hep-th0601066/GHB.tex
1: \documentclass[11pt,a4paper,twoside]{article}
2: 
3: \setlength{\textwidth}{6.0 in}
4: %\setlength{\oddsidemargin}{1.5 cm}
5: \setlength{\evensidemargin}{0.0cm}
6: \def\theequation{\thesection.\arabic{equation}}
7: \usepackage{axodraw}
8: %\usepackage[italian]{babel}
9: \usepackage[latin1]{inputenc}
10: \def\be{\begin{equation}}
11: \def\ee{\end{equation}}
12: \def\ba{\begin{array}}
13: \def\bacc{\begin{array} {cc}}
14: \def\ea{\end{array}}
15: \def\bea{\begin{eqnarray}}
16: \def\eea{\end{eqnarray}}
17: \def\bd{\begin{displaymath}}
18: \def\ed{\end{displaymath}}
19: \def\ha{\hat{\alpha}}
20: \def\hb{\hat{\beta}}
21: \def\D{\mathcal{D}}
22: \def\a{\alpha}
23: \def\b{\beta}
24: \def\c{\gamma}
25: \def\d{\delta}
26: \def\o{\omega}
27: \def\mv{\mathcal{V}}
28: \def\e{\mbox{det}\left(e^{\a}_m\right)}
29: \def\U{\mathcal{U}}
30: \renewcommand{\theequation}{\thesection.\arabic{equation}}
31: 
32: \begin{document}
33: \begin{center}
34: 
35: {\Large\bf On the Decoupling of Heavy Modes in Kaluza-Klein
36: Theories}
37: 
38: \vspace{1cm}
39: 
40: {\large S. Randjbar-Daemi$^a$\footnote{Email: seif@ictp.trieste.it},
41: A. Salvio$^b$\footnote{Email: salvio@sissa.it} and M.
42: Shaposhnikov$^c$\footnote{Email:
43: Mikhail.Shaposhnikov@epfl.ch}}\\
44: 
45: \vspace{.6cm}
46: 
47: {\it {$^a$ International Center for Theoretical Physics, \\Strada
48: Costiera 11, 34014 Trieste, Italy}}
49: 
50: \vspace{.4cm}
51: 
52: {\it {$^b$ International School for Advanced Studies,\\
53: Via Beirut 2-4, 34014 Trieste, Italy}}
54: 
55: \vspace{.4cm}
56: 
57: {\it {$^c$ Institut de Th\'eorie des Ph\'enom\`enes Physiques,\\
58:   \'Ecole Polytechnique F\'ed\'erale de Lausanne,\\
59:   CH-1015 Lausanne, Switzerland}} \vspace{.4cm}
60: 
61: \end{center}
62: 
63: \vspace{1cm}
64: 
65: 
66: \begin{abstract}
67: 
68: In this paper we examine the 4-dimensional effective theory for the
69: light Kaluza-Klein (KK) modes.  Our main interest is in the
70: interaction terms. We point out that the  contribution of the heavy
71: KK modes is generally needed  in order to reproduce the correct
72: predictions for the observable quantities involving the light modes.
73: As an example we study  in some detail a 6-dimensional
74: Einstein-Maxwell theory coupled to a charged scalar and  fermions.
75: In this case the contribution of the heavy KK modes are
76: geometrically interpreted as the deformation of the internal space.
77: 
78: \end{abstract}
79: 
80: 
81: \newpage
82: 
83: 
84: 
85: \tableofcontents
86: 
87: \newpage
88: 
89: \section{Introduction} \label{intro}
90: \setcounter{equation}{0}
91: 
92:  In studying the low energy physics of the
93: light modes of a 4+d dimensional theory the attention is usually
94: paid only to the spectral aspects. After determining the quantum
95: numbers of the light modes  the nature and the form of the
96: interaction terms are often assumed to be dictated by symmetry
97: arguments.  Such arguments fix the general form of all the
98: renormalilzable terms and if the effective theory is supersymmetric
99: certain relationship between the couplings can also be established
100: by supersymmetry. The masses are derived from the bilinear part of
101: the effective action and the role of the heavy modes in the actual
102: values of the masses and the couplings of the effective theory for
103: the light modes are seldom taken into account. It is, however, well
104: known from the study of the GUT's in 4-dimensions that the heavy
105: modes have an important role to play even at low energies
106: \cite{GUT}. This happens through their contributions to the
107: couplings  entering into the effective Lagrangians describing the
108: low energy physics of the light modes. According to Wilsonian
109: approach, in order to obtain an effective theory applicable in large
110: distances, the heavy modes should be integrated out \cite{Wilson,
111: Weinberg:1980wa}. The processes of "integrating out" has the effect
112: of modifying the couplings of the light modes or introducing
113: additional terms which are suppressed by inverse powers of the heavy
114: masses \cite{Appelquist:1974tg}.
115: 
116: 
117: The aim of the present paper is to examine the role of the heavy
118: modes in the low energy description of a higher dimensional theory.
119: To this end we shall basically perform two complementary
120: calculations. The first one will start from a solution of a higher
121: dimensional theory with a 4-dimensional Poincar\'e invariance and
122: develop an action functional for the light modes of the effective
123: 4-dimensional theory. This effective action generally has a local
124: symmetry which should be broken by  Higgs mechanism. Our interest is
125: in the spectrum of the broken theory. The procedure is essentially
126: what is adopted in the effective description of higher dimensional
127: theories including superstring and M-theory compactifications. In
128: this construction the heavy KK modes are generally ignored simply by
129: reasoning that their masses are of the order of the compactification
130: mass and this can be as heavy as the Planck mass. Therefore they can
131: not affect the low energy physics of the light modes.
132: 
133:  In the second approach which we shall call the $\it{ geometrical\ approach}$ we shall find a solution of the higher dimensional equations
134: with the same symmetry group as the one of the broken phase of the
135: effective 4-dimensional theory for the light modes. We shall then
136: study the physics of the 4-dimensional light modes around this
137: solution. The result for the effective 4-dimensional theory will
138: turn out to be different from the first approach.  The aim of this
139: paper is to show that the difference is precisely due to the fact
140: that in constructing the effective theory along the lines of the
141: first approach the contribution of the heavy KK modes have been
142: ignored. Indeed it will be argued - and demonstrated by working out
143: some explicit examples - that taking due care of the role of the
144: heavy modes a complete equivalence is established between the two
145: approaches.
146: 
147: To motivate the discussion in a simpler context, in section
148: \ref{simple4D} we shall work out a simple model of two coupled
149: scalar fields in 4-dimensions which will be generalized to a
150: multiplet of scalar fields in arbitrary dimensions in section
151: \ref{general}. The examples in sections \ref{simple4D} and
152: \ref{general} will clarify the relevance of the heavy modes in the
153: low energy description of the light modes. In sections \ref{def} and
154: \ref{u13} we shall study a higher dimensional (in this case six
155: dimensional) theory of Einstein-Maxwell system \cite{RSS} coupled to
156: a charged scalar and eventually also to charged fermions. Such a
157: model can arise in the compactification of string or M-theory to
158: lower dimensions.  The system has enough number of adjustable
159: parameters to allow us to go to various limits in order to establish
160: the main point of our paper. The result will of course confirm the
161: above mentioned expectation that in order to obtain a correct
162: 4-dimensional description of the physics of the light modes the
163: contribution of the heavy modes should be duly taken into
164: account\footnote{Of course this doesn't prove that the heavy modes
165: contribution never vanishes: for instance \cite{Gibbons:2003gp}
166: proves the decoupling of the heavy modes in the $(Minkowski)_4\times
167: S^2$ compactification of the 6-dimensional chiral supergravity
168: \cite{Salam:1984cj}, which is basically the supersymmetric version
169: of our 6-dimensional theory.}. This example is particularly
170: interesting because the first kind of solution will produce an
171: effective 4-dimensional gauge theory with a $SU(2)\times U(1)$
172: symmetry which will be broken to $U(1)$ by a complex triplet of
173: Higgs fields. The geometrical approach, on the other hand, will take
174: us directly to the unbroken $U(1)$ phase by deforming a round sphere
175: into an ellipsoid\footnote{ This corresponds to the magnetic
176: monopole charge of 2 as explained in section 5.  A monopole charge
177: of unity will produce a Higgs doublet of SU(2). }. In the
178: geometrical approach the W and the Z masses originate from the
179: deformation of the internal space. In this sense the standard Higgs
180: mechanism acquires a geometrical origin\footnote{It should be
181: mentioned that all of our discussion is ( semi-) classical.  To
182: include quantum and renormalization effects is beyond the scope of
183: the present paper.}. We elaborate a little more on this point in
184: section \ref{conclusions} which summarizes our results.  Some
185: technical aspects of various derivations have been detailed in the
186: appendices.
187: 
188: 
189: 
190: 
191: \section{A Simple 4D Theory} \label{simple4D}
192: \setcounter{equation}{0}
193: 
194:  Let's consider a 4-dimensional theory,
195: which contains two real scalar fields $\varphi$ and $\chi$ and with
196: the lagrangian
197: %
198: $$ \mathcal{L}=
199: -\frac{1}{2}\partial_{\mu}\varphi\partial^{\mu}\varphi-\frac{1}{2}\partial_{\mu}\chi\partial^{\mu}\chi
200: -\frac{1}{2}m^2_{\varphi}\varphi^2-\frac{1}{2}m^2\chi^2-\frac{1}{4}\lambda_{\varphi}\varphi^4
201: -\frac{1}{4}\lambda_{\chi}\chi^4-a\varphi^2\chi^2, $$
202: %
203: where $m^2_{\varphi}$, $m^2$, $\lambda_{\varphi}$, $\lambda_{\chi}$
204: and $a$ are real parameters\footnote{Of course we consider only the
205: values of these parameters such that the scalar potential is bounded
206: from below.}. Here we have the symmetry:
207: %
208: \bea &&Z_2: \varphi\rightarrow \pm \varphi, \nonumber \\
209: &&Z_2': \chi \rightarrow \pm \chi. \eea
210: %
211:  This is a very particular
212: example and of course we don't want to present any general result in this section, we just want to
213: provide a framework in which the general equivalence that we spoke about in the introduction emerges in a simple
214: way and is not obscured by technical difficulties.
215: 
216: For $m^2_{\varphi}<0$ we have the following solution of the
217: equations of motion (EOM):
218: %
219: \be \chi=0, \quad \quad
220: \varphi=\sqrt{\frac{-m^2_{\varphi}}{\lambda_{\varphi}}}\equiv
221: \varphi_{eff}, \label{simplefirstbac}\ee
222: %
223: which breaks $Z_2$ but preserves $Z_2'$. We can express the
224: lagrangian in terms of the fluctuation $\d \varphi $ and $\chi$
225: around this background:
226: %
227:  \bea \mathcal{L}&=&
228: -\frac{1}{2}\partial_{\nu}\d\varphi\partial^{\nu}\d\varphi-\frac{1}{2}\partial_{\nu}\chi\partial^{\nu}\chi
229: +m^2_{\varphi}\left(\d\varphi\right)^2-\frac{1}{2}\mu^2\chi^2
230: \nonumber
231: \\
232: &&-\sqrt{-m^2_{\varphi}\lambda_{\varphi}}\left(\d\varphi\right)^3-\frac{1}{4}\lambda_{\varphi}\left(\d\varphi\right)^4
233: -\frac{1}{4}\lambda_{\chi}\chi^4-2a\sqrt{\frac{-m^2_{\varphi}}{\lambda_{\varphi}}}\d\varphi\chi^2
234: \nonumber \\&&
235: -a\left(\d\varphi\right)^2\chi^2+constants,\label{lagexp} \eea
236: %
237: where
238: %
239: \be \mu^2\equiv m^2-2a\frac{m^2_{\varphi}}{\lambda_{\varphi}}. \ee
240: %
241: 
242: If $|\mu^2|\ll |m^2_{\varphi}|$, we expect that the heavy mode
243: $\d\varphi$ can be integrated out and an effective theory for
244: $\chi$ can be constructed for both the signs of $\mu^2$. However
245: it's important to note that $\d\varphi$ cannot be simply neglected
246: because it gives a contribution, because of the
247: trilinear\footnote{Also the quartic coupling
248: $\left(\d\varphi\right)^2\chi^2$ gives a contribution to the
249: operator $\chi^4$, but this is negligible in the classical limit.}
250: coupling $\d\varphi \chi^2$ in (\ref{lagexp}), to the operator
251: $\chi^4$ in the effective theory, through the diagram
252: \ref{simplediagram}. This is similar to what is usually done in GUT theories \cite{GUT}, where, for instance, four fermions
253: effective interactions emerge by integrating out the heavy gauge fields \cite{Buras:1977yy}.
254: %
255: \begin{figure}[t]
256: \begin{center}
257: \begin{picture}(300,100)(0,0)
258: 
259: \DashLine(50,80)(190,80){4}\Text(120,100)[]{Heavy scalar}
260: \Line(190,80)(230,50)\Text(210,55)[]{$\chi$}
261: \Line(190,80)(230,110)\Text(210,115)[]{$\chi$}
262: \Line(10,50)(50,80)\Text(30,55)[]{$\chi$}
263: \Line(10,110)(50,80)\Text(30,115)[]{$\chi$}
264: 
265: 
266: \end{picture}
267: \end{center}
268: \caption{\footnotesize A tree diagram which describes the scattering
269: of two light $\chi$, through the exchange of an heavy scalar. This
270: kind of diagram gives a contribution to the quartic term in the
271: effective theory potential.} \label{simplediagram}
272: \end{figure}
273: %
274: At the classical level the effective lagrangian for $\chi$ is
275: %
276: \be
277: \mathcal{L}_{eff}=-\frac{1}{2}\partial_{\nu}\chi\partial^{\nu}\chi-\frac{1}{2}\mu^2\chi^2
278: -\frac{1}{4}\left(\lambda_{\chi}-\frac{4a^2}{\lambda_{\varphi}}\right)\chi^4+...\,,\label{effsimple}
279: \ee
280: %
281: where the dots represent higher dimensional operators. The term
282: $a^2\chi^4/\lambda_{\varphi}$ is the contribution of the heavy
283: mode. The result (\ref{effsimple}) was originally derived in \cite{Chan:1979ce}, but here we
284: want also to study the effective theory with spontaneous symmetry breaking and we want to compare it with the
285: low energy limit of the fundamental theory.
286: 
287:  For $\mu^2>0$, the minimum of the effective theory potential
288: is for $\chi=0$. Instead for $\mu^2<0$ we have
289: %
290: \be
291: \chi=\sqrt{\frac{-\mu^2}{\lambda_{\chi}-\frac{4a^2}{\lambda_{\varphi}}}}\label{chieff}\ee
292: %
293: and the fluctuation $\d\chi$ over this background has the
294: following mass squared:
295: %
296: \be M^2(\d\chi)=-2\mu^2. \label{Meff}\ee
297: %
298: This results will be not modified by the higher dimensional operator
299: at the leading order\footnote{The mass $\mu$ is small in the sense
300: $|\mu|\ll|m_{\varphi}|$.} in $\mu$. The equations (\ref{chieff}) and
301: (\ref{Meff}) represent the effective theory prediction for the VEV
302: and the spectrum in the phase where $Z_2'$ is broken.
303: 
304: On the other hand, a solution of the fundamental EOM, namely the EOM
305: derived from the fundamental lagrangian $\mathcal{L}$, is
306: %
307: \bea
308: \chi^2&=&\frac{-\mu^2}{\lambda_{\chi}-\frac{4a^2}{\lambda_{\varphi}}}+O(\mu^3),\nonumber
309: \\
310: \varphi^2&=&-\frac{m^2_{\varphi}}{\lambda_{\varphi}}+\frac{2a\mu^2}{\lambda_{\varphi}\lambda_{\chi}-4a^2}+O(\mu^3)
311: \label{correctsimpleVEV}\eea
312: %
313: which is a small deformation of (\ref{simplefirstbac}) at the
314: leading non trivial order in $\mu$ and breaks the $Z_2'$ symmetry.
315: Moreover the light mode which corresponds to this solution has a
316: mass squared $-2\mu^2$.
317: 
318: Therefore the effective theory prediction for the light mode VEV and
319: spectrum is correct, at the order $\mu$, in this simple framework,
320: but the heavy mode contribution is necessary in order the effective
321: theory prediction to be correct.
322: 
323: \section{A More General Case}\label{general}
324: \setcounter{equation}{0}
325: 
326: Now we want to extend the result of section \ref{simple4D} and ref
327: \cite{Chan:1979ce} to a more general class of theories. We consider
328: a set of real D-dimensional scalars $\Phi_i$ with a general
329: potential $V$: the lagrangian is
330: 
331: %
332: \be \mathcal{L}=-\frac{1}{2}\partial_M\Phi_i\partial^M\Phi_i -
333: V(\Phi), \ee
334: %
335: where $M,N,...$ run over all the space-time dimensions, while
336: $\mu,\nu,...$ and $m,n,...$ are respectively the 4-dimensional and
337: the internal coordinates indices. The EOM are
338: %
339: \be
340: \partial_M\partial^M\Phi_i-\frac{\partial V}{\partial\Phi_i}(\Phi)=0.\label{generalEOM}
341: \ee
342: %
343: We consider now a solution $\Phi_{eff}$ of (\ref{generalEOM})
344: which preserves the 4-dimensional Poincaré invariance and some
345: internal symmetry group $\mathcal{G}$; the corresponding mass squared eigenvalue
346: problem for the 4-dimensional states is
347: %
348: \be -\partial_m\partial_m\d \Phi_i+\frac{\partial^2V}{\partial\Phi_i\partial
349: \Phi_j}(\Phi_{eff})\d \Phi_j=M^2\d\Phi_i,\label{generaleigen}
350: \ee
351: %
352: where $\d \Phi$ is the fluctuation around $\Phi_{eff}$. We assume
353: that there are $n$ normalizable solutions $\D_l$ with small
354: eigenvalues ($M^2\sim \mu^2$), other, in principle infinite,
355: solutions\footnote{In principle $h$ can be a discrete or a
356: continuous variable.} $\tilde{\D}_h$with large eigenvalues ($M^2\gg
357: |\mu^2|$) and nothing else. These hypothesis are needed in order to
358: define the concept of light KK modes.
359: 
360: We can expand the scalars $\Phi_i$
361: as follows
362: %
363: \be
364: \Phi_i=\left(\Phi_{eff}\right)_i+\chi_l(x)\D_{li}(y)+\tilde{\chi}_h(x)\tilde{\D}_{hi}(y),\label{expansion}
365: \ee
366: %
367: where $\chi_l$ and $\tilde{\chi}_h$ are respectively the light and
368: heavy KK modes. We choose the $\D_l$ and $\tilde{\D}_h$ in order
369: that they form an orthonormal basis for the functions over the
370: internal space:
371: %
372: \bea \left<\D_l|\D_{l'}\right>&\equiv& \int
373: d^{D-4}y\D_{li}(y)\D_{l'i}(y)=\d_{ll'},\nonumber \\
374: \left<\tilde{\D}_h|\tilde{\D}_{h'}\right>&\equiv& \int
375: d^{D-4}y\tilde{\D}_{hi}(y)\tilde{\D}_{h'i}(y)=\d_{hh'},\nonumber
376: \\
377: \left<\D_l|\tilde{\D}_{h}\right>&\equiv& \int
378: d^{D-4}y\D_{li}(y)\tilde{\D}_{hi}(y)=0. \eea
379: %
380: We note that $\chi_l$ and $\tilde{\chi}_h$ could both belong to some
381: non trivial representation of the internal symmetry group
382: $\mathcal{G}$.
383: 
384: 
385: \subsection{The Effective Theory Method}\label{EFFGeneral}
386: 
387: 
388: We construct now some relevant terms in the effective theory for the
389: light KK modes $\chi_l$. Here "relevant terms" mean relevant terms
390: in the classical limit and in case we have a small point of minimum
391: of the order $\mu$ of the effective theory potential: we want to
392: compare the results of the effective theory for the light KK modes
393: with the low energy limit of the fundamental theory expanded around
394: a vacuum which is a small perturbation of $\Phi_{eff}$. Further we
395: calculate everything at leading non trivial order\footnote{The $\mu$
396: mass scale is small in the sense $|\mu|$ is much smaller than the
397: heavy masses.} in $\mu$.
398:  The relevant terms can be
399: computed by putting just the light KK modes in the action and
400: performing the integration over the extra dimensions and then by
401: taking into account the effect of heavy KK modes through the
402: diagrams like figure \ref{simplediagram}. In order to calculate
403: those diagrams, we give the interactions between two light modes
404: $\chi_l$ and one heavy mode $\tilde{\chi}_h$:
405: %
406: \be -\frac{1}{2}\left(\int d^{D-4}y
407: V_{ijk}\D_{li}\D_{mj}\tilde{\D}_{hk}\right)\chi_l\chi_m\tilde{\chi}_h,
408: \ee
409: %
410: where we have used the notation
411: %
412: \be V_{i_1 ... i_N}\equiv \frac{\partial^N V}{\partial\Phi_{i_1}
413: ... \partial\Phi_{i_N} }\left(\Phi_{eff}\right). \ee
414: %
415: We get the following relevant terms in the effective theory
416: potential $\U$:
417: %
418: \be \U(\chi)=\frac{1}{2}c_l\mu^2\chi_l\chi_l
419: +\frac{1}{3}\lambda^{(3)}_{lmp}\chi_l\chi_m\chi_p+\frac{1}{4}\lambda_{lmpq}^{(4)}\chi_l\chi_m\chi_p\chi_q+...
420: \,,\label{Ugeneral}\ee
421: %
422: where the dots represent non relevant terms, $c_l$ are
423: dimensionless numbers and
424: %
425: \bea\lambda^{(3)}_{lmp}&\equiv&\frac{1}{2}\int d^{D-4}y
426: V_{ijk}\D_{li}\D_{mj}\D_{pk},\\
427: \lambda_{lmpq}^{(4)}&\equiv&\frac{1}{3!}\left(\int d^{D-4}y
428: V_{ijkk'}\D_{li}\D_{mj}\D_{pk}\D_{qk'}\right)+a_{lmpq},
429: \label{l4}\eea
430: %
431: where the quantities $a_{lmpq}$ represent the heavy modes
432: contribution and they are given by
433: %
434: \be a_{lmpq}=c_{lmpq}+c_{lpmq}+c_{lqpm} \ee
435: %
436: and
437: %
438: \be c_{lmpq}\equiv -\frac{1}{6}\int
439: d^{D-4}yd^{D-4}y'V_{ijk}(y)\D_{li}(y)\D_{mj}(y)
440: G_{kk'}(y,y')V_{i'j'k'}(y')\D_{pi'}(y')\D_{qj'}(y').\ee
441: %
442: The object $G_{kk'}$ is the Green's function for the mass squared operator at
443: the left hand side of (\ref{generaleigen}) and it's explicitly
444: given by
445: %
446: \be G_{kk'}(y,y')=\sum_h
447: \frac{1}{m_h^2}\tilde{\D}_{hk}(y)\tilde{\D}_{hk'}(y'),\ee
448: %
449: where $m^2_h$ is the eigenvalue associated to the eigenfunction
450: $\tilde{\D}_h$.
451: 
452: In the rest of this section we consider the predictions of the
453: effective theory with spontaneous symmetry breaking.  The potential
454: (\ref{Ugeneral}) has to be considered as a generalization of
455: (\ref{effsimple}), which was originally derived in
456: \cite{Chan:1979ce}. A non vanishing VEV breaks in general
457: $\mathcal{G}$ to some subgroup and it must satisfies
458: %
459: \bea \frac{\partial
460: \U}{\partial\chi_l}&=&c_l\mu^2\chi_l+\lambda^{(3)}_{lmp}\chi_m\chi_p+\lambda_{lmpq}^{(4)}\chi_m\chi_p\chi_q=
461: 0. \label{effvacua}\eea
462: %
463: Since we require that $\chi_l$ goes to zero as $\mu$ goes to zero we
464: have
465: %
466: \be \chi_l=\chi_{l1}+\chi_{l2}+... \ee
467: %
468: where $\chi_{l1}$ is proportional to $\mu$, $\chi_{l2}$ is
469: proportional to $\mu^2$ and so on. At the order $\mu^2$ the
470: equations (\ref{effvacua}) reduce to
471: %
472: \be \lambda^{(3)}_{lmp}\chi_{m1}\chi_{p1}=0 \ee
473: %
474: which implies
475: %
476: \be \lambda^{(3)}_{lmp}\chi_{p1}=0. \label{eff2vacua}\ee
477: %
478: While, at the order $\mu^3$, the equations (\ref{effvacua}) reduce
479: to
480: %
481: \be
482: c_l\mu^2\chi_{l1}+\lambda_{lmpq}^{(4)}\chi_{m1}\chi_{p1}\chi_{q1}=0,\label{eff3vacua}
483: \ee
484: %
485: where we have used the equations (\ref{eff2vacua}).
486: 
487: Finally the mass spectrum corresponding to a solution of
488: (\ref{effvacua}) is given by the eigenvalues of the hessian matrix
489: of $\U$ in that solution:
490: %
491: \be \frac{\partial^2 \U}{\partial\chi_l
492: \partial
493: \chi_{l'}}=c_l\mu^2\d_{ll'}+2\lambda^{(3)}_{ll'm}\chi_m+
494: 3\lambda_{ll'mq}^{(4)}\chi_m\chi_q. \ee
495: %
496: If we assume, for simplicity, $\lambda^{(3)}_{ll'm}=0$, which
497: corresponds to the absence of cubic terms in $\U$, the leading
498: order approximation of the hessian is simply given by
499: %
500: \be \frac{\partial^2 \U}{\partial\chi_l
501: \partial
502: \chi_{l'}}=c_l\mu^2\d_{ll'}+
503: 3\lambda_{ll'mq}^{(4)}\chi_{m1}\chi_{q1}+O(\mu^3).\label{hessian}
504: \ee
505: %
506: In subsection \ref{correct} we show that this matrix, which
507: represents the mass spectrum for the light KK modes, and the
508: equations (\ref{eff2vacua}) and (\ref{eff3vacua}) for the light
509: modes VEVs are exactly reproduced by a D-dimensional analysis.
510: 
511: 
512: \subsection{D-dimensional analysis}\label{correct}
513: 
514: Now we want to find a solution of (\ref{generalEOM}) which is a
515: small perturbation, of the order $\mu$, of $\Phi_{eff}$ and then we
516: want to find the low energy mass spectrum of the fluctuations around
517: this solution. In general this solution will break $\mathcal{G}$ to
518: some subgroup like a solution of (\ref{effvacua}) does in the
519: effective theory method. The explicit form of such solution in the
520: simple case of section \ref{simple4D} is given by
521: (\ref{correctsimpleVEV}) and the low energy mass spectrum in that
522: simple case is represented by the squared mass $-2\mu^2$; now we
523: want to generalize these results.
524: 
525: Let's consider the expansion (\ref{expansion}); we observe that the
526: statement that the solution is a small perturbation of $\Phi_{eff}$
527: means
528: %
529: \bea&&\chi_l=\chi_{l1}+\chi_{l2}+...\,, \nonumber\\
530: &&\tilde{\chi}_h=\tilde{\chi}_{h1}+\tilde{\chi}_{h2}+...\,,\label{analitic}
531: \eea
532: %
533: that is there are no big $\mu-$independent terms in $\chi_l$ and
534: $\tilde{\chi}_h$. We consider now a Taylor expansion of the
535: equations (\ref{generalEOM}) around $\Phi_{eff}$:
536: %
537: \bea
538: &&\partial_m\partial_m\left(\Phi_i-\left(\Phi_{eff}\right)_i\right)\nonumber
539: \\
540: &&-\sum_{k=1}^{N}\frac{1}{k!}V_{ii_1...i_k}
541: \left(\Phi_{i_1}-\left(\Phi_{eff}\right)_{i_1}\right)\cdot ... \cdot
542: \left(\Phi_{i_k}-\left(\Phi_{eff}\right)_{i_k}\right)\nonumber\\
543: &&+O(\mu^{N+1})=0.\label{pertEOM}\eea
544: %
545: At the order $\mu$ the equations (\ref{pertEOM}) reduce to
546: %
547: \be \left(\partial_m
548: \partial_m\d_{ij}-V_{ij}\right)\left(\Phi_{j}-\left(\Phi_{eff}\right)_{j}\right)+O(\mu^2)=0,
549: \ee
550: %
551: which simply states
552: %
553: \be \tilde{\chi}_{h1}=0.\label{heavyVEV}\ee
554: %
555:  Moreover at the order $\mu^2$ the
556: equations (\ref{pertEOM}) imply
557: %
558: \be
559: \tilde{\chi}_{h2}\left(\partial_m\partial_m\d_{ij}-V_{ij}\right)\tilde{\D}_{hj}=
560: \frac{1}{2}V_{ijk}\D_{lj}\D_{mk}\chi_{l1}\chi_{m1},
561: \label{pertEOM2}\ee
562: %
563: which has two consequences: the first one is
564: %
565: \be \lambda^{(3)}_{lmp}\chi_{p1}=0,\label{con1}\ee
566: %
567: which can be derived from (\ref{pertEOM2}) by projecting over
568: $\D_l$ and it exactly reproduces (\ref{eff2vacua}) of the
569: effective theory method; the second consequence is
570: %
571: \be
572: \tilde{\chi}_{h2}\tilde{\D}_{hi'}(y)=-\frac{1}{2}\chi_{l1}\chi_{m1}\int
573: d^{D-4}y'G_{i'i}(y,y')V_{ijk}(y')\D_{lj}(y')\D_{mk}(y'),\label{con2}\ee
574: %
575: where $G$ still represents the Green's function for the operator
576: at the left hand side of (\ref{generaleigen}). Now we can write
577: the $\mu^3$ part of the equation (\ref{pertEOM}) as follows
578: %
579: \bea
580: &&-c_l\mu^2\chi_{l1}\D_{li}-m^2_h\tilde{\chi}_{h3}\tilde{D}_{hi}\nonumber
581: \\
582: &&-\frac{1}{2}V_{ijk}\chi_{l1}\D_{lj}\left(\tilde{\chi}_{h2}\tilde{\D}_{hk}+
583: \chi_{m2}\D_{mk}\right)\nonumber\\
584: &&-\frac{1}{2}V_{ijk}\left(\chi_{l2}\D_{lj}+
585: \tilde{\chi}_{h2}\tilde{\D}_{hj}\right)\chi_{m1}\D_{mk}\nonumber\\
586: &&-\frac{1}{3!}V_{ijkk'}\D_{lj}\D_{mk}\D_{pk'}\chi_{l1}\chi_{m1}\chi_{p1}=0.\eea
587: %
588: If one projects this equation over $\D_l$ and uses the equations
589: (\ref{con1}) and (\ref{con2}) one gets exactly the equations
590: (\ref{eff3vacua}). Therefore, at the order $\mu$, all the solutions
591: of (\ref{effvacua}) are reproduced by the D-dimensional analysis and
592: viceversa. Moreover we observe that these light KK modes VEVs,
593: predicted by the effective theory, constitute approximate solutions
594: of the fundamental D-dimensional EOM at leading non trivial order
595: because of the equation (\ref{heavyVEV}), which states that the
596: heavy KK modes VEVs are higher order quantity with respect to the
597: light KK modes VEVs.
598: 
599: Now we consider the mass squared eigenvalue problem which
600: corresponds to a solution $\Phi$; moreover we assume for simplicity
601: $\lambda^{(3)}_{lmp}=0$, like in the effective theory method. This
602: eigenvalue problem is
603: 
604: %
605: \be \mathcal{O}_{ij}\d\Phi_j\equiv -\partial_m\partial_m\d
606: \Phi_i+\frac{\partial^2V}{\partial\Phi_i\partial \Phi_j}(\Phi)\d
607: \Phi_j=M^2\d\Phi_i,\label{perteigen} \ee
608: %
609: where $\d \Phi_i$ represents the fluctuations of the scalars around
610: the solution $\Phi$. We observe now that the equation (\ref{perteigen}) can
611: be considered a time-independent Schrodinger equation:
612: $\mathcal{O}$ is the hamiltonian and $M^2$ the generic energy
613: level. Moreover we can perform a Taylor expansion of $\mathcal{O}$
614: around $\mu=0$:
615: %
616: \be \mathcal{O}=\mathcal{O}_0+\mathcal{O}_1+\mathcal{O}_2+...
617: \,.\ee
618: %
619: The operators $\mathcal{O}_1$ and $\mathcal{O}_2$ can be easily
620: expressed just in terms of $\chi_{l1}$ and $\chi_{l2}$ by using
621: (\ref{expansion}), (\ref{analitic}) and the constraints (\ref{con2})
622: and (\ref{heavyVEV}) which come from the EOM. From the perturbation
623: theory of quantum mechanics we know that the leading value of the
624: low energy mass spectrum is given by the eigenvalues of the
625: following mass squared matrix:
626: %
627: \be M^2_{ll'}\equiv A_{ll'}+B_{ll'},\ee
628: %
629: where
630: %
631: \be A_{ll'}\equiv <\D_l|\mathcal{O}_2|\D_{l'}>\ee
632: %
633: and
634: %
635: \be B_{ll'}\equiv
636: -\sum_{h}\frac{1}{m^2_h}<\D_l|\mathcal{O}_1|\tilde{\D}_h><\tilde{\D}_h|\mathcal{O}_1|\D_{l'}>.\ee
637: %
638: If one express the matrices $A$ and $B$ in terms\footnote{The
639: dependence on $\chi_{l2}$ disappears because we assume
640: $\lambda^{(3)}_{lmp}=0$, as one can easily check.} of $\chi_{l1}$ , one finds exactly the
641: corresponding result (\ref{hessian}) predicted by the effective
642: theory.
643: 
644: So we have two equivalent (at least at the leading non trivial order
645: in $\mu$) approaches to study the breaking of $\mathcal{G}$: the
646: spontaneous symmetry breaking in the 4-dimensional effective theory
647: and the D-dimensional analysis. We stress that, like in the simple
648: model of section \ref{simple4D}, also in this more general case the
649: heavy KK modes contribution in the effective theory can't be
650: neglected if one wants to reproduce the D-dimensional result, even
651: at the classical level. In general this is true not only in scalar
652: theories but also in theories which involve gauge and gravitational
653: interactions, as we illustrate in sections \ref{def} and \ref{u13}.
654: 
655: 
656: 
657: 
658: 
659: \section{A 6D Gauge and Gravitational Theory} \label{def}
660: \setcounter{equation}{0}
661: 
662: Now we consider a 6-dimensional field theory of gravity with a
663: $U(1)$ gauge invariance, including a charged scalar field $\phi$ and
664: eventually fermions. The bosonic action is\footnote{Some conventions
665: are fixed in appendix \ref{GR}.}
666: %
667: \be S_B=\int d^6 X\sqrt{-G}
668: \left[\frac{1}{\kappa^2}R-\frac{1}{4}F_{MN}F^{MN} -(\nabla_M
669: \phi)^*\nabla^M \phi -V(\phi)  \right], \label{action} \ee
670: %
671: where $R$ is the Ricci scalar, $\kappa$ represents the 6-dimensional
672: Planck scale, $F_{MN}$ is the field strength of the $U(1)$ gauge
673: field $A_M$, defined by
674:  %
675: \be F_{MN}=\partial_M A_N- \partial_N A_M \ee
676: %
677: and
678: %
679: \be
680: \nabla_M \phi = \partial_M \phi +ie A_M \phi,\ee
681: %
682: where $e$ is the $U(1)$ gauge coupling. Moreover $V$ is a scalar
683: potential and we choose
684: %
685: \be V(\phi)=m^2\phi^*\phi +\xi (\phi^* \phi)^2 + \lambda,  \label{V} \ee
686: %
687: where $m^2$ and $\xi$ are generical real constants, with the
688: constraint $\xi > 0$ and $\lambda $ represents the 6-dimensional cosmological constant.
689: 
690: 
691: From the action (\ref{action}) we can derive the general bosonic
692: EOM. However we focus on the following class of backgrounds, which
693: are invariant under the 4-dimensional Poincar\'e group:
694: %
695: \bea ds^2 &=&\eta_{\mu \nu}dx^{\mu}dx^{\nu}
696: +g_{mn}(y)dy^m dy^n.\label{Bmetric}\\
697: A&=&A_m(y)dy^m, \label{BA}\\
698: \phi &=&\phi(y), \label{Bphi} \eea
699: %
700: where $g_{mn}$ is the metric of a  2-dimensional compact internal
701: manifold $K_2$; so the 6-dimensional space-time manifold is
702: $(Minkowski)_4 \times K_2$. By using (\ref{Bmetric}), (\ref{BA}) and
703: (\ref{Bphi}), we can write the bosonic EOM in the following form:
704: %
705: \bea &&\nabla^2 \phi -m^2\phi-2\xi(\phi^* \phi)\phi=0,\nonumber \\
706: && \nabla_mF^{mn}+ie\left[\phi^*\nabla^n\phi
707: -(\nabla^n \phi)^*\phi \right]=0,\nonumber \\
708: &&\frac{1}{\kappa^2}R_{mn}-\frac{1}{2}F_{mp}F_n^{\,\,p}
709: -\frac{1}{2}(\nabla_m\phi)^*\nabla_n\phi-
710: \frac{1}{2}(\nabla_n\phi)^*\nabla_m\phi=0,\nonumber \\
711: &&\frac{1}{4}F^2-\lambda-m^2\phi^*\phi-\xi(\phi^*\phi)^2=0,
712: \label{EOM}
713:  \eea
714:  where $\nabla^2\equiv\nabla_m\nabla^m$ is the covariant laplacian over
715: the internal manifold. The equations (\ref{EOM}) must be satisfied
716: by the bosonic VEV.
717: 
718: We introduce also fermions and gauge invariant coupling
719: with the scalar $\phi$. In order to do that it's necessary to introduce
720: at least a pair of 6-dimensional Weyl spinors $\psi_+$ and $\psi_-$, where $+$ and $-$ refer here to the
721: 6-dimensional chirality. We consider the following fermionic action:
722: %
723: \be S_F= \int d^6 X\sqrt{-G} \left(\overline{\psi_+}\Gamma^M \nabla_M \psi_+ +\overline{\psi_-}\Gamma^M \nabla_M \psi_-
724: +g_Y\phi^* \overline{\psi_+}\psi_- + g_Y\phi \overline{\psi_-}\psi_+\right), \label{fermionL}\ee
725: %
726: where $g_Y$ is a real Yukawa coupling constant. In (\ref{fermionL})
727: $\nabla_M$ represents the covariant derivative acting on spinor,
728: which includes the gauge and the spin connection; moreover our
729: conventions for $\Gamma^M$ are given in appendix \ref{GR}. The
730: $U(1)$ charge $e_+$ and $e_-$ of $\psi_+$ and $\psi_-$ have to
731: satisfy the condition $e_-=e_+ +e$ coming from the gauge invariance
732: of the Yukawa terms. In the following we consider the choice
733: $e_+=e/2$ and $e_-=3e/2$, corresponding to a simple harmonic
734: expansion for the compactification over $(Minkowski)_4\times S^2$.
735: From (\ref{fermionL}) we get the following EOM:
736: %
737: \be \Gamma^M \nabla_M \psi_+ +g_Y\phi^* \psi_-=0, \quad \Gamma^M \nabla_M \psi_- +g_Y\phi \psi_+=0.
738: \ee
739: %
740: Now we define the following 4-dimensional Weyl spinors:
741: %
742: \be \psi_{\pm L}=\frac{1-\gamma^5}{2}\psi_{\pm }, \quad \psi_{\pm R}=\frac{1+\gamma^5}{2}\psi_{\pm}, \ee
743: %
744: where $\gamma^5$ is the 4-dimensional chirality matrix. In terms of
745: $\psi_{\pm L}$ and $\psi_{\pm R}$ the EOM, for a
746: $(Minkowski)_4\times K_2$ background space-time, are\footnote{We
747: rearrange the equations in a way that the left handed and right
748: handed sector are split.}
749: 
750: %
751: \bea \left(\partial^2 + 2 \nabla_+ \nabla_- -g_Y^2|\phi|^2\right)\psi_{+L}
752: -\sqrt2 g_Y\left(\nabla_+\phi^* \right)\psi_{-L}=0, \nonumber \\
753: \left(\partial^2 + 2 \nabla_- \nabla_+ -g_Y^2|\phi|^2\right)\psi_{-L}-\sqrt2 g_Y\left(\nabla_-\phi \right)\psi_{+L}=0, \nonumber \\
754: \left(\partial^2 + 2 \nabla_- \nabla_+ -g_Y^2|\phi|^2\right)\psi_{+R}
755: +\sqrt2 g_{Y}\left(\nabla_-\phi^* \right)\psi_{-R}=0, \nonumber \\
756: \left(\partial^2 + 2 \nabla_+ \nabla_- -g_Y^2|\phi|^2\right)\psi_{-R}
757: +\sqrt2 g_{Y}\left(\nabla_+\phi \right)\psi_{+R}=0, \label{fermioneq}
758: \eea
759: %
760: where $\partial^2\equiv \eta^{\mu \nu}\partial_{\mu}\partial_{\nu}$,
761: %
762: \be \nabla_{\pm}= \frac{1}{\sqrt2}(\nabla_5 \pm i\nabla_6) \ee
763: %
764: and $\nabla_{5,6}$ are the covariant derivative components in an
765: orthonormal basis. The equations (\ref{fermioneq}) will be used in
766: order to compute the fermionic spectrum.
767: 
768: %
769: \subsection{The $SU(2)\times U(1)$ Background Solution} \label{su2 x u1}
770: %
771: 
772: An $SU(2)\times U(1)$-invariant  solution of (\ref{EOM}) is
773: \cite{RSS}
774: %
775: \bea ds^2 &=&\eta_{\mu \nu}dx^{\mu}dx^{\nu}+
776: a^2\left(d\theta^2+\sin^2\theta d\varphi^2\right), \label{sphere}\\
777: A&=&\frac{n}{2e}(\cos\theta -1)d\varphi
778: \equiv -\frac{n}{2e}e^3(y), \label{monopole} \\
779: \phi &=&0, \label{phi0} \eea
780: %
781: subject to the constraints
782: %
783: \be \lambda=\frac{n^2}{8e^2a^4}=\frac{1}{\kappa^2a^2},
784: \label{constraint}\ee
785: %
786: where $n$ is a (integer) monopole number. So here we have $K_2=S^2$,
787: and $a$ is the radius of $S^2$. We introduce also an orthonormal
788: basis in the internal cotangent space \cite{RSS}:
789: %
790: \be e^{\pm}(y)=\pm \frac{i}{\sqrt{2}}e^{\pm i\varphi}\left(d\theta
791: \pm i\sin\theta d\varphi \right). \label{epm}\ee
792: %
793: In the following we consider, just for simplicity, the case
794: %
795: \be n=2. \ee
796: %
797: In fact for this value of the monopole charge we can find a very
798: simple solution of the fundamental 6-dimensional EOMs (\ref{EOM}) which is invariant under a $U(1)$ subgroup of
799: $SU(2)\times U(1)$; this solution is discussed in section
800: \ref{u13}. Like in section \ref{general} our purpose is in fact to construct the 4-dimensional
801: $SU(2)\times U(1)$-invariant effective theory,
802: study the spontaneous symmetry breaking $SU(2)\times U(1)\rightarrow U(1)$ and the Higgs mechanism in the effective
803: theory and then compare the results with the corresponding quantities predicted by the 6-dimensional theory; therefore,
804: in order to do that, one has to find a 6-dimensional U(1)-invariant solution of the EOMs.
805: 
806: 
807: If $\Phi_{\lambda}$ is a field with an integer or half-integer
808: iso-helicity $\lambda$, we can perform an harmonic expansion
809: \cite{RSS}:
810: %
811: \be \Phi_{\lambda}(x,\theta, \phi)=\sum_{l\geq |\lambda|}
812: \sum_{|m|\leq l}\Phi_m^l(x)
813: \sqrt{\frac{2l+1}{4\pi}}\mathcal{D}^{(l)\lambda}_{m}(\theta , \varphi), \label{lexpansion}\ee
814: %
815: where, for a given $l$, $\mathcal{D}^{(l)\lambda}_{m}$ is a
816: $(2l+1)\times (2l+1)$ unitary matrix. The
817: $\mathcal{D}^{(l)\lambda}_{m}$ were originally introduced in
818: \cite{Wigner} and our conventions are given in appendix \ref{GR}.
819: For example $\phi$ has an expansion like (\ref{lexpansion}) with
820: $\lambda=1$.
821: 
822: The low energy 4-dimensional spectrum coming from the background
823: (\ref{sphere}), (\ref{monopole}) and (\ref{phi0}) is given in the
824: reference \cite{RSS} for the spin-1 and spin-2 sectors. The massless sector is the
825: following: there are a graviton (helicities
826: $\pm2$, $l=0$), a $U(1)$ gauge field (helicities $\pm1$, $l=0$)
827: coming from $\mathcal{V}_{\mu}$ and a Yang-Mills $SU(2)$ triplet (helicities $\pm 1$, $l=1$)
828: coming from $h_{\mu \a}$ and $\mathcal{V}_{\mu}$, where $\mathcal{V}_M$ and $h_{MN}$ are
829: the fluctuations of the gauge field and the metric around the solution (\ref{sphere}), (\ref{monopole}) and (\ref{phi0}).
830: Regarding the scalar spectrum all the scalars from $G_{MN}$ and $A_M$ have very large masses,
831: of the order $1/a$, and we can get only an
832: $SU(2)$-triplet from $\phi$ in the low energy spectrum
833: if we choose $m^2$ such
834: that
835: %
836: \be |\mu^2| \ll \frac{1}{a^2}, \label{assumption1}\ee
837: %
838: where
839: %
840: \be \mu^2 \equiv -\frac{1}{a^2}\eta \equiv m^2+\frac{1}{a^2}.
841: \label{edefin}\ee
842: %
843: 
844: In fact $-1/a^2$ is the eigenvalue of the laplacian operator acting
845: on the harmonic with $l=1$ and $\lambda=1$ , as one can check using
846: the related formula of \cite{RSS}. The parameter $\mu^2$ is in fact
847: the squared mass of the triplet from $\phi$, and it can be in
848: principle either positive or negative. If (\ref{assumption1}) holds
849: all the remaining scalars have masses at least of the order $1/a$
850: and they don't appear in the low energy theory. So we assume that
851: (\ref{assumption1}) holds. Finally in order to find the low energy
852: fermionic spectrum we have to calculate the associated
853: iso-helicities by using the explicit expression for the background
854: covariant derivative of $\psi_{\pm}$ along the internal space:
855: %
856: \be \nabla_m \psi_{\pm} = \left( \partial_m \pm\omega_m \frac{1}{2} \gamma^5 +ie_{\pm} A_m \right)\psi_{\pm}, \ee
857: %
858: where $\omega_\theta = 0$, $\omega_{\varphi}=\frac{i}{a}( \cos\theta
859: - 1)$, $e_+ = e/2$ and $e_- = 3e/2$. We get
860: %
861: \be \lambda_{+L}=0, \quad \lambda_{+R}=1, \quad \lambda_{-L}=2,
862: \quad \lambda_{-R}=1 \ee
863: %
864: and the corresponding expansions are given by (\ref{lexpansion}). So
865: the equations (\ref{fermioneq}) tell us that there are 4 zero-modes:
866: the $l=0$, $m=0$ mode in $\psi_{+L}$ and the $l=1$, $m=+1,-1,0$ in
867: $\psi_{-R}$. So we have a massless $SU(2)$ singlet from $\psi_{+L}$
868: and a massless $SU(2)$ triplet from $\psi_{-R}$.
869: 
870: 
871: 
872: \subsection{The 4D $SU(2)\times U(1)$ Effective Lagrangian \\ and the
873: Higgs Mechanism} \label{effective}
874: 
875: Now we want to study the 4D effective theory: which is the
876: 4-dimensional
877:  theory obtained
878: from the background (\ref{sphere}), (\ref{monopole}) and (\ref{phi0}) retaining only
879:  the low energy spectrum we discussed at the end of subsection \ref{su2 x u1}, that is the particles with
880: masses much smaller than $1/a$, and integrating out all the heavy modes, namely those with mass at least of the
881: order $1/a$.
882: This is an $SU(2)\times U(1)$-invariant theory,
883: which includes a charged scalar, that we call $\chi$, in the $3$-dimensional representation of $SU(2)$, and, if we want, two
884: Weyl spinors in the $1_{1/2}$ and $3_{3/2}$ of $SU(2)\times U(1)$.
885:  The background (\ref{sphere}), (\ref{monopole}) and (\ref{phi0}) is the analogous of what we
886:  called $\Phi_{eff}$ in section \ref{general}. In this section we give
887: only some relevant terms\footnote{Here ``relevant terms'' has the same meaning
888: as in the subsection \ref{EFFGeneral}.} appearing in the lagrangian of this theory.
889: In particular we calculate the scalar potential, we study the Higgs mechanism,
890: which is active only for $\mu^2 <0$,
891: and we give in this case the masses of the spin-1, spin-0 and spin-1/2 particles.
892: 
893: Like in the general scalar theory of section \ref{general}, in the
894: following we perform all the calculations at the order $\eta$. If we
895: use the information regarding the low-energy spectrum which we
896: discussed at the end of subsection \ref{su2 x u1}, we can construct
897: some relevant terms of the 4D effective theory through the following
898: ansatz\footnote{The ansatz (\ref{0ansatz}) is a generalization of
899: the zero-mode ansatz of \cite{RSS}, which doesn't include scalar
900: fields.}
901: %
902: \bea E^a(x)&=& E^a_{\mu}(x)dx^{\mu}, \nonumber \\
903: E^{\alpha}(x,y)&=&e^{\alpha}(y)-\frac{\kappa}{a\sqrt{4\pi}}W_{\mu}^{\hat{\alpha}}(x)dx^{\mu}
904: \mathcal{D}^{\alpha}_{\hat{\alpha}}(y),
905: \nonumber  \\
906: A(x,y)&=& -\frac{n}{2ea}e^3(y) \nonumber\\
907: &&+\frac{1}{a\sqrt{4\pi}}V_{\mu}(x)dx^{\mu}
908: -\frac{n\kappa}{2ea^2\sqrt{4\pi}}U_{\mu}^{\hat{\alpha}}(x)dx^{\mu}
909: \mathcal{D}_{\hat{\alpha}}^3(y), \nonumber \\
910:  \phi(x,y)&=&\frac{1}{a}\sqrt{\frac{3}{4\pi}}\chi^{\hat{\a}}(x)\mathcal{D}_{-,\hat{\a}}(y),\nonumber \\
911: \psi_{+R}&=&\psi_{-L}=0, \nonumber \\
912: \psi_{-R}&=&\frac{1}{a}\sqrt{\frac{3}{4\pi}}\psi_R^{\hat{\a}}(x)\mathcal{D}_{-,\hat{\a}}(y),\nonumber \\
913: \psi_{+L}&=&\frac{1}{a\sqrt{4\pi}}\psi_L(x),
914: \label{0ansatz} \eea
915: %
916: where $E^A$, $A=0,1,2,3,+,-$, are the 6-dimensional orthonormal
917: 1-form basis, $E^a_{\mu}$ is the 4-dimensional vielbein, $V_{\mu}$
918: is the 4-dimensional $U(1)$ gauge field coming from
919: $\mathcal{V}_{\mu}$, a linear combination\footnote{The orthogonal
920: linear combination has a large mass; we show this in appendix
921: \ref{S^2simm}.} of $W_{\mu}$ and $U_{\mu}$ is the Yang-Mills $SU(2)$
922: triplet \cite{RSS} coming from $h_{\mu \a}$ and $\mathcal{V}_{\mu}$;
923: finally $\psi_L$ and $\psi_R$ are the $SU(2)$ fermion singlet and
924: fermion triplet, respectively.
925:  Actually the ansatz (\ref{0ansatz}) is the
926: background (\ref{sphere}), (\ref{monopole}) and (\ref{phi0}) plus
927: some fluctuations, which include all the light KK states.
928: 
929: 
930: Now we want to write some relevant terms of the effective lagrangian
931: for $\chi$ by using the light-mode ansatz (\ref{0ansatz}) and by
932: taking into account the heavy modes contribution. The scalar
933: potential $\mathcal{U}$ in the 4D effective theory, including the
934: bilinears and the quartic interactions, is
935: %
936: \be \mathcal{U}(\chi)=\mu^2 \chi^{\dag} \chi
937: +\left(\lambda_H+c_1\lambda_G\right)\left(\chi^{\dag}\chi\right)^2
938: -\frac{\lambda_H+c_2\lambda_{G}}{3}\left|\chi^{\hat{\alpha}}g_{\hat{\alpha}\hat{\beta}}
939: \chi^{\hat{\beta}}\right|^2+...,\label{U} \ee
940: %
941: where $c_1$ and $c_2$ are dimensionless parameters,
942: %
943: \be \lambda_H\equiv \frac{9}{20 \pi a^2}\xi,\quad \quad \lambda_G\equiv \frac{9\kappa^2}{80\pi a^4} \label{lGlH}\ee
944: %
945: and the dots represent higher order non relevant terms, for example
946: terms with a product of 6 $\chi$ or 8 $\chi$. These terms don't
947: contribute to the VEV of $\chi$ as we want this VEV to be of the
948: order\footnote{The order $\eta^{1/2}$ corresponds to the order $\mu$
949: because of equation (\ref{edefin}).} $\eta^{1/2}$. In (\ref{U}) we
950: took into account that the quartic part of $\mathcal{U}$ comes from
951: the quartic term in the 6-dimensional potential $V$ in (\ref{V}) and
952: from the contribution of the heavy scalars, namely $h_{\a\b}$ and
953: $\mv_{\a}$, through diagrams like figure \ref{simplediagram}. The
954: latter contribution is represented by $c_1\lambda_G$ and
955: $c_2\lambda_G$, the analogous of $a_{lmpq}$ in the equation
956: (\ref{l4}). Moreover we give also the expression for the gauge
957: covariant derivative of $\chi$:
958: %
959: \be D_{\mu} \chi ^{\hat{\alpha}}=\partial_{\mu}\chi^{\hat{\alpha}}
960: +ig_1V_{\mu}\chi^{\hat{\alpha}}
961: +g_2\mathcal{A}_{\mu}^{\hat{\beta}}
962: \epsilon_{\hat{\beta}\hat{\gamma}}^{\,\,\,\,\,\,\hat{\alpha}}\chi^{\hat{\gamma}},\label{covphi}
963: \ee
964: %
965: where $\mathcal{A}_{\mu}$ is defined in appendix \ref{S^2simm} and it represents the $SU(2)$
966: Yang-Mills field, $\epsilon_{\hat{\gamma}\hat{\beta}\hat{\alpha}}$ is a
967: totally antisymmetric symbol with $\epsilon_{+-3}=i$, and
968: %
969: \be g_1=\frac{e}{\sqrt{4\pi}a}, \,\,\,\,
970: g_2=\sqrt{\frac{3}{16\pi}}\frac{\kappa}{a^2},\label{g12}\ee
971: %
972: are the 4-dimensional $U(1)$ and $SU(2)$ gauge couplings.
973: 
974: Therefore the complete lagrangian for $\chi$ is
975: %
976: \be \mathcal{L}_{\chi eff}=-\left(D_{\mu}\chi\right)^{\dagger}D^{\mu}\chi-\mathcal{U}(\chi).
977: \label{actionchi} \ee
978: %
979: 
980: Let's look for the points of minimum of the order $\eta^{1/2}$ of
981: the potential $\mathcal{U}$ in (\ref{U}). We have a minimum, in the
982: case $\mu^2<0$, for
983: %
984: \be \chi_1=\chi_2=0, \,\,\,\,
985: \chi_3=v\equiv\sqrt{\frac{-3\mu^2}{4\left[\lambda_H+\frac{1}{2}\left(3c_1-c_2\right)\lambda_G\right]}},\label{minimum}\ee
986: %
987: which corresponds to the global minimum
988: %
989: \be \mathcal{U}_0=0 \label{Umin}\ee
990: %
991: at the order $\eta$. This fact states that, at leading order, the
992: 4-dimensional flatness condition in the background is compatible
993: with the procedure of the 4D effective theory. In fact
994: $\mathcal{U}_0$ can be interpreted as a 4-dimensional cosmological
995: constant and the flatness implies $\mathcal{U}_0=0$. Instead for
996: $\mu^2>0$ we don't have any order parameter because the global
997: minimum $\mathcal{U}_0=0$ corresponds to $\chi=0$.
998: 
999: 
1000: 
1001: If we take, for $\mu^2<0$,
1002: the vacuum (\ref{minimum}),
1003:  $SU(2)\times U(1)$ breaks to $U(1)_3$, where $U(1)_3$ is the $U(1)$-subgroup
1004: of $SU(2)$ generated by its third generator. The gauge field of $U(1)$ and $SU(2)$ are respectively
1005: $V_{\mu}$ and $\mathcal{A}_{\mu}$; before Higgs mechanism these gauge field are of course massless
1006: as one can see by looking at their bilinear lagrangian given
1007: in appendix \ref{S^2simm}. From (\ref{actionchi}) and
1008: (\ref{covphi}) we can calculate the masses of these vector fields in the 4D effective theory after the Higgs mechanism.
1009: We get a massless vector field $\mathcal{A}_{\mu}^3$, which corresponds
1010: to the unbroken $U(1)_3$ gauge symmetry. Instead $V_{\mu}$ and $\mathcal{A}_{\mu}^{\pm}$ acquire respectively
1011: the following squared masses
1012: %
1013: \be M^2_{V}  = \frac{3e^2}{8\pi
1014: a^2}\frac{-\mu^2}{\lambda_H+\frac{1}{2}\left(3c_1-c_2\right)\lambda_G},\label{MV}
1015: \ee
1016: %
1017: \be M^2_{V\pm}=\frac{9e^2}{16\pi
1018: a^2}\frac{-\mu^2}{\lambda_H+\frac{1}{2}\left(3c_1-c_2\right)\lambda_G},
1019: \label{MA} \ee
1020: %
1021: where the subscript $V$ indicates that we're dealing with vector
1022: particles. Moreover, in the spin-0 sector, we have two physical
1023: scalar fields: a real scalar and a complex one, which is charged
1024: under the residual $U(1)_3$ symmetry. Their squared masses are
1025: respectively
1026: %
1027: \be M^2_S=-2\mu^2,\ee
1028: %
1029: \be
1030: M^2_{S\pm}=-\mu^2\frac{\lambda_H+c_2\lambda_G}{\lambda_H+\frac{1}{2}\left(3c_1-c_2\right)\lambda_G}.
1031: \label{MS+-}\ee
1032: %
1033: Finally we can determine the fermionic spectrum by examining the
1034: fermionic lagrangian in the effective theory:
1035: %
1036: \be \mathcal{L}_{F eff}=\overline{\psi_L}\gamma^{\mu} D_{\mu} \psi_L+\overline{\psi_R}\gamma^{\mu} D_{\mu} \psi_R
1037: + g_4\overline{\psi_L}\chi^{\dag}\psi_R + g_4\overline{\psi_R}\chi\psi_L,\ee
1038: %
1039: where
1040: %
1041: \be g_4=\frac{g_Y}{a\sqrt{4\pi}}. \ee
1042: %
1043: The result is a neutral Dirac fermion, with squared mass
1044: %
1045: \be M_F^2=\frac{3g_Y^2}{16\pi
1046: a^2}\frac{-\mu^2}{\lambda_H+\frac{1}{2}\left(3c_1-c_2\right)\lambda_G},
1047: \label{efffermion}\ee
1048: %
1049: and a pair of massless right-handed Weyl fermions.
1050: We observe that the mass spectrum that we gave here is parametrized by the $c_i$.
1051: Of course these constants are not free parameters but they can be in principle computed
1052: by evaluating explicitly the heavy modes contribution. In the rest of this paper we don't compute the $c_i$ but we prove that the
1053: 4D effective theory without heavy modes contribution, that is $c_i=0$, is not correct because it predicts a wrong VEV
1054: of the light KK scalars and a wrong mass spectrum.
1055: 
1056: 
1057: \section{Symmetry Breaking in the 6D Theory} \label{u13}\setcounter{equation}{0}
1058: 
1059: 
1060: Now we perform a 6-dimensional (or geometrical) analysis of
1061: spontaneous symmetry breaking: this method corresponds to the
1062: contents of section \ref{general} for scalar theories. Of course we
1063: perform all the calculations at the order $\eta$, as in the
1064: effective theory method.
1065: 
1066: The most simple solution, up to higher order terms in $\eta$, that
1067: we find is similar to the background which appears in the reference
1068: \cite{S} \footnote{This solution was discussed in reference
1069: \cite{S}, but incorrectly.} :
1070: %
1071: \bea ds^2 &=&\eta_{\mu \nu}dx^{\mu}dx^{\nu}+ a^2\left[(1+|\eta |\beta \sin^{2}\theta)d\theta^2+
1072: \sin^2\theta d\varphi^2\right], \nonumber \\
1073: A&=& -\frac{1}{e}e^3, \nonumber \\
1074: \phi &=&\eta^{1/2}\alpha\exp\left(i\varphi \right)
1075: \sin\theta,
1076: \label{solution2} \eea
1077: %
1078: where $\beta\equiv \kappa^2 |\alpha|^2$. As required, for $\eta=0$
1079: this background reduces to the background of subsection \ref{su2 x
1080: u1}. The value of $\phi$ in (\ref{solution2}) is proportional to the
1081: harmonic $\D_{-,0}$, that is the harmonic with $l=1$, $\lambda=1$
1082: and $m=0$. In order that (\ref{solution2}) is a solution, up to
1083: $O(\eta^{3/2})$, it is necessary that (\ref{constraint}) holds and
1084: $|\alpha |^2$ is given by the following equation:
1085: %
1086: \be \frac{1}{a^2}\eta \int \,D^* \phi + \int \,D^*L_2\phi=
1087: 2\xi\int \, D^*|\phi|^2\phi, \label{alpha}\ee
1088: %
1089: where $D$ is $\D_{-,0}$ and $L_2\phi$ is the function proportional
1090: to $\eta^{3/2}$ in the expansion of $\nabla^2 \phi$ in powers of
1091: $\eta^{1/2}$. Further in (\ref{alpha}) the integrals are performed
1092: with the round $S^2$ measure. The equation (\ref{alpha}) can be
1093: derived by putting (\ref{solution2}) in the Klein-Gordon equation.
1094: For $\mu^2<0$ the equation (\ref{alpha}) has a solution for
1095: %
1096: \be \lambda_H>\lambda _G, \label{geometricalstability}\ee
1097: %
1098: where $\lambda_H$ and $\lambda _G$ are defined by (\ref{lGlH}),
1099: while, for $\mu^2>0$, we have a solution for
1100: %
1101: \be \lambda_H<\lambda _G. \ee
1102: %
1103: Whether $\mu^2>0$ or $\mu^2<0$, the solution of (\ref{alpha}) is
1104: %
1105: \be |\alpha|^2=\frac{5}{|8\xi a^2-2\kappa^2|}=\frac{9}{32 \pi
1106: a^4}\frac{1}{|\lambda_H-\lambda_G|}.\label{malpha} \ee
1107: %
1108: Note that here we have symmetry breaking for both signs of
1109: $\mu^2$.
1110:  This is not so interesting because the solution with
1111: $\mu^2>0$ is unstable, as it is discussed in subsection
1112: \ref{spin-0}. We want to stress that the value of $|\alpha|^2$
1113: predicted by the 4D effective theory is not equal to (\ref{malpha})
1114: if we neglect the heavy modes contribution to the effective theory,
1115: namely for $c_i=0$: indeed in this case the effective theory
1116: predicts a value of $|\a|^2$ equal to
1117: %
1118: \be |\alpha|_{eff}^2=\frac{9}{32 \pi
1119: a^4}\frac{1}{\lambda_H},\label{malpha2} \ee
1120: %
1121: which is equal to (\ref{malpha}) only for $\lambda_G=0$. However
1122: from (\ref{lGlH}) it's clear that $\lambda_G$ cannot be taken equal
1123: to zero. Therefore we have already proved that the heavy modes
1124: contribution is needed at least for the light mode VEV. We shall
1125: prove that this is the case also for the mass spectrum.
1126: 
1127: As required the background (\ref{solution2}) has the symmetry
1128: %
1129: \be U(1)_3 \subset SU(2).\label{symmetrybr}\ee
1130: %
1131:   So the 4-dimensional effective low energy
1132: theory, which follows from this background, is $U(1)_3$-invariant
1133: and comparing these results with the effective theory predictions
1134: makes sense.
1135: 
1136: We note that the symmetry breaking (\ref{symmetrybr}) is associated,
1137: in the 6-dimensional theory, to a geometrical deformation of the
1138: internal space. Further we observe that (\ref{solution2}) tell us
1139: the heavy modes VEVs are higher order corrections with respect to
1140: the light modes VEVs like in the scalar theories of section
1141: \ref{general}.
1142: 
1143: Now we calculate the low energy vector, scalar and fermion spectrum
1144: by analyzing the 4-dimensional bilinear lagrangian for the
1145: fluctuations around the solution (\ref{solution2}).
1146: 
1147: 
1148: 
1149: \subsection{Spin-1 Spectrum}
1150: 
1151: The spin-1 spectrum can be calculated in a way similar to the
1152: light mode ansatz (\ref{0ansatz}). However it must be noted that
1153: the sectors with different $l$ no longer decouple for $\eta\neq
1154: 0$, but the mixing terms are of the order $\eta$ and they give
1155: negligible corrections of the order $\eta^2$ to the vector boson
1156: masses. These facts are evident from the general formula of
1157: \cite{RS}. So we can neglect the modes with $l>1$ in the
1158: calculation of spin-1 spectrum.
1159:  Therefore we can compute the vector boson masses
1160: by putting the following
1161: ansatz in the action and integrating over the extra dimensions:
1162: 
1163: %
1164: \bea E^a(x)&=& E^a_{\mu}(x)dx^{\mu}, \nonumber \\
1165: E^{\alpha}(x,y)&=&e^{\alpha}(y,\eta)-\frac{\kappa}{a\sqrt{4\pi}}W_{\mu}^{\hat{\alpha}}(x)dx^{\mu}
1166: \mathcal{D}^{\alpha}_{\hat{\alpha}}(y) ,
1167: \nonumber  \\
1168: A(x,y)&=& -\frac{1}{ea}e^3(y) \nonumber\\
1169: &&+\frac{1}{a\sqrt{4\pi}}V_{\mu}(x)dx^{\mu}
1170: -\frac{\kappa}{ea^2\sqrt{4\pi}}U_{\mu}^{\hat{\alpha}}(x)dx^{\mu}
1171: \mathcal{D}_{\hat{\alpha}}^3(y), \nonumber \\
1172:  \phi(x,y)&=&\eta^{1/2}\alpha\exp\left(i\varphi \right)
1173: \sin\theta,
1174: \label{ansatz} \eea
1175: %
1176: where $e^{\a}(y,\eta)$ is the orthonormal basis for the 2-dimensional metric in (\ref{solution2}):
1177: :
1178: %
1179: \be e^{\pm}(y,\eta)=\pm \frac{i}{\sqrt{2}}e^{\pm i\varphi}
1180: \left[\left(1+|\eta|\frac{\beta}{2}\sin^{2}\theta\right)d\theta
1181: \pm i\sin\theta d\varphi \right]. \label{epmeps}\ee
1182: %
1183: In (\ref{ansatz}) we consider the spin-1 fluctuations but we don't
1184: consider the spin-0 fluctuations, because they are not necessary
1185: for the calculation of vector boson masses. It's important to note
1186: that in (\ref{ansatz}) the VEV of $E^{\alpha}$ is
1187: $e^{\alpha}(y,\eta)$, it's not $e^{\alpha}(y)$ as in
1188: (\ref{0ansatz}).
1189: 
1190: 
1191: From (\ref{ansatz}) it follows that some of the previous ($\eta=0$)
1192: massless states acquire masses for $\eta\neq 0$. Up to
1193: $O(\eta^{3/2})$, the $U(1)$ gauge boson ($l=0$) has the mass squared
1194: %
1195: \be M^2_V=\eta \frac{20}{3}\frac{e^2}{8\xi
1196: a^2-2\kappa^2}=\frac{3e^2}{8\pi
1197: a^2}\frac{-\mu^2}{\lambda_H-\lambda_G},\label{M0} \ee
1198: %
1199: while the Yang-Mills triplet $\mathcal{A}$ ($l=1$) is separated in
1200: a massless gauge boson, which is associated to $U(1)_3$ gauge
1201: invariance, and a couple of massive vector fields with the same
1202: mass squared
1203: %
1204: \be M^2_{V\pm}=\eta \frac{10e^2}{8\xi
1205: a^2-2\kappa^2}=\frac{9e^2}{16\pi a^2}\frac{-\mu^2}{\lambda_H-\lambda_G}.
1206: \label{M1} \ee
1207: %
1208: By comparing (\ref{M0}) and (\ref{M1}) with (\ref{MV}) and
1209: (\ref{MA}), we get that the heavy modes contribution is needed in
1210: the effective theory. However we observe that the ratio
1211: $M^2_V/M^2_{V\pm}$ is correctly predicted by the 4D effective theory
1212: for every $c_i$.
1213: 
1214:  Since the computation of vector bosons masses is complicated we
1215: present it explicitly. In order to prove (\ref{M0}) and (\ref{M1})
1216: it's useful to split the action in four terms:
1217: %
1218: \be S_B=S_R+S_F+S_{\lambda}+S_{\phi}, \ee
1219: %
1220