In contrast, in a
796: non-extremal black hole background, the horizon is still a regular
797: singular point for the  first order perturbation equation, but
798: $Q(x)$ has only a single pole.  It turns out that the resulting
799: non-singular solution can go to any constant value at the horizon and
800: does not vanish in general.
801: 
802: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
803: \subsubsection{Second Order Solution}
804: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
805: 
806: 
807: The first order perturbation of the dilaton sources a second order
808: correction in the metric.  We turn to calculating this correction
809: next.
810: 
811: Let us write,
812: \begin{eqnarray}
813:   b & = & b_{0}+\epsilon^{2}b_{2}\label{pert2}\\
814:   a^{2} & = & a_{0}^{2}+\epsilon^{2}a_{2}\nonumber \\
815:   b^{2} & = & b_{0}^{2}+2\epsilon^{2}b_{2}b_{0},\nonumber 
816: \end{eqnarray}
817: where $b_{0}$ and $a_{0}$ are the zeroth order extremal Reissner
818: Nordstrom solution eq.(\ref{ern}).
819: 
820: 
821: Equation (\ref{eq1}) gives,
822: \begin{equation}
823:   \label{sa1}
824:   a^2b^2=(r-r_H)^2 + d_1 r + d_2. 
825: \end{equation}
826: The two integration constants, $d_1,d_2$ can be determined by imposing
827: boundary conditions.  We are interested in extremal black hole
828: solutions with vanishing surface gravity. These should have a horizon
829: where $b$ is finite and $a^2$ has a ``double-zero'', i.e., both $a^2$
830: and its derivative $(a^2)'$ vanish. By a gauge choice we can always
831: take the horizon to be at $r=r_H$.  Both $d_1$ and $d_2$ then vanish.
832: Substituting eq.(\ref{pert2}) in the equation(\ref{eq1}) we get to
833: second order in $\epsilon$,
834: \begin{equation}
835:   2a_{0}^{2}b_{0}b_{2}+b_{0}^{2}a_{2}=0.
836: \end{equation}
837: Substituting for $a_0,b_0$ then determines, $a_2$ in terms of $b_2$,
838: \begin{equation}
839:   a_{2}=-2\left(1-\frac{r_{H}}{r}\right)^{2}\frac{b_{2}}{r}. \label{a2}
840: \end{equation}
841: From eq.(\ref{eq2}) we find next that,
842: \begin{equation}
843:   b_{2}(r)=-\sum_i \frac{c_{1i}^{2}\gamma}{2(2\gamma_i-1)}
844:   r \left(\frac{r-r_H}{r}\right)^{2\gamma_i} +A_{1}r +A_2r_H
845:   \label{b2}
846: \end{equation}
847: $A_1,A_2$ are two integration constants. The two terms proportional to
848: these integration constant solve the equations of motion for $b_2$ in
849: the absence of the $O(\epsilon)^2$ source terms from the dilaton.
850: This shows that the freedom associated with varying these constants is
851: a gauge degree of freedom.  We will set $A_1=A_2=0$ below.  Then,
852: $b_2$ is,
853: \begin{equation}
854:   b_{2}(r)=-\sum_i\frac{c_{1i}^{2}\gamma_i}{2(2\gamma_i-1)}
855:   r \left(\frac{r-r_H}{r}\right)^{2\gamma_i} 
856:   \label{finalb2}
857: \end{equation}
858: It is easy to check that this solves the constraint
859: eq.(\ref{constraint}) as well.
860: 
861: To summarise, the metric components to second order in $\epsilon$ are
862: given by eq.(\ref{pert2}) with $a_0,b_0$ being the extremal Reissner
863: Nordstrom solution and the second order corrections being given in
864: eq.(\ref{a2}) and eq.(\ref{finalb2}). Asymptotically, as $r
865: \rightarrow \infty$, $b_2 \rightarrow c \times r$, and,
866: $a_2\rightarrow -2 \times c$, so the solution continues to be
867: asymptotically flat to this order.  Since $\gamma_i>0$ we see from
868: eq.(\ref{a2}, \ref{finalb2}) that the second order corrections are
869: well defined at the horizon.  In fact since $b_2$ goes to zero at the
870: horizon, $a_2$ vanishes at the horizon even faster than a double-zero.
871: Thus the second order solution continues to be a double-zero horizon
872: black hole with vanishing surface gravity.  Since $b_2$ vanishes the
873: horizon area does not change to second order in perturbation theory
874: and is therefore independent of the asymptotic value of the dilaton.
875: % Note that the vanishing of $b_2$ at the horizon is in agreement with
876: % the general argument given after eq.(\ref{constraint}) above.
877: 
878:  
879: 
880: The scalars also gets a correction to second order in $\epsilon$.
881: This can be calculated in a way similar to the above analysis.  We
882: will discuss this correction along with higher order corrections, in
883: one simple example, in the next subsection.
884: 
885: Before proceeding let us calculate the mass of the black hole to
886: second order in $\epsilon$.  It is convenient to define a new
887: coordinate,
888: \begin{equation}
889:   \label{by}
890:   y\equiv b(r)
891: \end{equation}
892: Expressing $a^2$ in terms of $y$ one can read off the mass from the
893: coefficient of the $1/y$ term as $y \rightarrow \infty$, as is
894: discussed in more detail in Appendix \ref{sec:appa}.  This gives,
895: \begin{equation}
896:   \label{mass2}
897:   M=r_H+\epsilon^{2}\sum_i\frac{r_Hc_{i1}^2\gamma_i}{2}
898: \end{equation}
899: where $r_H$ is the horizon radius given by (\ref{horizon}).  Since
900: $\gamma_i$ is positive, eq.(\ref{deft}), we see that as $\epsilon$
901: increases, with fixed charge, the mass of the black hole increases.
902: The minimum mass black hole is the extremal RN black hole solution,
903: eq.(\ref{ern}), obtained by setting the asymptotic values of the
904: scalars equal to their critical values.
905: 
906: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
907: \subsubsection{An Ansatz to All Orders}
908: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
909: 
910: Going to higher orders in perturbation theory is in principle
911: straightforward.  For concreteness we discuss the simple example,
912: eq.(\ref{se}), below.  We show in this example that the form of the
913: metric and dilaton can be obtained to all orders in perturbation
914: theory analytically.  We have not analysed the coefficients and
915: resulting convergence of the perturbation theory in great detail.  In
916: a subsequent section we will numerically analyse this example and find
917: that even the leading order in perturbation theory approximates the
918: exact answer quite well for a wide range of charges.  This discussion
919: can be generalised to other more complicated cases in a
920: straightforward way, although we will not do so here.
921: 
922: 
923: Let us begin by noting that eq.(\ref{eq1}) can be solved in general to
924: give,
925: \begin{equation}
926:   a^2b^2=(r-r_H)^2 + d_1r +d_2
927:   \label{gena}
928: \end{equation}
929: As in the discussion after eq(\ref{sa1}) we set $d_1=d_2=0,$ since we
930: are interested in extremal black holes.  This gives,
931: \begin{equation}
932:   a^2b^2=(r-r_H)^2, 
933:   \label{finala}
934: \end{equation}
935: where $r_H$ is the horizon radius given by eq.(\ref{horizon}).  This
936: can be used to determine $a$ in terms of $b$.
937: 
938: Next we expand $b$, $\phi$ and $a^2$ in a power series in $\epsilon$,
939: \begin{eqnarray}
940:   b & = & b_0+\sum_{n=1}^{\infty}\epsilon^{n}b_{n} \label{pertb}\\
941:   \phi & = & \phi_0 + \sum_{n=1}^\infty \epsilon^{n}\phi_{n} \label{pertphi} \\
942:   a^2 &    = & a_0^2 + \sum_{n=1}^\infty \epsilon^n a_n \label{perta}
943: \end{eqnarray}
944: where $b_0$, $a_0$ are given by eq.(\ref{ern}) and $\phi_0$ is given
945: by eq.(\ref{attractorexample}).
946: 
947: The ansatz which works to all orders is that the $n^{th}$ order terms
948: in the above two equations take the form,
949: \begin{equation}
950:   \phi_{n}(r)=c_{n}\left(\frac{r-r_{H}}{r}\right)^{n\gamma}\label{diln}
951: \end{equation}
952: \begin{equation}
953:   b_{n}(r)=d_{n}r\left(\frac{r-r_{H}}{r}\right)^{n\gamma}\label{bn}, \end{equation}
954: and, 
955: \begin{equation}
956:   a_n=e_n \left({r-r_H \over r}\right)^{n\gamma+2}, \label{an}
957: \end{equation}
958: where $\gamma$ is given by eqs.(\ref{deft}) and in this case takes the
959: value,
960: \begin{equation}
961:   \label{valgammase}
962:   \gamma={1\over 2}\left(\sqrt{1-2\alpha_1\alpha_2} -1\right). 
963: \end{equation} 
964: The discussion in the previous two subsections is in agreement with
965: this ansatz.  We found $b_1=0$, and from eq.(\ref{finalb2}) we see
966: that $b_2$ is of the form eq.(\ref{bn}). Also, we found $a_1=0$ and
967: from eq.(\ref{a2}) $a_2$ is of the form eq.(\ref{an}). And from
968: eq.(\ref{soldil1}) we see that $\phi_1$ is of form eq.(\ref{diln}).
969: We will now verify that this ansatz consistently solves the equations
970: of motion to all orders in $\epsilon$.  The important point is that
971: with the ansatz eq.(\ref{diln}, \ref{bn}) each term in the equations
972: of motion of order $\epsilon^n$ has a functional dependence $({(r-r_H)
973:   \over r})^{2\gamma n}$.  This allows the equations to be solved
974: consistently and the coefficients $c_n, d_n$ to be determined.
975: 
976: Let us illustrate this by calculating $c_2$.  From eq.(\ref{eq2}) and
977: eq.(\ref{finala}) we see that the equation of motion for $\phi$ can be
978: written in the form,
979: \begin{eqnarray}
980:   2b(r)^{2}\partial_{r}((r-r_{H})^{2}\partial_{r}\phi) & = & e^{\alpha_{i}\phi}Q_{i}^{2}\alpha_{i}\label{eqphi}
981: \end{eqnarray}
982: To $O(\epsilon^2)$ this gives,
983: \begin{equation}
984:   \label{int2}
985:   \left(\frac{r-r_H}{r}\right)^{2\gamma}
986:   \left(
987:     2c_{2}(e^{\alpha_{i}\phi_{0}}Q_{i}^{2}\alpha_{i}^2-4r_H^{2}\gamma(1+2\gamma))
988:     +e^{\alpha_{i}\phi_{0}}Q_{i}^{2}\alpha_{i}^3c_{1}^{2}
989:   \right)=0 
990: \end{equation}
991: Notice that the term $(\frac{r-r_H}{r})^{2\gamma}$ has factored out.
992: Solving eq.(\ref{int2}) for $c_2$ we now get,
993: \begin{eqnarray}
994:   c_{2}
995:   &=& \ha c_1^2 (\alpha_1+\alpha_2)\frac{(\gamma+1)}{(3\gamma+1)}
996:   \label{e2}
997: \end{eqnarray}
998: 
999: More generally, as discussed in Appendix \ref{sec:appa}, working to
1000: the required order in $\epsilon$ we can recursively find, $c_n,d_n,
1001: e_n$.
1002: 
1003: 
1004: One more comment is worth making here. We see from eq.(\ref{finalb2})
1005: that $b_2$ blows up when when $\gamma=1/2$. Similarly we can see from
1006: eq.(\ref{dk_ans}) that $b_n$ blows up when $\gamma={1\over n}$ for
1007: $b_n$.  So for the values, $\gamma={1\over n},$ where $n$ is an
1008: integer, our perturbative solution does not work.
1009: 
1010: Let us summarise. We see in the simple example studied here that a
1011: solution to all orders in perturbation theory can be found.  $b$,
1012: $\phi$ and $a^2$ are given by eq.(\ref{bn}), eq.(\ref{diln}) and
1013: eq.(\ref{an}) with coefficients that can be determined as discussed in
1014: Appendix \ref{sec:appa}.  In the solution, $a^2$ vanishes at $r_H$ so
1015: it is the horizon of the black hole.  Moreover $a^2$ has a double-zero
1016: at $r_H$, so the solution is an extremal black hole with vanishing
1017: surface gravity.  One can also see that $b_n$ goes linearly with $r$
1018: as $r \rightarrow \infty$ so the solution is asymptotically flat to
1019: all orders. It is also easy to see that the solution is non-singular
1020: for $r\geqslant r_H$.  Finally, from eq.(\ref{diln}) we see that $\phi_n=0$,
1021: for all $n>0$, so all corrections to the dilaton vanish at the
1022: horizon.  Thus the attractor mechanism works to all orders in
1023: perturbation theory.  Since all corrections to $b$ also vanish at the
1024: horizon we see that the entropy is uncorrected in perturbation theory.
1025: This is in agreement with the general argument given after
1026: eq.(\ref{BH}). Note that no additional conditions had to be imposed,
1027: beyond eq.(\ref{critical}, \ref{positive}), which were already
1028: appeared in the lower order discussion, to ensure the attractor
1029: behaviour \footnote{In our discussion of exact solutions in section 4
1030:   we will be interested in the case, $\alpha_1=-\alpha_2$.  From
1031:   eq.(\ref{e2}, \ref{dk_ans}) we see that the expressions for $c_{2}$
1032:   and $d_{3}$ become,
1033:   \begin{eqnarray}
1034:     c_{2} & = & 0\\
1035:     d_{3} & = & 0\end{eqnarray}
1036:   It follows that in the perturbation series for $\phi$ and $b$ only the 
1037:   $c_{2n+1}$(odd) terms and  $d_{2n}$(even) terms are non-vanishing respectively.}. 
1038: 
1039: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1040: \section{Numerical Results}\label{sec:num}
1041: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1042: 
1043: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1044: \begin{figure}
1045:   \begin{center}
1046:     \epsfig{file=numerical_a_1.7.eps,height=0.8\textheight}
1047:   \end{center}
1048:   \caption{\label{cap:num1}Comparison of numerical integration of
1049:     $\phi$ with $1^{\mathrm{st}}$ order perturbation result. The upper
1050:     graph is a close up of the lower one near the horizon. The
1051:     perturbation result is denoted by a dashed line. We chose
1052:     $\alpha_1,-\alpha_2=1.7$, $Q_1=3$, $Q_2=3$, $\delta r = 2.3\times
1053:     10^{-8}$ and $c_1$ in the range $[-\frac{1}{2},\frac{1}{2}]$ }.
1054:   $\phi_0 = 0 $.
1055: \end{figure}
1056: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1057: \begin{figure}
1058:   \begin{center}
1059:     \epsfig{file=numerical_a_3.1.eps,height=0.8\textheight}
1060:   \end{center}
1061:   \caption{\label{cap:num4}Comparison of numerical integration of
1062:     $\phi$ with $1^{\mathrm{st}}$ order perturbation result. The upper
1063:     graph is a close up of the lower one near the horizon. The
1064:     perturbation result is denoted by a dashed line. We chose
1065:     $\alpha_1,-\alpha_2=3.1$, $Q_1=2$, $Q_2=3$, $\delta r = 2.9\times
1066:     10^{-8}$ and $c_1$ is in the range $[-\frac{1}{2},\frac{1}{2}]$ }.
1067:   $\phi_0 = 0.13 $.
1068: \end{figure}
1069: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1070: 
1071: There are two purposes behind the numerical work we describe in this
1072: section.  First, to check how well perturbation theory works. Second,
1073: to see if the attractor behaviour persists, even when $\epsilon$,
1074: eq.(\ref{defepsilon}), is order unity or bigger so that the deviations
1075: at asymptotic infinity from the attractor values are big.  We will
1076: confine ourselves here to the simple example introduced near
1077: eq.(\ref{se}), which was also discussed in the higher orders analysis
1078: in the previous subsection.
1079: 
1080: In the numerical analysis it is important to impose the boundary
1081: conditions carefully.  As was discussed above, the scalar has an
1082: unstable mode near the horizon.  Generic boundary conditions imposed
1083: at $r\rightarrow \infty$ will therefore not be numerically stable and
1084: will lead to a divergence.  To avoid this problem we start the
1085: numerical integration from a point $r_i$ near the horizon.  We see
1086: from eq.(\ref{diln}, \ref{bn}) that sufficiently close to the horizon
1087: the leading order perturbative corrections \footnote{We take the
1088:   $O(\epsilon)$ correction in the dilaton, eq.(\ref{soldil1}), and the
1089:   $O(\epsilon^2)$ correction in $b, a^2$, eq.(\ref{a2}, \ref{b2}).
1090:   This consistently meets the constraint eq.(\ref{constraint}) to
1091:   $O(\epsilon^2)$.}  becomes a good approximation.  We use these
1092: leading order corrections to impose the boundary conditions near the
1093: horizon and then numerically integrate the exact equations,
1094: eq.(\ref{eq1},\ref{eq2}), to obtain the solution for larger values of
1095: the radial coordinate.
1096: 
1097: The numerical integration is done using the Runge-Kutta method.  We
1098: characterise the nearness to the horizon by the parameter
1099: \begin{equation}
1100:   \label{eq:def:deltar}
1101:   \delta r = \frac{r_i-r_H}{r_i}
1102: \end{equation}
1103: where $r_i$ is the point at which we start the integration.  $c_1$
1104: refers to the asymptotic value for the scalar, eq.(\ref{soldil1}).
1105: 
1106: In figs. (\ref{cap:num1},\ref{cap:num4}) we compare the numerical and
1107: $1^\mathrm{st}$ order correction. The numerical and perturbation
1108: results are denoted by solid and dashed lines respectively.  We see
1109: good agreement even for large $r$.  As expected, as we increase the
1110: asymptotic value of $\phi$, which was the small parameter in our
1111: perturbation series, the agreement decreases.
1112: 
1113: Note also that the resulting solutions turn out to be singularity free
1114: and asymptotically flat for a wide range of initial conditions.  In
1115: this simple example there is only one critical point,
1116: eq.(\ref{attractorexample}).  This however does not guarantee that the
1117: attractor mechanism works. It could have been for example that as the
1118: asymptotic value of the scalar becomes significantly different from
1119: the attractor value no double-zero horizon black hole is allowed and
1120: instead one obtains a singularity.  We have found no evidence for
1121: this. Instead, at least for the range of asymptotic values for the
1122: scalars we scanned in the numerical work, we find that the attractor
1123: mechanism works with attractor value, eq.(\ref{attractorexample}).
1124: 
1125: It will be interesting to analyse this more completely, extending this
1126: work to cases where the effective potential is more complicated and
1127: several critical points are allowed. This should lead to multiple
1128: basins of attraction as has already been discussed in the
1129: supersymmetric context in e.g., \cite{Denefa, Denef:2001xn}.
1130: 
1131: 
1132: 
1133: 
1134: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1135: \section{Exact Solutions }\label{sec:exact_sol}
1136: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1137: %%%%%%%%%%%%%%%%
1138: % new short cuts
1139: %%%%%%%%%%%%%%%%
1140: \newcommand{\deltaa}{\alpha_1-\alpha_2}
1141: \newcommand{\bovera}{\frac{\alpha_2}{\alpha_1}}
1142: \newcommand{\fooa}{\frac{\alpha_1}{\deltaa}}
1143: \newcommand{\foob}{\frac{-\alpha_2}{\deltaa}}
1144: 
1145: In certain cases the equation of motion can be solved exactly
1146: \cite{Gibbons:1987ps}.  In this section, we shall look at some
1147: solvable cases and confirm that the extremal solutions display
1148: attractor behaviour. In particular, we shall work in $4$ dimensions
1149: with one scalar and two gauge fields, taking $V_{eff}$ to be given by
1150: eq.(\ref{seffpot}),
1151: \begin{equation}
1152:   \label{seffpot2}
1153:   V_{eff}=e^{\alpha_1 \phi} (Q_1)^2 + e^{\alpha_2 \phi} (Q_2)^2.
1154: \end{equation} 
1155: We find that at the horizon the scalar field relaxes to the attractor
1156: value (\ref{attractorexample})
1157: \begin{equation}
1158:   e^{(\alpha_1-\alpha_2)\phi_{0}}
1159:   =-\frac{\alpha_2 Q_{2}^{2}}{\alpha_1 Q_{1}^{2}}\label{eq:attractor_again}
1160: \end{equation}
1161: which is the critical point of $V_{eff}$ and independent of the
1162: asymptotic value, $\phi_\infty$.  Furthermore, the horizon area is
1163: also independent of $\phi_\infty $ and, as predicted in section
1164: \ref{sec:cond:attr}, it is proportional to the effective potential
1165: evaluated at the attractor point.  It is given by
1166: \begin{eqnarray}
1167:   \mathrm{Area} &=& 4\pi b_H^2 = 4\pi V_{eff}(\phi_0) \\
1168:   &=& 4\pi \spadesuit (Q_1)^{2\foob}  (Q_2)^{2\fooa}\label{eq:horizon_area}
1169: \end{eqnarray} 
1170: where
1171: \begin{equation}
1172:   \spadesuit = \textstyle 
1173:   \left(-\bovera\right)^\fooa +
1174:   \left(-\bovera\right)^\foob
1175:   \label{eq:def_spade}
1176: \end{equation}
1177: is a numerical factor. It is worth noting that when
1178: $\alpha_1=-\alpha_2$, one just has
1179: \begin{equation}
1180:   \label{eq:simple_area}
1181:   \textstyle{1\over4}\mathrm{Area}= 2\pi  |Q_1 Q_2|
1182: \end{equation}
1183: 
1184: Interestingly, the solvable cases we know correspond to $\gamma=1,2,3$
1185: where $\gamma$ is given by (\ref{deft}). The known solutions for
1186: $\gamma=1,2$ are discussed in \cite{Gibbons:1987ps} and references
1187: therein (although they fixed $\phi_\infty=0$). We found a solution for
1188: $\gamma=3$ and it appears as though one can find exact solutions as
1189: long as $\gamma$ is a positive integer. Details of how these solutions
1190: are obtained can be found in the references and appendix
1191: \ref{sec:details}. % See also appendix \ref{otherref}.
1192:  
1193: For the cases we consider, the extremal solutions can be written in
1194: the following form
1195: \begin{eqnarray}
1196:   e^{(\alpha_1-\alpha_2)\phi} & = & \left(-\frac{\alpha_2}{\alpha_1}\right)
1197:   \left(\frac{\PP}{\Q}\right)^2
1198:   \left(\frac{f_2}{f_1}\right)^{-\ha\alpha_1\alpha_2}\label{eq:exact_scalar_sol}\\
1199:   b^{2} & = &
1200:   \spadesuit \left((Q_1 f_1)^{-\alpha_2}  (Q_2 f_2)^{\alpha_1}\right)^{{\frac{2}{\deltaa}}}
1201:   \label{eq:exact_b_sol}\\
1202:   a^{2} &=& \rho^2/b^2
1203: \end{eqnarray}
1204: where $\rho=r-r_H$ and the $f_i$ are polynomials in $\rho$ to some
1205: fractional power. In general the $f_i$ depend on $\phi_\infty$ but
1206: they have the property
1207: \begin{equation}
1208:   \label{eq:poly_prop}
1209:   f_i|_\mathrm{Horizon}=1.
1210: \end{equation}
1211: Substituting (\ref{eq:poly_prop}) into
1212: (\ref{eq:exact_scalar_sol},\ref{eq:exact_b_sol}), one sees that that
1213: at the horizon the scalar field takes on the attractor value
1214: (\ref{eq:attractor_again}) and the horizon area is given by
1215: (\ref{eq:horizon_area}).
1216: 
1217: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1218: \begin{figure}
1219:   \begin{center}
1220:     \epsfig{file=attr1.eps,width=\linewidth}
1221:   \end{center}
1222:   \caption{\label{cap:nice_graphs}Attractor behaviour for the case
1223:     $\gamma=1$; $\alpha_1,-\alpha_2=2$}
1224: \end{figure}
1225: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1226: \begin{figure}
1227:   \begin{center}
1228:     \epsfig{file=attr2.eps,width=\linewidth}
1229:   \end{center}
1230:   \caption{\label{cap:nice_graphs2}Attractor behaviour for the case
1231:     $\gamma=2$; $\alpha_1,-\alpha_2=2\sqrt{3}$}
1232: \end{figure}
1233: %%%%%%%%%%%%%%%%%%%%%%%%%%%%
1234: 
1235: Notice that, when $\alpha=|\alpha_i|$,
1236: (\ref{eq:exact_scalar_sol},\ref{eq:exact_b_sol}) simplify to
1237: \begin{eqnarray}
1238:   e^{\alpha\phi} & = & \frac{|\PP|}{|\Q|}\left(\frac{f_2} {f_1}\right)^{\frac{1}{4}\alpha^2}\\
1239:   b^{2} & = & 2 |Q_1| |Q_2| \left( f_1  f_2\right)
1240: \end{eqnarray}
1241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1242: % Discuss solutions
1243: 
1244: \subsection{Explicit Form of the $f_i$}
1245: \label{sec:explicit}
1246: 
1247: In this section we present the form of the functions $f_i$ mainly to
1248: show that, although they depend on $\phi_\infty$ in a non trivial way,
1249: they all satisfy (\ref{eq:poly_prop}) which ensures that the attractor
1250: mechanism works. It is convenient to define
1251: \begin{equation}
1252:   \bar{Q}_i^2 = e^{\alpha_i\phi_\infty}Q_i^2\nosum
1253:   \label{eq:def_barQ}
1254: \end{equation}
1255: which are the effective $U(1)$ charges as seen by an asymptotic
1256: observer. For the simplest case, $\gamma=1$, we have
1257: \begin{equation}
1258:   f_{i}=1+\textstyle\left(\bar{Q}_i^{-1}|\alpha_i|{(4+\alpha_i^2)}^{-\ha}\right)\rho
1259: \end{equation}
1260: Taking $\gamma=2$ and $\alpha_1=-\alpha_2=2\sqrt{3}$ one finds
1261: \begin{equation}
1262:   f_i  = 
1263:   \left( 
1264:     1+
1265:     (\bar{Q}_1 \bar{Q}_2)^{-\frac{2}{3}} (\bar{Q}_1^{\frac{2}{3}} + \bar{Q_2}^{\frac{2}{3}})^\ha \rho
1266:     +{\textstyle\ha} (\bar{Q}_i\bar{Q}_1 \bar{Q}_2)^{-\frac{2}{3}}\rho^{2}
1267:   \right)^\ha
1268: \end{equation}
1269: Finally for $\gamma=3$ and $\alpha_1=4$ $, \alpha_2=-6$ we have
1270: \begin{equation}
1271:   f_{1} =  
1272:   \left(
1273:     1-6a_{2}\rho + 12a_{2}^{2}\rho^{2}-6a_{0}\rho^{3}
1274:   \right)^{\frac{1}{3}}
1275: \end{equation}
1276: \begin{equation}
1277:   f_{2}  =  
1278:   \left(
1279:     1-
1280:     {\textstyle \frac{24}{3}}a_{2}\rho
1281:     +24a_{2}\rho^{2}
1282:     -(48a_{2}^{3}-12a_{0})\rho^{3}
1283:     +\left(48a_{2}^{4}-24a_{0}a_{2}\right)\rho^{4}
1284:   \right)^{\frac{1}{4}}
1285: \end{equation}
1286: where $a_0$ and $a_2$ are non-trivial functions of $\bar{Q}_i$.  Further details are
1287: discussed in section \ref{sec:nonex} and appendix \ref{sec:details}. The scalar field
1288: solutions for $\gamma=1$ and $2$ are illustrated in figs. \ref{cap:nice_graphs} and
1289: \ref{cap:nice_graphs2} respectively.
1290: 
1291: \subsection{Supersymmetry and the Exact Solutions}
1292: 
1293: As mentioned above, the first two cases ($\gamma=1,2$) have been
1294: extensively studied in the literature.
1295: 
1296: The SUSY of the extremal $\alpha_1=-\alpha_2=2$ solution is discussed
1297: in \cite{Kallosh:1992ii}. They show that it is supersymmetric in the
1298: context of ${\cal N}=4$ SUGRA. It saturates the BPS bound and
1299: preserves $1\over4$ of the supersymmetry - ie. it has ${\cal N}=1$
1300: SUSY. There are BPS black-holes in this context which carry only one
1301: $U(1)$ charge and preserve $\ha$ of the supersymmetry. The
1302: non-extremal blackholes are of course non-BPS.
1303: % (while purely electric/magnetic solutions with only one gauge field,
1304: % which we have not discussed, only break half the SUSY).
1305: 
1306: On the other hand, the extremal $\alpha_1=-\alpha_2=2\sqrt3$ blackhole
1307: is non-BPS \cite{Gibbons:1994ff}. It arises in the context of
1308: dimensionally reduced $5D$ Kaluza-Klein gravity \cite{Dobiasch:1981vh}
1309: and is embeddable in ${\cal N}=2$ SUGRA. There however are BPS
1310: black-holes in this context which carry only one $U(1)$ charge and
1311: once again preserve $\ha$ of the supersymmetry \cite{Gibbons:1993xt}.
1312: 
1313: We have not investigated the supersymmetry of the $\gamma=3$ solution,
1314: we expect that it is not a BPS solution in a supersymmetric theory.
1315:   
1316: %
1317: % \begin{itemize}
1318: %
1319: % \item $\alpha=2$
1320: %
1321: % 
1322: %   \begin{itemize}
1323: %   \item Extremal solution:\begin{eqnarray*}
1324: %       \exp(2\phi) & = & e^{2\phi_{\infty}}\frac{A}{B}\\
1325: %       a^{2} & = & \frac{(r-M)^{2}}{AB}\\
1326: %       b^{2} & = & AB\end{eqnarray*}
1327: 
1328: %   \end{itemize}
1329: %   \begin{eqnarray*}
1330: %     A & = & (r-M)+\sqrt{2}Pe^{-\phi_{\infty}}\\
1331: %     B & = & (r-M)+\sqrt{2}Qe^{\phi_{\infty}}\end{eqnarray*}
1332: %   \[
1333: %   M=\frac{|\bar{P}|+|\bar{Q}|}{\sqrt{2}}\qquad\Sigma=\frac{|\bar{P}|-|\bar{Q}|}{\sqrt{2}}\]
1334: %
1335: %
1336: %   \begin{itemize}
1337: %   \item see \cite{Kallosh:1992ii} page 23 onwards for a discussion
1338: %     of the SUSY of this solution
1339: %
1340: %     \begin{itemize}
1341: %     \item solution is supersymmetric in the context of $N=4$ SUGRA
1342: %     \item - Central charges :
1343: %       $z_{1}=\frac{1}{\sqrt{2}}\left(Q-P\right)$
1344: %       $z_{2}=\frac{1}{\sqrt{2}}\left(Q+P\right)$
1345: %     \item non-extreme black holes break all the supersymmetry
1346: %     \item pure electric/magnetic solutions breaks 1/2 the SUSY
1347: %       $\rightarrow$ $N=2$
1348: %     \item the above dyonic solution:
1349: %     \item - saturates the BPS bound
1350: %     \item - preserves 1/4 of SUSY : has $N=1$SUSY
1351: %     \end{itemize}
1352: %   \end{itemize}
1353: % \item $\alpha=2\sqrt{3}$
1354: %
1355: %   \begin{itemize}
1356: %   \item see \cite{Gibbons:1994ff}
1357: %   \item obtainable by dimensional reduction of 5d Einstein gravity
1358: %   \item the scalar field corresponds to $g_{55}$
1359: %   \item Extremal solution:\begin{eqnarray*}
1360: %       \exp(4\phi/\sqrt{3}) & =e^{4\phi_{\infty}/\sqrt{3}} & \frac{A}{B}\\
1361: %       a^{2} & = & \frac{(r-r_{+})(r-r_{-})}{\sqrt{AB}}\\
1362: %       b^{2} & = & \sqrt{AB}\end{eqnarray*}
1363: %     \begin{eqnarray*}
1364: %       A & = &
1365: %       (r-M)^{2}+2^{\frac{4}{3}}\bar{P}^{\frac{2}{3}}M^{\frac{1}{3}}(r-M)
1366: %       +2\bar{P}^{\frac{4}{3}}\bar{Q}\\
1367: %       B & = & 
1368: %       (r-M)^{2}+2^{\frac{4}{3}}\bar{Q}^{\frac{2}{3}}M^{\frac{1}{3}}(r-M)
1369: %       +2\bar{Q}^{\frac{4}{3}}\bar{P}^{\frac{2}{3}}
1370: %     \end{eqnarray*}
1371: %     \[
1372: %     M=\frac{1}{2}\left(\bar{P}^{2/3}+\bar{Q}^{2/3}\right)^{3/2}
1373: %     \]
1374: %
1375: %   \item this solution exceeds the BPS condition (hep-th 9407118 and
1376: %     hep-th/9310118)
1377: %
1378: %     \begin{itemize}
1379: %     \item solutions with unbroken symmetry must saturate the bound
1380: %       \[
1381: %       M\geq\frac{1}{\sqrt{1+\alpha^{2}/4}}\sqrt{\bar{Q}^{2}+\bar{P}^{2}}
1382: %       =\frac{1}{2}\sqrt{\bar{Q}^{2}+\bar{P}^{2}}\]
1383: %
1384: %     \item the above solution exceeds this bound\end{itemize}
1385: %   \end{itemize}
1386: % \end{itemize}
1387: 
1388: % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1389: 
1390: % \begin{itemize}
1391: % \item SUSY of solution
1392: % \item BPS conditions
1393: 
1394: % \end{itemize}
1395: % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1396: % \subsection{$\gamma=2$, $\alpha=-\ta=2\sqrt{3}$}
1397: % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1398: 
1399: 
1400: % \begin{itemize}
1401: % \item SUSY of L
1402: % \item non-BPS nature
1403: 
1404: % \end{itemize}
1405: % \subsection{$\gamma=3$, $\alpha=4 $ $\ta=-6$}
1406: 
1407: 
1408: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1409: \section{General Higher Dimensional Analysis}
1410: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1411: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1412: \subsection{The Set-Up}
1413: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1414: 
1415: 
1416: 
1417: 
1418: It is straightforward to generalise our results above to higher
1419: dimensions.  We start with an action of the form,
1420: \begin{equation}
1421:   \label{higherdaction}
1422:   S=\frac{1}{\kappa^{2}}\int d^{d}x\sqrt{-G}(R-2(\partial\phi_i)^{2}-
1423:   f_{ab}(\phi_i)F^a F^b)
1424: \end{equation}
1425: %%%%%%%%%%%%%%%%%%%%%%%%
1426: % less general ansatz
1427: %%%%%%%%%%%%%%%%%%%%%%%
1428: % \begin{equation}
1429: %S=\frac{1}{\kappa^{2}}\int %d^{d}x\sqrt{-G}(R-2(\partial\phi)^{2}-e^{\alpha\phi}F^{2}-e^{\tilde{\alpha}%\phi}\tilde{F}^{2}-V(\phi))
1430: %\end{equation}
1431: Here the field strengths, $F_a$ are $(d-2)$ forms which are magnetic
1432: dual to $2$-form fields.
1433: 
1434: We will be interested in solution which preserve a $SO(d-2)$ rotation
1435: symmetry. Assuming all quantities to be function of $r$, and taking
1436: the charges to be purely magnetic, the ansatz for the metric and gauge
1437: fields is \footnote{Black hole which carry both electric and magnetic
1438:   charges do not have an $SO(d-2)$ symmetry for general $d$ and we
1439:   only consider the magnetically charged case here. The analogue of
1440:   the two-form in 4 dimensions is the $d/2$ form in $d$ dimensions. In
1441:   this case one can turn on both electric and magnetic charges
1442:   consistent with $SO(d/2)$ symmetry. We leave a discussion of this
1443:   case and the more general case of $p$-forms in $d$ dimensions for
1444:   the future.}
1445: \begin{eqnarray}
1446:   \label{metrichd}
1447:   ds^{2}=-a(r)^{2}dt^{2}+a(r)^{-2}dr^{2}+b(r)^{2}d\Omega_{d-2}^{2}\\
1448:   F^a=Q^a\sin^{d-3}\theta\sin^{d-4}\phi\cdots d\theta\wedge d\phi\wedge\cdots\\
1449:   \tilde{F}^a=Q^{a}\sin^{d-3}\theta\sin^{d-4}\phi\cdots d\theta\wedge d\phi\wedge\cdots
1450: \end{eqnarray}
1451:   
1452:   
1453: The equation of motion for the scalars is
1454: \begin{equation}
1455:   \partial_{r}(a^{2}b^{d-2}\partial_{r}\phi_i)  = \frac{(d-2)!
1456:     \partial_iV_{eff}} {4b^{d-2}}.
1457:   \label{dilatoneqhd}
1458: \end{equation}
1459: Here $V_{eff}$, the effective potential for the scalars, is given by
1460: \begin{equation}
1461:   \label{eact}
1462:   V_{eff}=f_{ab}(\phi_i) Q^aQ^b.
1463: \end{equation}
1464: 
1465: From the $(R_{rr}-{G^{tt} \over G^{rr}} R_{tt})$ component of the
1466: Einstein equation we get,
1467: \begin{eqnarray}
1468:   \sum_i(\phi_i')^{2}= -{(d-2) b''(r) \over 2 b(r)}.
1469:   \label{Rrreq}
1470: \end{eqnarray}
1471: The $R_{rr}$ component gives the constraint,
1472: \begin{equation}
1473:   \label{constrainthd}
1474:   \begin{array}{l}
1475:     -(d-2)\{ (d-3)-ab'(2a'b+(d-3)ab')\}     
1476:     =  2\phi_i'^{2}a^{2}b^{2} -\frac{(d-2)!}{b^{2(d-3)}} V_{eff}(\phi_i)
1477:   \end{array}
1478: \end{equation}
1479: 
1480: 
1481: In the analysis below we will use eq.(\ref{dilatoneqhd}) to solve for
1482: the scalars and then eq.(\ref{Rrreq}) to solve for $b$. The constraint
1483: eq.(\ref{constrainthd}) will be used in solving for $a$ along with one
1484: extra relation, $R_{tt}=(d-3) {a^2 \over b^2} R_{\theta\theta}$, as is
1485: explained in appendix \ref{sec:higherdim}.  These equations (aside
1486: from the constraint) can be derived from a one-dimensional action
1487: \begin{equation}
1488:   \label{aohdv}
1489:   \begin{array}{cccc}
1490:     S&=&\frac{1}{\kappa^{2}}\int dr &
1491:     \Big((d-3)(d-2)b^{d-4}(1+a^{2}b^{'2})+(d-2)b^{d-3}(a^{2})^{'}b^{'} \\
1492:     &&&   -2a^{2}b^{d-2}(\partial_{r
1493:     }\phi)^{2}-\frac{(d-2)!}{b^{d-2}}V_{eff}\Big)
1494:   \end{array}
1495: \end{equation}
1496: 
1497: As the analysis below shows if the potential has a critical point at
1498: $\phi_i=\phi_{i0}$ and all the eigenvalues of the second derivative
1499: matrix $\partial_{ij}V(\phi_{i0})$ are positive then the attractor
1500: mechanism works in higher dimensions as well.
1501: 
1502: \subsection{Zeroth and First Order Analysis}
1503: Our starting point is the case where the scalars take asymptotic
1504: values equal to their critical value, $\phi_i=\phi_{i0}$. In this case
1505: it is consistent to set the scalars to be a constant, independent of
1506: $r$.  The extremal Reissner Nordstrom black hole in $d$ dimensions is
1507: then a solution of the resulting equations.  This takes the form,
1508: \begin{eqnarray}
1509:   \label{ernhd}
1510:   a_0(r)=\left(1-\frac{r_{H}^{d-3}}{r^{d-3}}\right)\quad
1511:   b_0(r)=r
1512: \end{eqnarray}
1513: where $r_{H}$ is the horizon radius.  From the eq.(\ref{constrainthd})
1514: evaluated at $r_H$ we obtain the relation,
1515: \begin{eqnarray}
1516:   r_{H}^{2(d-3)} =  (d-4)! V_{eff}(\phi_{i0})
1517: \end{eqnarray}
1518: Thus the area of the horizon and the entropy of the black hole are
1519: determined by the value of $V_{eff}(\phi_{i0})$, as in the
1520: four-dimensional case.
1521: 
1522: Now, let us set up the first order perturbation in the scalar fields,
1523: \begin{equation}
1524:   \phi_i=\phi_{i0}+\epsilon\phi_{i1}
1525: \end{equation}
1526: The first order correction satisfies,
1527: \begin{eqnarray}
1528:   \partial_{r}(a_0^{2}b_0^{d-2}\partial_{r}\phi_{i1})=\frac{\kkappa_i^{2}}{b^{d-2}}\phi_{i1}
1529: \end{eqnarray}
1530: where, $\kkappa_i^2$ is the eigenvalue of the second derivative matrix
1531: ${(d-2)! \over 4} \partial_{ij}V_{eff}(\phi_{i0})$ corresponding to
1532: the mode $\phi_i$.  This equation has two solutions. If $\kkappa_i^2
1533: >0$ one of these solutions blows up while the other is well defined
1534: and goes to zero at the horizon.  This second solution is the one we
1535: will be interested in. It is given by,
1536: \begin{equation}
1537:   \phi_{i1}=c_{i1}(1-{r_{H}^{d-3}/r^{d-3}})^{\gamma_i}
1538: \end{equation}
1539: where $\gamma$ is given by
1540: \begin{eqnarray}
1541:   \gamma_i={\frac{1}{2}}\left(-1+{\sqrt{1+4\kkappa_i^{2}r_{H}^{6-2d}/(d-3)^{2}}}\right)
1542: \end{eqnarray}
1543: 
1544: 
1545: 
1546: 
1547: 
1548: 
1549: 
1550: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1551: \subsubsection{Second order calculations (Effects of backreaction)}
1552: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1553: 
1554: The first order perturbation in the scalars gives rise to a second
1555: order correction for the metric components, $a,b$.  We write,
1556: \begin{eqnarray}
1557:   b(r)=b_{0}(r)+\epsilon^{2}b_{2}(r)\label{pertstart}\\
1558:   a(r)^{2}=a_{0}(r)^{2}+\epsilon^{2}a_{2}(r)\\
1559:   b(r)^{2}=b_{0}(r)^{2}+2\epsilon^{2}b_{2}(r)b_{0}(r)\label{pertend}
1560: \end{eqnarray}
1561: where $a_0,b_{0}$ are given in eq.(\ref{ernhd}).
1562: 
1563: From (\ref{Rrreq}) one can solve for the second order perturbation
1564: $b_{2}(r)$. For simplicity we consider the case of a single scalar
1565: field, $\phi$. The solution is given by double-integration form,
1566: \begin{eqnarray}
1567:   {\partial_{r}^{2}b_{2}(r)} & = & 
1568:   -{\frac{2}{(d-2)}}r(\partial_{r}\phi_{1})^{2}
1569:   =-c_{1}'{\frac{1}{r^{2d-5}}}({\frac{{r^{d-3}-r_{H}^{d-3}}}{{r^{d-3}}}})^{2\gamma-2}\nonumber \\
1570:   \Rightarrow b_{2}(r) & = &  d_{1}r+d_{2}
1571:   -{\frac{c_{1}' r}{{2(d-3)(d-4)\gamma(2 \gamma-1)r_H^{2d}}}}\times \nonumber\\
1572:   && \Bigl(-(d-4)
1573:   F\left[{\textstyle{\frac{1}{{3-d}}},1-2\gamma,{\frac{{d-4}}{{d-3}}};({\frac{r_H}{r}})^{d-3}}\right]\nonumber \\
1574:   & & +(2\gamma-1){\textstyle \left(\frac{r_H}{r}\right)^{d-3}}F\left[{\textstyle{\frac{{d-4}}{{d-3}}},1-2\gamma,{\frac{{2d-7}}{{d-3}}};({\frac{r_H}{r}})^{d-3}}\right]\Bigr)\,,
1575:   \label{btwoexact}
1576: \end{eqnarray}
1577: where $c_{1}'\equiv2(d-3)^{2}c_{1}^{2}\gamma^{2}r_H^{d}/(d-2)$, a
1578: positive definite constant, and $F$ is Gauss's Hypergeometric
1579: function.  More generally, for several scalar fields, $b_2$ is
1580: obtained by summing over the contributions from each scalar field.
1581: The integration constants $d_1, d_2$, in eq.(\ref{btwoexact}), can be
1582: fixed by coordinate transformations and requiring a double-zero
1583: horizon solution. We will choose coordinate so that the horizon is at
1584: $r=r_H$, then as we will see shortly the extremality condition
1585: requires both $d_1,d_2$ to vanish.  As $r \rightarrow r_H$ we have
1586: from eq.(\ref{btwoexact}) that
1587: \begin{equation}
1588:   b_{2}(r)\propto-\left({\frac{{r^{d-3}-r_{H}^{d-3}}}{{r^{d-3}}}}\right)^{2\gamma}
1589: \end{equation}
1590: Since $\gamma>0$, we see that $b_2$ vanishes at the horizon and thus
1591: the area and the entropy are uncorrected to second order.  At large
1592: $r$,
1593: $b_{2}(r)\propto{\mathcal{O}}(r)+{\mathcal{O}}(1)+{\mathcal{O}}(r^{7-2d})$
1594: so asymptotic behaviour is consistent with asymptotic flatness of the
1595: solution.
1596: 
1597: The analysis for $a_2$ is discussed in more detail in appendix
1598: \ref{sec:higherdim}.  In the vicinity of the horizon one finds that
1599: there is one non-singular solution which goes like, $a_2(r)
1600: \rightarrow C (r-r_H)^{(2\gamma +2)}$.  This solution smoothly extends
1601: to $r \rightarrow \infty$ and asymptotically, as $r\rightarrow
1602: \infty$, goes to a constant which is consistent with asymptotic
1603: flatness.
1604: 
1605: Thus we see that the backreaction of the metric is finite and well
1606: behaved.  A double-zero horizon black hole continues to exist to
1607: second order in perturbation theory. It is asymptotically flat.  The
1608: scalars in this solution at the horizon take their attractor values
1609: irrespective of their values at infinity..
1610: 
1611: Finally, the analysis in principle can be extended to higher orders.
1612: Unlike four dimensions though an explicit solution for the higher
1613: order perturbations is not possible and we will not present such an
1614: higher order analysis here.
1615: 
1616: We end with Fig 5.  which illustrates the attractor behaviour in
1617: asymptotically flat $4+1$ dimensional space. This figure has been
1618: obtained for the example, eq.(\ref{se}, \ref{seffpot}). The parameter
1619: $\delta r$ is defined in eq.(\ref{eq:def:deltar}).
1620:  
1621: \begin{figure}[htb]
1622:   \begin{center}
1623:     \epsfig{file=higherdim.eps,width=0.8\linewidth}
1624:   \end{center}
1625:   \caption{\label{cap:$4+1$ Dimensional Attractor}Numerical plot of
1626:     $\phi(r)$ with $\alpha_1=-\alpha_2=2$ for the an extremal black
1627:     hole in $4+1$ dimensions displaying attractor behaviour.}
1628: \end{figure}
1629: 
1630: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1631: \section{Attractor in $AdS_4$}
1632: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1633: Next we turn to the case of Anti-de Sitter space in four dimensions.
1634: Our analysis will be completely analogous to the discussion above for
1635: the four and higher dimensional case and so we can afford to be
1636: somewhat brief below.
1637: 
1638: The action in 4-dim. has the form
1639: \begin{equation}
1640:   \label{fourdact}
1641:   S=\frac{1}{\kappa^{2}}\int d^{4}x\sqrt{-G}
1642:   (R-2\Lambda-2(\partial\phi_i)^{2}-f_{ab}(\phi_i) F^a F^b
1643:   -{\textstyle {1 \over 2}} {\tilde f}_{ab}
1644:   (\phi_i) \epsilon^{\mu \nu \rho \sigma} F^a_{\mu \nu} F^b_{\rho \sigma} ) 
1645: \end{equation}
1646: where $\Lambda=-3/L^{2}$ is the cosmological constant.  For simplicity
1647: we will discuss the case with only one scalar field here. The
1648: generalisation to many scalars is immediate and along the lines of the
1649: discussion for asymptotically flat four-dimensional case. Also we take
1650: the coefficient of the scalar kinetic energy term to be field
1651: independent.
1652:  
1653:  
1654: For spherically symmetric solutions the metric takes the form,
1655: eq.(\ref{metric2}). The field strengths are given by
1656: eq.(\ref{fstrenghtgen}).  This gives rise to a one dimensional action
1657: \begin{equation}
1658:   \label{actoned}
1659:   S=\frac{1}{\kappa^{2}}\int dr
1660:   \left(2-(a^{2}b^{2})^{''}-2a^{2}bb^{''}-2a^{2}b^{2}(\partial_{r}\phi)^{2}-2\frac{V_{eff}}{b^{2}}
1661:     +\frac{3b^{2}}{L^{2}}\right),
1662: \end{equation} 
1663: where $V_{eff}$ is given by eq.(\ref{defpotgen}).  The equations of
1664: motion, which can be derived either from eq.(\ref{actoned}) or
1665: directly from the action, eq.(\ref{fourdact}) are now given by,
1666: \begin{eqnarray}
1667:   \partial_{r}(a^{2}b^{2}\partial_{r}\phi)=\frac{\partial_\phi V_{eff}(\phi)}{2b^{2}}\\
1668:   \frac{b^{''}}{b}=-(\partial_{r}\phi)^{2},\label{eq2ads}\end{eqnarray}
1669: which are unchanged from the flat four-dimensional case,
1670: and,
1671: \begin{equation}
1672:   (a^{2}(r)b^{2}(r))^{''}=2(1-2\Lambda b^{2}),\label{eq1ads}\end{equation}
1673: \begin{equation}
1674:   -1+a^{2}b^{'2}+\frac{a^{2'}b^{2'}}{2}=\frac{-1}{b^{2}}(V_{eff}(\phi))+a^{2}b^{2}(\partial_{r}\phi)^{2}+\frac{3b^{2}}{L^2},\label{constraintads}\end{equation}
1675: where the last equation is the first order constraint.
1676:  
1677:  
1678:  
1679:  
1680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1681: \subsection{Zeroth and First Order Analysis for $V$}
1682: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1683: The zeroth order solution is obtained by taking the asymptotic values
1684: of the scalar field to be its critical values, $\phi_{0}$ such that
1685: $\partial_iV_{eff}(\phi_{0})=0$.
1686: 
1687: The resulting metric is now the extremal Reissner Nordstrom black hole
1688: in AdS space, \cite{Chamblin:1999tk}, given by,
1689: \begin{eqnarray}
1690:   a_{0}(r)^{2}&=&{\frac{{(r-r_{H})^{2}(L^2+3r_{H}^{2}+2r_{H}r+r^{2})}}{{L^2r^{2}}}}\\
1691:   b_{0}(r)&=&r\end{eqnarray}
1692:   
1693: 
1694: The horizon radius $r_H$ is given by evaluating the constraint
1695: eq.(\ref{constraintads}) at the horizon,
1696: \begin{eqnarray}
1697:   {\frac{{(L^2r_H^{2}+2r_H^{4})}}{{L^2}}}=V_{eff}(\phi_{0}).\nonumber \\
1698: \end{eqnarray}
1699: 
1700: The first order perturbation for the scalar satisfies the equation,
1701: \begin{eqnarray}
1702:   \label{seads}
1703:   \partial_{r}(a_{0}^{2}b_0^{2}\partial_{r}\phi_{1})=\frac{\kkappa^{2}}{b^{2}}\phi_{1}\end{eqnarray}
1704: where,
1705: \begin{equation}
1706:   \label{valkappaads}
1707:   \kkappa^2={1 \over 2}\partial_\phi^2V_{eff}(\phi_0).
1708: \end{equation}
1709: This is difficult to solve explicitly.
1710:  
1711: In the vicinity of the horizon the two solutions are given by
1712: \begin{equation}
1713:   \label{sfoads}
1714:   \phi_1=C_{\pm} (r-r_H)^{t_{\pm}}
1715: \end{equation}
1716: If $V_{eff}''(\phi_0)>0$ one of the two solutions vanishes at the
1717: horizon.  We are interested in this solution. It corresponds to the
1718: choice,
1719: \begin{equation}
1720:   \label{sfoadsb}
1721:   \phi_1=C (r-r_H)^\gamma,
1722: \end{equation}
1723: where,
1724: \begin{equation}
1725:   \label{valtpads}
1726:   \gamma=
1727:   \frac{\sqrt{1+\frac{4\kkappa^{2}}{\delta r_H^{2}}}-1}{2},
1728: \end{equation}
1729: and, $\delta={(L^2+6r_H^2)\over L^2}$.  As discussed in the appendix
1730: \ref{sec:ads} this solution behaves at $r \rightarrow \infty$ as
1731: $\phi_1 \rightarrow C_1 +C_2/r^3$.  Also, all other values of $r$,
1732: besides the horizon and $\infty$, are ordinary points of the second
1733: order equation eq.(\ref{seads}). All this establishes that there is
1734: one well-behaved solution for the first order scalar perturbation. In
1735: the vicinity of the horizon it takes the form eq.(\ref{sfoads}) with
1736: eq.(\ref{valtpads}), and vanishes at the horizon. It is non-singular
1737: everywhere between the horizon and infinity and it goes to a constant
1738: asymptotically at $r \rightarrow \infty$.
1739:  
1740:  
1741: 
1742: We consider metric corrections next.  These arise at second order. We
1743: define the second order perturbations as in eq.(\ref{pert2}).  The
1744: equation for $b_2$ from the second order terms in eq.(\ref{eq2ads})
1745: takes the form,
1746: \begin{eqnarray}
1747:   b_{2}''=-r(\phi_{1}'(r))^{2},\label{adsb2eq}
1748: \end{eqnarray}
1749: and can be solved to give,
1750: \begin{equation}
1751:   \label{b2ads}
1752:   b_2(r)=-\int_{r_H}\int_{r_H} [r (\phi_{1}'(r))^{2}] 
1753: \end{equation}
1754: We fix the integration constants by taking take the lower limit of
1755: both integrals to be the horizon. We will see that this choice gives
1756: rise to an double-zero horizon solution. Since $\phi_{1}$ is well
1757: behaved for all $r_H\le r \le \infty$ the integrand above is well
1758: behaved as well. Using eq.(\ref{sfoads}) we find that in the near
1759: horizon region
1760: \begin{equation}
1761:   \label{nhb2ads}
1762:   b_2 \sim (r-r_H)^{(2 \gamma)}
1763: \end{equation}
1764: 
1765: At $r \rightarrow \infty$ using the fact that $\phi_1 \rightarrow C_1
1766: +C_2/r^3$ we find
1767: \begin{equation}
1768:   \label{asbads}
1769:   b_2 \sim D_1 r + D_2 + D_3/r^6.
1770: \end{equation}
1771: This is consistent with an asymptotically AdS solution.
1772: 
1773: Finally we turn to $a_2$. As we show in appendix \ref{sec:ads} a
1774: solution can be found for $a_2$ with the following properties.  In the
1775: vicinity of the horizon it goes like,
1776: \begin{equation}
1777:   \label{nhads}
1778:   a_2 \propto (r-r_H)^{(2\gamma + 2)},
1779: \end{equation}
1780: and vanishes faster than a double-zero.  As $r \rightarrow \infty$,
1781: $a_2 \rightarrow d_1 r $ and grows more slowly than $a_0^2$.  And for
1782: $r_H<r<\infty$ it is well-behaved and non-singular.
1783: 
1784: This establishes that after including the backreaction of the metric
1785: we have a non-singular, double-zero horizon black hole which is
1786: asymptotically AdS. The scalar takes a fixed value at the horizon of
1787: the black hole and the entropy of the black hole is unchanged as the
1788: asymptotic value of the scalar is varied.
1789: 
1790: Let us end with two remarks.  In the AdS case one can hope that there
1791: is a dual description for the attractor phenomenon.  Since the
1792: asymptotic value of the scalar is changing we are turning on a
1793: operator in the dual theory with a varying value for the coupling
1794: constant. The fact that the entropy, for fixed charge, does not change
1795: means that the number of ground states in the resulting family of dual
1796: theories is the same.  This would be worth understanding in the dual
1797: description better.  Finally, we expect this analysis to generalise in
1798: a straightforward manner to the AdS space in higher dimensions as
1799: well.
1800: 
1801: Fig. 6 illustrates the attractor mechanism in asymptotically $AdS_4$
1802: space.  This Figure is for the example, eq.(\ref{se}, \ref{seffpot}).
1803: The cosmological constant is taken to be, $\Lambda =-2.91723$, in
1804: $\kappa=1$ units.
1805:  
1806:  
1807: \begin{figure}[htb]
1808:   \begin{center}
1809:     \epsfig{file=ads.eps,width=0.8\linewidth}
1810:   \end{center}
1811:   \caption{\label{cap:ads}Numerical plot of $\phi(r)$ with
1812:     $\alpha_1=-\alpha_2=2$ for the an extremal black hole in $AdS_4$
1813:     displaying attractor behaviour. }
1814: \end{figure}
1815: 
1816: 
1817: 
1818: 
1819: \section{Additional Comments}
1820: The theories we considered in the discussion of asymptotically flat
1821: space-times and AdS spacetimes have no potential for the scalars.  We
1822: comment on this further here.
1823: 
1824: Let us consider a theory with ${\cal N}=1$ supersymmetry containing
1825: chiral superfields whose lowest component scalars are,
1826: \begin{equation}
1827:   \label{defcs}
1828:   S_i=\phi_i+i a_i
1829: \end{equation}
1830: We take these scalars to be uncharged under the gauge symmetries.
1831: These can be coupled to the superfields $W_\alpha^a$ by a coupling
1832: \begin{equation}
1833:   \label{gcoup}
1834:   L_{gauge kinetic}=\int d^2\theta f_{ab}(S_i) W_\alpha^a W_\alpha^b
1835: \end{equation}
1836: Such a coupling reproduces the gauge kinetic energy terms in and
1837: eq.(\ref{fourdact}), eq.(\ref{actoned}), (we now include both $\phi_i,
1838: a_i$ in the set of scalar fields which we denoted by $\phi_i$ in the
1839: previous sections).
1840: 
1841: An additional potential for the scalars would arise due to F-term
1842: contributions from a superpotential.  If the superpotential is absent
1843: we get the required feature of no potential for these scalar.  Setting
1844: the superpotential to be zero is at least technically natural due to
1845: its non-renormalisability.
1846: 
1847: In a theory with no supersymmetry there is no natural way to suppress
1848: a potential for the scalars and it would arise due to quantum effects
1849: even if it is absent at tree-level.  In this case we have no good
1850: argument for not including a potential for the scalar and our analysis
1851: is more in the nature of a mathematical investigation.
1852: 
1853: The absence of a potential is important also for avoiding no-hair
1854: theorems which often forbid any scalar fields from being excited in
1855: black hole backgrounds \cite{Nohair}. In the presence of a mass $m$ in
1856: asymptotically flat four dimensional space the two solutions for first
1857: order perturbation at asymptotic infinity go like,
1858: \begin{equation}
1859:   \label{assmass}
1860:   \phi \sim C_1 e^{mr}/r, \phi \sim C_2 e^{-mr}/r.
1861: \end{equation}
1862: We see that one of the solutions blows up as $r \rightarrow \infty$.
1863: Since one solution to the equation of motion also blows up in the
1864: vicinity of the horizon, as discussed in section 2, there will
1865: generically be no non-singular solution in first order perturbation
1866: theory. This argument is a simple-minded way of understanding the
1867: absence of scalar hair for extremal black holes under discussion here.
1868: In the absence of mass terms, as was discussed in section 2, the two
1869: solutions at asymptotic infinity go like $ \phi \sim {\rm const}$ and
1870: $\phi \sim 1/r$ respectively and are both acceptable.  This is why one
1871: can turn on scalar hair. The possibility of scalar hair for a massless scalar 
1872: is of course well known. See \cite{Gibbons:1985ac}, \cite{Gibbons:1987ps},
1873:  for some early examples of solutions with scalar hair, \cite{Masood-ul-Alam:1993ea, Mars:2001pz, 
1874: Gibbons:2002ju,Gibbons:2002av},
1875:  for theorems on uniqueness in the presence of such hair, and \cite{Gibbons:1996af}  
1876: for a discussion of 
1877: resulting thermodynamics.  
1878: 
1879: In asymptotic AdS space the analysis is different. Now the $(mass)^2$
1880: for scalars can be negative as long as it is bigger than the BF bound.
1881: In this case both solutions at asymptotic infinity decay and are
1882: acceptable. Thus, as for the massless case, it should be possible to
1883: turn on scalar fields even in the presence of these mass terms and
1884: study the resulting black holes solutions.  Unfortunately, the
1885: resulting equations are quite intractable. For small $(mass)^2$ we
1886: expect the attractor mechanism to continue to work.
1887:  
1888: If the $(mass)^2$ is positive one of the solutions in the asymptotic
1889: region blows up and the situation is analogous to the case of a
1890: massive scalar in flat space discussed above.  In this case one could
1891: work with AdS space which is cut off at large $r$ (in the infrared)
1892: and study the attractor phenomenon.  Alternatively, after
1893: incorporating back reaction, one might get a non-singular geometry
1894: which departs from AdS in the IR and then analyse black holes in this
1895: resulting geometry.  In the dual field theory a positive $(mass)^2$
1896: corresponds to an irrelevant operator.  The growing mode in the bulk
1897: is the non-normalisable one and corresponds to turning on a operator
1898: in the dual theory which grows in the UV. Cutting off AdS space means
1899: working with a cut-off effective theory. Incorporating the
1900: back-reaction means finding a UV completion of the cut-off theory.
1901: And the attractor mechanism means that the number of ground states at
1902: fixed charge is the same regardless of the value of the coupling
1903: constant for this operator.
1904: 
1905: 
1906: 
1907: 
1908:  
1909: 
1910: 
1911: 
1912: 
1913: 
1914: 
1915: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1916: \section{Asymptotic de Sitter Space}
1917: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1918: In de Sitter space the simplest way to obtain a double-zero horizon is
1919: to take a Schwarzschild black hole and adjust the mass so that the de
1920: Sitter horizon and the Schwarzschild horizon coincide. The resulting
1921: black hole is the extreme Schwarzschild-de Sitter spacetime
1922: \cite{exds}.  We will analyse the attractor behaviour of this black
1923: hole below.  The analysis simplifies in 5-dimensions and we will
1924: consider that case, a similar analysis can be carried out in other
1925: dimensions as well.  Since no charges are needed we set all the gauge
1926: fields to zero and work only with a theory of gravity and scalars.  Of
1927: course by turning on gauge charges one can get other double-zero
1928: horizon black holes in dS, their analysis is left for the future.
1929: 
1930: We start with the action of the form,
1931: \begin{equation}
1932:   \label{actds}
1933:   S=\frac{1}{\kappa^{2}}\int d^{5}x\sqrt{-G}(R-2(\partial\phi)^{2}-V(\phi))
1934: \end{equation}
1935: Notice that the action now includes a potential for the scalar,
1936: $V(\phi)$, it will play the role of $V_{eff}$ in our discussion of
1937: asymptotic flat space and AdS space.  The required conditions for an
1938: attractor in the dS case will be stated in terms of $V$. A concrete
1939: example of a potential meeting the required conditions will be given
1940: at the end of the section.  For simplicity we have taken only one
1941: scalar, the analysis is easily extended for additional scalars.
1942: 
1943: The first condition on $V$ is that it has a critical point,
1944: $V'(\phi_0)=0$.  We will also require that $V(\phi_0) >0$.  Now if the
1945: asymptotic value of the scalar is equal to its critical value,
1946: $\phi_0$, we can consistently set it to this value for all times $t$.
1947: The resulting equations have a extremal black hole solution mentioned
1948: above. This takes the form
1949: \begin{equation}
1950:   \label{nariamet}
1951:   ds^2=-{t^2 \over (t^2/L-L/2)^2 } dt^2 +{(t^2/L-L/2)^2 \over t^2} dr^2 + t^2 d\Omega_3^2
1952: \end{equation}
1953: Notice that it is explicitly time dependent.  $L$ is a length related
1954: to $V(\phi_0)$ by , $V(\phi_0)={20 \over L^2}$. And $t=\pm {L \over
1955:   \sqrt{2}}$ is the location of the double-zero horizon. A suitable
1956: near-horizon limit of this geometry is called the Nariai solution,
1957: \cite{nariai}.
1958: 
1959: \subsection{Perturbation Theory}
1960: Starting from this solution we vary the asymptotic value of the
1961: scalar. We take the boundary at $t \rightarrow -\infty$ as the initial
1962: data slice and investigate what happens when the scalar takes a value
1963: different from $\phi_0$ as $t \rightarrow -\infty$. Our discussion
1964: will involve part of the space-time, covered by the coordinates in
1965: eq.(\ref{nariamet}), with $-\infty\le t \le t_H= -{L \over \sqrt{2}}$.
1966: We carry out the analysis in perturbation theory below.
1967: 
1968: Define the first order perturbation for the scalar by, $$
1969: \phi=\phi_0
1970: + \epsilon \phi_1$$
1971: This satisfies the equation,
1972: \begin{equation}
1973:   \label{fods}
1974:   \partial_t(a_0^2b_0^3\partial_t\phi_1)={b^3\over 4} V''(\phi_0)\phi_1
1975: \end{equation}
1976: where $a_0={(t^2/L-L/2)\over t}$, $b_0=t$.  This equation is difficult
1977: to solve in general.
1978: 
1979: In the vicinity of the horizon $t=t_H$, we have two solutions which go
1980: like,
1981: \begin{equation}
1982:   \label{twosolds}
1983:   \phi_1=C_{\pm} (t-t_H)^{-1+\sqrt{1+\kappa^2} \over 2}
1984: \end{equation}
1985: where
1986: \begin{equation}
1987:   \label{valkds}
1988:   \kappa^2=-{1\over 4} V''(\phi_0)
1989: \end{equation}
1990: We see that one of the two solutions in eq.(\ref{twosolds}) is
1991: non-divergent and in fact vanishes at the horizon if
1992: \begin{equation}
1993:   \label{condpotds}
1994:   V''(\phi_0) <0.
1995: \end{equation}
1996: We will henceforth assume that the potential meets this condition.
1997: Notice this condition has a sign opposite to what was obtained for the
1998: asymptotically flat or AdS cases.  This reversal of sign is due to the
1999: exchange of space and time in the dS case.
2000: 
2001: In the vicinity of $t \rightarrow -\infty$ there are two solutions to
2002: eq.(\ref{fods}) which go like,
2003: \begin{equation}
2004:   \label{foinds}
2005:   \phi_1={\tilde C}_{\pm}|t|^{p_{\pm}}
2006: \end{equation}
2007: where
2008: \begin{equation}
2009:   \label{valpm}
2010:   p_{\pm}=2(-1 \pm \sqrt{1+\kappa^2/4}).
2011: \end{equation}
2012: If the potential meets the condition, eq.(\ref{condpotds}) then
2013: $\kappa^2>0$ and we see that one of the modes blows up at $t
2014: \rightarrow -\infty$.
2015: 
2016: \subsection{Some Speculative  Remarks}
2017: In view of the diverging mode at large $|t|$ one needs to work with a
2018: cutoff version of dS space \footnote{This is related to some comments
2019:   made in the previous section in the positive $(mass)^2$ case in AdS
2020:   space.}.  With such a cutoff at large negative $t$ we see that there
2021: is a one parameter family of solutions in which the scalar takes a
2022: fixed value at the horizon. The one parameter family is obtained by
2023: starting with the appropriate linear combination of the two solutions
2024: at $t \rightarrow -\infty$ which match to the well behaved solution in
2025: the vicinity of the horizon.  While we will not discuss the metric
2026: perturbations and scalar perturbations at second order these too have
2027: a non-singular solution which preserves the double-zero nature of the
2028: horizon.  The metric perturbations also grow at the boundary in
2029: response to the growing scalar mode and again the cut-off is necessary
2030: to regulate this growth. This suggests that in the cut-off version of
2031: dS space one has an attractor phenomenon. Whether such a cut-off makes
2032: physical sense and can be implemented appropriately are question we
2033: will not explore further here.
2034: 
2035: One intriguing possibility is that quantum effects implement such a
2036: cut-off and cure the infra-red divergence. The condition on the
2037: potential eq.(\ref{condpotds}) means that the scalar has a negative
2038: $(mass)^2$ and is tachyonic. In dS space we know that a tachyonic
2039: scalar can have its behaviour drastically altered due to quantum
2040: effects if it has a $(mass)^2<H^2$ where $H$ is the Hubble scale of dS
2041: space. This can certainly be arranged consistent with the other
2042: conditions on the potential as we will see below. In this case the
2043: tachyon can be prevented from ``falling down'' at large $|t|$ due to
2044: quantum effects and the infrared divergences can be arrested by the
2045: finite temperature fluctuations of dS space.  It is unclear though if
2046: any version of of the attractor phenomenon survives once these quantum
2047: effects became important.
2048: 
2049: We end by discussing one example of a potential which meets the
2050: various conditions imposed above.  Consider a potential for the
2051: scalar,
2052: \begin{equation}
2053:   \label{potds}
2054:   V=\Lambda_1e^{\alpha_1 \phi} + \Lambda_2 e^{\alpha_2 \phi}.
2055: \end{equation}
2056: We require that it has a critical point at $\phi=\phi_0$ and that the
2057: value of the potential at the critical point is positive. The critical
2058: point for the potential eq.(\ref{potds}) is at,
2059: \begin{equation}
2060:   \label{critds}
2061:   e^{\phi_{0}}=-\left( {\alpha_2 \Lambda_2 \over \alpha_1 \Lambda_1}\right)^{1\over \alpha_1-\alpha_2}
2062: \end{equation} 
2063: Requiring that $V(\phi_0)>0$ tells us that
2064: \begin{equation}
2065:   \label{tcds}
2066:   V(\phi_0)
2067:   =\Lambda_2 e^{\alpha_2 \phi_0}\left(1-{\alpha_2 \over \alpha_1}\right) >0
2068: \end{equation}
2069: Finally we need that $V''(\phi_0) <0$ this leads to the condition,
2070: \begin{equation}
2071:   \label{sdds}
2072:   V''(\phi_0)=\Lambda_2e^{\alpha_2\phi_0}\alpha_2(\alpha_2-\alpha_1)<0
2073: \end{equation}
2074: These conditions can all be met by taking both $\alpha_1, \alpha_2>0$,
2075: $\alpha_2<\alpha_1$, $\Lambda_2>0$ and $\Lambda_1<0$.  In addition if
2076: $\alpha_2\alpha_1 \gg 1$ the resulting $-(mass)^2 \gg H^2$.
2077: 
2078: 
2079: 
2080: 
2081: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2082: \section{Non-Extremal = Unattractive }\label{sec:nonex}
2083: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2084: 
2085: We end the paper by examining the case of an non-extremal black hole
2086: which has a single-zero horizon.  As we will see there is no attractor
2087: mechanism in this case. Thus the existence of a double-zero horizon is
2088: crucial for the attractor mechanism to work.
2089: 
2090: Our starting point is the four dimensional theory considered in
2091: section 2 with action eq.(\ref{actiongen}).  For simplicity we
2092: consider only one scalar field.  We again start by consistently
2093: setting this scalar equal to its critical value, $\phi_0$, for all
2094: values of $r$, but now do not consider the extremal Reissner Nordstrom
2095: black hole.  Instead we consider the non-extremal black hole which
2096: also solves the resulting equations.  This is given by a metric of the
2097: form, eq.(\ref{metric2}), with
2098: \begin{equation}
2099:   \label{metne}
2100:   a^2(r)=\left(1-{r_+\over r}\right)\left({1-{r_-\over r}}\right),\qquad b(r)=r
2101: \end{equation}
2102: where $r_\pm$ are not equal. We take $r_+>r_-$ so that $r_+$ is the
2103: outer horizon which will be of interest to us.
2104:  
2105: The first order perturbation of the scalar field satisfies the
2106: equation,
2107: \begin{equation}
2108:   \label{fone}
2109:   \partial_r(a^2b^2\partial_r\phi_i)={V_{eff}''(\phi_0)\over 4 b^2} \phi_1
2110: \end{equation}
2111: In the vicinity of the horizon $r=r_+$ this takes the form,
2112: \begin{equation}
2113:   \label{vhne}
2114:   \partial_y(y \partial_y \phi_1)=\alpha \phi_1
2115: \end{equation}
2116: where $\alpha$ is a constant dependent on $V''(\phi_0), r_+,r_-$, and
2117: $y \equiv r-r_+$.
2118: 
2119: This equation has one non-singular solution which goes like,
2120: \begin{equation}
2121:   \label{nhsne}
2122:   \phi_1 = C_0 + C_1 y + \cdots
2123: \end{equation}
2124: where the ellipses indicate higher terms in the power series expansion
2125: of $\phi_1$ around $y=0$.  The coefficients $C_1, C_2 , \cdots$ are
2126: all determined in terms of $C_0$ which can take any value.  Thus we
2127: see that unlike the case of the double-horizon extremal black hole,
2128: here the solution which is well-behaved in the vicinity of the horizon
2129: does not vanish.
2130: 
2131: Asymptotically, as $r \rightarrow \infty$ both solutions to
2132: eq.(\ref{fone}) are well defined and go like $1/r, constant$
2133: respectively. It is then straightforward to see that one can choose an
2134: appropriate linear combination of the two solutions at infinity and
2135: match to the solution, eq.(\ref{nhsne}) in the vicinity of the
2136: horizon. The important difference here is that the value of the
2137: constant $C_0$ in eq.(\ref{nhsne}) depends on the asymptotic values of
2138: the scalar at infinity and therefore the value of $\phi$ does not go
2139: to a fixed value at the horizon. The metric perturbations sourced by
2140: the scalar perturbation can also be analysed and are non-singular. In
2141: summary, we find a family of non-singular black hole solutions for
2142: which the scalar field takes varying values at infinity.  The crucial
2143: difference is that here the scalar takes a value at the horizon which
2144: depends on its value at asymptotic infinity.  The entropy and mass for
2145: these solutions also depends on the asymptotic value of the scalar
2146: \footnote{An intuitive argument was given in the introduction in
2147:   support of the attractor mechanism.  Namely, that the degeneracy of
2148:   states cannot vary continuously. This argument only applies to the
2149:   ground states. A non-extremal black hole corresponds to excited
2150:   states.  Changing the asymptotic values of the scalars also changes
2151:   the total mass and hence the entropy in this case.}.
2152: 
2153: It is also worth examining this issue in a non-extremal black holes
2154: for an exactly solvable case.
2155: 
2156: If we consider the case $|\alpha_i|=2$, section \ref{sec:exact_sol},
2157: the non-extremal solution takes on a relatively simple form.  It can
2158: be written\cite{Kallosh:1992ii}
2159: \begin{eqnarray}
2160:   \label{eq:nonex}
2161:   \exp(2\phi) & = & e^{2\phi_{\infty}}\frac{(r+\Sigma)}{(r-\Sigma)}\nonumber \\
2162:   a^{2} & = & \frac{(r-r_{+})(r-r_{-})}{(r^{2}-\Sigma^{2})}\label{eq:alpha2_solution}\\
2163:   b^{2} & = & (r^{2}-\Sigma^{2})\nonumber 
2164: \end{eqnarray}
2165: where\footnote{The radial coordinate $r$ in
2166:   eq.(\ref{eq:alpha2_solution}) is related to our previous one by a
2167:   constant shift.}
2168: \begin{equation}
2169:   \label{rhne}
2170:   r_{\pm}=M\pm r_{0}\qquad r_{0}=\sqrt{M^{2}+\Sigma^{2}-\bar{Q}_2^{2}-\bar{Q}_1^{2}}.
2171: \end{equation}
2172: and the Hamiltonian constraint becomes
2173: \begin{equation}
2174:   \Sigma^{2}+M^{2}-\bar{Q}_{1}^{2}-\bar Q_{2}^{2}=\frac{1}{4}(r_+-r_-)^{2}.
2175: \end{equation}
2176: The scalar charge, $\Sigma$, defined by
2177: $\phi\sim\phi_\infty+\frac{\Sigma}{r} $, is not an independent
2178: parameter.  It is given by
2179: \begin{equation}
2180:   \Sigma=\frac{\bar{Q}_2^{2}-\bar{Q}_1^{2}}{2M}.
2181: \end{equation} 
2182: There are horizons at $r=r_{\pm}$, the curvature singularity occurs at
2183: $r=\Sigma$ and $r_{0}$ characterises the deviation from extremality.
2184: We see that the non-extremal solution does not display attractor
2185: behaviour.
2186: 
2187: 
2188: Fig. \ref{cap:unattractive} shows the behaviour of the scalar field
2189: \footnote{for $\alpha_1=-\alpha_2=2$} as we vary $\phi_\infty$ keeping
2190: $M$ and $Q_i$ fixed. The location of the horizon as a function of $r$
2191: depends on $\phi_\infty$, eq.(\ref{rhne}).  The horizon as a function
2192: of $\phi_\infty$ is denoted by the dotted line. The plot is terminated
2193: at the horizon.
2194: 
2195: In contrast, for the extremal black hole,
2196: \begin{equation}
2197:   \label{eq:alpha2:Mass}
2198:   M=\frac{|\bar{Q}_2|+|\bar{Q}_1|}{\sqrt{2}}\qquad\Sigma=\frac{|\bar Q_2|-|\bar Q_1|}{\sqrt{2}},
2199: \end{equation}
2200: so (\ref{eq:nonex}) gives
2201: \begin{equation}
2202:   \label{eq:phi:oncemore}
2203:   e^{2\phi_0}=e^{2\phi_\infty}\frac{M+\Sigma}{M-\Sigma}
2204:   \stackrel{(\ref{eq:alpha2:Mass})}{=}
2205:   \frac{|Q_2|}{|Q_1|},
2206: \end{equation}
2207: which is indeed the attractor value.
2208: % **** Still to do - write down some explicit formulae for $\phi_0
2209: % (T)$ etc. and relate to 1st law of Black hole thermodynamics etc.
2210: \begin{figure}[htb]
2211:   \begin{center}
2212:     \epsfig{file=unattractive.new.eps,width=0.8\linewidth}
2213:   \end{center}
2214:   \caption{\label{cap:unattractive}Plot $\phi(r)$ with
2215:     $\alpha_1=-\alpha_2=2$ for the non-extremal black hole with $M,
2216:     Q_i$ held fixed while varying $\phi_\infty$.  The dotted line
2217:     denotes the outer horizon at which we terminate the plot.  It is
2218:     clearly unattractive.}
2219: \end{figure}
2220: 
2221: 
2222: \bigskip \goodbreak \centerline{\bf Acknowledgements}
2223: \noindent
2224: We thank A.  Dabholkar, R. Gopakumar, G. Gibbons, S. Minwalla, A. Sen,
2225: M. Shigemori and A.  Strominger for discussions.  N.I. carried out
2226: part of this work while visiting Harvard University.  He would like to
2227: thank the Harvard string theory group for its hospitality and
2228: support.  We thank the organisers of ISM-04, held at Khajuraho, for a
2229: stimulating meeting. This research is supported by the Government of India. 
2230: S.T. acknowledges support from the Swarnajayanti Fellowship, DST, Govt. of India. 
2231: Most of all we thank the people of India for
2232: generously supporting research in String Theory.
2233: 
2234:  
2235: 
2236: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2237: \appendix
2238: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2239: \section{Perturbation Analysis}\label{sec:appa}
2240: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2241: \renewcommand{\theequation}{A.\arabic{equation}}
2242: \setcounter{equation}{0}
2243: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2244: 
2245: \subsection{Mass}
2246: \label{Mass_calc}
2247: 
2248: Here, we first calculate the mass of the extremal black hole discussed
2249: in section 2.2.  From eq.(\ref{finalb2}), for large r, $b_{2}$ is
2250: given by,
2251: \begin{equation}
2252:   b_{2}=cr+d\label{assb2}
2253: \end{equation}
2254: where
2255: \begin{eqnarray}
2256:   c & = & -\frac{c_{1}^{2}\gamma}{2(2\gamma-1)}\\
2257:   d & = & \frac{r_H c_{1}^{2}\gamma^2}{(2\gamma-1)}
2258: \end{eqnarray}
2259: Now, we can easily write down the expression for $a_{2}$ using
2260: eq.(\ref{a2}).  We choose coordinate $y$ as introduced in
2261: eq.(\ref{by}) such that at large r,
2262: \begin{eqnarray}
2263:   r^{2}+2\epsilon^{2}(cr^{2}+dr) & = & y^{2}\\
2264:   \frac{1}{r} & = & \frac{1}{y}(1+\epsilon^{2}(c+\frac{d}{y}))\end{eqnarray}
2265: We use the extremality condition (\ref{finala}) to find,
2266: \begin{eqnarray}
2267:   a(r)=\left(\frac{r-r_H}{y}\label{assa2}\right)\end{eqnarray}
2268: 
2269: Using, eq.(\ref{assb2}, \ref{assa2}) one finds that asymptotically, as
2270: $r\rightarrow \infty$ the metric takes the form,
2271: \begin{equation}
2272:   \label{assmetr}
2273:   ds^2=-\left(1-{2(r_H+\epsilon^2 (c r_H+d)) \over y} \right) d{\tilde t}^2 + 
2274:   {1 \over (1-{2(r_H+\epsilon^2 (c r_H+d)) \over y} )} dy^2 + y^2 d\Omega^2
2275: \end{equation} 
2276: where ${\tilde t}$ is obtained by rescaling $t$ and $d\Omega^2$
2277: denotes the metric of $S^2$.  The mass $M$ of the black hole is then
2278: given by the $1/y$ term in the $g_{yy}$ component of the metric.  This
2279: gives,
2280: \begin{equation}
2281:   M=r_H+\epsilon^{2}\frac{r_Hc_1^2\gamma}{2}. \end{equation}
2282: 
2283: 
2284: \subsection{Perturbation Series to All Orders}
2285: \label{sec:all_orders}
2286: 
2287: 
2288: Next we go on to discuss the perturbation series to all orders, Using
2289: (\ref{pertb}) for $b$ and (\ref{pertphi}) for $\phi$ in
2290: eq.(\ref{eq2})and eq.(\ref{eqphi}), we get,
2291: \begin{eqnarray}
2292:   b_{k}''=-\sum_{i=0}^{k}\sum_{j=0}^{k-i}b_{i}\phi_{j}'\phi_{k-i-j}'
2293: \end{eqnarray}
2294: \begin{eqnarray}
2295:   \sum_{i+j\leqslant k }
2296:   2b_{j}b_{k-i-j}((r-r_H)^{2}\phi_{i}')'
2297:   = Q_i^2 e^{\alpha_i\phi_0}\alpha_i  {\cal V}_{ik}
2298: \end{eqnarray}
2299: where
2300: \begin{eqnarray}
2301:   \label{eq:def:s_k}
2302:   {\cal V}_{ik} &=& \sum_{\stackrel{\{n_1, n_2 \ldots n_{k}\}}{\sum m n_m = k}}
2303:   \frac{\phi_1^{n_1}\phi_2^{n_2}\ldots \phi_k^{n_{k}}}{n_1!n_2!\ldots n_{k}!}\alpha^{n_1+n_2+\ldots+n_{k}}_i.\\
2304: \end{eqnarray}
2305: After substituting our ansatz (\ref{diln})and (\ref{bn}), the above
2306: equations give,
2307: \begin{equation}
2308:   \label{eq:dk_rel}
2309:   k (k \gamma-1)d_k = -\gamma\sum_{{i+j< k} } j(k-i-j)d_i c_j c_{k-i-j}  
2310: \end{equation}
2311: and
2312: \begin{equation}
2313:   \label{eq:aftersubs}
2314:   k(k\gamma+1)c_k+
2315:   T_k= (\gamma+1)(c_k+ S_k)
2316: \end{equation}
2317: where $S_k$ and $T_k$ are given by
2318: \begin{eqnarray}
2319:   \label{eq:def:Sk}
2320:   S_k &=&
2321:   \sum_{\stackrel{\{n_1, n_2 \ldots n_{k-1}\}}{\sum m n_m = k}}
2322:   \frac{c_1^{n_1}c_2^{n_2}\ldots c_{k-1}^{n_{k-1}}}{n_1!n_2!\ldots n_{k-1}!}
2323:   \left(
2324:     \alpha^{\sum n_l-1}_1 +\alpha^{\sum n_l-1}_2
2325:   \right)
2326: \end{eqnarray}
2327: and
2328: \begin{equation}
2329:   \label{eq:def:Tk}
2330:   T_k=\sum_{\stackrel{j+l\leqslant k}{l<k}}l(l\gamma+1)d_j d_{k-l-j}c_l.
2331: \end{equation}
2332: Then solving for $d_k$ and $c_k$ gives
2333: \begin{eqnarray}
2334:   d_{k}&=&-\frac{ \gamma}{k(k\gamma-1)}\sum_{{i+j< k} } j(k-i-j)d_i e_j e_{k-i-j}
2335:   \label{dk_ans} \\  
2336:   c_k &=&\frac{(\gamma+1)S_k-T_k}{((k+1)\gamma+1)(k-1)}
2337:   \label{ek_ans}
2338: \end{eqnarray}
2339: Finally, $e_k$ can be obtained using eq.(\ref{finala}), eq.(\ref{bn}).
2340: It can be verified that the ansatz, eq.(\ref{diln}, \ref{bn},
2341: \ref{an}) with the coefficients eq.(\ref{dk_ans}, \ref{ek_ans}) also
2342: solves the constraint eq.(\ref{constraint}).
2343: 
2344: 
2345: % \begin{eqnarray}
2346: %   \sum_{\stackrel{j=0\ldots k-i}{\scriptscriptstyle i=0\ldots k}}
2347: %   2b_{j}b_{k-i-j}((r-r_{1})^{2}\phi_{i}')'=\textstyle{\mathcal{A}}((\alpha_1-\alpha_2)\phi_{k}+\ldots+\frac{(\alpha_1^{k}-\alpha_2^{k})}{k!}\phi_{1}^{k})\end{eqnarray} 
2348: %   % as the resulting equations. 
2349: %   % After substituting our ansatz (\ref{diln})and (\ref{bn}),the above equations at order k gives,
2350: %   \begin{eqnarray}
2351: %   d_{k}=-\frac{\sum_{i=0}^{k}\sum_{j=0}^{k-i}d_{i}\gamma^{2}(j)(k-i-j)e_{j}e_{k-i-j}}{(k\gamma)(k\gamma-1)}\label{dk}
2352: % \end{eqnarray}
2353: % \begin{eqnarray}
2354: %   e_{k}=
2355: %   -\frac{{\mathcal{A}}(\frac{(\alpha_{1}^{k}-\alpha_{2}^{k})}{k!}e_{1}^{k}+.......)-\sum_{i=0}^{k-2}\sum_{j=0}^{k-i}2d_{j}d_{k-i-j}((i\gamma)(1+i\gamma)r_{1}^{2}e_{i})}
2356: %   {{\mathcal{A}}(\alpha_1-\alpha_2)-2k\gamma r_{1}^{2}(1+k\gamma)}\label{ek}
2357: % \end{eqnarray}
2358:  
2359: 
2360: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% 
2361: \section{Exact Analysis }\label{sec:details}
2362: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2363: \renewcommand{\theequation}{B.\arabic{equation}}
2364: \setcounter{equation}{0}
2365: 
2366: 
2367: Exact solutions can be found by writing the equations of motion as
2368: generalised Toda equations \cite{Toda}, which may, in certain special
2369: cases, be solved exactly \cite{Gibbons:1987ps} - we rederive this
2370: result in slightly different notation below. As noted in
2371: \cite{Lu:1996jr}, in a marginally different context, the extremal
2372: solutions, are, in appropriate variables, polynomial solutions of the
2373: Toda equations.  The polynomial solutions are much easier to find and
2374: are related to the functions $f_i$ mentioned in section
2375: \ref{sec:exact_sol}. For ease of comparison we occasionally use
2376: notation similar to \cite{Lu:1996jr}.
2377: 
2378: 
2379: 
2380: 
2381: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2382: \subsection{New Variables}
2383: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2384: 
2385: 
2386: To recast the equations of motion into a generalised Toda equation we
2387: define the following new variables
2388: \begin{equation}
2389:   u_1=\phi
2390:   \qquad u_2=\log a
2391:   \qquad z=\log ab
2392:   \qquad\cdot=\partial_{\tau}=a^{2}b^{2}\partial_{r}
2393: \end{equation}
2394: In terms of $r$, $\tau$ is given by
2395: \begin{equation}
2396:   \tau=\int\frac{dr}{a^{2}b^{2}}=\frac{1}{(r_{+}-r_{-})}\log\left(\frac{r-r_{+}}{r-r_{-}}\right)
2397: \end{equation}
2398: where $r_\pm$ are the integration constants of (\ref{eq1}). In general
2399: (\ref{eq1}) implies
2400: \begin{equation}
2401:   \label{eq:def:rpm}
2402:   a^2 b^2 = (r-r_+)(r-r_-).
2403: \end{equation}
2404: Notice that
2405: \begin{eqnarray}
2406:   \tau\rightarrow0 & \mbox{as} & r\rightarrow\infty\\
2407:   \tau\rightarrow-\infty & \mbox{as} & r\rightarrow r_{+}
2408: \end{eqnarray}
2409: When we have a double-zero horizon, $r_{H}=r_{\pm}$, $\tau$ takes the
2410: simple form
2411: \begin{equation}
2412:   \label{eq:tau_double}
2413:   \tau^{-1}=-{\left(r-r_{H}\right)}.
2414: \end{equation}
2415: Since we are mainly interested in solutions with double-zero horizons,
2416: in what follows it will be convenient to work with a new radial
2417: coordinate, $\rho$, defined by
2418: \begin{equation}
2419:   \label{eq:def:rho}
2420:   \rho=-\tau^{-1}.
2421: \end{equation}
2422: which has the convenient property that $\rho_H=0$.
2423: 
2424: \subsection{Equivalent Toda System}
2425: In terms of these new variables the equations of motion become
2426: \begin{eqnarray}
2427:   \ddot{u}_1 &=&
2428:   \textha\alpha_1 e^{2u_2+\alpha_1 u_1}Q_{1}^{2}+\textha\alpha_2e^{2 u_2 +\alpha_2u_1}Q_{2}^{2} \label{eq:phi_eom_3}\\
2429:   \ddot{u}_2 &=& e^{2u_2+\alpha_1 u_1}Q_{1}^{2}+e^{2u_2+\alpha_2u_1}Q_{2}^{2} \label{eq:v_eom_1}\\
2430:   \ddot{z}&=&e^{2z}  \label{eq:z_EOM}
2431: \end{eqnarray}
2432: \begin{equation}
2433:   \dot{u_1}^{2}+\dot{u_2}^{2}-\dot{z}^{2}
2434:   +e^{2z}-e^{2u_2+\alpha_1 u_1}Q_{1}^{2}-e^{2u_2+\alpha_2 u_1}Q_{2}^{2}=0
2435:   \label{eq:toda_E}
2436: \end{equation}
2437: (\ref{eq:z_EOM}) decouples from the other equations and is equivalent
2438: to (\ref{eq1}).  Finally making the coordinate change
2439: \begin{equation}
2440:   \label{eq:def:X}
2441:   X_i = n_{ij}^{-1}u_j +m_{ij}^{-1}\log\left((\alpha_1-\alpha_2)Q_{j}^{2}\right)
2442: \end{equation}
2443: where
2444: \begin{equation}
2445:   \label{eq:def:n}
2446:   n^{-1}=\left(
2447:     \begin{array}{cc}
2448:       2 & -\alpha_2 \\
2449:       -2 & \alpha_1
2450:     \end{array}\right)
2451: \end{equation}
2452: % \begin{equation}
2453: %   \label{eq:n2}
2454: %   n=\frac{1}{2(\alpha_1-\alpha_2)}\left(
2455: %     \begin{array}{cc}
2456: %       \alpha_1 & \alpha_2 \\
2457: %       2 & 2
2458: %     \end{array}\right)
2459: % \end{equation}
2460: and
2461: \begin{equation}
2462:   m_{ij}=\frac{1}{2\left(\alpha_1-\alpha_2\right)}
2463:   \left(
2464:     4+\alpha_i\alpha_j
2465:   \right)
2466: \end{equation}
2467: % \begin{equation}
2468: %   m^{-1}=\frac{1}{2\left(\alpha_1-\alpha_2\right)}\left(
2469: %     \begin{array}{cc}
2470: %       4+\alpha_2^{2} & -(4+\alpha_1\alpha_2)\\
2471: %       -(4+\alpha_1\alpha_2) & 4+\alpha_1^{2}
2472: %     \end{array}\right)
2473: % \end{equation}
2474: we obtain the generalised 2 body Toda equation
2475: \begin{equation}
2476:   \ddot{X}_{i}=e^{m_{ij}X_{j}},
2477: \end{equation}
2478: together with
2479: \begin{equation}
2480:   \label{eq:toda_ham}
2481:   \sum_{ij}\left(\textstyle\ha \dot X_i m_{ij}\dot X_j 
2482:     -e^{m_{ij} X_j}\right)=(\alpha_1-\alpha_2)\mathcal{E}
2483: \end{equation}
2484: where ${\cal E}=\frac{1}{4}(r_+-r_-)^2$.
2485: % and $\det m = 1$. One may check that $n^T n=
2486: % m/(2(\alpha_1-\alpha_2))$.
2487: %
2488: % As special case we see that if $\alpha_1=-\alpha_2$, and lettting
2489: % $2e^{-y}=\alpha_1$ then we have
2490: % \begin{equation}
2491: %   \ddot{X}_{\mu}=e^{\  m_{\mu}^{\phantom{\mu}\nu}X_{\nu}}\qquad m=\left(
2492: %     \begin{array}{cc}
2493: %       \cosh y & \sinh y\\
2494: %       \sinh y & \cosh y
2495: %     \end{array}\right)
2496: % \end{equation}
2497: After solving the above, the original fields will be given by
2498: \begin{eqnarray}
2499:   \label{eq:final_sol}
2500:   e^{(\alpha_1-\alpha_2)\phi}&=&\frac{Q_2^2}{Q_1^2}e^{\ha(\alpha_1 X_1 +\alpha_2 X_2)} \\
2501:   a^2 &=& e^{\frac{2}{\alpha_1-\alpha_2}(X_1+X_2)}/\diamondsuit \\
2502:   b^2 &\stackrel{(\ref{eq1})}{=}& (r-r_+)(r-r_-)/a^2
2503: \end{eqnarray}
2504: where
2505: \begin{equation}
2506:   \label{eq:defdiamond}
2507:   \diamondsuit = (\alpha_1-\alpha_2)Q_1^{2\foob} Q_2^{2\fooa}
2508: \end{equation}
2509: 
2510: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2511: \subsection{Solutions}
2512: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2513: 
2514: 
2515: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2516: \subsubsection{Case I: $\gamma=1\Leftrightarrow \alpha_1\alpha_2=-4$}
2517: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2518: 
2519: In this case, $m_{ij}$ is diagonal
2520: \begin{equation}
2521:   \label{eq:M_gamma1}
2522:   m=\mathrm{diag}(\alpha_1/2,-\alpha_2/2),
2523: \end{equation}
2524: so the equations of motion decouple:
2525: \begin{equation}
2526:   \label{eq:eom:gamma1}
2527:   \ddot{X}_{i}=e^{\frac{|\alpha_i|}{2} X_{i}}.
2528: \end{equation}
2529: (\ref{eq:eom:gamma1}) has solutions
2530: \begin{equation}
2531:   X_i = \frac{2}{|\alpha_i|}
2532:   \log\left(
2533:     \frac{4c_i^2}{|\alpha_i|\sinh^2(c_i(\tau-d_i))}
2534:   \right)
2535: \end{equation}
2536: The integration constants are fixed by imposing asymptotic boundary
2537: conditions and requiring that the solution is finite at the horizon.
2538: Letting
2539: \begin{equation}
2540:   F_{i}=\sinh(c_{i}\left(\tau-d_{i}\right))
2541: \end{equation}
2542: in terms of $\phi$ and $a$ we get
2543: \begin{equation}
2544:   \label{eq:exact_sol_case1}
2545:   \begin{array}{cll}
2546:     e^{(\alpha_1-\alpha_2)\phi}&=&\displaystyle\frac{Q_2^2}{Q_1^2}e^{\ha(\alpha_1 X_1 +\alpha_2 X_2)} \\
2547:     &=& \displaystyle
2548:     \left(
2549:       -\frac{\alpha_2}{\alpha_1}
2550:     \right) 
2551:     \left(
2552:       \frac{Q_2 F_2 c_1}{Q_1 F_1 c_2}
2553:     \right)^2 \\
2554:     a^2 &=&\displaystyle\left.e^{\frac{2}{\alpha_1-\alpha_2}(X_1+X_2)}\right/ \diamondsuit\\
2555:     &=& 
2556:     \displaystyle
2557:     \left.
2558:       \left(\frac{ c_1}{Q_1 F_1}\right)^{2\foob} 
2559:       \left(
2560:         \frac{ c_2}{Q_2 F_2}
2561:       \right)^{2\aod}
2562:     \right/\spadesuit
2563:   \end{array}
2564: \end{equation}
2565: As $r\rightarrow r_{+}$(ie. $\tau\rightarrow-\infty$) the scalar field
2566: goes like
2567: \begin{equation}
2568:   e^{(\alpha_1-\alpha_2)\phi}  \sim  e^{2(c_{1}-c_{2})\tau}\\
2569: \end{equation}
2570: so we require
2571: \begin{equation}
2572:   c:= c_{1}  =  c_{2}
2573: \end{equation}
2574: for a finite solution at the horizon. Also at the horizon
2575: \begin{equation}
2576:   \label{eq:b_at_horiz}
2577:   b^{2} \sim (r-r_+)/a^2 \sim  e^{\left((r_{+}-r_{-})-2c\right)\tau}  
2578: \end{equation}
2579: which necessitates
2580: \begin{equation}
2581:   (r_{+}-r_{-})  =  2c
2582: \end{equation}
2583: To find the extremal solutions we take the limit $c\rightarrow0$ which
2584: gives
2585: \begin{eqnarray}
2586:   e^{(\alpha_1-\alpha_2)\phi} 
2587:   & = & \left(
2588:     -\frac{\alpha_2}{\alpha_1}\right)\frac{\left(Q_{2}f_{2}\right)^{2}}{\left(Q_{1}f_{1}
2589:     \right)^{2}}\\
2590:   b^{2} 
2591:   & = & \spadesuit \left(
2592:     Q_{1}f_{1}\right)^{\frac{-2{\alpha_2}}{\alpha_1-\alpha_2}}\left(Q_{2}p_{2}
2593:   \right)^{\frac{2\alpha_1}{\alpha_1-\alpha_2}}\\
2594:   a^{2} & = & \rho^{2}/b^{2}
2595: \end{eqnarray}
2596: where
2597: \begin{equation}
2598:   \label{eq:fi}
2599:   f_{i}  =  1+d_{i}\rho.
2600: \end{equation}
2601: % and as before $\rho=r-r_+$.
2602: Requiring $\phi\rightarrow\phi_\infty$ and $a\rightarrow1$ as
2603: $r\rightarrow\infty$ fixes
2604: \begin{equation}
2605:   d_{i}=\bar{Q_{i}}^{-1}\sqrt{\frac{|\alpha_{i}|}{\alpha_1-\alpha_2}}
2606: \end{equation}
2607: where as before
2608: \begin{equation}
2609:   \bar{Q}_{i}^2=e^{\alpha_{i}\phi_{\infty}}Q_{i}^2\nosum.
2610: \end{equation}
2611: 
2612: For comparison with the non-extremal solution in this case see section
2613: \ref{sec:nonex}.
2614: % In the simplest subcase, $\alpha_1=-\alpha_2=2$, the non-extremal
2615: % solution (\ref{eq:exact_sol_case1}) can be written
2616: % \cite{Kallosh:1992ii}
2617: % \begin{eqnarray}
2618: %   \exp(2\phi) & = & e^{2\phi_{\infty}}\frac{(r+\Sigma)}{(r-\Sigma)}\\
2619: %   a^{2} & = & \frac{(r-r_{+})(r-r_{-})}{(r^{2}-\Sigma^{2})}\\
2620: %   b^{2} & = & (r^{2}-\Sigma^{2})
2621: % \end{eqnarray}
2622: % The scalar charge, not an independent quantity, is given by
2623: % \begin{equation}
2624: %   \Sigma=\frac{\bar{Q}_{2}^{2}-\bar{Q}_{1}^{2}}{2M}
2625: % \end{equation}
2626: % and the Hamiltonian constraint becomes
2627: % \begin{equation}
2628: %   \Sigma^{2}+M^{2}-\bar{Q}_{1}^{2}-\bar{Q_{2}}^{2}=\frac{1}{4}(r_+-r_-)^{2}
2629: % \end{equation}
2630: 
2631: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2632: \subsubsection{Case II: $\gamma=2$ and $\alpha_1=-\alpha_2=2\sqrt{3}$
2633: }
2634: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2635: In this case, $m_{ij}$ becomes
2636: \begin{equation}
2637:   \label{eq:M_gamma2}
2638:   m=\left(
2639:     \matrix{ \frac{2}
2640:       {{\sqrt{3}}} & -
2641:       \frac{1}{{\sqrt{3}}}
2642:       \cr -
2643:       \frac{1}{{\sqrt{3}}}
2644:       & \frac{2}
2645:       {{\sqrt{3}}} \cr  }
2646:   \right).
2647: \end{equation}
2648: It is convenient to use the coordinates
2649: \begin{equation}
2650:   \label{eq:def:q_gamma2}
2651:   q_i = \frac{1}{\sqrt 3}X_i - \sqrt3\log\sqrt3 
2652: \end{equation}
2653: so the equations of motion are the two particle Toda equations
2654: \begin{eqnarray}
2655:   \label{eq:eom:gamma2}
2656:   \ddot{q}_{1}&=&e^{2 q_{1}-q_2}\\
2657:   \ddot{q}_{2}&=&e^{2 q_{2}-q_1}.
2658: \end{eqnarray}
2659: These maybe integrated exactly but the explicit form is, in general, a
2660: little complicated. Fortunately we are mainly interested in extremal
2661: solutions which have a simpler form \cite{Lu:1996jr}. As in,
2662: \cite{Lu:1996jr}, taking the ansatz that $e^{-q_i}$ is a second order
2663: polynomial one finds
2664: \begin{eqnarray}
2665:   \label{eq:gamma2_anzatz}
2666:   e^{-q_1}&=&a_0+a_1\tau+\ha\tau^2 \\
2667:   e^{-q_2}&=&a_1^2-a_0+a_1\tau+\ha\tau^2
2668: \end{eqnarray}
2669: Finally, returning to the original variables and imposing the
2670: asymptotic boundary conditions gives the solution
2671: \begin{eqnarray}
2672:   e^{4\sqrt{3}\phi} & = & \left(\frac{\PP}{\Q}\right)^2
2673:   \left(\frac{f_2}{f_1}\right)^{6}\label{eq:exact_scalar_sol_gamma2}\\
2674:   b^{2} & = & 2Q_1 Q_2 f_1 f_2\label{eq:exact_b_sol_gamma2}\\
2675:   a^{2} &=& \rho^2/b^2
2676: \end{eqnarray}
2677: where
2678: \begin{equation}
2679:   f_i  = \left( 1+
2680:     (\bar{Q}_1 \bar{Q}_2)^{-\frac{2}{3}} (\bar{Q}_1^{\frac{2}{3}} + \bar{Q_2}^{\frac{2}{3}})^\ha \rho
2681:     +{\textstyle\ha} (\bar{Q}_i\bar{Q}_1 \bar{Q}_2)^{-\frac{2}{3}}\rho^{2}\right)^\ha
2682: \end{equation} 
2683: as quoted in section \ref{sec:exact_sol}.
2684: 
2685: For completeness we note that the general, non-extremal solution of
2686: \cite{Dobiasch:1981vh,Gibbons:1985ac}, modified for a non-zero
2687: asymptotic value of $\phi$, is
2688: \begin{eqnarray}
2689:   \exp(4\phi/\sqrt{3}) & =e^{4\phi_{\infty}/\sqrt{3}} & \frac{p_2}{p_1}\\
2690:   a^{2} & = & \frac{(r-r_{+})(r-r_{-})}{\sqrt{p_1 p_2}}\\
2691:   b^{2} & = & \sqrt{p_1 p_2}
2692: \end{eqnarray}
2693: where
2694: \begin{equation}
2695:   p_i=(r-r_{i+})(r-r_{i-})
2696: \end{equation}
2697: \begin{equation}
2698:   r_{i\pm}=\frac{2}{(-\alpha_i)}\Sigma\pm\bar{Q}_i\sqrt{\frac{4\Sigma}{2\Sigma+\alpha_i M}}
2699: \end{equation}
2700: and scalar charge, $\Sigma$, which is again not an independent
2701: parameter, is given by
2702: \begin{equation}
2703:   \frac{1}{\sqrt{3}}\Sigma=\frac{\bar{Q}_2^{2}}{2M(\lambda-1)}+\frac{\bar{Q}_1^{2}}{2M(\lambda+1)}\qquad\lambda=\frac{\Sigma}{\sqrt{3}M}
2704: \end{equation}
2705: 
2706: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2707: \subsubsection{Case III: $\gamma=3$ and $\alpha_1=4\qquad\alpha_2=-6$
2708: }
2709: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2710: In this case, $m_{ij}$ becomes
2711: \begin{equation}
2712:   m=\left(
2713:     \begin{array}{cc}
2714:       1 & -1\\
2715:       -1 & 2
2716:     \end{array}\right)
2717: \end{equation}
2718: Making the coordinate change
2719: \begin{eqnarray}
2720:   \label{eq:def:q_gamma3}
2721:   q_1 &=& \textha X_1 - \log 2 \\
2722:   q_2 &=& X_2 - \log 2
2723: \end{eqnarray}
2724: The equations of motion are
2725: \begin{eqnarray}
2726:   \label{eq:gamma3_eom}
2727:   \ddot q_1 &=& e^{2q_1-q_2} \\
2728:   \ddot q_2 &=& e^{2q_2-2q_2} 
2729: \end{eqnarray}
2730: Now consider the three particle Toda system
2731: \begin{eqnarray}
2732:   \ddot q_1 &=& e^{2q_1-q_2} \label{eq:toda_gamma3_1}\\
2733:   \ddot q_2 &=& e^{2q_2-q_1-q_3} \\
2734:   \ddot q_3 &=& e^{2q_3-q_2} \label{eq:toda_gamma3_2}
2735: \end{eqnarray} 
2736: which may be integrated exactly. Notice that by identifying $q_1$ and
2737: $q_3$ we obtain (\ref{eq:toda_gamma3_1}-\ref{eq:toda_gamma3_2}). Once
2738: again the general solution is slightly complicated but taking the
2739: ansatz that $e^{-q_i}$ is a polynomial one finds
2740: \begin{eqnarray}
2741:   e^{-q_{1}} & = & a_{0}+2a_{2}^{2}\tau+a_{2}\tau^{2}+\frac{1}{6}\tau^{3}\\
2742:   e^{-q_{2}} 
2743:   & = & 4a_{2}^{4}-2a_{0}a_{2}
2744:   +(4a_{2}^{3}-a_{0})\tau+2a_{2}^{2}\tau^{2}+\frac{2}{3}a_{2}\tau^{3}+\frac{1}{12}\tau^{4}
2745: \end{eqnarray}
2746: 
2747: Rewriting in terms of the original fields we get
2748: \begin{eqnarray}
2749:   e^{10\phi} & = & \left(\frac{Q_{2}}{Q_{1}}\right)^{2}\exp\left(2X_{1}-3X_{2}\right)\\
2750:   & = & \frac{6}{4}\left(\frac{Q_{2}}{Q_{1}}\right)^{2}\left(\frac{f_{2}}{f_{1}}\right)^{12}\\
2751:   b^{2} 
2752:   & = & \rho^2 10 Q_{1}^{\frac{6}{5}}Q_{2}^{\frac{4}{5}}
2753:   \exp\left(-\frac{1}{5}X_{1}-\frac{1}{5}X_{2}\right)\\
2754:   &= &\frac{5}{2}
2755:   \left(\frac{2}{3}\right)^{\frac{3}{5}}Q_{1}^{\frac{6}{5}}Q_{2}^{\frac{4}{5}}f_1 f_2 \\
2756:   a^2 &=& \rho^2/b^2
2757: \end{eqnarray} 
2758: where
2759: \begin{equation}
2760:   f_{1} =  \left(
2761:     1-6a_{2}\rho + 12a_{2}^{2}\rho^{2}-6a_{0}\rho^{3}
2762:   \right)^{\frac{1}{3}}
2763: \end{equation}
2764: \begin{equation}
2765:   f_{2}  =  \left(
2766:     1-{\textstyle \frac{24}{3}}a_{2}\rho
2767:     +24a_{2}\rho^{2}-(48a_{2}^{3}-12a_{0})\rho^{3}+\left(48a_{2}^{4}-24a_{0}a_{2}\right)\rho^{4}
2768:   \right)^{\frac{1}{4}}
2769: \end{equation}
2770: At the horizon we do indeed have $\phi$ at the critical point of
2771: $V_{eff}$:
2772: \begin{equation}
2773:   e^{10\phi_0}=\frac{3}{2}\frac{Q_{2}^{2}}{Q_{1}^{2}}
2774: \end{equation}
2775: and $b^2$ given by $V_{eff}(\phi_0)$:
2776: \begin{equation}
2777:   b_H^{2}=
2778:   \frac{5}{2}\left(\frac{2}{3}\right)^{\frac{3}{5}}
2779:   Q_{1}Q_{2}\left(\frac{Q_{2}}{Q_{1}}\right)^{\frac{1}{5}}.
2780: \end{equation}
2781: Imposing the asymptotic boundary conditions we get
2782: % \begin{eqnarray}
2783: %   1 & = & 
2784: %   2\frac{\bar{Q}_{2}^{2}}{\bar{Q}_{1}^{2}}\frac{\left(4a_{2}^{4}-2a_{0}a_{2}\right)^{3}}{a_{0}^{4}}\\
2785: %   1 & = & 
2786: %   \frac{2^{3}10^{-5}\bar{Q}_{1}^{-6}\bar{Q}_{2}^{-4}}{a_{0}^{2}\left(4a_{2}^{4}-2a_{0}a_{2}\right)}
2787: % \end{eqnarray}
2788: % which implies
2789: % 
2790: \begin{equation}
2791:   a_{0}=\pm\frac{2^{\frac{5}{7}}}{\bar{Q}_{1}^{\frac{10}{7}}\bar{Q}_{2}^{\frac{5}{7}}}
2792:   \qquad
2793:   \left(4a_{2}^{4}-2a_{0}a_{2}\right)
2794:   =\frac{2^{\frac{11}{7}}}{\bar{Q}_{1}^{\frac{22}{7}}\bar{Q}_{2}^{\frac{18}{7}}}
2795: \end{equation}
2796: so letting
2797: \begin{eqnarray}
2798:   \clubsuit & = & \frac{2^{\frac{11}{7}}}{Q_{1}^{\frac{22}{7}}Q_{2}^{\frac{18}{7}}}\\
2799:   \Delta_{1} & = & 3\sqrt[3]{a_{0}^{3}+\sqrt{a_{0}^{6}+\frac{64}{3}a_{0}^{3}\clubsuit^{3}}}\\
2800:   \Delta_{2} & = & \sqrt{\frac{3^{\frac{1}{3}}\Delta_{1}}{a_0}-\frac{3^{\frac{2}{3}}4}{\Delta_1}}
2801: \end{eqnarray}
2802: we may write $a_2$ as
2803: \begin{equation}
2804:   a_{2}
2805:   =\pm\frac{1}{2\sqrt{6}}\Delta_{2}\pm\frac{1}{2}
2806:   \sqrt{
2807:     \frac{2\clubsuit}{3^{\frac{1}{3}}\Delta_{1}}
2808:     -\frac{\Delta_{1}}{2\,3^{\frac{2}{3}}}+\frac{\sqrt{6}}{\Delta_{2}}
2809:   }
2810: \end{equation}
2811: Despite the non-trivial form of the solution we see that it still
2812: takes on the attractor value at the horizon.
2813: 
2814: In terms of the $U(1)$ charges (written implicitly in terms of $a_0$
2815: and $a_2$), the mass and scalar charge are expressed below
2816: 
2817: % \begin{equation}
2818: %   \phi\sim\mathrm{const}+\frac{3}{10}\log\left(1-\frac{(4a_{2}^{3}-a_{0})}{\left(4a_{2}^{4}-2a_{0}a_{2}\right)}\frac{1}{\rho}\right)-\frac{2}{5}\log\left(1-\frac{2a_{2}^{2}}{a_{0}}\frac{1}{\rho}\right)
2819: % \end{equation}
2820: 
2821: \begin{equation}
2822:   \Sigma=\frac{3a_{0}^{2}-28a_{0}a_{2}^{3}+32a_{2}^{6}}{40a_{0}a_{2}^{4}-20a_{0}^{2}a_{2}}
2823: \end{equation}
2824: 
2825: % \begin{equation}
2826: %   a^{2}\sim1-\frac{2^{\frac{3}{5}}}{50}Q_{1}^{-\frac{6}{5}}Q_{2}^{-\frac{4}{5}}\left(4\frac{a_{2}^{2}}{a_{0}}+\frac{4a_{2}^{3}-a_{0}}{4a_{2}^{4}-2a_{0}a_{2}}\right)\frac{1}{\rho}
2827: % \end{equation}
2828: 
2829: \begin{equation}
2830:   M=
2831:   \frac{
2832:     (a_{0}^{2}+4a_{0}a_{2}^{3}-16a_{2}^{6})
2833:   }
2834:   {
2835:     2^{\frac{3}{5}}50a_{0}^{\frac{7}{5}}a_{2}(a_{0}-2a_{2}^{3})
2836:     (2a_{2}^{4}-a_{0}a_{2})^{\frac{1}{5}}
2837:     Q_{1}^{\frac{6}{5}}Q_{2}^{\frac{4}{5}}
2838:   }
2839: \end{equation}
2840: 
2841: This solution is related to a 3 charge p-brane solution found in
2842: \cite{Lu:1996jr} - in this case we have identified two of the degrees
2843: of freedom.
2844: % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2845: % \section{Cheat Sheet}\label{sec:cheat}
2846: % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% \renewcommand{\theequation}{C.\arabic{equation}} \setcounter{equation}{0}
2847: % \begin{equation}
2848: %   \label{eq:noris_eqn}
2849: %   \alpha\ta <0 
2850: % \end{equation}
2851: % \begin{equation}
2852: %   e^{\phi_{0}}=\left(-\frac{\ta Q_{2}^{2}}{\alpha Q_{1}^{2}}\right)^{\frac{1}{(\alpha-\ta)}}
2853: %   \label{eq:appc:attractor}
2854: % \end{equation}
2855: 
2856: % \begin{equation}
2857: %   \label{eq:rel:alpha1}
2858: %   -\frac{\alpha}{\tilde\alpha} \exp^{\alpha\phi_0}Q_1^2 = \exp^{\tilde\alpha\phi_0}Q_2^2
2859: % \end{equation}
2860: % \begin{equation}
2861: %   \label{eq:rel:r1}
2862: %   r_1^2 = Q_1^2 e^{\alpha\phi_0}+Q_2^2 e^{\tilde\alpha\phi_0}
2863: %   =  Q_1^2 e^{\alpha\phi_0}
2864: %   (-\ta)^{-1}(\alpha  -\ta) 
2865: % \end{equation}
2866: % \begin{equation}
2867: %   \label{eq:rel:beta}
2868: %   \beta^2 =\ha (\alpha^{2}e^{\alpha\phi_{0}}Q_{1}^{2}+\ta^{2}e^{\ta\phi_{0}}Q_{2}^{2})
2869: %   = \ha   Q_1^2 e^{\alpha\phi_0} \alpha(\alpha-\ta )
2870: % \end{equation}
2871: % \begin{equation}
2872: %   \label{eq:rel:ration}
2873: %   \frac{\beta^2}{r_1^2}=\ha \alpha (-\ta)
2874: % \end{equation}
2875: % \begin{equation}
2876: %   \label{eq:gamma_def}
2877: %   \gamma=\ha (\sqrt{1+4\beta^2/r_1^2}-1)
2878: % \end{equation}
2879: % \begin{equation}
2880: %   \label{eq:rel:gamma}
2881: %   \gamma(\gamma+1)= \beta^2/r_1^2
2882: % \end{equation}
2883: % \begin{eqnarray}
2884: %   r_1^2&
2885: %   Q_1^2 \left(-\frac{\ta Q_{2}^{2}}{\alpha Q_{1}^{2}}\right)^{\frac{\alpha}{(\alpha-\ta)}}
2886: %   + Q_2^2 \left(-\frac{\ta Q_{2}^{2}}{\alpha Q_{1}^{2}}\right)^{\frac{\ta}{(\alpha-\ta)}}\\
2887: %   &=& Q_1^{2\taod}Q_2^{2\aod}
2888: %   \underbrace{
2889: %     \left(
2890: %       \left(-\frac{\ta}{\alpha}\right)^\aod +
2891: %       \left(-\frac{\ta}{\alpha}\right)^\frac{\ta}{\alpha-\ta}
2892: %     \right)}
2893: %   _{=\spadesuit=\left(
2894: %       -\frac{\alpha}{\ta}\right)^\taod 
2895: %     \left(
2896: %       1-\frac{\ta}{\alpha}
2897: %     \right)
2898: %     = (\alpha-\ta)/(-\ta)^{\taod}\alpha^{\aod}
2899: %   } \label{eq:rel:r1_2}
2900: % \end{eqnarray}
2901: 
2902: 
2903: 
2904: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2905: % \section{Comparison with other references}\label{otherref}
2906: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2907: % \renewcommand{\theequation}{C.\arabic{equation}}
2908: % \setcounter{equation}{0}
2909: 
2910: % \begin{itemize}
2911: % \item Extremality condition:\[
2912: %   M^{2}+\Sigma^{2}=P^{2}e^{\tilde{\alpha}\phi_{\infty}}+Q^{2}e^{\alpha\phi_{\infty}}=\bar{P}^{2}+\bar{Q}^{2}\]
2913: 
2914: 
2915: %   \begin{itemize}
2916: %   \item $P^{2}e^{\tilde{\alpha}\phi_{\infty}}:=\bar{P}^{2}\qquad
2917: %     Q^{2}e^{\alpha\phi_{\infty}}:=\bar{Q}^{2}$
2918: %   \end{itemize}
2919: %   {\bf Role of SUSY and integrability}
2920: 
2921: 
2922: % \end{itemize}
2923:  
2924:  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2925: \section{Higher Dimensions }\label{sec:higherdim}
2926: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2927: \renewcommand{\theequation}{C.\arabic{equation}}
2928: \setcounter{equation}{0}
2929: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2930: Here we give some more details related to our discussion of the higher
2931: dimensional attractor in section 5.  The Ricci components calculated
2932: from the metric, eq.(\ref{metrichd}) are,
2933: \begin{eqnarray}
2934:   R_{tt} & = & a^{2}\left(a'^{2}+{\frac{{(d-2)aa'b'}}{{b}}}+aa''\right) \\
2935:   R_{rr} & = & -\{ b\left(a'^{2}+aa''\right)+(d-2)a\left(a'b'+ab''\right)\}/a^{2}b  \\
2936:   R_{\theta\theta} & = &
2937:   (d-3)-2aba'b'-a^{2}\left((d-3)b'^{2}+bb''\right)
2938: \end{eqnarray}
2939: 
2940: The Einstein equations from the action eq.(\ref{higherdaction}), take
2941: the form,
2942: \begin{eqnarray}
2943:   R_{tt} & = & \frac{(d-3)(d-3)!a(r)^{2}}{b(r)^{2(d-2)}}V_{eff}(\phi_i) \\
2944:   R_{rr} & = & 2(\partial_{r}\phi)^{2}-\frac{(d-3)(d-3)!}{b(r)^{2(d-2)}a(r)^{2}}V_{eff}(\phi_i) \nonumber \\
2945:   \\
2946:   R_{\theta\theta} & = & \frac{(d-3)!}{b^{2(d-3)}}V_{eff}(\phi_i),
2947: \end{eqnarray}
2948: where $V_{eff}$ is given by eq.(\ref{eact}).
2949: 
2950: Taking the combination, ${1\over 2}(R_{rr}-{G_{rr}\over
2951:   G_{tt}}R_{tt})$ gives, eq.(\ref{Rrreq}).  Similarly we have,
2952: \begin{eqnarray}
2953:   \label{Hconstraint}
2954:   &  & {\frac{b(r)^{2}}{a(r)^{2}}}R_{tt}+a(r)^{2}b(r)^{2}R_{rr}-(d-2)R_{\theta\theta}\nonumber \\
2955:   & = & -(d-2)\{ d-3-a(r)b'(r)(2a'(r)b(r)+(d-3)a(r)b'(r))\}\nonumber \\
2956:   & = & 2(\partial_{r}\phi_i)^{2}a(r)^{2}b(r)^{2}-\frac{(d-2)(d-3)!}{b^{2(d-3)}}V_{eff}(\phi_i)\
2957: \end{eqnarray}
2958: This gives eq.(\ref{constrainthd}).  Finally the relation,
2959: $R_{tt}=(d-3){a^2\over b^2}R_{\theta\theta}$ yields,
2960: \begin{eqnarray}
2961:   &  & (d-3)^{2}(-1+a(r)^{2}b'(r)^{2})+b(r)^{2}(a'(r)^{2}+a(r)a''(r))\nonumber \\
2962:   &  & +a(r)b(r)((-8+3d)a'(r)b'(r)+(d-3)a(r)b''(r))=0. \label{leqhd}\end{eqnarray}
2963:  
2964: We now discuss solving for $a_2$, the second order perturbation in the
2965: metric component $a$, in some more detail. We restrict ourselves to
2966: the case of one scalar field, $\phi$.  The constraint,
2967: eq.(\ref{constrainthd}), to $O(\epsilon^2)$ is,
2968: \begin{eqnarray}
2969:   (d-2) ra_{2}'+(d-2)(d-3)a_{2}-2(\phi_{1}')^{2}r^{2}(1-(\frac{r_{H}}{r})^{d-3})^{2}\label{constrainthigher1storder}\\
2970:   -2(d-2)(d-3)^{2}\frac{r_{H}^{2(d-3)}}{r^{2(d-3)+1}}b_{2}+2(d-3)^{2}\frac{\gamma(\gamma+1)\phi_{1}^{2}}{r^{2(d-3)}r_{H}^{6-2d}}\nonumber \\
2971:   +2(d-2)\frac{(d-3)(r_{H}^{3}r^{d}-r_{h}^{d}r^{3})}{r_{H}^{6}r^{2d}}\left\{ r_{H}^{d}r^{2}b_{2}+r_{h}^{3}r^{d}b_{2}'\right\}  =  0\nonumber \end{eqnarray}
2972: This is a first order equation for $a_2$ of the form,
2973: \begin{equation}
2974:   f_{1}a_{2}'+f_{2}a_{2}+f_{3}=0,
2975: \end{equation}
2976: where,
2977: \begin{eqnarray}
2978:   f_{1} & = & (d-2) r \nonumber \\
2979:   f_{2} & = & (d-2)(d-3)\nonumber \\
2980:   f_{3} & = & -2(\phi_{1}')^{2}r^{2}(1-(\frac{r_{H}}{r})^{d-3})^{2}-2(d-2)(d-3)^{2}\frac{r_{H}^{2(d-3)}}{r^{2(d-3)+1}}b_{2}\nonumber \\
2981:   &  & +2(d-3)^{2}\frac{t(t+1)\phi_{1}^{2}}{r^{2(d-3)}r_{H}^{6-2d}} \nonumber \\
2982:   & & +2(d-2)\frac{(d-3)(r_{H}^{3}r^{d}-r_{H}^{d}r^{3})}{r_{H}^{6}r^{2d}}\left\{ r_{H}^{d}r^{2}b_{2}+r_{H}^{3}r^{d}b_{2}'\right\} 
2983:   \label{f1f2f3}\end{eqnarray}
2984: 
2985: The solution to this equation is given by,
2986: \begin{equation}
2987:   \label{sola2hd}
2988:   a_{2}(r)=Ce^{\mathcal{F}}-e^{\mathcal{F}}\int e^{-\mathcal{F}}\frac{f_{3}}{f_{1}}dr
2989: \end{equation}
2990: where $\mathcal{F}=-\int\frac{f_{2}}{f_{1}}dr$.  It is helpful to note
2991: that $e^{\mathcal{F}}= {1\over r^{(d-3)}}$ and,
2992: $\frac{e^{-\mathcal{F}}}{f_{1}}={r^{d-4} \over (d-2)}$.
2993:  
2994: Now the first term in eq.(\ref{sola2hd}), proportional to $C$, blows
2995: up at the horizon. We will omit some details but it is easy to see
2996: that the second term in eq.(\ref{sola2hd}) goes to zero. Thus for a
2997: non-singular solution we must set $C=0$.  One can then extract the
2998: leading behaviour near the horizon of $a_2$ from eq.(\ref{sola2hd}),
2999: however it is slightly more convenient to use eq.(\ref{leqhd}) for
3000: this purpose instead. From the behaviour of the scalar perturbation
3001: $\phi_1$, and metric perturbation, $b_2$, in the vicinity of the
3002: horizon, as discussed in the section on attractors in higher
3003: dimensions, it is easy to see that
3004: \begin{equation}
3005:   a_{2}(r)=A_{2}(r^{d-3}-r_{H}^{d-3})^{2\gamma+2}
3006: \end{equation}
3007: where, $A_2$ is an appropriately determined constant.
3008: % \begin{equation}
3009: %A_{2}=\frac{-r_{H}^{2d-4}B_{2}(t+1)(2t+1)-(d-3)r_{1}}{2r_{1}+2(3d-8)tr_{1}^{2}}\end{equation}
3010: Thus we see that the non-singular solution in the vicinity of the
3011: horizon vanishes like $(r-r_H)^{(2\gamma+2)}$ and the double-zero
3012: nature of the horizon persists after including back-reaction to this
3013: order.
3014: 
3015: Finally, expanding eq.(\ref{sola2hd}) near $r\rightarrow \infty$ (with
3016: $C=0$) we get that $a_2 \rightarrow {\rm Const} +
3017: {\mathcal{O}}(1/r^{d-3})$.  The value of the constant term is related
3018: to the coefficient in the linear term for $b_2$ at large $r$ in a
3019: manner consistent with asymptotic flatness.
3020: 
3021: 
3022: In summary we have established here that the metric perturbation $a_2$
3023: vanishes fast enough at the horizon so that the black hole continues
3024: to have a double-zero horizon, and it goes to a constant at infinity
3025: so that the black hole continues to be asymptotically flat.
3026: 
3027: \section{More Details on Asymptotic AdS Space} \label{sec:ads}
3028: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3029: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3030: \renewcommand{\theequation}{D.\arabic{equation}}
3031: \setcounter{equation}{0}
3032: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3033: We begin by considering the asymptotic behaviour at large $r$ of
3034: $\phi_1$, eq.(\ref{seads}).  One can show that this is given by
3035: \begin{equation}
3036:   \label{asads}
3037:   \phi_{1}(r)\rightarrow 
3038:   c_{+}{\frac{1}{r^{3/2}}}I_{3/4}\left(\frac{\kkappa L}{2r^{2}}\right)
3039:   +c_{-}{\frac{1}{r^{3/2}}}I_{-3/4}\left(\frac{\kkappa L}{2r^{2}}\right)
3040: \end{equation}
3041: 
3042: Here $I_{3/4}$ stands for a modified Bessel function
3043: \footnote{Modified Bessel function $I_{\nu}(r),K_{\nu}(r)$ does
3044:   satisfy following differential eq.
3045:   \begin{equation}
3046:     z^{2}I_{\nu}''(z)+zI_{\nu}'(z)-(z^{2}+\nu^{2})I(z)=0.
3047:   \end{equation}}
3048: Asymptotically, $I_\nu \propto r^{-2\nu}$. Thus $\phi_r$ has  two solutions 
3049: which go asymptotically to a constant and as $1/r^3$ respectively. 
3050: 
3051: Next, we consider values of r, $r_H<r<\infty$.  These are all ordinary
3052: points of the differential equation eq.(\ref{seads}).  Thus the
3053: solution we are interested is well-behaved at these points.  For a
3054: differential equation of the form,
3055: \begin{equation}
3056:   \label{deord}
3057:   \mathcal{L}(\psi)=\frac{d^{2}\psi}{dz^{2}}+p(z)\frac{d\psi}{dz}+q(z)\psi=0,
3058: \end{equation}
3059: all values of $z$ where $p(z),q(z)$ are analytic are ordinary points.
3060: About any ordinary point the solutions to the equation can be expanded
3061: in a power series, with a radius of convergence determined by the
3062: nearest singular point \cite{MF}.
3063:  
3064: We turn now to discussing the solution for $a_2$.  The constraint
3065: eq.(\ref{constraintads}) takes the form,
3066: \begin{equation}
3067:   2a_{0}^{2}b_{2}'+a_{2}+ (a_0^2)'(r b_2)'+ra_{2}'
3068:   =\frac{-1}{r^{2}}\kkappa^{2}\phi_{1}^{2}+a_{0}^{2}
3069:   r^{2}(\partial_{r}\phi_{1})^{2}+{2 b_2 \over r^3}(r_H^2+{2r_H^4 \over L^2}) +\frac{6rb_{2}}{L^2}
3070: \end{equation}
3071: 
3072: The solution to this equation is given by,
3073: \begin{equation}
3074:   \label{sola2ads}
3075:   a_{2}(r)=\frac{c_{2}}{r}-\frac{1}{r}\int_{r_H} f_{3}dr
3076: \end{equation}
3077: where
3078: \begin{equation}
3079:   f_{3}=2a_{0}^{2}b_{2}'+ (a_0^2)'(rb_2)'
3080:   +\frac{1}{r^{2}}\beta^{2}\phi_{1}^{2}-a_{0}^{2}r^{2}(\partial_{r}\phi_{1})^{2}
3081:   -{2 b_2 \over r^3}(r_H^2+{2 r_H^4 \over L^2})-\frac{6rb_{2}}{L^2}
3082: \end{equation}.
3083: We have set the lower limit of integration in the second term at
3084: $r_H$.  We want a solution the preserves the double-zero structure of
3085: the horizon. This means $c_2$ must be set to zero.
3086: 
3087: To find an explicit form for $a_2$ in the near horizon region it is
3088: slightly simpler to use the equation, eq.(\ref{eq1ads}).  In the near
3089: horizon region this can easily be solved and we find the solution,
3090: \begin{equation}
3091:   \label{asha2ads}
3092:   a_2 \propto (r-r_H)^{(2 \gamma + 2)}.
3093: \end{equation}
3094: 
3095: 
3096: At asymptotic infinity one can use the integral expression,
3097: eq.(\ref{sola2ads}) (with $c_2=0$).  One finds that $f_3 \rightarrow r
3098: $ as $r \rightarrow \infty$.  Thus $a_2 \rightarrow d_2 r$.  This is
3099: consistent with the asymptotically AdS geometry.
3100: 
3101: In summary we see that that there is an attractor solution to the
3102: metric equations at second order in which the double-zero nature of
3103: the horizon and the asymptotically AdS nature of the geometry both
3104: persist.
3105: 
3106: 
3107: 
3108: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3109: \begin{thebibliography}{99}
3110: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3111: \addcontentsline{toc}{section}{References}
3112: 
3113: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3114: %\cite{Ferrara:1995ih}
3115: \bibitem{Ferrara:1995ih} S.~Ferrara, R.~Kallosh and A.~Strominger,
3116:   % ``N=2 extremal black holes,''
3117:   Phys.\ Rev.\ D {\bf 52}, R5412 (1995) [arXiv:hep-th/9508072].
3118:   %% CITATION = HEP-TH 9508072;%%
3119: 
3120: 
3121: %\cite{Cvetic:1995bj}
3122: \bibitem{Cvetic:1995bj}
3123:   M.~Cvetic and A.~A.~Tseytlin,
3124:   %``Solitonic strings and BPS saturated dyonic black holes,''
3125:   Phys.\ Rev.\ D {\bf 53}, 5619 (1996)
3126:   [Erratum-ibid.\ D {\bf 55}, 3907 (1997)] [arXiv:hep-th/9512031].
3127:   %%CITATION = HEP-TH 9512031;%%
3128: 
3129: 
3130: %\cite{Strominger:1996kf}
3131: \bibitem{Strominger:1996kf} A.~Strominger,
3132:   % ``Macroscopic Entropy of $N=2$ Extremal Black Holes,''
3133:   Phys.\ Lett.\ B {\bf 383}, 39 (1996) [arXiv:hep-th/9602111].
3134:   %% CITATION = HEP-TH 9602111;%%
3135: 
3136: 
3137: %\cite{Ferrara:1996dd}
3138: \bibitem{Ferrara:1996dd} S.~Ferrara and R.~Kallosh,
3139:   % ``Supersymmetry and Attractors,''
3140:   Phys.\ Rev.\ D {\bf 54}, 1514 (1996) [arXiv:hep-th/9602136].
3141:   %% CITATION = HEP-TH 9602136;%%
3142: 
3143: 
3144: %\cite{Ferrara:1996um}
3145: \bibitem{Ferrara:1996um} S.~Ferrara and R.~Kallosh,
3146:   % ``Universality of Supersymmetric Attractors,''
3147:   Phys.\ Rev.\ D {\bf 54}, 1525 (1996) [arXiv:hep-th/9603090].
3148:   %% CITATION = HEP-TH 9603090;%%
3149: 
3150: 
3151: 
3152: %\cite{Cvetic:1996zq}
3153: \bibitem{Cvetic:1996zq}
3154:   M.~Cvetic and C.~M.~Hull,
3155:   %``Black holes and U-duality,''
3156:   Nucl.\ Phys.\ B {\bf 480}, 296 (1996) [arXiv:hep-th/9606193].
3157:   %%CITATION = HEP-TH 9606193;%%
3158: 
3159: 
3160: %\cite{Ferrara:1997tw}
3161: \bibitem{Ferrara:1997tw} S.~Ferrara, G.~W.~Gibbons and R.~Kallosh,
3162:   % ``Black holes and critical points in moduli space,''
3163:   Nucl.\ Phys.\ B {\bf 500}, 75 (1997) [arXiv:hep-th/9702103].
3164:   %% CITATION = HEP-TH 9702103;%%
3165: 
3166: 
3167: 
3168: %\cite{Gibbons:1996af}
3169: \bibitem{Gibbons:1996af} G.~W.~Gibbons, R.~Kallosh and B.~Kol,
3170:   % ``Moduli, scalar charges, and the first law of black hole
3171:   % thermodynamics,''
3172:   Phys.\ Rev.\ Lett.\ {\bf 77}, 4992 (1996) [arXiv:hep-th/9607108].
3173:   %% CITATION = HEP-TH 9607108;%%
3174: 
3175: 
3176: %\cite{Denefa}
3177: \bibitem{Denefa} F.~Denef,
3178:   % ``Supergravity flows and D-brane stability,''
3179:   JHEP {\bf 0008}, 050 (2000) [arXiv:hep-th/0005049].
3180:   %% CITATION = HEP-TH 0005049;%%
3181: 
3182: 
3183: %\cite{Denef:2001xn}
3184: \bibitem{Denef:2001xn} F.~Denef, B.~R.~Greene and M.~Raugas,
3185:   % ``Split attractor flows and the spectrum of BPS D-branes on the
3186:   % quintic,''
3187:   JHEP {\bf 0105}, 012 (2001) [arXiv:hep-th/0101135].
3188:   %% CITATION = HEP-TH 0101135;%%
3189: 
3190: 
3191: 
3192: 
3193: %\cite{Ooguri:2004zv}
3194: \bibitem{Ooguri:2004zv} H.~Ooguri, A.~Strominger and C.~Vafa,
3195:   % ``Black hole attractors and the topological string,''
3196:   Phys.\ Rev.\ D {\bf 70}, 106007 (2004) [arXiv:hep-th/0405146].
3197:   %% CITATION = HEP-TH 0405146;%%
3198: 
3199: 
3200: %\cite{LopesCardoso:1998wt}
3201: \bibitem{LopesCardoso:1998wt} G.~Lopes Cardoso, B.~de Wit and
3202:   T.~Mohaupt,
3203:   % ``Corrections to macroscopic supersymmetric black-hole entropy,''
3204:   Phys.\ Lett.\ B {\bf 451}, 309 (1999) [arXiv:hep-th/9812082].
3205:   %% CITATION = HEP-TH 9812082;%%
3206: 
3207: 
3208: 
3209: 
3210: %\cite{Dabholkar:2004yr}
3211: \bibitem{Dabholkar:2004yr} A.~Dabholkar,
3212:   % ``Exact counting of black hole microstates,''
3213:   [arXiv:hep-th/0409148].
3214:   %% CITATION = HEP-TH 0409148;%%
3215: 
3216: 
3217: 
3218: 
3219: %\cite{Ooguri:2005vr}
3220: \bibitem{Ooguri:2005vr} H.~Ooguri, C.~Vafa and E.~P.~Verlinde,
3221:   % ``Hartle-Hawking wave-function for flux compactifications,''
3222:   [arXiv:hep-th/0502211].
3223:   %% CITATION = HEP-TH 0502211;%%
3224: 
3225: 
3226: %\cite{Dijkgraaf:2005bp}
3227: \bibitem{Dijkgraaf:2005bp} R.~Dijkgraaf, R.~Gopakumar, H.~Ooguri and
3228:   C.~Vafa,
3229:   % ``Baby universes in string theory,''
3230:   [arXiv:hep-th/0504221].
3231:   %% CITATION = HEP-TH 0504221;%%
3232: 
3233: 
3234: %\cite{KKLT}
3235: \bibitem{KKLT} S.~Kachru, R.~Kallosh, A.~Linde and S.~P.~Trivedi,
3236:   % ``De Sitter vacua in string theory,''
3237:   Phys.\ Rev.\ D {\bf 68}, 046005 (2003) [arXiv:hep-th/0301240].
3238:   %% CITATION = HEP-TH 0301240;%%
3239: 
3240: 
3241: %\cite{TT3}
3242: \bibitem{TT3}
3243:  P.~K.~Tripathy and S.~P.~Trivedi,
3244:   %``Non-Supersymmetric Attractors in String Theory,''
3245:  [arXiv:hep-th/0511117].
3246:   %%CITATION = HEP-TH 0511117;%%
3247: 
3248: 
3249: 
3250: %\cite{Sen:2005kj}
3251: \bibitem{Sen:2005kj}
3252:   A.~Sen,
3253:   %``Stretching the horizon of a higher dimensional small black hole,''
3254:   JHEP {\bf 0507}, 073 (2005)
3255:   [arXiv:hep-th/0505122].
3256:   %%CITATION = HEP-TH 0505122;%%
3257: 
3258: 
3259: %\cite{Sen:2005wa}
3260: \bibitem{Sen:2005wa}
3261:   A.~Sen,
3262:   %``Black hole entropy function and the attractor mechanism in higher
3263:   %derivative gravity,''
3264:   JHEP {\bf 0509}, 038 (2005)
3265:   [arXiv:hep-th/0506177].
3266:   %%CITATION = HEP-TH 0506177;%%
3267: 
3268: 
3269: %\cite{Kraus:2005vz}
3270: \bibitem{Kraus:2005vz} P.~Kraus and F.~Larsen,
3271:   % ``Microscopic black hole entropy in theories with higher
3272:   % derivatives,''
3273:   [arXiv:hep-th/0506176].
3274:   %% CITATION = HEP-TH 0506176;%%
3275: 
3276: 
3277: %\cite{IJ}
3278: \bibitem{IJ}
3279: Norihiro Iizuka and  Rudra P. Jena,  In Preparation.
3280: 
3281: 
3282: %\cite{MF}
3283: \bibitem{MF} P. M. Morse and H. Feshbach, ``Methods of Theoretical
3284:   Physics'' (McGraw-Hill Book Company, New York, 1953), Chapter 5, page 530-532.
3285: 
3286: 
3287: %\cite{Gibbons:1987ps}
3288: \bibitem{Gibbons:1987ps} G.~W.~Gibbons and K.~i.~Maeda,
3289:   % ``Black Holes And Membranes In Higher Dimensional Theories With
3290:   % Dilaton Fields,''
3291:   Nucl.\ Phys.\ B {\bf 298}, 741 (1988).
3292:   %% CITATION = NUPHA,B298,741;%%
3293: 
3294: 
3295: %\cite{Kallosh:1992ii}
3296: \bibitem{Kallosh:1992ii} R.~Kallosh, A.~D.~Linde, T.~Ortin, A.~W.~Peet
3297:   and A.~Van Proeyen,
3298:   % ``Supersymmetry as a cosmic censor,''
3299:   Phys.\ Rev.\ D {\bf 46}, 5278 (1992) [arXiv:hep-th/9205027].
3300:   %% CITATION = HEP-TH 9205027;%%
3301: 
3302: 
3303: %\cite{Gibbons:1994ff}
3304: \bibitem{Gibbons:1994ff} G.~W.~Gibbons and R.~E.~Kallosh,
3305:   % ``Topology, entropy and Witten index of dilaton black holes,''
3306:   Phys.\ Rev.\ D {\bf 51}, 2839 (1995) [arXiv:hep-th/9407118].
3307: 
3308: 
3309: %\cite{Dobiasch:1981vh}
3310: \bibitem{Dobiasch:1981vh} P.~Dobiasch and D.~Maison,
3311:   % ``Stationary, Spherically Symmetric Solutions Of Jordan's Unified
3312:   % Theory Of Gravity And Electromagnetism,''
3313:   Gen.\ Rel.\ Grav.\ {\bf 14}, 231 (1982).
3314:   %% CITATION = GRGVA,14,231;%%
3315: 
3316: 
3317: %\cite{Gibbons:1993xt}
3318: \bibitem{Gibbons:1993xt} G.~W.~Gibbons, D.~Kastor, L.~A.~J.~London,
3319:   P.~K.~Townsend and J.~H.~Traschen,
3320:   % ``Supersymmetric selfgravitating solitons,''
3321:   Nucl.\ Phys.\ B {\bf 416}, 850 (1994) [arXiv:hep-th/9310118].
3322:   %% CITATION = HEP-TH 9310118;%%
3323: 
3324: 
3325: 
3326: 
3327: 
3328: %\cite{Chamblin:1999tk}
3329: \bibitem{Chamblin:1999tk} A.~Chamblin, R.~Emparan, C.~V.~Johnson and
3330:   R.~C.~Myers,
3331:   % ``Charged AdS black holes and catastrophic holography,''
3332:   Phys.\ Rev.\ D {\bf 60}, 064018 (1999) [arXiv:hep-th/9902170];
3333:   A.~Chamblin, R.~Emparan, C.~V.~Johnson and R.~C.~Myers,
3334:   % ``Holography, thermodynamics and fluctuations of charged AdS black
3335:   % holes,''
3336:   Phys.\ Rev.\ D {\bf 60}, 104026 (1999) [arXiv:hep-th/9904197];
3337:   S.~W.~Hawking and H.~S.~Reall,
3338:   % ``Charged and rotating AdS black holes and their CFT duals,''
3339:   Phys.\ Rev.\ D {\bf 61}, 024014 (2000) [arXiv:hep-th/9908109].
3340:   %% CITATION = HEP-TH 9908109;%%
3341: 
3342: 
3343: %\cite{Nohair}
3344: \bibitem{Nohair} J. D. Bekenstein, %``Nonexistence of Baryon Number
3345:   % for Static Black Holes,''
3346:   Phys.\ Rev.\ D {\bf 5}, 1239-1246 (1972); J. B. Hartle. ``Can a
3347:   Schwarzschild Black Hole Exert Long-Range Neutrino Forces?''  in
3348:   Magic without Magic, ed. J. Kaluder (Freeman, San Francisco,1972); C.
3349:   Teitelboim, %``Nonmeasurability of the Quantum Numbers of a Black
3350:   % Hole,''
3351:   Phys.\ Rev.\ D {\bf 5}, 2941-2954 (1972).
3352: 
3353: 
3354: %\cite{Gibbons:1985ac}
3355: \bibitem{Gibbons:1985ac} G.~W.~Gibbons and D.~L.~Wiltshire, %``Black
3356:   % Holes In Kaluza-Klein Theory,''
3357:   Annals Phys.\ {\bf 167}, 201 (1986) [Erratum-ibid.\ {\bf 176}, 393
3358:   (1987)].
3359:   %% CITATION = APNYA,167,201;%%
3360: 
3361: 
3362: 
3363: %\cite{Masood-ul-Alam:1993ea}
3364: \bibitem{Masood-ul-Alam:1993ea}
3365:   A.~K.~M.~Masood-ul-Alam,
3366:   %``Uniqueness of a static charged dilaton black hole,''
3367:   Class.\ Quant.\ Grav.\  {\bf 10}, 2649 (1993).
3368: 
3369: 
3370: %\cite{Mars:2001pz}
3371: \bibitem{Mars:2001pz}
3372:   M.~Mars and W.~Simon,
3373:   %``On uniqueness of static Einstein-Maxwell-dilaton black holes,''
3374:   Adv.\ Theor.\ Math.\ Phys.\  {\bf 6}, 279 (2003)
3375:   [arXiv:gr-qc/0105023].
3376:   %%CITATION = GR-QC 0105023;%%
3377: 
3378: %\cite{Gibbons:2002ju}
3379: \bibitem{Gibbons:2002ju}
3380:   G.~W.~Gibbons, D.~Ida and T.~Shiromizu,
3381:   %``Uniqueness of (dilatonic) charged black holes and black p-branes in  higher
3382:   %dimensions,''
3383:   Phys.\ Rev.\ D {\bf 66}, 044010 (2002)
3384:   [arXiv:hep-th/0206136].
3385:   %%CITATION = HEP-TH 0206136;%%
3386: 
3387: %\cite{Gibbons:2002av}
3388: \bibitem{Gibbons:2002av}
3389:   G.~W.~Gibbons, D.~Ida and T.~Shiromizu,
3390:   %``Uniqueness and non-uniqueness of static black holes in higher
3391:   %dimensions,''
3392:   Phys.\ Rev.\ Lett.\  {\bf 89}, 041101 (2002)
3393:   [arXiv:hep-th/0206049].
3394:   %%CITATION = HEP-TH 0206049;%%
3395: 
3396: %\cite{exds}
3397: \bibitem{exds} K. Lake and R. C. Roeder, Phys.\ Rev.\ D {\bf 15},
3398:   3513 (1977); K. H. Geyer, Astron. Nachr. {\bf 301}, 135 (1980); J.
3399:   Podolosky, %``The Structure of the Extreme Schwarzschild-deSitter
3400:   % Space-time,  
3401:   Gen.\ Rel.\ Grav.\  {\bf 31}, 1703 (1999)
3402:   [arXiv:gr-qc/9910029]
3403: 
3404: 
3405: %\cite{nariai}
3406: \bibitem{nariai} H. Nariai, Sci. Rep. Res. Inst. Tohuku Univ. A {\bf
3407:     35}, 62 (1951).
3408: 
3409: 
3410: %\cite{Toda}
3411: \bibitem{Toda} M. Toda, ``Theory of Nonlinear Lattices (2nd Ed.),''
3412:   (Springer-Verlang, Berlin, 1988).
3413: 
3414: 
3415: %\cite{Lu:1996jr}
3416: \bibitem{Lu:1996jr} H.~Lu and C.~N.~Pope, %``SL(N+1,R) Toda solitons
3417:   % in supergravities,''
3418:   Int.\ J.\ Mod.\ Phys.\ A {\bf 12}, 2061 (1997)
3419:   [arXiv:hep-th/9607027].
3420:   %% CITATION = HEP-TH 9607027;%%
3421: 
3422: 
3423: 
3424: \end{thebibliography}
3425: \end{document}
3426: 
3427: