hep-ph0603026/dcs.tex
1: \documentclass[aps,prd,showpacs]{revtex4}
2: \textheight=25cm 
3: \input epsf.sty
4: \usepackage[dvips]{graphicx}
5: \def\beq{\begin{eqnarray}}
6: \def\eeq{\end{eqnarray}}
7: 
8: 
9: %\textwidth=17cm 
10: \begin{document} 
11: 
12: \newcommand{\ep}{\epsilon}
13: \newcommand{\fr}{\frac}
14: \newcommand{\reals}{\mbox{${\rm I\!R }$}}
15: \newcommand{\nats}{\mbox{${\rm I\!N }$}}
16: \newcommand{\intgs}{\mbox{${\rm Z\!\!Z }$}}
17: \newcommand{\cam}{{\cal M}}
18: \newcommand{\caz}{{\cal Z}}
19: \newcommand{\cao}{{\cal O}}
20: \newcommand{\cac}{{\cal C}}
21: \newcommand{\aaa}{\int\limits_{mR}^{\infty}dk\,\,}
22: \newcommand{\bbb}{\left[\left(\frac k R\right)^2-m^2\right]^{-s}}
23: \newcommand{\ccc}{\frac{\partial}{\partial k}}
24: \newcommand{\fff}{\frac{\partial}{\partial z}}
25: \newcommand{\iikma}{\aaa \bbb \ccc}
26: \newcommand{\ddd}{\int\limits_{mR/\nu}^{\infty}dz\,\,}
27: \newcommand{\eee}{\left[\left(\frac{z\nu} R\right)^2-m^2\right]^{-s}}
28: \newcommand{\lll}{\frac{(-1)^j}{j!}}
29: \newcommand{\iinma}{\ddd\eee\fff}
30: \newcommand{\cah}{{\cal H}}
31: \newcommand{\nn}{\nonumber}
32: \renewcommand{\theequation}{\mbox{\arabic{section}.\arabic{equation}}}
33: \newcommand{\komplex}{\mbox{${\rm I\!\!\!C }$}}
34: \newcommand{\sip}{\frac{\sin (\pi s)}{\pi}}
35: \newcommand{\numr}{\left(\frac{\nu}{mR}\right)^2}
36: \newcommand{\mzs}{m^{-2s}}
37: \newcommand{\rzs}{R^{2s}}
38: \newcommand{\abl}{\partial}
39: \newcommand{\g}{\Gamma\left(}
40: \newcommand{\ikma}{\int\limits_{\gamma}\frac{dk}{2\pi i}\,\,(k^2+m^2)^{-s}\frac{\partial}{\partial k}}
41: \newcommand{\ead}{e_{\alpha}(D)}
42: \newcommand{\sual}{\sum_{\alpha =1}^{D-2}}
43: \newcommand{\sulnu}{\sum_{l=0}^{\infty}}
44: \newcommand{\sujnu}{\sum_{j=0}^{\infty}}
45: \newcommand{\suani}{\sum_{a=0}^i}
46: \newcommand{\suanzi}{\sum_{a=0}^{2i}}
47: \newcommand{\zend}{\zeta_D^{\nu}}
48: \newcommand{\amed}{A_{-1}^{\nu ,D}(s)}
49: \renewcommand{\and}{A_{0}^{\nu ,D}(s)}
50: \newcommand{\aid}{A_{i}^{\nu ,D}(s)}
51: \newcommand{\res}{{\rm res}\,\,\,}
52: \newcommand{\sn}{\frac{\sin \pi s}{\pi}}
53: \newcommand{\ha}{\frac{3}{2}}
54: \newcommand{\smj}{\sum_{j=\ha}^\infty d(j)}
55: \newcommand{\smN}{\sum_{k=1}^N}
56: \newcommand{\smk}{\sum_{k=0}^\infty \frac{(-1)^k}{k!}}
57: \newcommand{\sma}{\sum_{a=0}^{2i} x_{i,a}}
58: \newcommand{\itz}{\int_{\frac{mR}j}^\infty dz}
59: \newcommand{\paran}{\left[\left(\frac{zj}R \right)^2-m^2\right]}
60: \newcommand{\dz}{\frac{\partial }{\partial z}}
61: \newcommand{\bee}{\begin{equation}}
62: \newcommand{\bea}{\begin{eqnarray}}
63: \newcommand{\eea}{\end{eqnarray}}
64: \newcommand{\app}[1]{\section{#1}\renewcommand{\theequation}
65:         {\mbox{\Alph{section}.\arabic{equation}}}\setcounter{equation}{0}}
66: \def\beq{\begin{eqnarray}}
67: \def\eeq{\end{eqnarray}}
68: 
69: %NOTAZIONI KIRSTEN REVIEW
70: 
71: \newcommand{\De}{\Delta}
72: \newcommand{\Om}{\Omega}
73: \newcommand{\Si}{\Sigma}
74: \newcommand{\G}{\Gamma}
75: \newcommand{\Ga}{\Gamma}
76: \newcommand{\Gb}{\Gamma^b}
77: \newcommand{\La}{\Lambda}
78: \newcommand{\Th}{\Theta}
79: 
80: 
81: \hspace{-10mm} 
82: %\leftline{\epsfbox{mark.eps}} 
83: \vspace*{-10.0mm} % for revtex 
84: \thispagestyle{empty} 
85: {\baselineskip-4pt 
86: \font\yitp=cmmib10 scaled\magstep2 
87: \font\elevenmib=cmmib10 scaled\magstep1  \skewchar\elevenmib='177 
88: 
89: \vskip20mm 
90: \begin{center}{\large \bf 
91: Neutrino Dark Energy and Moduli Stabilization in a BPS Braneworld Scenario}
92: \end{center} 
93: \vspace*{4mm} 
94: \centerline{Andrea Zanzi} 
95: \vspace*{2mm}
96: 
97: 
98: \centerline{{\it Dipartimento di Fisica, Universit\`a di Padova,  via Marzolo 8, I-35131, Padova, Italy}}
99: %\centerline{{\it and}}
100: \centerline{{\it INFN, Sezione di Padova, via Marzolo 8, I-35131, Padova, Italy}}
101: 
102: \centerline{\rm e--mail: zanzi@pd.infn.it}
103: 
104: \vspace*{4mm} 
105: 
106: 
107: \begin{abstract}
108: 
109: A braneworld model for neutrino Dark Energy (DE) is presented. We consider a five dimensional two-branes set up
110: with a bulk scalar field motivated by supergravity. Its low-energy effective theory is derived with a moduli space approximation
111: (MSA). The position of the two branes are parameterized by two scalar degrees of freedom (moduli). After detuning
112: the brane tensions a classical potential for the moduli is generated. This potential is unstable for $dS_4$ branes
113: and we suggest to consider as a stabilizing contribution the Casimir energy of bulk fields. In particular we 
114: add a massive spinor (neutrino) field in the bulk and then evaluate the Casimir contribution
115: of the bulk neutrino with the help of zeta function regularization techniques. 
116: We construct an explicit form of the 4D neutrino mass as function of the two moduli. To recover the correct 
117: DE scale for the moduli potential the usual cosmological constant fine-tuning
118: is necessary, but, once accepted, 
119: this model suggests a stronger connection 
120: between DE and neutrino physics.
121: 
122: 
123: \vspace*{8mm} 
124: \hspace{111mm} DFPD/06/TH/04
125: 
126: \end{abstract}
127: \pacs{04.50.+h, 13.15.+g, 11.25.-w} 
128: \maketitle
129: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
130: \section{Introduction}
131: 
132: There is an increasing observational evidence that our Universe is in a state of accelerated expansion \cite{1}.
133: The mixing of data coming from CMB, supernovae Ia and abundance measurements illustrate a surprising 
134: picture of the cosmic energy budget: the dominant energy component (dark energy, DE, for reviews see \cite{quint-review}) has negative enough pressure and its 
135: contribution is comparable, at present, with that of the pressureless dark matter component (DM), although 
136: their dynamical evolution 
137: during the cosmological history can be totally different ("coincidence problem"). To remove, at least 
138: partially, this problem, a direct interaction between DM and DE has been discussed
139: \cite{4,5,5a,5b,amen} and the "old" idea of varying-mass particles \cite{vamps} has been taken into account again. 
140: These models are also well-motivated from theoretical point of view by the identification
141: of the DE with the string-theory dilaton (\cite{6,PT,7}, but also \cite{gaspstab} for DE as a stabilized dilaton). These 
142: so-called "coupled DE models"  can show 
143: a late-time attractor solution, i.e. the late-time cosmology is insensitive to the initial conditions for DE.
144: For a phenomenologically acceptable scenario, in these models the present coupling of the 
145: scalar field with baryons must be sufficiently weak 
146: while the couplings with DM must be sufficiently strong (baryo-phobic quintessence). This is a very general 
147: problem: allowing a direct interaction between matter and a quintessence (ultra-light $m_{\phi} \sim 10^{-33}$eV)
148: field could be phenomenologically dangerous (violations of the equivalence principle, time dependence of 
149: couplings - for reviews see \cite{uzan}). 
150: 
151: A possible way-out of these problems is guaranteed by 
152: scalar-tensor (ST) theories of gravitation \cite{bd}, where by construction matter has a
153: purely metric coupling with gravity (for reviews see \cite{maeda},\cite{damour},\cite{will}). The existence of an attraction
154: mechanism of ST towards Einstein's General Relativity (GR) has been discussed in \cite{dam3a, dam3b, max}. Simultaneously 
155: the presence of a second attractor is possible (\cite{max}) and the correct evolution of $\rho_{DE}$ along a so-called `tracking'
156: \cite{tracker} solution is rescued. However deviations from standard cosmology and General Relativity
157: may be present at nucleosynthesis, CMB and solar system tests of gravity \cite{catena,caffe2}. Another possible way-out is to 
158: consider "chameleon fields" (\cite{justin}): the mass of the scalar fields is determined by the density of the environment. On cosmological distances,
159: where the densities are very tiny, the mass can be of the order of the Hubble constant and the fields can roll on cosmological time scales.
160: On the Earth instead, the density is much higher and the field acquires a large enough mass to evade experimental bounds mantaining relevant coupling with matter. 
161: 
162: Recently Fardon, Nelson and Weiner (FNW) exploited the notion of "tracking" in a completely new way (\cite{fnw}): they 
163: suggested that DE tracks neutrinos. They established a (not so!) firm connection between the 
164: dark energy scale and the neutrino mass scale: 
165: the similarity between the two scales becomes a true relation.
166: The authors imagined furthermore that DE and neutrinos are directly coupled: a new dark sector arises, taking contributions
167: from a scalar field $\phi$ (a sort of quintessence) and neutrinos. In this scenario the dark energy density depends 
168: on the neutrino masses, which are not 
169: fixed but varying as functions of the neutrino density. The coupling between DE and neutrinos is encoded in the variable neutrino
170:  mass $m_{\nu}(\phi)$. During the cosmological expansion the temperature $T$ decreases. 
171: When $T< m_{\nu}$ neutrinos become non-relativistic and their contribution to the energy density becomes mass-dependent.
172: In this way the coupling $\phi-\nu$ brings to an effective potential of the form:
173: \begin{equation}\label{prima}
174:  V_{eff}=V_{0}(\phi)+n_{\nu} m_{\nu}(\phi),
175: \end{equation}
176: where $V_0$ is the scalar field potential and $n_{\nu}$ is the neutrino number density. If we assume that $V_0$ is minimized for large $m_{\nu}$,
177: a competition between the terms in (\ref{prima}) is created, a minimum in $V_{eff}$ arises and the neutrino mass is determined 
178: {\it dynamically} during
179: the cosmological expansion. For works on mass varying neutrinos see also \cite{peccei}, \cite{neut}.
180: 
181: Several points are unsatisfactory about this approach (see also \cite{peccei}): 
182: \begin{itemize}
183:  \item The FNW model is characterized by $V_{0}(\phi)$ (the potential for the scalar field) and $m(\phi)$ (the neutrino mass).
184:  The explicit form of these functions is inserted by-hand. It would be interesting to obtain a firmer 
185: theoretical derivation of these functions in a "top-down" scenario.
186:  \item The connection between dark energy and neutrino mass is not well established unless the scalar field gets an appropriate
187: value. 
188: \end{itemize}
189: As we will show, an extradimensional scenario addresses 
190: both drawbacks of the FNW model: \\
191: 1) a 4D neutrino mass as a function of two scalar 
192: degrees of freedom (moduli) and the potential for these scalar fields are not inserted by hand but rather explicitly
193: calculated.\\
194: 2) A firmer relationship is established between neutrino mass scale and Dark Energy scale.
195: 
196: It must be stressed that solving these problems led us to a model that is significantly different:
197: the only dynamic that we contemplate for the scalar fields is a chameleontic behaviour for one of the two moduli (the radion). 
198: Then, we have to distinguish between different distance scales: on cosmological distances, the densities are tiny and the 
199: radion can roll on cosmological time scales (the neutrino mass is {\it variable} with time and possible deviations
200: from a pure cosmological constant contribution can be observed in the DE sector); on smaller scales instead, 
201: the chameleon acquires a large mass (radion stabilization)
202: and {\it both} moduli are simply sitting on the minimum of the potential (for a discussion of possible manifestations 
203: of DE at short scales
204: see \cite{caffe}).
205: As a final result, the connection between DE and neutrino mass will be established in the form
206: \beq
207: (\frac{V^{1/4}}{m_{\nu}})_{min}\sim 1,
208: \eeq
209: where $V$ is the total moduli potential. 
210: 
211: Before getting into the detailed presentation of the model, 
212: here is a summary of our approach. We consider 
213: a supersymmetric extension of Randall-Sundrum two branes model (\cite{RS1,RS2}, while 
214: for reviews on extra-dimensions see \cite{extra}).
215: This kind of approach finds its theoretical firmer justification in 
216: string theory (see \cite{HW,witten}).  
217: Our general set-up will follow \cite{brax,brax1}: a five-dimensional braneworld model 
218: with two boundary branes and a scalar field in the bulk motivated
219: by supergravity.
220: As  we will see in the forecoming sections, the most
221: distinguishing feature of this (BPS) class of models is that the brane and bulk dynamics are related to each 
222: other and under control. Once we have solved bulk equations (Einstein equations and Klein-Gordon equation) the 
223: boundary conditions do not
224: give any additional information, so we can put the branes without any obstacle into the background. The positions of the 
225: two branes are specified by two independent constants (no-force condition; for a discussion on BPS condition and connections between 
226: brane physics and black-hole physics see  \cite{gionni}). 
227: We will discuss than the moduli space approximation to derive 
228: the low energy action and a bi-scalar-tensor theory is recovered, 
229: with the two scalar fields $\lambda$ and $\phi$ (moduli) parametrizing the positions of the branes: 
230: the two constants are promoted to scalar degrees
231: of freedom. To render this model more realistic, 
232: we can leave the BPS configuration in two different ways: i) putting matter on the branes; ii) perturbing 
233: the brane tensions (i.e. parametrizing the supersymmetry breaking in a phenomenological way). It has been 
234: shown (\cite{gonz}) that it is 
235: possible to stabilize both moduli putting matter on branes without perturbing the brane tensions. However, since we want 
236: to describe a Universe with a positive cosmological constant, the detuning of the brane tensions is unavoidable and a 
237: classical potential for the moduli is generated (\cite{baggerredi, marsiglia}).
238: Unfortunately the potential is unstable and, to stabilize the system, we suggest to consider as a stabilization
239: mechanism the quantum (Casimir) contribution 
240: from bulk fields. We will then add a massive spinor (neutrino) in the bulk and we will evaluate the Casimir 
241: energy using zeta function regularization techniques (for reviews on Casimir energy see \cite{casimir}, \cite{kirsten};
242: for zeta function techniques see \cite{TEN}, \cite{kirsten}, \cite{MOSS2}). Remarkably the potential
243: $V=V(\lambda,\phi)$ is not inserted by-hand in our model and the geometrical configuration of the system is stabilized 
244: (a long-standing problem in extra-dimensional 
245: and string-inspired scenarios, see for example \cite{buchbinder} and
246: references therein for a general treatment, but also \cite{gonz, webster} for moduli stabilization in BPS braneworlds).
247: 
248: As far as the 4d neutrino mass is concerned, following \cite{grossman}, we will obtain a small 4d neutrino mass
249: without recurring to a see-saw mechanism.
250: 
251:  
252: 
253: About the organization of this article, in section 2 we present the action and the moduli space approximation; in section 3 we study
254: the chameleon mechanism for the radion; in section 4 the explicit form of the neutrino mass is derived as function of the moduli;
255: in section 5 the moduli potential is analyzed. In section 6 we will discuss the parameters of the model and the moduli stabilization. In section 7 we will draw some concluding remarks. The appendix contains some technical details.
256: 
257: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
258: \section{The model}
259: 
260: The bulk action is the sum of two main contributions: the 
261: first one (SUGRA) is motivated by supergravity and follows \cite{brax, brax1}, the
262: second one (SPINOR) is the bulk-neutrino part. We have
263: 
264: \begin{equation}
265: S_{\rm bulk} = S_{\rm SUGRA} + S_{\rm SPINOR}
266: \end{equation}
267: 
268: where the SUGRA action consists of two terms which describe 
269: gravity and the bulk scalar field ($C$) dynamics:
270: \begin{equation}
271: S_{\rm SUGRA} = \frac{1}{2\kappa_5^2} \int d^5 x \sqrt{-g_5}
272: \left( {\cal R} - \frac{3}{4}\left( (\partial C)^2 + U \right)\right);
273: \end{equation}
274: 
275: while the SPINOR action, which takes into account a Dirac bulk fermion with mass $m$ 
276: of order the fundamental scale $M$, is written in the form: 
277: 
278: \begin{equation}
279: \label{action}
280:    S = \int\!\mbox{d}^4x\!\int\!\mbox{d}z\,\sqrt{-g_5}
281:    \left\{ E_a^A \left[ \frac{i}{2}\,\bar\Psi\gamma^a
282:    (\partial_A-\overleftarrow{\partial_A})\Psi
283:    + \frac{\omega_{bcA}}{8}\,\bar\Psi \{\gamma^a,\sigma^{bc}\} \Psi
284:    \right] - m\,\mbox{sgn}(\alpha)\,\bar\Psi\Psi \right\} \,.
285: \end{equation}
286: We use capital
287: indices $A,B,\dots$ for objects defined in curved space, and 
288: lower-case indices $a,b,\dots$ for objects defined in the tangent 
289: frame. The matrices $\gamma^a=(\gamma^\mu,i\gamma_5)$ provide a 
290: four-dimensional representation of the Dirac matrices in 
291: five-dimensional flat space. The quantity
292: $E_a^A$ is the
293: inverse vielbein and $\omega_{bcA}$ is the spin connection. Because 
294: in our case the metric is diagonal the connection gives no contribution to
295: the action (2.5).
296: 
297: The sign change of the mass term under $\alpha\to-\alpha$ is necessary in 
298: order to conserve $\alpha$-parity, as required by the $Z_2$ orbifold 
299: symmetry we impose.
300: 
301: 
302: Further, our setup contains two branes. One of these branes has a
303: positive tension, the other brane has a negative tension.  They are
304: described by the action
305: \begin{eqnarray}
306: S_{\rm brane 1} &=& -\frac{3}{2\kappa_5^2}\int d^5x \sqrt{-g_5} U_B
307: \delta(z_1), \label{b1} \\
308: S_{\rm brane 2} &=& +\frac{3}{2\kappa_5^2}\int d^5x \sqrt{-g_5} U_B \delta(z_2) \label{b2}.
309: \end{eqnarray}
310: In these expressions, $z_1$ and $z_2$ are the (arbitrary) positions of
311: the two branes, $U_B$ is the superpotential; $U$, the bulk potential
312: energy of the scalar field, is given by (BPS relation)
313: \begin{equation}
314: U = \left(\frac{\partial U_B}{\partial C}\right)^2 - U_B^2.
315: \end{equation}
316: This specific configuration, when no fields other than the bulk scalar field are present and when the
317: bulk and brane potentials are unperturbed (no susy breaking effects), is the BPS configuration (\cite{gonz}). When $U_B$ 
318: is the constant potential, the Randall-Sundrum model is recovered with a bulk cosmological constant
319: $ \Lambda_5=(3/8) U=-(3/8) U_B^2$.
320: 
321: We will also include the Gibbons--Hawking boundary term for each
322: brane, which have the form
323: \begin{equation}
324: S_{\rm GH} = \frac{1}{\kappa_5^2}\int d^4 x \sqrt{-g_4} K,
325: \end{equation}
326: where $K$ is the extrinsic curvature of the individual branes.
327: 
328: 
329: The solution of the system above can be derived from 
330: BPS--like equations of the form
331: \begin{equation}
332: \frac{a'}{a}=-\frac{U_B}{4},\ C'=\frac{\partial U_B}{\partial C},
333: \end{equation}
334: where $'=d/dz$ for a metric of the form
335: \begin{equation}\label{background}
336: ds^2 = dz^2 + a^2(z)\eta_{\mu\nu}dx^\mu dx^\nu.
337: \end{equation}
338: We will particularly focus on the case where the superpotential is an 
339: exponential function:
340: \begin{equation}\label{potential}
341: U_B=4k e^{\alpha C}.
342: \end{equation}
343: The solution for the scale factor reads
344: \begin{equation}\label{scale}
345: a(z)=(1-4k\alpha^2z)^{\frac{1}{4\alpha^2}},
346: \end{equation}
347: while the scalar field solution is
348: \begin{equation}\label{psi}
349: C = -\frac{1}{\alpha}\ln\left(1-4k\alpha^2z\right).
350: \end{equation}
351: In the $\alpha\to 0$ we retrieve the AdS profile
352: \begin{equation}
353: a(z)=e^{-kz}.
354: \end{equation}
355: In that case the scalar field decouples and the singular point (for which the scale factor vanishes) 
356: at $z_* = 1/(4k\alpha^2)$ 
357: is removed (singularities in braneworld scenarios are analyzed in \cite{naked}). 
358: 
359: 
360: In the following we will discuss the moduli space approximation. Two of the moduli of the
361: system are the brane positions.  That is, in the solution above the
362: brane positions are arbitrary.  In the moduli space approximation,
363: these moduli are assumed to be space-time dependent since in connection to matter distribution on branes.  We denote the
364: position of ("visible") brane 1 with $z_1 = \phi(x^\nu)$ and the position of ("hidden")
365: brane 2 with $z_2 = \lambda(x^\mu)$. We consider the case where the
366: evolution of the brane is slow. This means that in constructing the
367: effective four--dimensional theory we neglect terms like $(\partial
368: \phi)^3$.
369: 
370: 
371: In addition to the brane positions, we need to include the graviton 
372: zero mode, which can be done by replacing $\eta_{\mu\nu}$ with a 
373: space--time dependent tensor $g_{\mu\nu}(x^\mu)$. Note that the moduli 
374: space approximation is only a good approximation
375: if the time--variation of the moduli fields is small. This should be
376: the case for late--time cosmology well after nucleosynthesis, which we 
377: are interested in.
378: 
379: \subsection{Moduli Space Approximation: the gravitational sector}
380: Let us first consider the SUGRA action. Replacing $\eta_{\mu\nu}$ 
381: with $g_{\mu\nu}(x^{\mu})$ in (\ref{background}) we have for the 
382: Ricci scalar ${\cal R} = {\cal R}^{(4)}/a^2 + \tilde {\cal R}$, where $\tilde {\cal R}$ is the 
383: Ricci--scalar of the background (\ref{background}). We explicitly use
384: the background solution (\ref{scale}) and (\ref{psi}), so that 
385: $\tilde R$ will not contribute to the low--energy effective action. 
386: Also, in this coordinate system, where the branes move, there is no 
387: contribution from the part of the bulk scalar field. Collecting 
388: everything we therefore have
389: \begin{equation}
390: S_{\rm bulk} = \frac{1}{2\kappa_5^2} \int dz d^4 x a^4
391: \sqrt{-g_4}\frac{1}{a^2}{\cal R}^{(4)} = \int d^4 x 
392: \sqrt{-g_4} f(\phi,\lambda) {\cal R}^{(4)},
393: \end{equation}
394: with 
395: \begin{equation}
396: f(\phi,\lambda) = \frac{1}{2 \kappa_5^2} \int^{\lambda}_{\phi} dz a^2 (z).
397: \end{equation}
398: We remind the reader that $a(z)$ is given by (\ref{scale}). 
399: 
400: We now turn to the boundary terms. The integrals 
401: (\ref{b1}) and (\ref{b2}) do not contribute to the effective action 
402: for the same reason that $\tilde R$ does not contribute. Let us 
403: therefore turn our attention to the Gibbons--Hawking boundary terms.
404: 
405: First, it is not difficult to construct the normal vectors to the
406: brane:
407: \begin{equation}
408: n^\mu = \frac{1}{\sqrt{1+ (\partial \phi)^2/a^2}}\left( \partial^\mu
409: \phi/a^2,1\right).
410: \end{equation}
411: 
412: Then the induced metric on each brane is given by
413: \begin{eqnarray}
414: g_{\mu\nu}^{\rm ind,1} &=& a^2(\phi) g_{\mu\nu}^4 - \partial_{\mu}\phi
415: \partial_{\nu} \phi, \\
416: g_{\mu\nu}^{\rm ind,2} &=& a^2(\lambda) g_{\mu\nu}^4 - \partial_{\mu}\lambda
417: \partial_{\nu} \lambda.
418: \end{eqnarray}
419: Thus, 
420: \begin{eqnarray}
421: \sqrt{-g^{\rm ind,1}} &=& a^4(\phi) \sqrt{-g_4}\left[ 1 -
422: \frac{1}{2a^2(\phi)}(\partial \phi)^2\right], \\
423: \sqrt{-g^{\rm ind,2}} &=& a^4(\lambda) \sqrt{-g_4}\left[ 1 -
424: \frac{1}{2a^2(\lambda)}(\partial \lambda)^2\right].
425: \end{eqnarray}
426: So the Gibbons--Hawking boundary terms take the form 
427: \begin{eqnarray}
428: \frac{1}{\kappa_5^2} \int d^4 x a^4 \sqrt{-g_4}\left[ 1 -
429: \frac{1}{2a^2(\phi)} (\partial \phi)^2 \right] K, \\
430: \frac{1}{\kappa_5^2} \int d^4 x a^4 \sqrt{-g_4}\left[ 1 -
431: \frac{1}{2a^2(\lambda)} (\partial \lambda)^2 \right] K.
432: \end{eqnarray}
433: The trace of the extrinsic curvature tensor can be calculated from 
434: \begin{equation}
435: K = \frac{1}{\sqrt{-g_5}}\partial_\mu \left[ \sqrt{-g_5}n^\mu \right].
436: \end{equation}
437: Neglecting higher order terms this gives
438: \begin{eqnarray}
439: K = 4\frac{a'}{a}\left[1 - \frac{(\partial \phi)^2}{4a^2}\right].
440: \end{eqnarray}
441: The terms for the second brane can be obtained analogously.
442: Using the BPS conditions and keeping only the kinetic terms, we get
443: for the Gibbons--Hawking boundary terms
444: \begin{eqnarray}
445: &+&\frac{3}{4\kappa_5^2}\int d^4x \sqrt{-g_4} a^2(\phi) U_B(\phi)
446: (\partial \phi)^2, \\
447: &-&\frac{3}{4\kappa_5^2}\int d^4x \sqrt{-g_4} a^2(\lambda) U_B(\lambda)
448: (\partial \sigma)^2.
449: \end{eqnarray}
450: Collecting all terms we find
451: \begin{eqnarray}
452: S_{\rm MSA} = \int d^4 x \sqrt{-g_4}\left[ f(\phi,\lambda) {\cal R}^{(4)} 
453: + \frac{3}{4}a^2(\phi)\frac{U_B(\phi)}{\kappa_5^2}(\partial \phi)^2 
454: - \frac{3}{4} a^2(\lambda)\frac{U_B}{\kappa_5^2}(\lambda)(\partial \lambda)^2 \right].
455: \end{eqnarray}
456: Note that the kinetic term of the field $\phi$ has the wrong
457: sign. This is an artifact of the frame we use here. It is possible to go to the Einstein frame with a simple
458: conformal transformation, in which the sign in front of the kinetic
459: term is correct for both fields. We observe that for a BPS system the moduli potential is totally absent: the fields describe flat directions.
460: 
461: Coming back to the action above, we redefine the fields in the following way:
462: \begin{eqnarray}
463: \tilde \phi^2 &=& \left(1 - 4k\alpha^2 \phi\right)^{2\beta}, \label{posia1}\\
464: \tilde \lambda^2 &=& \left(1-4k\alpha^2 \lambda\right)^{2\beta} \label{posia2},
465: \end{eqnarray}
466: with 
467: \begin{equation}
468: \beta = \frac{2\alpha^2 + 1}{4\alpha^2};
469: \end{equation}
470: then, the gravitational sector can be written as
471: \begin{eqnarray}
472: S_{\rm MSA} &=&\frac{1}{2k\kappa_5^2(2\alpha^2 + 1)}\int d^4 x \sqrt{-g_4}\left[ \left(\tilde\phi^2 -
473: \tilde\lambda^2 \right) {\cal R}^{(4)} +
474: \frac{6}{2\alpha^2 + 1}\left( (\partial \tilde\phi)^2
475: -(\partial \tilde\lambda)^2\right)\right].
476: \end{eqnarray}
477: 
478: This is an action of the form of a multi--scalar tensor theory, in
479: which one scalar field has the wrong sign in front of the kinetic
480: term. Furthermore, in this frame there is a peculiar point where the
481: factor in front of ${\cal R}$ can vanish, namely when
482: $\tilde\phi=\tilde\lambda$, which corresponds to colliding branes. 
483: 
484: In order to avoid mixed
485: terms like $(\partial_\mu \tilde\phi)(\partial^\mu \tilde\lambda)$, we
486: shall define two new fields\footnote{Do not confuse the Ricci scalar
487: ${\cal R}$ with the new field $R$.}:
488: \begin{eqnarray}
489: \tilde \phi &=& Q \cosh R, \label{posib1} \\
490: \tilde \lambda &=& Q \sinh R \label{posib2}.
491: \end{eqnarray}
492: To go to the Einstein frame we perform a conformal transformation:
493: \begin{equation}
494: \tilde g_{\mu\nu} = Q^2 g_{\mu\nu}.
495: \end{equation}
496: Then 
497: \begin{equation}
498: \sqrt{-g} Q^2 {\cal R} = \sqrt{-\tilde g} \left( \tilde {\cal R} -
499: \frac{6}{Q^2}(\tilde\partial Q)^2 \right) .
500: \end{equation}
501: Collecting everything we get the action in the Einstein frame
502: (where we now drop the tilde):
503: \begin{eqnarray}
504: S_{\rm EF} &=& \frac{1}{2k\kappa^2_5(2\alpha^2 + 1)} 
505: \int d^4x \sqrt{-g}\left[ {\cal R} -  \frac{12\alpha^2}{1+2\alpha^2}
506: \frac{(\partial Q)^2}{Q^2} - \frac{6}{2\alpha^2 + 1}(\partial R)^2\right].
507: \end{eqnarray}
508: Clearly, in this frame both fields have the correct sign in front of
509: the kinetic terms. Note that for $\alpha \rightarrow 0$ 
510: (i.e.\ the Randall--Sundrum case) the $Q$--field decouples. 
511: In this case, the field $R$ plays the role of the radion, i.e. it 
512: measures the distance between the branes. Furthermore, we can identify 
513: the gravitational constant:
514: \begin{equation}
515: 16\pi G = 2k\kappa_5^2 (1+2\alpha^2).
516: \end{equation}
517: 
518: \subsection{Moduli space approximation: the matter sector}
519: In this section we are going to leave the BPS configuration. Let's suppose we put some matter on the branes:
520: Standard Model fields on one brane ("visible") and Dark Matter particles on the other brane ("hidden").
521: When the cosmological evolution of the branes is analyzed in the absence 
522: of supersymmetry breaking potentials, it has been shown (\cite{gonz}) that the branes are driven by their matter
523: content: one brane is driven towards the minimum of $U_B$, while the other brane towards the maximum. 
524: In this way it is possible to achieve a full stabilization of the system, granted that we choose the superpotential
525: of the theory in a proper way. If we don't localize matter on branes and don't break supersymmetry, than the BPS
526: configuration is respected, the no-force condition between the branes is present and the system is static.
527: 
528: However it is our intention to describe a Universe with a positive cosmological constant and detuning is necessary in order to
529: achieve de Sitter geometry. We therefore introduce matter as well as supersymmetry breaking 
530: potentials $V(Q,R)$ and $W(Q,R)$ on each branes. We begin with the potentials:
531: to first order in the moduli space approximation we get
532: \begin{equation}
533: \int d^4 x \sqrt{-g_4} \left[ a^4(\phi)V(\phi) \right]
534: \end{equation}
535: with $a^4(\phi) = \tilde \phi^{4/(1+2\alpha^2)}$. The expression for 
536: a  potential $W$ on the second brane is similar with 
537: $a(\phi)$ is replaced by $a(\lambda)$. In the Einstein frame we have
538: (dropping the tilde from the metric):
539: \begin{equation}
540: \int d^4 x\sqrt{-g} Q^{-8\alpha^2/(1+2\alpha^2)}(\cosh
541: R)^{4/(1+2\alpha^2)} V(Q,R) \equiv \int d^4 x\sqrt{-g} V_{\rm eff}(Q,R),
542: \end{equation}
543: where we have defined 
544: \begin{equation}
545: V_{\rm eff}(Q,R) = Q^{-8\alpha^2/(1+2\alpha^2)}(\cosh
546: R)^{4/(1+2\alpha^2)} V(Q,R).
547: \end{equation}
548: The expression for $W(Q,R)$ in the Einstein frame is
549: \begin{equation}
550: \int d^4 x\sqrt{-g} Q^{-8\alpha^2/(1+2\alpha^2)}(\sinh
551: R)^{4/(1+2\alpha^2)} W(Q,R) \equiv \int d^4 x\sqrt{-g} W_{\rm eff}(Q,R),
552: \end{equation}
553: where
554: \begin{equation}
555: W_{\rm eff}(Q,R) = Q^{-8\alpha^2/(1+2\alpha^2)}(\sinh
556: R)^{4/(1+2\alpha^2)} W(Q,R).
557: \end{equation}
558: For matter, the action has the form
559: \begin{equation}
560: S_m^{(1)} = S_m^{(1)}(\Psi_1,g^{ind,1}_{\mu\nu}) \hspace{0.5cm} {\rm and}
561: \hspace{0.5cm} S_m^{(2)} = S_m^{(2)}(\Psi_2,g^{ind,2}_{\mu\nu}),
562: \end{equation}
563: where $g^{ind}$ denotes the {\it induced} metric on each branes and
564: $\Psi_i$ the matter fields on each branes. Note that we do not 
565: couple the matter fields $\Psi_i$ to the bulk scalar field, and thus 
566: not to the fields $Q$ and $R$. In going to the Einstein frame we get 
567: \begin{equation}
568: S_m^{(1)} = S_m^{(1)}(\Psi_1,A^2(Q,R)g_{\mu\nu}) \hspace{0.5cm} {\rm and}
569: \hspace{0.5cm} S_m^{(2)} = S_m^{(2)}(\Psi_2,B^2(Q,R)g_{\mu\nu}),
570: \end{equation}
571: In this expression we have used the fact that, in going to the
572: Einstein frame, the induced metrics on each branes transform with a
573: different conformal factor, which we denote with $A$ and $B$. We have neglected the derivative terms in the moduli fields when considering the coupling to matter on the brane. They lead to higher order operators which can be easily incorporated.  
574: The energy--momentum tensor in the Einstein frame is defined as
575: \begin{equation}
576: T_{\mu\nu}^{(1)} = 2 \frac{1}{\sqrt{-g}} \frac{\delta
577: S_m^{(1)}(\Psi,A^2(Q,R)g)}{\delta g^{\mu\nu}}
578: \end{equation}
579: with an analogous definition for the energy--momentum tensor for
580: matter on the second brane.
581: 
582: In the Einstein frame, the total action is therefore, 
583: \begin{eqnarray}
584: S_{\rm EF} &=& \frac{1}{16\pi G} 
585: \int d^4x \sqrt{-g}\left[ {\cal R} -  \frac{12\alpha^2}{1+2\alpha^2}
586: \frac{(\partial Q)^2}{Q^2} - \frac{6}{2\alpha^2 + 1}(\partial
587: R)^2\right] \nonumber \\
588: &-& \int d^4 x\sqrt{-g} (V_{\rm eff}(Q,R)+W_{\rm eff}(Q,R)) 
589: + S_m^{(1)}(\Psi_1,A^2(Q,R)g_{\mu\nu}) + S_m^{(2)}(\Psi_2,B^2(Q,R)g_{\mu\nu}).
590: \end{eqnarray}
591: 
592: We point out that in this braneworld scenario we naturally obtain a multimetric theory. 
593: Particles localized in different branes "feel" the respective induced metric ($g_{\mu\nu}^{ind}$).
594: In the next section we will discuss an interesting phenomenological application of this fact: the chameleon 
595: mechanism.
596: The moduli potential will be presented in a forecoming section. For more details on the moduli 
597: space approximation see \cite{gonz}.
598: 
599: 
600: 
601: 
602: 
603: 
604: 
605: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
606: \section{Chameleon fields}
607: 
608: As we already mentioned in the introduction, the coupling of the moduli with matter is severely 
609: constrained by phenomenological tests. Introducing massless or ultra-light scalar fields with generic 
610: coupling with matter is prohibited by tests on the equivalence principle and on variations of the 
611: fundamental constants. Different ways to avoid this conflict have been considered
612: in literature. Basically the fields introduced in the model have either a small coupling today or
613: they have a large enough mass (stabilization). Recently a new stabilization scenario has been introduced in \cite{justin}. 
614: According to these authors, the mass of the scalar fields is determined by the density of the environment. On cosmological distances,
615: where the densities are very tiny, the mass can be of the order of the Hubble constant and the fields can roll on cosmological time scales.
616: On the Earth instead, the density is much higher and the field acquires a large enough mass to evade all current experimental bounds
617: on deviations from GR. In other words, the scalar field shows a "chameleontic behaviour".
618: In this section we will discuss this kind of scalar fields. 
619: We will start by reviewing the mechanism with a basic example; then
620: we will discuss the role of the radion as possible chameleon field. 
621: 
622: 
623: \subsection{The chameleon mechanism}
624: 
625: Let's consider a multi-metric model with the following scalar-tensor action:
626: 
627: \begin{equation}
628: S =\int d^4 x\sqrt{-g} \left\{\frac{M_{Pl}^2R}{2} - \frac{(\partial \phi)^2}{2} - 
629:  V(\phi) + {\cal L}_m(\psi_m^{i},g^{(i)}_{\mu\nu})\right\}
630: \end{equation}
631: where $\phi$ is a generic scalar field with potential $V(\phi)$, $\psi_m^{(i)}$ denotes
632: matter fields which follow geodesics of a metric $g^{(i)}_{\mu\nu}=A^2_i(\phi)g_{\mu\nu}$. The scalar field interacts with matter through the conformal factor $A^2(\phi)$, so
633: the dynamics of the chameleon $\phi$ is described by an effective potential that for pressureless (non-relativistic)
634: matter takes the form (with general multi-metric theory a sum over $i$ is present)
635: 
636: \begin{equation}
637: V_{\rm eff}(\phi)=V(\phi) +\rho_m A(\phi)\,.
638: \end{equation}
639: In other words, the conformal coupling modifies the Klein-Gordon equation for the chameleon
640: in the following way:
641: \begin{equation}
642: \nabla^2 \phi = V_{,\phi} + \alpha_\phi\rho_m A(\phi)\,,
643: \label{KG2}
644: \end{equation}
645: %
646: where $\rho_m$ is conserved with respect to the Einstein frame metric ($g_{\mu\nu}$).
647: Supposing a run-away behaviour for the "bare" potential $V(\phi)$, the effective 
648: potential will show a minimum if a "competition" between $V(\phi)$ and the density-dependent term is
649: realized (i.e. $A(\phi)$ is an increasing function).  In this way the chameleontic properties of the field are recovered:
650: the location of the minimum and the mass of the fluctuactions $m^2=V_{eff}''$ both depend on $\rho_m$. 
651: 
652: Another effect which suppresses the chameleon-mediated force is the following: for large enough objects, the $\phi$-force 
653: on a test particle is almost totally due to a thin shell just below the surface of the object, while the contribution of core-matter
654: is negligible. Only a small fraction of the total mass affects the motion of a test particle outside.
655: This mechanism has been named "thin-shell mechanism". To give a more detailed derivation, we consider a chameleon solution
656: for a spherically symmetric object of radius R and homogeneous density $\rho$.
657: We take $V(\phi) = M^{4+n}/\phi^n$, where $M$ has units of mass, and an exponential coupling of the form $A(\phi)= e^{\beta\phi}$ with
658: $\beta={\cal O}(1)$. Next we choose boundary conditions requiring that the solution be non-singular at the origin and that the 
659: chameleon tends to its environment value $\phi_0$ far from the object.
660: 
661: For sufficiently large objects the field assumes inside the object a value $\phi_c$ which minimizes the effective
662: potential, i.e. $V_{,\phi}(\phi_c) + \beta\rho_c e^{\beta\phi_c/M_{Pl}}/M_{Pl} = 0$. This holds everywhere inside the object except
663: within a thin shell of thickness $\Delta R$ below the surface where the field grows. Outside the object, the profile for
664: $\phi$ is essentially that of a massive scalar, $\phi \sim \exp(-m_0r)/r$, where $m_0$ is the mass of the chameleon in the ambient medium.
665: 
666: The thickness of the shell is related to $\phi_0$, $\phi_c$,
667: and the Newtonian potential of the object, $\Phi_N=M/8\pi M_{Pl}^2R$, by
668: %
669: \begin{equation}
670: \frac{\Delta R}{R} \approx \frac{\phi_\infty-\phi_c}{6\beta
671: M_{Pl}\Phi_N}\,.
672: \label{DR}
673: \end{equation}
674: %
675: The exterior solution can then be written explicitly as~\cite{justin}
676: %
677: \begin{eqnarray}
678: \phi(r) &\approx& -\left(\frac{\beta}{4\pi M_{Pl}}\right)\left(\frac{3\Delta
679: R}{R}\right)\frac{M e^{-m_0 (r-R)}}{r} + \phi_0\,. \label{thinsoln}
680: \end{eqnarray}
681: 
682: This equation gives the correction to Newton's law at short distances in the form
683: $F= (1+\theta) F_N$ where $\theta = 2\beta^2 \left(\frac{3\Delta R}{R}\right)$. Hence a thin-shell $\Delta R/R<<1$
684: guarantees a small deviation from Newton's law. 
685: 
686: 
687: \subsection{The radion as a chameleon}
688: In the general set-up we described in the first section, the possible interpretation of the radion $R$ (interbrane distance) 
689: as a dark energy chameleon field has been investigated in \cite{radion}.
690: These authors introduced by-hand a run-away bare potential for the radion, to avoid the falling
691: of the hidden brane into the naked singularity, in the form
692: \begin{equation}
693: V(R) = \Lambda^4 R^{-\gamma}.
694: \end{equation}
695: Their investigation studied the efficiency of the thin-shell mechanism for the radion. The coupling to matter
696: was written by the expansion
697: 
698: \begin{equation}\label{approx}
699: A(R) \approx 1 + \frac{1}{6}\frac{R^2}{2} +...,
700: \end{equation}
701: where the neglection of higher-order terms was allowed by the cosmological evolution, that drove
702: the radion to small field value. Furthermore, the authors calculated a radion-mediated force, with
703: a correction to Newton's law given by
704: \begin{equation}\label{radioncoupling}
705: \theta=\frac{1}{18} R_\infty^2.
706: \end{equation}
707: The strength of the radion-mediated force is specified by the cosmological value of the radion.
708: Since $\theta$ must be small the parameters of the theory must be adjusted to guarantee a small field value.
709: In particular, imposing the gravitational constraints implied $\Lambda \sim 10^{-3} $ eV
710: and $\gamma\le 10^{-5}$. In other words the model must be extremely fine-tuned if we want simultaneously
711: (1) to interpret the radion as a dark energy candidate and (2) to satisfy local gravitational constraints.
712: The fine-tuning of $\gamma$ can be avoided by considering the potential $ V(R)= \Lambda ^4 e^{(\frac{\Lambda}{R})^n}$, but
713: the fine-tuning on $\Lambda$ remains as a cosmological constant fine-tuning.
714: 
715: We are going to exploit the chameleon mechanism to stabilize the radion (R-modulus) following the approach
716: of \cite{radion}. A detailed description of the moduli stabilization will be given in section 6.  
717: 
718: To proceed further, our first step will be the creation of a "naturally small"
719: neutrino mass in this braneworld set-up (see next section). The second step will be the construction of the potential
720: for the scalar field(s) using a physical mechanism and not simply including potentials by-hand (see section 5).
721: 
722: 
723: 
724: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
725: \section{The neutrino mass}
726: 
727: In this section we describe how the sterile bulk fermion can provide a mechanism (different from the see-saw)
728: to obtain a small Dirac neutrino mass in four dimensions. In our scenario all standard model matter and gauge fields are localized on the visible brane.
729: We begin with a brief review of the scenario with mass-varying neutrinos recently proposed in \cite{fnw}. 
730: We proceed with the description of the Kaluza-Klein reduction, then we show that it is possible to 
731: localize the sterile fermion near the hidden brane following \cite{grossman}.
732: In the end of the section we will discuss the role of the conformal transformation to the Einstein frame. 
733: 
734: 
735: \subsection{Mass varying neutrinos and dark energy: the need for moduli stabilization}
736:  
737: We are now going to describe in more detail the idea proposed recently by Fardon, Nelson and Weiner \cite{fnw}
738: that connects the dark energy sector with neutrino's one. 
739: 
740: The main assumption in the FNW scenario is that the dark energy $sector$ has two components: 
741: the dark energy contribution of some (for example) quintessence field $\phi$ plus the neutrino energy density
742: \begin{equation}
743: \rho_{{\rm dark}}=\rho_{\nu} +\rho_{\rm{dark~ energy}}.
744: \end{equation}
745: 
746: An interaction is present between the two components, due to the dependence of the neutrino mass on
747: the quintessence field. The form of this dependence is obtained via a see-saw mechanism with a variable mass $M=M(\phi)$
748: for the right-handed neutrino. Remarkably the functional dependence $M(\phi)$ has been put in by-hand. 
749: It could be $M(\phi)=h\phi$ or $M(\phi)=M_0e^{\phi^2/f^2}$ with $h$ and $f$ some constants or something else. 
750: In the model considered in this article, however, we will be able to derive explicitly a form of the neutrino mass as function of the two
751: moduli. In our model variable neutrino mass is not put in by-hand.
752: During the expansion of the Universe, the temperature falls down till $T<m_{\nu}$.
753: At that time the neutrinos become non-relativistic particles and an effective potential is generated in the form
754: \begin{equation}
755: \rho_{{\rm dark}}=m_{\nu}n_{\nu} +\rho_{\rm{dark ~energy}}(m_{\nu}).
756: \end{equation}
757: 
758: 
759: 
760: FNW assumed that $\rho_{dark}$ is stationary with respect to variations in the neutrino mass. In other words 
761: the neutrino mass is fixed by a competition between $\rho_{dark energy}$ and $\rho_{\nu}$. The hypothesis of stationarity
762: implies that
763: 
764: 
765: \begin{equation}
766: \frac{\partial \rho_{\rm{dark}}}{\partial m_{\nu}}=n_{\nu} +\frac{\partial\rho_{\rm{dark ~energy}}(m_{\nu})}{\partial m_{\nu}} =0.
767: \end{equation}
768: This last equation determines the temperature dependence of the neutrino mass in a specific dark energy model.
769: Adding to this scheme the conservation of energy equation
770: \begin{equation}
771: \dot{\rho}=- 3H(\rho + p),
772: \end{equation}
773: it is possible to show that
774: \begin{equation}
775: \omega +1= \frac {m_{\nu}n_{\nu}}{\rho_{\rm{dark}}}= \frac{m_{\nu}n_{\nu}}{
776: m_{\nu}n_{\nu}+\rho_{\rm{dark ~energy}}}.
777: \end{equation}
778: where
779: \begin{equation}
780: \omega = \frac{p_{dark }}{\rho_{dark}}.
781: \end{equation}
782: Equation (61) inform us that the neutrino contribution to $\rho_{dark}$ is a small fraction
783: of $\rho_{dark}$.
784: Another important result can be obtained observing that the continuity equation guarantees $\rho_{dark} \sim R^{-3(1+w)}$.
785: Since $n_{\nu} \sim R^{-3}$, the neutrino mass is nearly inversely proportional to the neutrino density:
786: 
787: \begin{equation}
788: m_{\nu} \sim n_{\nu}^{\omega}\simeq 1/n_{\nu}.
789: \end{equation}
790: 
791: For further details we refer the reader to references \cite{peccei},\cite{neut}.
792: 
793: As pointed-out by Peccei \cite{peccei}, in the FNW scenario the intrinsic scale of dark energy is not obtained 
794: naturally since the value of the potential in the present epoch is fixed by hand:
795: \begin{equation}
796: V(m_{\nu}^0) \sim T^3_0 m_{\nu}^0 \sim \rho_{DE} \sim (2 \times 10^{-3} eV)^4.
797: \end{equation}
798: Thus, the connection of the dark energy scale with the neutrino mass scale is translated into a stabilization 
799: problem for the field $\phi$. When neutrinos become non-relativistic and $m_{\nu}n_{\nu}$ switches on, is it possible to 
800: guarantee the correct minimum?
801: 
802: Remarkably, in our approach we will obtain a neutrino mass dependent on the moduli $\lambda$ and $\phi$. 
803: In other words the neutrino mass depends on the geometrical configuration of the braneworld. As we already mentioned in 
804: the introduction, the only dynamic that we contemplate for the scalar fields is a 
805: chameleontic behaviour for the radion. 
806: On cosmological distances, 
807: the densities are tiny and the 
808: radion can roll on cosmological time scales (the neutrino mass is {\it variable} 
809: with time and possible deviations
810: from a pure cosmological constant contribution can be observed in the DE sector); 
811: on smaller scales instead, densities are higher and 
812: {\it both} moduli are simply sitting on the minimum of the potential.
813: As a final result, the connection between DE and neutrino mass will be established in the form
814: \beq
815: (\frac{V^{1/4}}{m_{\nu}})_{min}\sim 1,
816: \eeq
817: where $V$ is the total moduli potential. 
818: 
819: 
820: \subsection{Kaluza-Klein reduction}
821: 
822: Let's consider again the SPINOR action
823: 
824: 
825: \begin{equation}\label{SPIN}
826:    S_{SPINOR} = \int\!\mbox{d}^4x\!\int\!\mbox{d}z\,\sqrt{-g_5}
827:    \left\{ E_a^A \left[ \frac{i}{2}\,\bar\Psi\gamma^a
828:    (\partial_A-\overleftarrow{\partial_A})\Psi
829:    + \frac{\omega_{bcA}}{8}\,\bar\Psi \{\gamma^a,\sigma^{bc}\} \Psi
830:    \right] - m\,\mbox{sgn}(\alpha)\,\bar\Psi\Psi \right\} \,.
831: \end{equation}
832: We use capital
833: indices $A,B,\dots$ for objects defined in curved space, and 
834: lower-case indices $a,b,\dots$ for objects defined in the tangent 
835: frame. The matrices $\gamma^a=(\gamma^\mu,i\gamma_5)$ provide a 
836: four-dimensional representation of the Dirac matrices in 
837: five-dimensional flat space. The quantity
838: $E_a^A$ is the
839: inverse vielbein and $\omega_{bcA}$ is the spin connection. Because 
840: in our case the metric is diagonal the connection gives no contribution to
841: the action (\ref{SPIN}).
842: 
843: The sign change of the mass term under $\alpha\to-\alpha$ is necessary in 
844: order to conserve $\alpha$-parity, as required by the $Z_2$ orbifold 
845: symmetry we impose.
846: 
847: 
848: Using an integration by parts and defining left- and right-handed
849: spinors $\Psi_{L,R}\equiv\frac12(1\mp\gamma_5)\Psi$, the action can 
850: be written as
851: \begin{eqnarray}\label{Sdel}
852:    S_{SPINOR} &=& \int\!\mbox{d}^4x\,\sqrt{-g_4}\int\!\mbox{d}z\,\bigg\{
853:     e^{-3\sigma} \left( \bar\Psi_L\,i\rlap/\partial\,\Psi_L
854:     + \bar\Psi_R\,i\rlap/\partial\,\Psi_R \right) 
855:     - e^{-4\sigma}\,m\,\mbox{sgn}(\alpha) \left( 
856:     \bar\Psi_L \Psi_R + \bar\Psi_R \Psi_L \right) \nonumber\\
857:    &&\mbox{}- \frac{1}{2} \left[ \bar\Psi_L \left(
858:     e^{-4\sigma} \partial_z + \partial_z\,e^{-4\sigma}
859:     \right) \Psi_R - \bar\Psi_R \left(
860:     e^{-4\sigma} \partial_z + \partial_z\,e^{-4\sigma}
861:     \right) \Psi_L \right] \bigg\} \,,
862: \end{eqnarray}
863: where we considered the redefinition $a(z)=e^{-\sigma}$ and we observe that $\sqrt{-g_5}=e^{-4\sigma}\sqrt{-g_4}$.
864: The action is even under the 
865: $Z_2$ orbifold symmetry if $\Psi_L(x,z)$ is an odd function of 
866: z and $\Psi_R(x,z)$ is even, or vice versa. To perform the 
867: Kaluza--Klein decomposition we write
868: \begin{equation}\label{KK}
869:    \Psi_{L,R}(x,z) = \sum_n \psi_n^{L,R}(x)\,
870:    e^{2\sigma}\,\hat f_n^{L,R}(z) \,.
871: \end{equation}
872: Because of the $Z_2$ symmetry of the action it is sufficient to 
873: restrict the integration over $z$ from $\phi$ to $\lambda$. Inserting the ansatz (\ref{KK}) 
874: into the action and requiring that the result takes the form of the 
875: usual Dirac action for massive fermions in four dimensions,
876: \begin{equation}\label{Sferm}
877:    S = \sum_n \int\!\mbox{d}^4x\,\sqrt{-g_4}\Big\{
878:    \bar\psi_n(x)\,i\rlap/\partial\,\psi_n(x)
879:    - m_n\,\bar\psi_n(x)\,\psi_n(x) \Big\} \,,
880: \end{equation}
881: where $\psi\equiv\psi_L+\psi_R$ (except for possible chiral modes) 
882: and $m_n\ge 0$ , we find that the functions $\hat f_n^{L,R}(\phi)$ 
883: must obey the conditions
884: \begin{eqnarray}\label{cond}
885:    \int\limits_\phi^\lambda\!\mbox{d}z\,e^\sigma
886:    \hat f_m^{L*}(z)\,\hat f_n^L(z)
887:    &=& \int\limits_\phi^\lambda\!\mbox{d}z\,e^\sigma
888:     \hat f_m^{R*}(z)\,\hat f_n^R(z) = \delta_{mn} \,,
889:     \nonumber\\
890:    \left( \pm\,\partial_z - m \right)
891:    \hat f_n^{L,R}(z) 
892:    &=& - m_n\,e^\sigma \hat f_n^{R,L}(z) \,.
893: \end{eqnarray}
894: The boundary conditions $\hat f_m^{L*}(\phi)\,\hat f_n^R(\phi)
895: =\hat f_m^{L*}(\lambda)\,\hat f_n^R(\lambda)=0$, which follow since either 
896: all left-handed or all right-handed functions are $Z_2$-odd, ensure
897: that the differential operators $(\pm \partial_z -m)$ 
898: are hermitian and their eigenvalues $m_n$ real. Since the equations 
899: are real, the functions $\hat f_n^{L,R}(\phi)$ could be chosen real 
900: without loss of generality.
901: 
902: 
903: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
904: \subsection{The right-handed zero mode: wave function localization}
905: 
906: We now choose boundary conditions such that all left-handed modes are odd under orbifold parity. 
907: We specialize our investigation to the right-handed massless zero-mode. (\ref{cond}) now takes the form
908: \begin{eqnarray}\label{cond2}
909:    \int\limits_\phi^\lambda\!\mbox{d}z\,e^\sigma
910:    \hat f_0^{R*}(z)\,\hat f_0^R(z)
911:    & = 1 \,,
912:     \nonumber\\
913:    \left( \partial_z + m \right)
914:    \hat f_0^{R}(z) 
915:    &=& 0 \,.
916: \end{eqnarray}
917: 
918: Let's consider, for example, the limit $\alpha\Rightarrow 0$, then $a(z)=e^{-kz}$ and we can write
919: \begin{eqnarray}
920:    \hat f^2(z)=\frac{k'e^{-2mz}}{e^{k'\lambda}-e^{k'\phi}},
921: \end{eqnarray}
922: where we defined $k'=k-2m$ and $\hat f=\hat f_0^R$.
923: 
924: The zero-mode wave function on the visible brane is thus given by
925: \begin{eqnarray}\label{massa}
926:    \hat f^2(\phi)=\frac{k'e^{-2m\phi}}{e^{k'\lambda}-e^{k'\phi}}.
927: \end{eqnarray}
928: 
929: Remarkably it is possible to localize the wave function on the hidden brane; in particular for well-separated brane we distinguish the cases:
930: 
931: \begin{itemize}
932:   \item $k'>0$ ................. $\hat f^2(\phi)<<1\Longleftrightarrow k'\lambda>>-2m\phi$
933:   \item $k'<0$ ................. $\hat f^2(\phi)<<1\Longleftrightarrow e^{-k\phi}<<1$.
934:   \end{itemize}
935:   
936:   
937:   
938: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
939: \subsection{The Higgs sector and the neutrino mass}
940: We now describe the realization of a small Dirac neutrino mass in four dimensions, using the localization of the wave 
941: function mentioned above. 
942: Omitting gauge interactions, the action for a Higgs doublet 
943: $H=(\phi_1,\phi_2)$, a left-handed lepton doublet $L=(\nu_L,e_L)$ 
944: and a right-handed lepton $e_R$ localized on the visible brane is 
945: \begin{eqnarray}
946:    S &=& \int\!\mbox{d}^4x\,\sqrt{-g^{\rm ind,1}} \left\{
947:     g_{\rm ind,1}^{\mu\nu}\,\partial_\mu H_B^\dagger\,\partial_\nu H_B
948:     - \lambda \left( |H_B|^2 - v_B^2 \right)^2 \right\} \nonumber\\
949:    &+& \int\!\mbox{d}^4x\,\sqrt{-g^{\rm ind,1}} \left\{
950:     \bar L_B\hat\gamma^\mu\partial_\mu L_B
951:     + \bar e_{RB}\hat\gamma^\mu\partial_\mu e_{RB}
952:     - \left( y_e \bar L_B H_B e_{RB} + \mbox{h.c.} \right)
953:     \right\} \,,
954: \end{eqnarray}
955: where $g_{\rm ind,1}^{\mu\nu}$ is the 
956: induced metric on the visible brane, and the superscript over the gamma 
957: matrices is used to make the vielbein implicit.
958: 
959: We now introduce a Yukawa coupling of the bulk fermion with the Higgs 
960: and lepton fields. With our choice of boundary conditions all 
961: left-handed Kaluza--Klein modes vanish at the visible brane, so only 
962: the right-handed modes can couple to the Standard Model fields on the 
963: brane. We consider a coupling of the form:
964: \begin{equation}
965:    S_Y = - \int\!\mbox{d}^4x\,\sqrt{-g_{\rm ind,1}} \left\{
966:    \hat Y_5 \bar L_B(x) \widetilde H_B(x) \Psi_R(x,\phi)
967:    + \mbox{h.c.} \right\} \,,
968: \end{equation}
969: where $\widetilde H=i\sigma_2 H^*$, and the Yukawa coupling $\hat Y_5$ 
970: is naturally of order $M^{-1/2}$, with $M$ the fundamental Planck 
971: scale of the theory. Rescaling the Standard Model fields and inserting for the bulk fermion the Kaluza--Klein 
972: ansatz (\ref{KK}), we find
973: \begin{equation}
974:    S_Y = - \sum_{n\ge 0} \int\!\mbox{d}^4x \sqrt{-g_{\rm ind,1}}\left\{ \frac{Y_5}{a^2(\phi)} 
975:    \bar L_B(x) \widetilde H_B(x) \psi_n^R(x) f_n^R(\phi) + \mbox{h.c.} \right\}
976: \end{equation}
977: where $Y_5$ is naturally of order unity.
978: 
979: Specializing this formula to the right-handed zero mode we have
980: \begin{equation}
981:    S_Y = - \int\!\mbox{d}^4x \sqrt{-g_{\rm ind,1}}\left\{ \frac{Y_5}{a^2(\phi)} 
982:    \bar L_B(x) \widetilde H_B(x) \psi_0^R(x) f_0^R(\phi) + \mbox{h.c.} \right\}
983: \end{equation}
984: with a mass for the 4d neutrino that can be expressed as a function of the moduli in the form
985: \begin{equation}
986:    m_{\nu} = m_{\nu}(\phi,\lambda)=\frac{Y_5f_0^R(\phi)v_b}{a^2(\phi)},
987: \end{equation}
988: where $f_0^R(\phi)$ is given by (\ref{massa}) after the rescaling.
989: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
990: \subsection{Conformal transformation}
991: Now we want to study the effect of a conformal transformation to the Einstein frame ($\tilde g$) on the neutrino mass. In order 
992: to do that, it is necessary to determine the rescaling laws of the various fields in the Yukawa term.
993: We have to take into account two different conformal transformations.
994: \begin{itemize}
995:  \item $g^{ind}\Longrightarrow  g^4$  
996:  \end{itemize}
997: The transformations is completely defined by equations (19) and (20).
998: 
999: \begin{itemize}
1000:  \item $g^4\Longrightarrow \tilde g$ ..............  $ \tilde g_{\mu\nu} = Q^2 g_{\mu\nu}^4$
1001:  \end{itemize}
1002: 
1003: Collecting the two transformations together and remembering $\beta = \frac{2\alpha^2 + 1}{4\alpha^2}$ we have:
1004: \begin{eqnarray}
1005: g_{\mu\nu}^{ind,1}=a^2(\phi)g_{\mu\nu}^4=a^2(\phi) Q^{-2} \tilde g_{\mu\nu} \equiv A^2 \tilde g_{\mu\nu}; \\
1006: g_{\mu\nu}^{ind,2}=a^2(\lambda)g_{\mu\nu}^4=a^2(\lambda) Q^{-2} \tilde g_{\mu\nu} \equiv B^2 \tilde g_{\mu\nu};
1007: \end{eqnarray}
1008: where 
1009: 
1010: \begin{equation}\label{fattori}
1011: A=Q^{-\frac{1}{2 \beta}}(\cosh R)^{\frac{1}{1+2 \alpha^2}},\ B=Q^{-\frac{1}{2 \beta}}(\sinh R)^{\frac{1}{1+2 \alpha^2}}.
1012: \end{equation}
1013: 
1014: From these formulas, requiring a canonical normalization of the kinetic terms of the fields, we find the following conformal rescalings for the fields:
1015: \begin{itemize}
1016:  \item CHIRAL ZERO MODE: $\Psi_0^R\Longrightarrow \tilde \Psi_0^R=Q^{-3/2} \Psi_0^R$  
1017:  \item HIGGS FIELD: $H_B\Longrightarrow H_E=\frac{a(\phi)}{Q}  H_B$
1018:  \item STANDARD MODEL'S SPINOR FIELD: $\Psi\Longrightarrow \tilde \Psi=(\frac{a(\phi)}{Q})^{3/2} \Psi$.
1019:  \end{itemize}
1020: 
1021: Using these last formulas in the Yukawa term, we obtain the "correction" to 
1022: the neutrino mass induced by the conformal transformation to the Einstein frame:
1023: \begin{equation}
1024: \tilde m_{\nu}(\phi,\lambda)=\frac{a(\phi)^{5/2}}{Q} m_{\nu}(\phi,\lambda).
1025: \end{equation}
1026: 
1027: 
1028: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1029: \section{The moduli potential}
1030: 
1031: In the original two-branes model proposed by Randall and Sundrum (\cite{RS1},\cite{RS2}) the brane tensions were 
1032: tuned in order to have Minkowski branes.
1033: The separation between the branes becomes a field (radion) in the process of 
1034: compactification to four dimensions. With flat branes the classical potential 
1035: for the radion is absent: the interbrane distance is parameterized by a massless scalar field. Since 
1036: this field has not been observed experimentally, the authors were faced with the so-called "radion stabilization problem". One 
1037: possible solution is 
1038: provided by quantum vacuum energy of bulk fields: Casimir contribution of bulk fields generates a potential for the radion. 
1039: The possibility of radion stabilization by this mechanism has been considered by several authors using scalar 
1040: fields (\cite{scalarf},\cite{GPT}, \cite{FT}), spinor fields (\cite{FMT,FMT2}). 
1041: The result is that it is generally not possible to give an acceptable mass to the radion and solve the hierarchy problem simultaneously. 
1042: 
1043: When the branes are $dS_4$ or $AdS_4$ a classical potential for the radion is present. 
1044: It is stabilizing for the $AdS_4$ case and unstable for $dS_4$ branes \cite{baggerredi,detuned}. 
1045: The Casimir effect is then a quantum correction to the classical potential (\cite{baggerredi}). In the $dS_4$ 
1046: brane case the Casimir contribution has been calculated for conformally 
1047: coupled scalar fields \cite{WN}, for massless spinor fields (\cite{fermionc}) and for massive scalar fields (\cite{ENOO}). 
1048: The $AdS_4$ case has been studied in \cite{norman}.
1049: 
1050: In this section we will present the classical potential for the moduli (Q, R) in the $dS_4$ case.
1051: Then we proceed with the evaluation of the quantum correction induced by a massive spinor field in the bulk. 
1052: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1053: \subsection{The classical potential}
1054: 
1055: In order to write the classical potential it is necessary to review the main characteristics of the
1056: supersymmetric Randall-Sundrum model (\cite{susyrs}) with detuned brane tensions. The detuning process will render our model more
1057: realistic: since we want to describe an Universe with a positive cosmological constant, a de Sitter geometry for branes is 
1058: the correct one and detuning is necessary. 
1059: 
1060: Denoting $T_{1,2}$ and $T_f$ respectively
1061: the tension of the first (second) brane and the fine-tuned tension, supersymmetry imposes the bound $|T_{1,2}| \le T_f$.
1062: The detuning process converts Minkowski branes in curved branes. The $AdS_4$ solution arises when $|T_{1,2}| < T_f$; the $dS_4$
1063: solution corresponds to $|T_{1,2}|> T_f$.
1064: 
1065: When one detunes the brane tension $U_B\to T U_B$, the moduli
1066: pick up a potential \cite{brax, marsiglia}:
1067: 
1068: \begin{equation}
1069: V = \frac {6(T-1)k}{\kappa_5^2} e^{\alpha \psi}.
1070: \end{equation}
1071: Taking into account both branes, the classical potential expressed as a function of $S=lnQ$ and $R$ in the Einstein frame is
1072: 
1073: \begin{equation}\label{poti}
1074: V_{class}(S,R) = V_{eff} + W_{eff},
1075: \end{equation}
1076: where
1077: \begin{equation}\label{poti}
1078: V_{eff}(S,R) = \frac{6(T-1)k}{\kappa_5^2} e^{-12\alpha^2 S/(1+2\alpha^2)}
1079: \left(\cosh R \right)^{(4-4\alpha^2)/(1+2\alpha^2)}
1080: \end{equation}
1081: and
1082: \begin{equation}\label{poti2}
1083: W_{eff}(S,R) = \frac{6(T-1)k}{\kappa_5^2} e^{-12\alpha^2 S/(1+2\alpha^2)}
1084: \left(\sinh R \right)^{(4-4\alpha^2)/(1+2\alpha^2)}.
1085: \end{equation}
1086: Note that $T=1$ corresponds to the BPS case.
1087: 
1088: 
1089: 
1090: As mentioned in \cite{baggerredi} the potential is unstable for $dS_4$ and we proceed with the analysis of the 
1091: Casimir energy as a stabilizing contribution for the system.
1092: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1093: \subsection{Casimir contribution: the 5-dimensional bag}
1094: We will calculate the Casimir contribution of the bulk spinor on a Euclideanised form of the metric. On the 
1095: Euclidean section the de 
1096: Sitter branes become concentric four spheres \cite{GS},
1097: \begin{equation}
1098: ds^2 =dz^2+a^2(z)d\Omega^2_4,
1099: \end{equation}
1100: where $d\Omega^2_4$ is the 4-sphere metric.
1101: The metric is conformal to a cylinder $I\times S^4$ \cite{GS,GEN}. Thus,
1102: \begin{equation}
1103: ds^2 =a^2(z)(dy^2+d\Omega^2_4)
1104: \quad\qquad a(z)=(1-4k\alpha^2 z)^{\frac{1}{4 \alpha^2}}\,,
1105: \label{metric}
1106: \end{equation}
1107: where the coordinates are dimensionful and we defined $dy=\frac{dz}{a(z)}$.
1108: The dimensional length $L$ is given by
1109: \begin{equation}
1110: L(\phi,\lambda)=\int_{\phi}^{\lambda}\frac{dz}
1111: {a(z)}=-\frac{1}{k} \frac{1}{4 \alpha^2 -1} [(1-4k \alpha^2 \lambda)^{1-\frac{1}{4 \alpha^2}}-(1-4k \alpha^2 \phi)^{1-\frac{1}{4 \alpha^2}}].
1112: \end{equation}
1113: 
1114: Let's consider a second transformation given by
1115: 
1116: \begin{equation}\label{trconf}
1117: y=-R lnr,
1118: \end{equation}
1119: with $0<r<1$ and $R$ is a constant introduced for dimensional reasons. In this way (\ref{metric}) becomes
1120: 
1121: \begin{equation}
1122: ds^2 =\frac{a^2(z)}{r^2}(dr'^2 + r'^2 d\Sigma^2) \equiv \beta^2 [dr'^2 + r'^2 d \Sigma^2],
1123: \label{metric2}
1124: \end{equation}
1125: where $d \Sigma^2 \equiv R^{-2} d \Omega_4^2$ is the metric for the 4-sphere of radius one, $\beta^2 \equiv a^2/r^2$ is the total conformal
1126: factor and $r' \equiv Rr$ is the dimensionful radial coordinate ($0<r'<R$). The metric (\ref{metric}) is thus conformally related
1127: to a 5-dimensional generalized cone endowed with 4-sphere as a base: a 5-ball of radius R. 
1128: 
1129: In order to keep track of the geometrical "reconfiguration" of the system under the total conformal 
1130: transormation, let's consider the following formulas:
1131: 
1132: \begin{itemize}
1133: \item $z<z^* \equiv \frac{1}{4k \alpha^2}$ where $z^*$ corresponds to the position of the bulk singularity,
1134: \item $y \equiv -\frac{1}{k} \frac{1}{4 \alpha^2 -1} (1-4k \alpha^2 z)^{1-\frac{1}{4 \alpha^2}}$,
1135: \item $y=-Rlnr$.
1136: \end{itemize}
1137: The first one corresponds to the assumption that the branes can't fall into the bulk singularity.
1138: Since we are interested in the description of a low-redshift Universe, we will consider well-separated branes:
1139: we assume that the hidden brane is close to the bulk singularity ($z_{hidden} \sim z^*$, remarkably this is a general behaviour 
1140: in the low-energy regime of 
1141: BPS-braneworlds \cite{palma}) while the visible one 
1142: is located far away in the bulk ($z_{vis}<<z_{hidden}$). 
1143: The remaining formulas modify the branes coordinate 
1144: in the following way:
1145: \begin{eqnarray} 
1146: z \rightarrow -\infty  \Longleftrightarrow y=0 \Longleftrightarrow r=1,  VISIBLE \\
1147: z \rightarrow z^* \Longleftrightarrow y\rightarrow +\infty \Longleftrightarrow r=0, SINGULARITY.
1148: \end{eqnarray}
1149: 
1150: Remarkably the total conformal transformation translated the cosmological two-branes
1151: set up into a 5-ball. The ball radius is connected to the moduli $\lambda$ and $\phi$ by 
1152: 
1153: \begin{equation}
1154: \frac{L(\lambda,\phi)}{R}=-ln \epsilon,
1155: \end{equation}
1156: where $\epsilon$ is a dimensionless parameter that is small for well-separated branes. In this geometrical configuration 
1157: the evaluation of the effective action will be easier, as we now deduce.
1158: 
1159: 
1160: We will start recalling the SPINOR action and studying the effect of the conformal transformation on 
1161: the spinor field $\Psi$. The action is
1162: 
1163: 
1164: \begin{equation}\label{action}
1165:    S = \int\!\mbox{d}^4x\!\int\!\mbox{d}z\,\sqrt{-g_5}
1166:    \left\{ E_a^A \left[ \frac{i}{2}\,\bar\Psi\gamma^a
1167:    (\partial_A-\overleftarrow{\partial_A})\Psi
1168:    + \frac{\omega_{bcA}}{8}\,\bar\Psi \{\gamma^a,\sigma^{bc}\} \Psi
1169:    \right] - m\,\mbox{sgn}(\alpha)\,\bar\Psi\Psi \right\} \,.
1170: \end{equation}
1171: 
1172: In the transformation $g_{AB}=\beta^2 g'_{AB}$ the action is rewritten as
1173: 
1174: \begin{equation}\label{apice}
1175:    S = \int\!\mbox{d}^5x\,\sqrt{-g'_5}
1176:    \left\{ E_a^{A'} \left[ \frac{i}{2}\,\bar\Psi'\gamma^a
1177:    (\partial_A-\overleftarrow{\partial_A})\Psi'
1178:    + \frac{\omega_{bcA}}{8}\,\bar\Psi' \{\gamma^a,\sigma^{bc}\} \Psi'
1179:    \right] - m'\,\mbox{sgn}(\alpha)\,\bar\Psi'\Psi' \right\} \,.
1180: \end{equation}
1181: where the primed quantities are referred to the ball metric and they are given by
1182: $\Psi'=\beta^{2} \Psi$, $m'=\beta m$ and $E_a^{A'}=\beta E_a^{A}$.
1183: We will now use a more compact notation and, omitting the prime, we write (\ref{apice}) as
1184: \begin{equation}
1185:    S = i \int\!\mbox{d}^5x\ \Psi^* D \Psi
1186: \end{equation}
1187: where $D \equiv \slash{\!\nabla}+ im$.
1188: 
1189: To proceed further we need to specify boundary conditions for the bulk fermion. 
1190: Since we are interested in a chiral theory on the branes, we choose the "option-L" of \cite{grossman}, namely: all the left handed modes are 
1191: odd under orbifold parity. In this way the correct boundary condition is 
1192: 
1193: \begin{eqnarray} 
1194: z=\phi :\quad P_-\psi=0,
1195: z=\lambda :\quad P_-\psi=0,
1196: \end{eqnarray}
1197: where $P_{\pm}=\frac12(1 \pm \gamma_5)$. To complete the specification of the 
1198: boundary conditions we remember the existence condition for the operator $D^*$: if $P_-\psi=0$, 
1199: this requires that the normal derivative $(\partial_y - m)P_+ \Psi=0$
1200: should vanish. In summary, we have defined two subspaces in direct sum generated by 
1201: the operators $P_{\pm}$; while on the first space (-) we have Dirichlet condition, on the second one (+) 
1202: we impose Robin boundary condition. In other words the bulk fermion satisfies mixed boundary conditions (for 
1203: further details see \cite{FMT,Mth,MOSS}).
1204: 
1205: In this way our cosmological set up is the 5-dimensional "extension" of the MIT bag model \cite{mit}. In these 
1206: systems quarks and gluons
1207: are free inside the bag, but they are unable to cross the boundary. This last condition 
1208: corresponds precisely to the mixed boundary
1209: conditions discussed above: a chiral theory on the branes corresponds to the condition 
1210: that no quark current is lost through the boundary. The zeta function for massive fermionic fields inside
1211: the bag has been considered in \cite{bag}. Analogous calculations have been developed for massive scalar field (\cite{bk,bek,bekl96}).
1212: Functional determinants were discussed in \cite{begk,bdk,abdk,eli1,eli2,stuart,Mth}. 
1213: For moduli stabilization with zeta function regularization see also \cite{GPT2,FGPT}. 
1214: 
1215: In the next section we will discuss the zeta function of the system following the approach of \cite{bag}.
1216: 
1217: 
1218:  
1219: \subsection{Casimir contribution: the $\zeta$ function}
1220: 
1221: We start recalling the general set-up for the MIT bag model in three dimensions. The setting we 
1222: consider first is the Dirac spinor inside a spherically symmetric bag
1223: confined to it by the appropriate boundary conditions. 
1224:  Thus, we must solve the equation:
1225: \beq
1226: H\phi_n (\vec r) = E_n \phi_n (\vec r) , \label{2.1}
1227: \eeq
1228: $H$ being the Hamiltonian, with the boundary conditions
1229: \beq
1230: \left[1+i \left(\frac{\vec r } r \vec \gamma \right) \right] \phi_n 
1231: \left|_{r=R} \right. =0 .\label{2.3}
1232: \eeq
1233: 
1234: 
1235: The  separation to be carried out
1236:  in the eigenvalue equation (\ref{2.1}) is rather
1237: standard and will not be given here in detail. Let $\vec J$ be the 
1238: total angular momentum operator and $K$ the spin projection operator. 
1239: Then there exists a simultaneous set of eigenvectors of $H,\vec J^2,J_3,K$
1240: and the parity $P$. The eigenfunctions for positive eigenvalues $\kappa
1241: =j+1/2$ of $K$ read
1242: \beq
1243: \phi_{jm} = \frac A {\sqrt{r} } \left(
1244:    \begin{array}{l} i J_{j+1} (\omega r ) \Omega_{jlm} \left( \frac {\vec r} r 
1245: \right)  \\
1246:     - \sqrt{\frac{E-m}{E+m} } J_j (\omega r ) \Omega_{jl'm } 
1247:          \left( \frac {\vec r} r \right)
1248:    \end{array} 
1249: \right), \label{2.4}
1250: \eeq
1251: whereas, for $\kappa = -(j+1/2) $, one finds
1252: \beq
1253: \phi_{jm} = \frac A {\sqrt{r} } \left(
1254:    \begin{array}{l} i J_{j} (\omega r ) \Omega_{jlm} \left( \frac {\vec r} r
1255: \right)  \\
1256:      \sqrt{\frac{E-m}{E+m} } J_{j+1} (\omega r ) \Omega_{jl'm }
1257:          \left( \frac {\vec r} r \right)
1258:    \end{array}
1259: \right). \label{2.5}
1260: \eeq
1261: Here 
1262: $\omega=\sqrt{E^2-m^2}$, $A$ is a normalization constant and 
1263: $\Omega_{jlm} (\vec r /r)$ are the well known spinor harmonics. 
1264: In order to obtain eigenfunctions of the parity operator we must set
1265:  $l'=l-1$ in
1266: (\ref{2.4}) and $l'=l+1$ in (\ref{2.5}). In both cases, $j=1/2,3/2,...,
1267: \infty$, and the eigenvalues are degenerate in $m=-j,...,+j$.
1268: 
1269: Imposing the boundary conditions (\ref{2.3}) on the solutions 
1270: (\ref{2.4}) and (\ref{2.5}), respectively, one easily finds the 
1271: corresponding implicit eigenvalue
1272: equation. For $\kappa > 0$, it reads
1273: \beq
1274: \sqrt{\frac{E+m}{E-m}} J_{j+1} (\omega R) +J_j (\omega R) =0 , \label{2.6}
1275: \eeq
1276: and for $\kappa < 0$, on its turn, 
1277: \beq
1278: J_j (\omega R)- \sqrt{\frac{E-m}{E+m}} J_{j+1} (\omega R) =0 .\label{2.7}
1279: \eeq
1280: 
1281: We define the zeta function of the system as
1282: \beq
1283: \zeta (s) = \sum_k (E_k^2) ^{-s} \label{relzeta}
1284: . 
1285: \eeq
1286: The power of the method lies in the well defined prescriptions 
1287: and procedures that we have at our hand to analytically continue the series
1288: to the rest of the complex $s$-plane, even when the spectrum $E_k$ is
1289: not known explicitly. In particular we are interested in the calculation of the Casimir energy for
1290: massive spinors in the bag and we will employ zeta function to regularize this ground state energy. Briefly we consider
1291: \beq\label{introcas}
1292: E_0 (s) &=& - \frac{1}{2} \sum_k 
1293: \left( E_k^2 \right)^{1/2 -s} \mu^{2s}, \qquad
1294: \mbox{Re}\ s >s_0= 2\nonumber\\
1295: &=& -\frac 1 2 \zeta (s-1/2) \mu^{2s} \label{grounden},
1296: \eeq
1297: and later analytically continue to the value $s=0$ in the complex plane.
1298: Here $s_0$ is the abscissa of convergence of the series, $\mu$ the usual 
1299: mass parameter.
1300: 
1301: 
1302: 
1303: 
1304: Let's apply this QCD analysis to the cosmological branes system. When we pass from three to five dimensions
1305: we have to take into account new degeneracy factors for the eigenvalues, while the implicit eigenvalue equation 
1306: remains formally untouched. In particular the degeneracy $d(j)$ for a spinor field on the ball is
1307: \begin{equation}
1308: d(j)=d_s {n+D-2\choose D-2},
1309: \end{equation}
1310: where $d_s$ is the spinor dimension, $D$ is the manifold dimension (i.e. the dimension of the generalized cone), $j=n+ \frac{D}{2} -1$ and $n=0,1,2,...$.
1311: Since a closed analytical form for the eigenvalues is not available for this system, we will 
1312: exploit the residue theorem (for a pedagogical description of these techniques see \cite{kirsten}):
1313: \beq
1314: \zeta (s) &=&  \sum_{j=3/2,\ldots}^\infty d(j) \int_\gamma \frac{dk}{2\pi i} 
1315: (k^2 +m^2)^{-s}  \nn \\ && \hspace{-10mm} 
1316: \times \, \frac{\partial}{\partial k}
1317:  \ln \left[ J_j^2 (kR) - J_{j+1}^2 (kR) 
1318: + \frac{2m}{k} J_j (kR) J_{j+1} (kR) \right].
1319: \eeq
1320: 
1321: Using the method ---ordinarily employed in this situation--- of deforming the 
1322: contour which originally encloses the singular points on the real axis,
1323: until it covers the imaginary axis, 
1324: after some manipulations we obtain the following equivalent expression for 
1325: $\zeta$:
1326: \beq
1327: \zeta   (s) &=& \frac{ \sin \pi s}{\pi} 
1328: \sum_{j=3/2,\ldots}^\infty d(j) 
1329: \int_{mR/j}^\infty dz \, \left[ \left( \frac{zj}{R}\right)^2 -m^2
1330: \right]^{-s} \nn  \\
1331: && \times \frac{\partial}{\partial z} \ln\left\{z^{-2j} \left[ I_j^2 (zj)
1332:  \left( 1 + \frac{1}{z^2}- \frac{2mR}{z^2j} \right) + {I_j'}^2 (zj)
1333: \right. \right. \nn \\ && \left. \left.
1334:  + \frac{2R}{zj} \left( m - \frac{j}{R} \right) I_j (zj) {I_j}' (zj)
1335: \right]\right\},
1336: \eeq
1337: where we defined $z=k/j$.
1338: As is usual, we will now split the zeta function into two parts:
1339: \beq
1340: \zeta (s)=Z_N(s)+\sum_{i=-1}^N A_i(s),
1341: \eeq 
1342: namely a regular one, $Z_N$, and a remainder that contains the 
1343: contributions of the $N$ first terms
1344: of the Bessel functions $I_\nu (k)$ as $\nu, k \rightarrow \infty$ with 
1345: $\nu /k$ fixed.
1346: The regular part of the zeta function is:
1347: \beq
1348: Z_N(s)&=& \sn \smj \itz \paran^{-s}  \nn\\
1349: && \hspace{-17mm}\times \frac{\partial}{\partial z}
1350:  \left\{ \ln \left[I_j^2(zj)(1+\frac{1}{z^2}-\frac{2mR}{z^2j})+
1351: I'_j\!{}^2(zj)+\frac{2R}{zj} (m-\frac{j}{R}) I_j(zj)I'_j(zj)\right] \right. 
1352:  \nn\\
1353: &&-\left.\ln\left[\frac{e^{2j\eta}(1+z^2)^\frac{1}{2}(1-t)}{\pi j z^2}\right]
1354:    -\sum_{k=1}^N \frac{D_k(mR,t)}{j^k}\right\},
1355: \eeq
1356: 
1357: where $\eta = \sqrt{1+z^2} +\ln [z/(1+\sqrt{1+z^2})]$ and $t=1/\sqrt{1+z^2}$.
1358: After renaming $mR=x$, the relevant polynomials (see appendix A) are given by
1359: \beq
1360: D_1(t)&=&{\frac {{t}^{3}}{12}}+\left (x-1/4\right )t \nn\\
1361: D_2(t)&=&-{\frac {{t}^{6}}{8}}-{\frac {{t}^{5}}{8}}+\left (-{\frac {x}{2}}+1/8
1362: \right ){t}^{4}+\left (-{\frac {x}{2}}+1/8\right ){t}^{3}-{\frac {{t}^
1363: {2}{x}^{2}}{2}} \nn\\
1364: D_3(t)&=&{\frac {179\,{t}^{9}}{576}}+{\frac {3\,{t}^{8}}{8}}+\left (-{\frac {23
1365: }{64}}+{\frac {7\,x}{8}}\right ){t}^{7}+\left (x-1/2\right ){t}^{6}+
1366: \left ({\frac {9}{320}}-{\frac {x}{4}}+{\frac {{x}^{2}}{2}}\right ){t}
1367: ^{5}    \nn\\
1368: &&+\left ({\frac {{x}^{2}}{2}}+1/8-{\frac {x}{2}}\right ){t}^{4}+
1369: \left (-{\frac {x}{8}}+{\frac {5}{192}}+{\frac {{x}^{3}}{3}}\right ){t
1370: }^{3} \nn\\
1371: D_4(t)&=& -\frac{71}{64} t^{12} - \frac{179}{128} t^{11} + (\frac{57}{32} - \frac{21}{8} x)t^{10} + (\frac{327}{128} - \frac{49}{16} x) t^{9} 
1372: + (-\frac{37}{64} + 2x - x^2) t^{8}    \nn\\
1373: &&+ (-\frac{165}{128} + 3x - \frac{5 x^2}{4})t^7 + (-\frac{1}{8} + \frac{x}{4} +\frac{x^2}{8} -\frac{x^3}{2})t^6 + (\frac{17}{128} - \frac{7x}{16} 
1374: + \frac{x^2}{2}- \frac{x^3}{2})t^5    \nn\\
1375: &&+ (\frac{1}{32} - \frac{x}{8} + \frac{x^2}{8} - \frac{x^4}{4})t^4  \nn\\
1376: D_5(t)&=& \frac{40573}{7680} t^{15} + \frac{213}{32} t^{14} + ( -\frac{5853}{512} + \frac{1463x}{128})t^{13} + (\frac{105x}{8}-\frac{507}{32})t^{12}  \nn\\
1377: && + (\frac{1835}{256}- \frac{487x}{32} + \frac{49 x^2}{16}) t^{11} + (\frac{397}{32}-\frac{159x}{8}+ 4 x^2 )t^{10} + (-\frac{1441}{2304}+\frac{217x}{64}-
1378: \frac{13x^2}{8} + \frac{9 x^3}{8}) t^{9} \nn\\
1379: && + (-\frac{109}{32} + \frac{31x}{4} - \frac{27x^2}{8} + \frac{3x^3}{2}) t^{8} + 
1380: (-\frac{1567}{3584} + \frac{33x}{32} - \frac{9x^2}{16} + \frac{x^4}{2}) t^{7}  \nn\\
1381: && + (-\frac{x}{2} + \frac{3x^2}{8} - \frac{x^3}{2} + \frac{3}{16} +  \frac{x^4}{2}) t^{6} + (\frac{107}{2560} - \frac{17x}{128} + \frac{x^2}{8} - \frac{x^3}{8} + \frac{x^5}{5})t^5.
1382: \label{asympol}
1383: \eeq
1384: The number N of terms to be subtracted must be high enough in order to absorb all possible divergent
1385: contributions into the groundstate energy. In our case N=5.
1386: The asymptotic contributions $A_i (s)$, $i=-1,...,5$ are defined as
1387: \beq
1388: A_{-1} (s) &=& 
1389: \frac{\sin (\pi s)} {\pi} \sum_{j=3/2}^{\infty} 
1390: 2jd(j) \int_{mR/j} ^{\infty} dz \left( \left( \frac{zj} R \right)^2 -m^2 
1391: \right)^{-s}
1392: \frac{\sqrt{1+z^2} -1} z \nonumber\\
1393: A_0 (s) &=& 
1394: \frac{\sin (\pi s)} {\pi} \sum_{j=3/2}^{\infty}
1395: d(j) \int_{mR/j} ^{\infty} dz \left( \left( \frac{zj} R \right)^2 -m^2 
1396: \right)^{-s}
1397: \frac{\partial} {\partial z} \ln \frac{\sqrt{1+z^2} (1-t) } {z^2} 
1398: \nn\\
1399: A_i (s) &=& 
1400: \frac{\sin (\pi s)} {\pi} \sum_{j=3/2}^{\infty}
1401: d(j) \int_{mR/j} ^{\infty} dz \left( \left( \frac{zj} R \right)^2 -m^2 
1402: \right)^{-s}
1403: \frac{\partial} {\partial z}
1404: \frac{D_i (t)} {j^i}.  \label{aisanf}
1405: \eeq
1406: This completes the description of the zeta function. In the next section we will consider the divergent contributions
1407: in the vacuum energy. After that we will move to the renormalization process.
1408: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1409: \subsection{Casimir contribution: Residue [$\zeta(-1/2)$]}
1410: 
1411: As a first step towards renormalization, let's start considering the divergent terms in the groundstate energy.
1412: By construction they are contained in the functions $A_i(s)$. We will discuss them separately following the approach 
1413: of \cite{bag},\cite{bekl96}.
1414: 
1415: 
1416: \subsubsection{Residue $A_{-1}(-1/2)$}
1417: 
1418: 
1419: The first asymptotic contribution reads
1420: \beq
1421: A_{-1} (s) &=& 
1422: \frac{\sin (\pi s)} {\pi} \sum_{j=3/2}^{\infty} 
1423: 2jd(j) \int_{mR/j} ^{\infty} dz \left( \left( \frac{zj} R \right)^2 -m^2 
1424: \right)^{-s}
1425: \frac{\sqrt{1+z^2} -1} z
1426: \eeq
1427: 
1428: of which we need the analytical continuation to $s=-1/2$.
1429: With the substitution $t=(zj /R)^2 -m^2$, this expression results into
1430: the following one
1431: \begin{eqnarray}
1432: A_{-1} (s)
1433: &=& \frac{\sin (\pi s)}{\pi} \sum_{j =3/2}^{\infty} d(j) \int\limits _0 ^{\infty}
1434: dt\,\,\frac{t^{-s}}{t+m^2}\left\{
1435: \sqrt{j^2 +R^2 (t+m^2)} -j\right\}\nonumber\\
1436: &=&-\frac 1 {2\sqrt{\pi}}
1437:  \frac{\sin (\pi s)}{\pi} \sum_{j=3/2}^{\infty} d(j) \int\limits _0 ^{\infty}
1438: dt\,\,t^{-s} \int\limits _0^\infty
1439: d\alpha \,\, e^{-\alpha (t+m^2)}\label{anhdaa2}\\
1440: & &\times\int\limits_0^{\infty}d\beta \,\,\beta ^{-3/2} \left\{
1441: e^{-\beta (j^2 +R^2 [t+m^2] ) } -e^{-\beta j^2}\right\},\nonumber
1442: \end{eqnarray}
1443: where the Mellin integral representation for the single factors has been
1444: used. As we see,
1445: the $\beta$-integral is well defined. In order to deal separately 
1446: with the two contributions in the $\beta$-integral, we introduce a regularization parameter
1447: $\delta$. In this way $A_{-1}(s)$ can be written as
1448: \begin{eqnarray}
1449: A_{-1} (s) =\lim_{\delta \to 0} \left[ A_{-1}^1 (s,\delta )
1450: +A_{-1} ^2 (s,\delta ) \right], \label{anhdaa3}
1451: \end{eqnarray}
1452: with
1453: \begin{eqnarray}
1454: A_{-1}^1 (s,\delta )
1455:  &=&-\frac 1 {2\sqrt{\pi}}
1456:  \frac{\sin (\pi s)}{\pi} \sum_{j=3/2}^{\infty} d(j)
1457: \int\limits_0^\infty  d\alpha \,\,
1458:  e^{-\alpha m^2}
1459: \int\limits_0^{\infty}d\beta \,\,\beta ^{-3/2+\delta}e^{-\beta (j^2 +R^2
1460: m^2 ) }
1461: \int\limits _0 ^{\infty}
1462: dt\,\,t^{-s}e^{-t(\alpha +\beta R^2)}\nonumber
1463: \end{eqnarray}
1464: and
1465: \begin{eqnarray}
1466: A_{-1} ^2 (s,\delta ) =    \frac 1 {2\sqrt{\pi}} \Gamma (1-s)
1467: \frac{\sin (\pi s)}{\pi}    \sum_{j=3/2}^{\infty} d(j)
1468: \int\limits_0^\infty  d\alpha \,\,
1469: e^{-\alpha m^2 }   \alpha^{s-1}\int\limits_0^{\infty}d\beta \,\,
1470: \beta ^{-3/2+\delta}e^{-\beta j^2}.
1471: \nonumber
1472: \end{eqnarray}
1473: 
1474: In $A_{-1}^1 (s,\delta
1475: )$ two of the integrals can be done, yielding
1476: \begin{eqnarray}
1477: A_{-1}^1 (s,\delta )
1478: &=& -\frac{R^{1-2\delta}}{2\sqrt{\pi}\Gamma (s)} \Gamma (s+\delta -1/2) \sum_{j=3/2}^{\infty} d(j) \int\limits  _1 ^{\infty}
1479: dy\,\, y^{s -1} \left[ m^2 y + \left(\frac{j} R\right)^2
1480: \right]^{1/2-s-\delta}. \label{anhdaa4}
1481: \end{eqnarray}
1482: 
1483: Exploiting the identity
1484: 
1485: \beq
1486: \int\limits_0^\infty  dx \,\,
1487: \frac{x^{\alpha -1}}{(y+zx)^{s+ \epsilon}} = \frac{ \Gamma(\alpha) 
1488: \Gamma( s + \epsilon - \alpha)}{\Gamma(s+ \epsilon)} z^{- \alpha} y^{- \epsilon -s + \alpha},
1489: \eeq
1490: it is possible to rewrite $A_{-1} ^2 (s,\delta )$ in the form
1491: \begin{eqnarray}
1492: A_{-1} ^2 (s,\delta )&=& \frac{R^{1-2\delta}}{2\sqrt{\pi}\Gamma (s)}
1493: \Gamma (s+\delta -1/2)
1494: \sum_{j=3/2}^{\infty} d(j) \int\limits  _0 ^{\infty}
1495: dx\,\, x^{s-1        }
1496: \left[ m^2 x +
1497: \left(\frac{j} R\right)^2 \right]^{1/2-s-\delta}.
1498: \label{anhdaa5}
1499: \end{eqnarray}
1500: 
1501: Adding up (\ref{anhdaa4}) and (\ref{anhdaa5}) yields
1502: \begin{eqnarray}
1503: A_{-1} (s) &=& \frac R {2\sqrt{\pi}\Gamma (s)}
1504: \Gamma (s -1/2) \sum_{j=3/2}^{\infty} d(j) \int\limits  _0 ^{1}
1505: dx\,\, x^{s-1        }
1506: \left[ m^2 x +
1507: \left(\frac{j} R\right)^2 \right]^{1/2-s} \nn\\
1508: &=& \frac{R^{2s}}{2\sqrt{\pi}} \frac{\G (s-1/2)}{\G (s+1)}
1509:       \sum_{j=3/2}^\infty d(j) \left\{ \frac 1
1510:      {\left[j^2 +(mR)^2\right] ^{s-1/2}} \right.\nn\\
1511: & &\left.\hspace{1cm}\quad+\left( s-\frac 1 2 \right) (mR)^2
1512:            \int_0^1 dx \,\, \frac{x^s}{(j^2 +(mR)^2 x)^{s+1/2}} \right\},
1513: \label{anhdaa6}
1514: \eeq
1515: where in the last step a partial integration has been performed such that
1516: the $x$-integral is well behaved at $s=-1/2$.
1517: 
1518: This is a form suited for the treatment of the sum,
1519: which is performed using
1520: 
1521: \beq
1522: \sum_{\nu =1/2,3/2,...}^{\infty} \nu^{2n+1} \left(1+\left(\frac{\nu} x
1523: \right) ^2 \right) ^{-s} &=& \frac 1 2 \frac {n! \Gamma (s-n-1)}{\Gamma (s)}
1524: x^{2n+2} \label{dispari}\\
1525: & &\hspace{-25mm} +(-1)^n 2 \int\limits_0^x d\nu \,\, \frac{\nu^{2n+1} }
1526: {1+e^{2\pi\nu}  } \left( 1-\left( \frac{\nu} x \right)^2 \right) ^{-s} \nn\\
1527: & & \hspace{-25mm} +(-1)^n 2\cos (\pi s)
1528:  \int\limits_x^{\infty} d\nu \,\, \frac{\nu^{2n+1} }
1529: {1+e^{2\pi\nu}  } \left( \left( \frac{\nu} x \right)^2 -1
1530: \right) ^{-s}\nn
1531: \eeq
1532: for odd powers of $\nu$ and
1533: 
1534: \beq \sum_{\nu =1/2,3/2,...}^{\infty}
1535: \nu^{2n} \left(1+\left(\frac{\nu} x \right) ^2 \right) ^{-s} &=&
1536: \frac 1 2 \frac { \Gamma (n+1/2) \Gamma (s-n-1/2)}{\Gamma (s)}
1537: x^{2n+1} \label{pari}\\ & &\hspace{-25mm} -(-1)^n 2\sin (\pi s)
1538:  \int\limits_x^{\infty} d\nu \,\, \frac{\nu^{2n} }
1539: {1+e^{2\pi\nu}  } \left( \left( \frac{\nu} x \right)^2 -1
1540: \right) ^{-s} \nn
1541: \eeq
1542: for even powers of $\nu$.
1543: This formula as well as (\ref{dispari}) can only be applied for $Re(s)<1$, otherwise the integral diverges at the integration
1544: limit $\nu=x$.
1545: 
1546: Since the prefactor in (\ref{anhdaa6}) contains a pole at s=-1/2 an expansion of the integrand is necessary and 
1547: remembering that $d(j)=(2/3) (j^3 + \frac{3}{2} j^2 - \frac{j}{4} - \frac{3}{8})$, our final expression for the residue is found to be
1548: \beq
1549: \res A_{-1} (-1/2) &=& - \frac{457}{241920 \pi R} - \frac{17 m^2 R}{2880 \pi} + \frac{m^4 R^3}{144 \pi} + \frac{m^6 R^5}{180 \pi}.   
1550: \eeq
1551: 
1552: 
1553: 
1554: \subsubsection{Residue $A_{i}(-1/2)$, $i\geq 1$}
1555: 
1556: Let's start remembering the form of the asymptotic contributions for $i\geq 1$:
1557: \beq
1558: A_i (s) &=& 
1559: \frac{\sin (\pi s)} {\pi} \sum_{j=3/2}^{\infty}
1560: d(j) \int_{mR/j} ^{\infty} dz \left( \left( \frac{zj} R \right)^2 -m^2 
1561: \right)^{-s}
1562: \frac{\partial} {\partial z}
1563: \frac{D_i (t)} {j^i}.  
1564: \eeq
1565: We need to introduce some notation as follows:
1566:  \beq 
1567: D_i (t) =\sum_{a=0}^{2i}
1568: x_{i,a} t^{i+a}
1569: \label{eq2.26} 
1570: \eeq
1571: (where the relevant coefficients $x_{i,a}$ are explicitly listed in (\ref{asympol}) and 
1572: \beq 
1573: f(s;c;b) =\sum_{\nu =
1574: 1/2,3/2,\ldots} ^{\infty} j ^c \left( 1+ \left( \frac{j} {mR}
1575: \right ) ^2 \right) ^{-s-b}.\nn 
1576: \eeq
1577: With this notation we write
1578: \beq
1579: A_i(s)&=& \frac{\sin (\pi s)} {\pi} \sum_{j=3/2}^{\infty}
1580: \frac{d(j)}{j^i} \sum_{a=0}^{2i} x_{i,a} \int_{mR/j} ^{\infty} dz \left( \left( \frac{zj} R \right)^2 -m^2 
1581: \right)^{-s}
1582: \frac{\partial} {\partial z}
1583: t^{a+i}=  \nn \\
1584: &=& -\frac{\sin (\pi s)} {\pi} \sum_{a=0}^{2i} \left\{\frac{x_{i,a}m^{-2s}}{(mR)^{a+i}} \frac{\Gamma (s+ \frac{a+i}{2}) \Gamma (1-s)}{\Gamma (\frac{a+i}{2})} \sum_{j=3/2}^{\infty} d(j) j^a [1+(\frac{j}{mR})^2]^{-s-\frac{a+i}{2}}\right\}= \nn \\
1585: &=& -\frac{\sin (\pi s) \Gamma (1-s)} {m^{2s} \pi} \frac{2}{3} \sum_{a=0}^{2i} \left\{
1586:  \frac{x_{i,a}}{(mR)^{a+i}} \frac{\Gamma (s+ \frac{a+i}{2})}{\Gamma (\frac{a+i}{2})} 
1587: \left[f(s;3+a;\frac{i+a}{2}) + \frac{3}{2} f(s;2+a;\frac{i+a}{2}) \right. \right. \nn \\
1588: & & \left. \left. -\frac{1}{4} f(s;1+a;\frac{i+a}{2})-\frac{3}{8} f(s;a;\frac{i+a}{2}) \right]\right\},
1589: \eeq
1590: where we performed the z-integration using the identity (\cite{kirsten})
1591: \beq 
1592: \iinma t^n &=& -\mzs
1593: \frac{n}{2(mR)^n}\frac{\g s+\frac n 2 \right) \Gamma (1-s)}{\g
1594: 1+\frac n 2\right)}  \nu ^n \left[1+\numr
1595: \right]^{-s-\frac n 2}
1596: \label{eq2.27} 
1597: \eeq
1598: and we exploited the vanishing of the degeneracy factor for $j=1/2$.
1599: In order to evaluate the $f(s;a;b)$ for the relevant values in $s=-1/2$ we can use the following formulas:
1600: \begin{itemize}
1601:  \item the recurrence formula
1602:  \beq
1603: f(s;a;b) = (mR)^2 \left[ f(s;a-2;b-1)- f(s;a-2;b)\right];
1604: \label{recur}
1605: \eeq
1606:  \item formulas \ref{dispari} and \ref{pari}, valid for $\Re{s}<1$;
1607:  \item an extension of formulas \ref{dispari} and \ref{pari} to $\Re{s}<k+1\leq n+2$ obtained using partial integrations:
1608:  \beq
1609:  \sum_{\nu = 1/2}^{\infty} \nu^{2n+1} \left( 1+(\frac{\nu}{x})^2 \right) ^{-s} = 
1610: \frac{n! \Gamma (s-n-1) x^{2n+2}}{2 \Gamma (s)} - \delta_{k,n+1} 
1611: \frac{n! \Gamma (s-n-1)}{2 \Gamma (s)} x^{2n+2} \nn \\
1612: + 2(-1)^{k+n} \frac{\Gamma (s-k)}{\Gamma (s)} \int\limits_0^{x} d\nu \,\, \left[ \left( \frac{d}{d\nu} \frac{x^2}{2 \nu} \right)^k  \frac{\nu^{2n+1} }
1613: {1+e^{2\pi\nu}  } \right] \left( 1- \left(  \frac{\nu} x \right)^2 
1614: \right) ^{-s +k} \nn \\
1615: +2(-1)^n cos(\pi s) \frac{\Gamma(s-k)}{\Gamma (s)} \int\limits_x^{\infty} d\nu \,\, \left[ \left( \frac{d}{d\nu} \frac{x^2}{2 \nu} \right)^k  \frac{\nu^{2n+1} }
1616: {1+e^{2\pi\nu}  } \right] \left( \left( \frac{\nu} x \right)^2 -1 
1617: \right) ^{-s +k}
1618:  \eeq
1619:  for odd powers of $\nu$ and
1620:  \beq
1621:   \sum_{\nu = 1/2}^{\infty} \nu^{2n} \left( 1+(\frac{\nu}{x})^2 \right) ^{-s} = 
1622: \frac{\Gamma (n+1/2) \Gamma (s-n-1/2)}{2 \Gamma (s)} x^{2n +1} \nn \\
1623: -2(-1)^n sin(\pi s) \frac{\Gamma(s-k)}{\Gamma (s)} \int\limits_x^{\infty} d\nu \,\, \left[ \left( \frac{d}{d\nu} \frac{x^2}{2 \nu} \right)^k  \frac{\nu^{2n} }
1624: {1+e^{2\pi\nu}  } \right] \left( \left( \frac{\nu} x \right)^2 -1 
1625: \right) ^{-s +k}
1626:  \eeq
1627:  for even powers of $\nu$.
1628: \end{itemize}
1629: 
1630: 
1631: In this way it is possible to construct an explicit form for the asymptotic contributions and, in particular,
1632: our result for the residues is given by
1633: 
1634: \beq
1635: \res A_{1} (-1/2) &=&  - \frac{17m}{2880 \pi} + \frac{17}{11520 \pi R} - \frac{m^2 R}{288 \pi} + \frac{m^3 R^2}{24 \pi} + \frac{m^4 R^3}{144 \pi} + \frac{m^5 R^4}{12 \pi}\nn\\
1636: \res A_2 (-1/2) &=&  m (\frac{1}{64} + \frac{1}{16 \pi}) - \frac{1}{R} (\frac{1}{1024} + \frac{1}{192 \pi}) \nn\\
1637: & &+ m^2 R (\frac{19}{512} + \frac{1}{48 \pi}) + 
1638: (\frac{3}{32} + \frac{1}{4 \pi})m^3 R^2 + \frac{m^4 R^3}{16} \nn\\ 
1639: \res A_{3} (-1/2) &=& -m (\frac{1}{192} + \frac{7}{480 \pi}) + \frac{1}{R} (\frac{1}{3072} + 
1640: \frac{97}{241920 \pi}) -m^2 R (\frac{25}{1536} + \frac{565}{12096 \pi}) \nn \\
1641: & & -m^3 R^2 (\frac{3}{32} + \frac{41}{120 \pi}) - m^4 R^3 (\frac{1}{16}+ \frac{2}{9 \pi}) - \frac{m^5 R^4}{9 \pi} \nn\\ 
1642: \res A_{4} (-1/2) &=& -m (\frac{63}{4096} + \frac{11}{240 \pi}) + \frac{1}{R} (\frac{35}{65536} + 
1643: \frac{13}{40320 \pi}) -m^2 R (\frac{13}{256} + \frac{1}{6 \pi}) \nn \\
1644: & & -m^3 R^2 (\frac{3}{64} + \frac{1}{6 \pi}) - \frac{1}{32} m^4 R^3 \nn \\
1645: \res A_{5} (-1/2) &=& m (\frac{61}{12288} + \frac{37}{2240 \pi})  + \frac{1}{R} (\frac{23}{196608}- \frac{31}{2661120 \pi}) + m^2 R (\frac{23}{768}+ \frac{359}{3780 \pi})\nn \\
1646: & & + m^3 R^2 (\frac{3}{64} + \frac{181}{1260 \pi})  + m^4 R^3 (\frac{1}{32} + \frac{4}{45 \pi}) + \frac{2}{45 \pi} m^5 R^4.
1647: \eeq
1648: 
1649: 
1650: \subsubsection{Residue $A_{0}(-1/2)$}
1651: 
1652: This calculation will exploit the basic formulas we showed above and the Mellin integral representation.
1653: With the substitution $t=(x \nu/a)^2 - m^2$ we write
1654: \beq
1655: A_0 (s) &=& 
1656: \frac{\sin (\pi s)} {\pi} \sum_{j=3/2}^{\infty}
1657: d(j) \int_{mR/j} ^{\infty} dz \left( \left( \frac{zj} R \right)^2 -m^2 
1658: \right)^{-s}
1659: \frac{\partial} {\partial z} \ln \frac{\sqrt{1+z^2} -1 } {z^2} = \nn \\
1660: &=& - \frac{1}{2} \frac{\sin (\pi s)}{\pi} \sum_{j =3/2}^{\infty} d(j) \int\limits _0 ^{\infty}
1661: dt\,\,\frac{t^{-s}}{t+m^2}\left\{\frac{
1662: \sqrt{j^2 +R^2 (t+m^2)} -j}{\sqrt{j^2 +R^2 (t+m^2)}}\right\} \nn \\
1663: &=& \frac 1 {2\sqrt{\pi}}
1664:  \frac{\sin (\pi s)}{\pi} \sum_{j=3/2}^{\infty} d(j)j \int\limits _0 ^{\infty}
1665: dt\,\,t^{-s} \int\limits _0^\infty
1666: d\alpha \,\, e^{-\alpha (t+m^2)}\label{anhdaa2}\\
1667: & &\times\int\limits_0^{\infty}d\beta \,\,\beta ^{-1/2} \left\{
1668: e^{-\beta (j^2 +R^2 [t+m^2] ) } -e^{-\beta j^2}\right\},\nonumber
1669: \eeq
1670: where in the last step the Mellin representation for the single factors has been used
1671: and, in particular, we exploited 
1672: \beq
1673: \frac{
1674: \sqrt{j^2 +R^2 (t+m^2)} -j}{\sqrt{j^2 +R^2 (t+m^2)}}  = -\frac{j}{\sqrt{\pi}} \int\limits_0^{\infty}d\beta \,\,\beta ^{-1/2} \left\{
1675: e^{-\beta (j^2 +R^2 [t+m^2] ) } -e^{-\beta j^2}\right\}.
1676: \eeq
1677: 
1678: In this way we can closely follow the calculation discussed previously for $\mbox{Res}\ A_{-1} (-1/2)$.
1679: The asymptotic contribution is rewritten as
1680: \beq
1681: A_{0} (s) &=&- \frac{R^{2s}}{2\sqrt{\pi}} \frac{\G (s+1/2)}{\G (s+1)}
1682:       \sum_{j=3/2}^\infty d(j)j \left\{ \frac 1
1683:      {\left[j^2 +(mR)^2\right] ^{s+1/2}} \right.\nn\\
1684: & &\left.\hspace{1cm}\quad+\left( s+\frac 1 2 \right) (mR)^2
1685:            \int_0^1 dx \,\, \frac{x^s}{(j^2 +(mR)^2 x)^{s+3/2}} \right\}
1686: \label{azero}
1687: \eeq
1688: and the relevant residue is given by
1689: \beq
1690: \mbox{Res}\ A_0 (-1/2) = \frac{17}{1920 \pi R} + \frac{m^2 R}{16 \pi} + \frac{ m^4 R^3}{24 \pi}.
1691: \label{poles}
1692: \eeq
1693: 
1694: 
1695: \subsection{Casimir contribution: renormalization}
1696: For the discussion of the renormalization let us look for the divergent
1697: terms in the groundstate energy. Adding up the contributions calculated previously, the residue 
1698: for the zeta function is
1699: \beq
1700: \mbox{Res}\ \zeta (-1/2) = \frac{4m}{315 \pi} + \frac{41}{10395 \pi R}  - \frac{2 m^2 R}{45 \pi} 
1701: -\frac{23 m^3 R^2}{315 \pi} - \frac{7 m^4 R^3}{90 \pi} + \frac{m^5 R^4}{60 \pi} + \frac{m^6 R^5}{180 \pi}
1702: \label{poles}
1703: \eeq
1704: These terms form the minimal set of counterterms necessary in order to
1705: renormalize our theory.
1706:  
1707: We are led into a physical system consisting of two parts:
1708: \begin{enumerate}
1709: \item A classical system consisting of a spherical surface ('bag') with radius
1710:  $R$. Its energy reads:
1711: \beq
1712: E_{class} = a R^5 + b R^4 + c R^3 + d R^2 + e R + f + \frac{g}{R}.
1713: \label{n1}
1714: \eeq
1715: This energy is determined by the parameters: $a, b, c, d, e, f, g$.
1716: \item A spinor quantum field $\Psi (x)$ obeying the Dirac 
1717: equation 
1718: and the MIT boundary conditions 
1719: (\ref{2.3}) on the surface.  
1720: The quantum field has a ground state energy given by $E_0$, Eq.
1721: (\ref{grounden}).    
1722: \end{enumerate}
1723: Thus, the complete energy of the physical system is
1724: \beq
1725: E= E_{class} +E_0 \label{eges}
1726: \eeq
1727: and in this context the renormalization can be achieved by shifting the 
1728: parameters in $E_{class}$ by an amount which
1729: cancels the divergent contributions. 
1730: 
1731: First we perform a kind of minimal subtraction, where only the 
1732: divergent contribution is eliminated,
1733: \beq
1734: a &\to &a + \frac{m^6} {360 \pi}
1735: \frac 1 {s}
1736: \quad
1737: b  \,\to  \, b +\frac{m^5}{120 \pi} \frac 1 s\nn\\
1738: c  &\to &  c - \frac{7 m^4}{180 \pi} 
1739:  \frac 1 {s}
1740: \quad
1741: d\, \to \, d-\frac{23 m^3}{630 \pi} \frac 1 s
1742: \label{n8}\\
1743: e &\to &e - \frac{m^2} {45 \pi}
1744: \frac 1 {s}
1745: \quad
1746: f  \,\to  \, f +\frac{2m}{315 \pi} \frac 1 s\nn\\
1747: g  &\to &  g + \frac{41}{20790 \pi} 
1748:  \frac 1 {s} \nn
1749: \eeq
1750: The quantities $\alpha = \{a, b, c, d, e, f, g \}$ are a set of free parameters of
1751: the theory to be determined experimentally. In principle we are free
1752: to perform finite renormalizations at our choice of all the parameters. 
1753: 
1754: We find natural to perform two further renormalizations.
1755: First it is possible to determine the asymptotic behavior of the $A_i$ for $m\to\infty$. The finite pieces not vanishing in the
1756: limit $m\to \infty$ are all of the same type appearing in the classical 
1757: energy. Our first finite renormalization is such that those pieces are 
1758: cancelled. As a result, only the "quantum contributions" are finally included, 
1759: because, physically, a quantum field of infinite mass is not expected 
1760: to fluctuate. The resulting $A_i$ will be  called $A_i ^{(ren)}$. 
1761: 
1762: Concerning $Z$ we have constructed a numerical fit of $Z$ by a polynomial 
1763: of the form
1764: \beq
1765: P(m) = \sum_{i=0}^6 c_i m^i, \nn
1766: \eeq
1767: and then subtracted this polynomial from $Z$. As explained above, 
1768: this is nothing
1769: else than an ulterior finite renormalization. The result will be 
1770: denoted by $Z^{(ren)}$.
1771: 
1772: Summing up, we can write the complete energy as
1773: \beq
1774: E=E_{class} +E_0^{(ren)}, \label{egesend}
1775: \eeq
1776: where $E_{class}$ is defined as in  
1777: (\ref{n1}) with the renormalized parameters $\alpha$ 
1778: and $E_0^{(ren)}= Z^{(ren)} + \sum_{i=-1}^5 A_i^{(ren)}$. 
1779: 
1780: Figure \ref{grafico} shows the quantum contribution $E_0^{(ren)}$ as a function of the bag radius. Remarkably the Casimir contribution 
1781: of the massive spinor is stabilizing: the energy exhibits a minimum corresponding to a stable geometrical configuration. 
1782: 
1783: In the next section we will discuss the parameters of the model and the stabilization mechanism in greater detail.
1784: 
1785: \begin{figure}[t]
1786: \begin{center}
1787: \includegraphics[width=12cm]{dcs.eps}
1788: \caption{$E_0^{(ren)}$ as a function of the bag radius R for a proper choice of parameters (in particular m=1 has been assumed). Only the contribution j=3/2 has been included as the leading one.}
1789: \label{grafico}
1790: \end{center}
1791: \end{figure}
1792: 
1793: \section{Moduli stabilization and parameters space}
1794: 
1795: Is it possible to achieve a realistic model with a "natural" choice of the parameters? 
1796: To answer this question it is necessary to remember that
1797: our purpose is to stabilize the entire geometrical configuration of the system which amounts to stabilizing {\it both} moduli.
1798: As we already said, it is not possible 
1799: to introduce in a model ultralight scalar fields with a generic coupling with matter, since phenomenological
1800: constraints must be faced. In this model the two moduli interact with matter in a known way. In fact, we could parametrize the strength of this coupling,
1801: since we know explicitly the conformal factors (4.78), but our approach will be different: dangerous couplings 
1802: are kept under control because we are giving a (large enough) mass
1803: to the moduli. It seems now worthwhile to focus on the stabilization mechanism.
1804: 
1805: \subsection{R-modulus stabilization: the chameleon mechanism}
1806: 
1807: As we already mentioned in section 3, the possible interpretation of the R-modulus as a chameleon field
1808: has been investigated in \cite{radion}. Following their approach we will add a run-away bare potential for the radion
1809: in the form:
1810: 
1811: \begin{equation}\label{champot}
1812: V_{add}(R)= M^4 e^{\left(\frac{M}{m_{pl} R}\right)^n}
1813: \end{equation}
1814: where $M=10^{-3}$ eV in order to trigger acceleration now. It is not a
1815: quintessence potential since it does not converge to zero at
1816: infinity. It should be considered as a very flat potential which may 
1817: appear from some non-perturbative physics for the radion.
1818: 
1819: For the potential above, the minimum of the effective potential
1820: \begin{equation}
1821: V_{eff}(R) = V_{add}(R) + \rho_m (1 + \frac{\xi}{2} R^2)
1822: \end{equation}
1823: is given by
1824: \begin{equation}
1825: R_{min}^{n+2}=n\left(\frac{M}{m_{pl}}\right)^{n} \frac{V_{add}(R_{min})}{\xi \rho_m},
1826: \end{equation}
1827: (see formulas \ref{approx},\ref{fattori}).
1828: 
1829: For $V_{add}=O( M^4)$ this leads to a cosmological value for the radion field given by
1830: \begin{equation}
1831: R_{\infty}^{n+2}=O\left(\frac{M}{m_{pl}}\right)^{n}.
1832: \end{equation}
1833: Notice that $R_{\infty}=O(10^{-10})$ for $n=1$ implying that solar tests are
1834: automatically satisfied.
1835: 
1836: The mass of the radion at the minimum reads
1837: \begin{equation}\label{massarad}
1838: \frac{m^2}{H^2}= 3\xi \Omega_m \left[ n(n+2)  + n^2 (\frac{M}{m_{pl}})^n
1839:   \frac{1}{R^n} \right].
1840: \end{equation}
1841: The interesting regime is obtained for small distances, then the matter density is high and $m \gg H$, the field "sits" on 
1842: the minimum of the effective potential (i.e. the minimum acts as an attractor): the radion is stabilized. 
1843: 
1844: Although the chameleon mechanism does work also without runaway potentials \cite{gubserk}, 
1845: the presence of the potential \ref{champot} is crucial in our model. Had we considered only
1846: the classical potential \ref{poti} for the radion, we would have been forced to face the following problems:
1847: \begin{itemize}
1848: \item It is not possible to achieve the competition between the 
1849: potential ($V_{class} \sim Cosh(R)^{\frac{4-4 \alpha^2}{1+2 \alpha^2}}$) and the matter coupling (see formula \ref{fattori}). 
1850: \item The value $R=0$ for the radion is not forbidden by the classical potential. 
1851: The falling of the hidden brane into the naked singularity ($R=0$) must be avoided if our intention is to maintain 
1852: a supergravity approximation \cite{brax1}. 
1853: \end{itemize}
1854: 
1855: It seems noteworthy that the Q-modulus is not a chameleon 
1856: since it is not possible to achieve the desired competition between the potential ($V_{class}(Q) \sim Q^{-\frac{3}{\beta}}$) 
1857: and the coupling to matter ($A(Q) \sim Q^{-\frac{1}{2 \beta}}$). To stabilize the Q-modulus we suggest to exploit the 
1858: Casimir energy of the bulk spinor evaluated in the previous section. 
1859: 
1860: 
1861: \subsection{Q-modulus stabilization: the Casimir energy}
1862: The stabilization of the system is based on the presence of (A) {\it two} 
1863: scalar degrees of freedom (the positions of the two branes are parametrized
1864: independently) and of (B) a cosmological attractor 
1865: that pushes the hidden brane towards the singularity.
1866: 
1867: 
1868: As mentioned above, the Casimir energy of the massive spinor 
1869: guarantees a stable geometrical configuration for the 
1870: 5D bag. The connection 
1871: between the bag and the braneworld scenario is the conformal transformation (\ref{trconf}). 
1872: Had we had only one scalar field in the low energy description of the model, than 
1873: only the distance between the two branes would have been physically 
1874: relevant (consider the radion in Randall-Sundrum two branes model) or, equivalently, 
1875: it would have been possible to shift the zero of the fifth dimension: moving the two branes away in the bulk while keeping fixed the
1876: distance between them. In this case the conformal transformation (\ref{trconf})
1877: would not have mapped the two branes set up into a ball, since only the bulk singularity corresponds to the centre of the ball.
1878: 
1879: The second crucial ingredient is the presence of a cosmological 
1880: attractor that pushes the hidden brane towards the bulk singularity \cite{palma}. The visible brane evolution
1881: becomes independent of the hidden brane and the braneworld system is described by a {\it single}-scalar-tensor theory. 
1882: For this reason 
1883: we will assume a very small value for the R-modulus (the hidden brane is 
1884: close to the singularity/bag centre) in a late-time cosmology
1885: we are interested in. In this way we can "forget" about the R modulus (it is fixed at small values and stabilized by the chameleon
1886: mechanism) if the 
1887: efficiency of the attractor is high enough. The Q-modulus corresponds to the remaining scalar degree of freedom (position of the 
1888: visible brane/bag surface) and the last step in the stabilization process will be to add the Casimir energy
1889: as a stabilizing contribution for the bag surface. 
1890: 
1891: To proceed further, it would be necessary to study in full detail the role of the conformal transformation to the Einstein
1892: frame (i.e. the cocycle function, for a pedagogical treatment see \cite{kirsten}) and write the full moduli potential. 
1893: However, we will consider
1894: a reasonable value for the Q-modulus ($Q \sim 1$) for the discussion of the neutrino mass.
1895: 
1896: 
1897: \subsection{Parameters Space: Neutrino Dark Energy}
1898: 
1899: Here is a possible choice of parameters.
1900: We will proceed stepwise: \\
1901: 1)set the fundamental scale of gravity as 
1902: \beq
1903: k_5^2 = M_{fund}^{-3}, \\
1904: M_{fund} \sim M_{Planck} \sim 10^{18} GeV.
1905: \eeq
1906: \\
1907: 2)Choose $ \alpha^2 \lesssim 10^{-6}$
1908: in order to face with phenomenological constraints (\cite{palma, gonz}).
1909: \\
1910: 3)Set the tension scale and the 5D neutrino mass near the fundamental scale of gravity:\\
1911: \beq
1912: k \sim m \sim 10^{18} GeV.
1913: \eeq
1914: 4) The Newton's constant as measured by 
1915: Cavendish experiments in the visible brane is given by (see also \cite{palma}): 
1916:  \\
1917: \beq
1918: \frac{Q^2}{k k_5^2 (1+2 \alpha^2) a^2(\phi)}=\frac{1}{8 \pi G}= (10^{18} GeV)^2.
1919: \eeq
1920: \\
1921: 
1922: For the screening potential of the form (\ref{champot})  
1923: the mass scale has to be chosen such that $M \approx 10^{-3}$~eV if  
1924: the radion must be a dark energy candidate. This tuning is no more
1925: than that required by a cosmological constant.
1926: 
1927: 
1928: The classical potential (\ref{poti}) shows a scale given by
1929: \beq
1930: V_{class} \simeq \frac{6 (T-1) k}{k_5^2} e^{\frac{-12 \alpha^2 S}{1 + 2 \alpha^2}} \simeq 6 (T-1) M_{Planck}^4.
1931: \eeq
1932: \\
1933: The correct dark energy scale is recovered imposing the usual (cosmological constant) fine-tuning $T-1 \sim 10^{-120}$.\\
1934: 
1935: On cosmological distances, if we consider $n=15$ in the potential we obtain $R_{\infty} \sim 10^{-27}$. In this way the 
1936: neutrino mass is strongly suppressed by the large separation of the branes. In fact, choosing 
1937: \beq
1938: m=\frac{k}{20}
1939: \eeq
1940: and evaluating the neutrino mass in the Einstein frame for $\bar Q \sim 1$ and $\bar R \sim 10^{-27}$ (corresponding
1941: to the minimum of the moduli potential after the discussed stabilization), we can write
1942: \beq
1943: \tilde m_{\nu}= Y_5 f_0^R(\phi) v_b \frac{\sqrt{a(\phi)}}{Q} \sim 10^{-3} eV.
1944: \eeq
1945: 
1946: Remarkably, it is possible to obtain a meV-mass for neutrinos with a reasonable branes configuration in five dimensions.
1947: Happily, in the minimum $(\bar Q, \bar R)$ of the total potential we can write
1948: \beq
1949: \frac{V_{tot}^{1/4}(\bar Q,\bar R)}{\tilde m_{\nu}(\bar Q,\bar R)} = f(T, k_5, k, m, \alpha, M, n ) \sim 1,
1950: \eeq
1951: without fine-tuning the parameters (with the only exception of the usual cosmological constant fine-tuning, a 
1952: proper choice of $n=15$ in the potential and a small value of $\alpha$).
1953: With this last configuration the radion mass on cosmological distances is $m \sim 10 H$, the 
1954: field can roll and the neutrino mass is {\it variable} with time. Also, notice that the neutrino mass is {\it space}-dependent: 
1955: this can be a source of phenomenological consequences.
1956: 
1957: %\begin{figure}[t]
1958: %\centering
1959: %\includegraphics[width=10cm]{proviamotesto}
1960: %\caption{Total potential (classical + Casimir) for the Q modulus. 
1961: %The energy scale is fixed imposing $k=10^{30}$ (meV scale). We have chosen $\epsilon \simeq 6 \times 10^{-3} $ in (5.94).}
1962: %\end{figure}
1963: 
1964: \section{Conclusions}
1965: In this paper a braneworld model for neutrino Dark Energy has been presented. The model is a supersymmetric extension
1966: of the Randall-Sundrum two branes model. After supersymmetry breaking the system 
1967: becomes dynamical and a (classical) moduli potential is generated. 
1968: 
1969: In this article we have achieved the following results: \\
1970: 1) In our model a 4D neutrino mass as a function of two scalar 
1971: degrees of freedom (moduli) has been explicitly calculated rather than inserted by-hand.
1972: 
1973: 2) The moduli stabilization issue is addressed with a de Sitter geometry for the branes. The chameleon 
1974: mechanism is exploited to stabilize the radion following the approach of \cite{radion}. As far as the stabilization of the remaining
1975: modulus is concerned, we suggest to exploit the
1976: Casimir energy of massive neutrinos as a stabilizing contribution.
1977: 
1978: 3) A direct relationship has been established between neutrino mass scale and Dark Energy scale.
1979: We stress again that the only dynamical behaviour associated to moduli is encoded into the chameleon field R.
1980: Namely, on cosmological distances, the densities are tiny and the 
1981: radion can roll on cosmological time scales, while on smaller scales, 
1982: the chameleon acquires a large mass (radion stabilization).
1983: The neutrino mass is {\it variable} with time on cosmological distances and possible deviations
1984: from a pure cosmological constant behaviour can be observed in the DE sector 
1985: (we will further discuss these issues in a future work). The neutrino mass is also space dependent and this can imply
1986: phenomenological consequences.
1987: 
1988: 
1989: 
1990: It seems to us that this model suggests a stronger connection 
1991: between DE and neutrino physics. To the best of our knowledge it is the first attempt 
1992: of connecting neutrino mass to Dark Energy
1993: from the standpoint of brane models. 
1994: 
1995: \acknowledgements{Special thanks are due to Antonio Masiero and Massimo Pietroni. 
1996: They patiently guided me and they were a constant source of inspiration during the development of this article. }
1997: 
1998: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1999: \appendix
2000: 
2001: \section{D-polynomials}
2002: 
2003: In this appendix we collect the basic formulas necessary to obtain the explicit form of
2004: the polynomials \ref{asympol}.
2005: We start remembering the uniform asymptotic expansion of the Bessel functions $I_\nu (k)$ as $\nu, k \rightarrow \infty$ with 
2006: $\nu /k$ fixed. One has 
2007: 
2008: \beq I_{\nu} (\nu z) \sim \frac
2009: 1 {\sqrt{2\pi \nu}}\frac{e^{\nu \eta}}{(1+z^2)^{\frac 1
2010: 4}}\left[1+\sum_{k=1}^{\infty} \frac{u_k (t)} {\nu
2011: ^k}\right], \nn
2012: \label{eq2.9} 
2013: \eeq 
2014: with $t=1/\sqrt{1+z^2}$ and $\eta=\sqrt{1+z^2}+\ln [z/(1+\sqrt{1+z^2})]$.
2015: The coefficients are determined by the recurrence formula \cite{abramo}
2016: \beq
2017: u_{k+1} (t) =\frac 1 2 t^2 (1-t^2) u'_k (t) +\frac 1 8
2018: \int\limits_0^t d\tau\,\, (1-5\tau^2 ) u_k (\tau ), \nn
2019: \label{eq2.10}
2020: \eeq 
2021: starting with $u_0 (t) =1$. As is clear, all the $u_k (t)$
2022: are polynomials in $t$.
2023: 
2024: Since boundary conditions involved Bessel functions and their derivatives, we need 
2025: also the expansion \cite{abramo}
2026: \beq
2027: I_{\nu} ' (\nu z )\sim
2028: \frac 1 {\sqrt{2\pi \nu}} \frac{e^{\nu \eta} (1+z^2)^{1/4}} z
2029: \left[1+\sum_{k=1}^{\infty}\frac{v_k(t)}{\nu^k}\right], \nn \label{eq2.48}
2030: \eeq
2031: with the $v_k (t)$ determined by
2032: \beq
2033: v_k (t) = u_k (t) +t (t^2 -1) \left[
2034: \frac 1 2 u_{k-1} (t) +t u_{k-1} ' (t) \right]. \nn \label{eq2.49}
2035: \eeq
2036: 
2037: Exploiting the previous formulas the relevant asymptotics can be found. With the notation
2038: \beq
2039: \Sigma_1 = \left[1+\sum_{k=1}^{\infty} \frac{u_k (t)} {\nu
2040: ^k}\right], \Sigma_2 = \left[1+\sum_{k=1}^{\infty} \frac{v_k (t)} {\nu
2041: ^k}\right], mR=x \nn
2042: \eeq
2043: it reads
2044: 
2045: \beq
2046: && \ln[I^2_{j}(zj)(1 + \frac{1}{z^2}-\frac{2x}{z^2 j}) + I'^2_j(zj) + \frac{2}{zj}(x-j)I_j(zj)I'_j(zj)]\simeq \nn \\
2047: && \simeq \ln\left\{\frac{e^{2j \eta}(1+z^2)^{1/2}}{2 \pi j z^2} [(1+\frac{1}{z^2} - \frac{2x}{z^2 j}) \frac{z^2}{1+z^2} \Sigma_1^2 + \Sigma_2^2 +
2048:  \frac{2}{j} (x-j) \frac{1}{(1+z^2)^{1/2}} \Sigma_1 \Sigma_2]  \right\} = \nn \\
2049: &&= \ln\left\{(\frac{e^{2j \eta}(1+z^2)^{1/2}}{2 \pi j z^2} 2(1-t))\frac{(1-\frac{2xt^2}{j}) \Sigma_1^2 + \Sigma_2^2 +
2050: 2t \frac{x-j}{j} \Sigma_1 \Sigma_2 }{2(1-t)}  \right\}. \nn
2051: \eeq
2052: 
2053: In this way we are led to the definition of D-polynomials as follows
2054: \beq
2055: \sum_{k=1}^{\infty} \frac{D_k(x,t)}{j^k}=\ln\left\{\frac{1}{2(1-t)}[(1-\frac{2xt^2}{j}) \Sigma_1^2 + \Sigma_2^2 +
2056: 2t \frac{x-j}{j} \Sigma_1 \Sigma_2] \right\}. \nn
2057: \eeq
2058: 
2059: 
2060: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2061: \begin{thebibliography}{4}
2062: 
2063: \bibitem{1}S. Perlmutter et al., Nature \textbf{391}, 51 (1998); A. G. Riess
2064: et al., Astron. J. \textbf{116}, 1009 (1998); P. de Bernardis et al.,
2065: Nature \textbf{404}, 955 (2000); S. Hanay et al., Astrophys. J. Lett.
2066: \textbf{545}, L5 (2000); N. W. Alverson et al., Astrophys. J. \textbf{568},
2067: 38 (2002); D. N. Spergel et al., Astrophys. J. Suppl. \textbf{148},
2068: 175 (2003).
2069: 
2070: \bibitem{quint-review}
2071: For recent reviews, see for example: 
2072: S.~M.~Carroll, Living Rev.\ Rel.\  {\bf 4} (2001) 1;
2073: P.~J.~E.~Peebles and B.~Ratra,
2074: %``The cosmological constant and dark energy,''
2075: Rev.\ Mod.\ Phys.\  {\bf 75} (2003) 559;
2076: T.~Padmanabhan, Phys.\ Rept.\  {\bf 380} (2003) 235.
2077: 
2078: \bibitem{4}L. Amendola, Phys. Rev. D \textbf{62}, 043511 (2000); L. Amendola
2079: and D. Tocchini-Valentini, Phys.Rev. D \textbf{64}, 043509 (2001);
2080: L. Amendola and D. Tocchini-Valentini, Phys.Rev. D \textbf{66}, 043528
2081: (2002) . 
2082: 
2083: \bibitem{5}L. Amendola, M. Gasperini, D. Tocchini-Valentini and C. Ungarelli, Phys.Rev. D \textbf{67}, 043512 (2003); L. Amendola, Mon. Not. R. Astron. Soc. \textbf{342}, 221 (2003); M. Gasperini, hep-th/0310293,  in Proc. of the Int. Conf. on {\em Thinking, Observing and Mining the Universe}, Sorrento 2003, eds. G. Longo and G. Miele (World Scientific, Singapore), in press. 
2084: 
2085: \bibitem{5a}D. Comelli, M. Pietroni and
2086: A. Riotto, Phys. Lett. B \textbf{571}, 115 (2003);
2087: M. Pietroni, Phys. Rev. D {\bf 67}, 103523 (2003); U. Franca and R.
2088: Rosenfeld, Phys.Rev.D {\bf 69} (2004) 063517;  G. R. Farrar and P. J. E. Peebles, Astrophys. J {\bf 604}, 1 (2004); A. V. Maccio et al. Phys. Rev. D \textbf{69}, 123516 (2004); M. Pietroni Phys. Rev. D {\bf 72} 043535 (2005).
2089: 
2090: \bibitem{5b}G. Huey and B. D. Wandelt, astro-ph/0407196.
2091: 
2092: \bibitem{amen} L. Amendola, M. Gasperini and F.Piazza, astro-ph/0407573.
2093: \bibitem{vamps} J.~A.~Casas, J.~Garcia-Bellido and M.~Quiros Class.\ Quant.\ Grav.\  {\bf 9}, 1371 (1992);
2094: Greg W. Anderson, Sean M. Carroll, astro-ph/9711288.
2095: \bibitem{6}M. Gasperini, F. Piazza and G. Veneziano, Phys. Rev. D \textbf{65},
2096: 023508 (2002).
2097: \bibitem{PT}F. Piazza and S. Tsujikawa, JCAP {\bf 0407},  004 (2004). 
2098: \bibitem{7}G. Veneziano, JHEP \textbf{0206}, 051 (2002). 
2099: \bibitem{gaspstab}M. Gasperini, Phys. Rev. D {\bf 64}, 043510 (2001).
2100: \bibitem{uzan} J.P. Uzan, Rev.Mod.Phys.75:403,2003; E. Fischbach and C.L. Talmadge "The search for non-newtonian gravity" / Springer-Verlag, New York 1999;
2101: E.G. Adelberger, B.R. Heckel and A.E. Nelson, Ann.Rev.Nucl.Part.Sci. {\bf 53} (2003) 77-121.
2102: \bibitem{bd}  P.~Jordan, {\em Schwerkaft und Weltall} (Vieweg, Braunschweig,
2103: 1955); M. Fierz, Helv. Phys. Acta {\bf 29}, 128 (1956); C.~Brans and
2104: R.H.~Dicke, Phys. Rev. {\bf 124}, 925 (1961).
2105: \bibitem{maeda}
2106: K.I. Maeda, Y. Fujii
2107: {\it The Scalar-Tensor theory of gravitation} - Cambridge University Press, 2003.
2108: 
2109: \bibitem{damour}
2110: T. Damour, gr-qc/9606079.
2111: 
2112: \bibitem{will}
2113: C.M. Will
2114: Living Rev.Rel.4:4,2001.
2115: 
2116: \bibitem{dam3a} T.~Damour and K.~Nordtvedt, Phys. Rev. {\bf D48}, 3436.
2117: (1993).
2118: 
2119: \bibitem{dam3b} T.~Damour and A.M.~Polyakov, Nucl.\ Phys. {\bf B423},
2120: 532 (1994).
2121: 
2122: \bibitem{max} 
2123: N.~Bartolo and M.~Pietroni, Phys.\ Rev.\ D {\bf 61} (2000) 023518.
2124: 
2125: \bibitem{tracker}
2126: A.~R.~Liddle and R.~J.~Scherrer,
2127:  %``A classification of scalar field potentials with cosmological scaling
2128: %solutions,''
2129: Phys.\ Rev.\ D {\bf 59} (1999) 023509;
2130: I.~Zlatev, L.~M.~Wang and P.~J.~Steinhardt,
2131: %``Quintessence, Cosmic Coincidence, and the Cosmological Constant,''
2132: Phys.\ Rev.\ Lett.\  {\bf 82} (1999) 896;
2133: P.~J.~Steinhardt, L.~M.~Wang and I.~Zlatev,
2134: %``Cosmological Tracking Solutions,''
2135: Phys.\ Rev.\ D {\bf 59} (1999) 123504;
2136: B.~Ratra and P.~J.~E.~Peebles,
2137: %``Cosmological Consequences Of A Rolling Homogeneous Scalar Field,''
2138: Phys.\ Rev.\ D {\bf 37} (1988) 3406;
2139: P.~J.~E.~Peebles and B.~Ratra,
2140: %``Cosmology With A Time Variable Cosmological 'Constant',''
2141: Astrophys.\ J.\  {\bf 325} (1988) L17.
2142: 
2143: \bibitem{catena}
2144: R. Catena, N. Fornengo, A. Masiero, M. Pietroni, F. Rosati
2145: Phys.Rev.D {\bf 70} (2004) 063519.
2146: 
2147: 
2148: \bibitem{caffe2}
2149: A. Coc, K. A. Olive, J.P. Uzan, E. Vangioni, astro-ph/0601299.
2150: 
2151: \bibitem{justin} J. Khoury and A. Weltman, Phys.Rev.Lett.{\bf 93}, 171104 (2004); J. Khoury and A. Weltman, 
2152: Phys.Rev.D{\bf 69}, 044026 (2004).
2153: 
2154: \bibitem{fnw}
2155: R. Fardon, A.E. Nelson and N. Weiner, JCAP 0410:005,2004  [astro-ph/0309800].
2156: 
2157: \bibitem{peccei}
2158: R. Peccei Phys.Rev. D {\bf 71} 023527 (2005) [hep-ph/0411137].
2159: 
2160: \bibitem{neut}
2161: P. Gu, X. Wang, X. Zhang, Phys. Rev. D {\bf 68}, 087301 (2003);
2162: P. Q. Hung, H. Pas Mod.Phys.Lett.A20:1209-1216,2005 [astro-ph/0311131];
2163: X. Bi, P. Gu, X. Wang, X. Zhang Phys.Rev. D {\bf 69} 113007 (2004) [hep-ph/0311022];
2164: P. Gu and X.J. Bi, Phys. Rev. D {\bf 70}, 063511 (2004);
2165: D.B. Kaplan, A.E. Nelson and N. Weiner Phys.Rev.Lett.{\bf 93} 091801 (2004) [hep-ph/0401099];
2166: H. Li, Z. Dai, X. Zhang, Phys. Rev. D {\bf 71}, 113003 (2005);
2167: X. Zhang, hep-ph/0410292;
2168: E.I. Guendelman and A.B. Kaganovich, hep-th/0411188;
2169: X. Bi, B. Feng, H. Li, X. Zhang hep-ph/0412002;
2170: R. Barbieri, L.J. Hall, S.J. Oliver, A. Strumia Phys.Lett.B {\bf 625} 189-195 (2005) [hep-ph/0505124];
2171: M. Cirelli, M.C. Gonzalez-Garcia, C. Pena-Garay, Nucl.Phys.B {\bf 719} 219-233 (2005) [hep-ph/0503028];
2172: R. Takahashi, M. Tanimoto hep-ph/0507142;
2173: R. Fardon, A. E. Nelson, N. Weiner hep-ph/0507235;
2174: N. Afshordi, M. Zaldarriaga, K. Kohri Phys.Rev.D {\bf 72} 065024 (2005) [astro-ph/0506663];
2175: A.W. Brookfield, C. van de Bruck, D.F. Mota, D. Tocchini-Valentini, astro-ph/0503349;
2176: M. Blennow, T. Ohlsson, W. Winter, hep-ph/0508175;
2177: H. Li, B. Feng, J. Q. Xia, X. Zhang, astro-ph/0509272;
2178: V. Barger, D. Marfatia, K. Whisnant, hep-ph/0509163;
2179: V. Barger, P. Huber, D. Marfatia, Phys.Rev.Lett. {\bf 95} 211802 (2005) [hep-ph/0502196];
2180: R. Horvat, astro-ph/0505507;
2181: N. Weiner, K. Zurek, hep-ph/0509201;
2182: M. Honda, R. Takahashi, M. Tanimoto, hep-ph/0510018;
2183: P. Gu, X.J. Bi, X. Zhang, hep-ph/0511027;
2184: M.C. Gonzalez-Garcia, P.C. de Holanda, R. Zukanovich Funchal, hep-ph/0511093;
2185: T. Schwetz, W. Winter, hep-ph/0511177;
2186: P. Gu, X.J. Bi, B. Feng, B.L. Young, X. Zhang, hep-ph/0512076;
2187: A. W. Brookfield, C. van de Bruck, D.F. Mota, D. Tocchini-Valentini, astro-ph/0512367;
2188: R. Takahashi, M. Tanimoto, astro-ph/0601119.
2189: 
2190: \bibitem{caffe}
2191: Y. V. Dumin, astro-ph/0507381; L. Iorio, gr-qc/0511137.
2192: 
2193: \bibitem{RS1} L.~Randall and R.~Sundrum,
2194:  Phys.\ Rev.\ Lett.\ {\bf 83} (1999) 3370.
2195: \bibitem{RS2} L.~Randall and R.~Sundrum,
2196:  Phys.\ Rev.\ Lett.\ {\bf 83} (1999) 4690.
2197:  
2198: \bibitem{extra}
2199: V.A. Rubakov Phys.Usp. 44 (2001) 871-893/ Usp.Fiz.Nauk 171 (2001) 913-938; G.Gabadadze hep-ph/0308112;
2200: C. Csaki hep-ph/0404096; R. Maartens, Living Rev.Rel. {\bf 7}, 7 (2004); 
2201: Ph. Brax, C. van de Bruck and A.-C. Davis, Rept.Prog.Phys. {\bf 67}, 2183 (2004).
2202: 
2203: \bibitem{HW}
2204: P. Horava and E. Witten, Nucl. Phys. B {\bf 460} (1996) 506; Nucl. Phys. B {\bf 475} (1996) 94.
2205: \bibitem{witten}
2206: E. Witten, Nucl. Phys. B {\bf 471} (1996) 135. 
2207: 
2208: \bibitem{brax} P. Brax, C. van de Bruck, A.C. Davis, C.S. Rhodes, Phys.Rev.D {\bf 67}, 023512 (2003). 
2209: \bibitem{brax1} Ph. Brax and A.-C. Davis, Phys.Lett.B{\bf 497}, 289 (2001).
2210: 
2211: \bibitem{gionni}
2212: C. Johnson, {\it D-branes}, Cambridge University Press (2003).
2213: \bibitem{gonz} G.A. Palma and A.-C. Davis, Phys.Rev.D {\bf 70}, 064021 (2004);
2214: G.A. Palma and A.-C. Davis, Phys.Rev.D {\bf 70}, 106003 (2004).
2215: \bibitem{baggerredi} J. Bagger and M. Redi, JHEP {\bf 0404} (2004) 031. 
2216: \bibitem{marsiglia}
2217: Ph. Brax, C. van de Bruck, A.C. Davis and C.S. Rhodes, e-Print Archive: hep-ph/0309180;
2218: Ph. Brax, C. van de Bruck, A.C. Davis and C.S. Rhodes, e-Print Archive: hep-ph/0309181.
2219: 
2220: \bibitem{casimir}
2221: K.A. Milton {\it The Casimir effect - Physical manifestation of the zero-point energy}, World Scientific 2001;
2222: K.A. Milton hep-th/0406024;
2223: D.V. Vassilevich, Phys. Rep. {\bf 388} (2003) 279;
2224: M. Bordag, U. Mohideen, V.M. Mostepanenko, Phys. Rep. {\bf 353} (2001) 1.
2225: 
2226: \bibitem{TEN} E.~Elizalde, {\it Ten physical applications of spectral zeta 
2227: functions}, Lect.\ Notes Phys.\ {\bf M35} (1995).
2228: \bibitem{kirsten}
2229: K. Kirsten, {\it Spectral functions in mathematics and physics},Chapman and Hall/CRC, Boca Raton, 2002.
2230: E. Elizalde, S.D. Odintsov, A. Romeo, A.A. Bytsenko, S. Zerbini, {\it Zeta regularization techniques with applications}, World Scientific, Singapore, 1994.
2231: \bibitem{MOSS2} I. G. Moss in {\it The Future of Theoretical Physics}, essays in 
2232: honour of S. W. Hawking's 60th birthday (Cambridge CUP).
2233: \bibitem{buchbinder}
2234: E.I. Buchbinder and B.A. Ovrut, Phys. Rev. D {\bf 69}, 086010 (2004).
2235: 
2236: \bibitem{webster}
2237: S.L. Webster and A-C. Davis, hep-th/0410042.
2238: 
2239: \bibitem{grossman} Y. Grossman and M. Neubert, Phys.Lett. B {\bf 474} (2000),361-371.
2240: \bibitem{naked} Ph. Brax and A.C. Davis, Phys. Lett. B {\bf 513}: 156-162 (2001).
2241: \bibitem{radion} Ph. Brax, C. van de Bruck and A.C. Davis, JCAP {\bf 0411} 004 (2004).
2242: \bibitem{scalarf}
2243: W.D. Goldberger, I.Z. Rothstein, Phys. Lett. B {\bf 491} 339 (2000);
2244: R. Hofmann, P. Kanti, M. Pospelov, Phys. Rev. D {\bf 63}, 124020 (2001);
2245: A.A. Saharian, M.R. Setare, Phys. Lett. B {\bf 552}, 119 (2004);
2246: A. Knapman, D.J. Toms, Phys. Rev. D {\bf 69}, 044023 (2004).
2247: \bibitem{GPT} J.~Garriga, O.~Pujol\`{a}s and T.~Tanaka,
2248:  Nucl.\ Phys.\ B {\bf 605} (2001) 192.
2249:  \bibitem{FT} A.~Flachi and D.~J.~Toms Nucl.\ Phys.\ B {\bf 610} (2001) 144.
2250: \bibitem{FMT} A.~Flachi, I.~G.~Moss and D.~J.~Toms, Phys.\ Lett.\ B {\bf 518} 
2251: (2001) 153.
2252: \bibitem{FMT2} A.~Flachi, I.~G.~Moss and D.~J.~Toms, Phys.\ Rev.\ D {\bf 64} 
2253: (2001) 105029.
2254: \bibitem{detuned}
2255: Ph. Brax, N. Chatillon, JHEP {\bf 0312} (2003) 026.
2256: \bibitem{WN} W.~Naylor and M.~Sasaki, Phys.\ Lett.\ B {\bf 542} (2002) 289.
2257: \bibitem{fermionc}
2258: I.G. Moss, W. Naylor, W. Santiago-German, M.Sasaki, Phys. Rev. D {\bf 67}, 125010 (2003).
2259: \bibitem{norman}
2260: J.P. Norman, Phys. Rev. D {\bf 69}, 125015 (2004).
2261: \bibitem{susyrs}
2262: R.~Altendorfer, J.~Bagger and D.~Nemeschansky, Phys.\ Rev. D {\bf 63} (2001) 125025
2263: [arXiv:hep-th/0003117]; A.~Falkowski, Z.~Lalak and S.~Pokorski, Phys.\ Lett.\
2264: B {\bf 491} (2000) 172 [arXiv:hep-th/0004093]; E.~Bergshoeff, R.~Kallosh and A.~Van Proeyen, JHEP {\bf 0010} (2000) 033
2265: [arXiv:hep-th/0007044]; M. Zucker, Phys. Rev. D {\bf 64} (2001) 024024 [arXiv:hep-th/0009083]; A. Falkowski, Z. Lalak and S. Pokorski, Phys. Lett. B {\bf 509} (2001) 337
2266: [arXiv:hep-th/0009167]; A.~Falkowski, Z.~Lalak and S.~Pokorski, Nucl. Phys. B
2267: {\bf 613} (2001) 189-217 [arXiv:hep-th/0102145]; J. Bagger, D. Nemeschansky and R.-J. Zhang, JHEP {\bf 0108} (2001)
2268: 057 [arXiv:hep-th/0012163]; J. Bagger and D. Belyaev, Phys. Rev. D {\bf 67} (2003) 025004
2269: [arXiv:hep-th/0206024]; P.~Brax and Z.~Lalak, Acta Phys.\ Polon.\ B {\bf 33} (2002)
2270: 2399 [arXiv:hep-th/0207102]; Z.~Lalak and R.~Matyszkiewicz, Nucl.\ Phys.\ B {\bf 649} (2003) 389
2271: [arXiv:hep-th/0210053]; Z. Lalak and R. Matyszkiewicz, Phys.Lett. B {\bf 562} (2003) 347-357
2272: [arXiv:hep-th/0303227]; J. Bagger and D. Belyaev, JHEP
2273: {\bf 0306} (2003) 013 [arXiv:hep-th/0306063].
2274: \bibitem{GS} J.~Garriga and M.~Sasaki, Phys.\ Rev.\ D {\bf 62} (2000) 043523.
2275: \bibitem{GEN} U.~Gen and M.~Sasaki, Prog.\ Theor.\ Phys.\ {\bf 105} (2001) 59.
2276: \bibitem{MOSS} I. G. ~Moss, ``Quantum theory, black holes and inflation", {\it 
2277: Wiley, (1996)}.
2278: \bibitem{mit}
2279: A. Chodos, R.L. Jaffe. K. Johnson, C.B. Thorn and V. Weisskopf,
2280: Phys. Rev. D {\bf 9} (1974) 3471; A. Chodos, R.L. Jaffe. K. Johnson and C.B. Thorn,
2281: Phys. Rev. D {\bf 10} (1974) 2599.
2282: \bibitem{bag}
2283: E. Elizalde, M. Bordag and K. Kirsten,  J.Phys.A31:1743-1759,1998
2284: e-Print Archive: hep-th/9707083.
2285: \bibitem{bk}
2286: M. Bordag and K. Kirsten, hep-th/9501064.
2287: \bibitem{bek} 
2288: M. Bordag, E. Elizalde and K. Kirsten, J. Math. Phys. {\bf 37} (1996) 895. 
2289: \bibitem{bekl96}
2290: M. Bordag, E. Elizalde, K. Kirsten and S. Leseduarte, Phys.Rev.D {\bf 56} 4896-4904,1997 
2291: e-Print Archive: hep-th/9608071.
2292: \bibitem{begk}
2293: M. Bordag, E. Elizalde, B. Geyer and K. Kirsten, Commun. Math. Phys. {\bf 179}
2294: (1996) 215.
2295: 
2296: \bibitem{bdk}
2297: M. Bordag, S. Dowker and K. Kirsten,  Commun. Math. Phys. {\bf 182} (1996) 371.
2298: 
2299: \bibitem{abdk}
2300: J. Apps, M. Bordag, S. Dowker and K. Kirsten, Class.Quant.Grav. {\bf 13} (1996) 2911-2920.
2301: 
2302: \bibitem{eli1}  
2303: E. Elizalde, M. Lygren and D.V. Vassilevich,
2304: %{\it Antisymmetric tensor fields on spheres: functional determinants
2305: %and non-local counterterms},
2306:  J. Math. Phys. {\bf 37} (1996) 3105.
2307: \bibitem{Mth}
2308: I. Moss, J.P. Norman, JHEP {\bf 0409} (2004) 020.
2309: 
2310: \bibitem{eli2}
2311: E. Elizalde, M. Lygren and D.V. Vassilevich,
2312: Commun. Math. Phys. {\bf 183} (1997) 645.
2313: 
2314: \bibitem{stuart}
2315: J. Dowker and K. Kirsten, Commun.Anal.Geom. 7 (1999) 641-679 [hep-th/9608189].
2316: \bibitem{GPT2} J.~Garriga, O.~Pujol\`{a}s and T.~Tanaka, Nucl.Phys.B {\bf 655} (2003) 127-169.
2317: \bibitem{FGPT}
2318: A.~Flachi, J.~Garriga, O.~Pujolas and T.~Tanaka, JHEP {\bf 0308} (2003) 053.
2319: \bibitem{palma} G.A. Palma, hep-th/0511170.
2320: \bibitem{gubserk} S.S. Gubser and J. Khoury,  Phys.Rev.D {\bf 70} (2004) 104001.
2321: \bibitem{abramo}
2322: M.~Abramowitz and I.A.~Stegun.
2323: {\em Handbook of Mathematical Functions}.
2324: Dover, New York,  (1970).
2325: 
2326: 
2327: 
2328: 
2329: 
2330: 
2331: 
2332: \bibitem{ENOO} E.~Elizalde, S.~Nojiri, S.~D.~Odintsov and S.~Ogushi, Phys.Rev. D {\bf 67} (2003) 063515.
2333: \bibitem{damournord} T. Damour and K. Nordtvedt, Phys.Rev.Lett.{\bf 70}, 2217 (1993).
2334: \bibitem{cosmocham} Ph. Brax, C. van de Bruck, A.C. Davis, J. Khoury and A. Weltman, Phys.Rev.D {\bf 70}, 123518 (2004).
2335: \bibitem{dpv} T.~Damour, F.~Piazza and G.~Veneziano, Phys.\ Rev.\ D {\bf 66} (2002) 046007;
2336: T.~Damour, F.~Piazza and G.~Veneziano, Phys.\ Rev.\ Lett.\  {\bf 89} (2002) 081601.
2337: 
2338: 
2339: 
2340: 
2341: \end{thebibliography}
2342: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2343: \end{document}\Documents\Documents
2344: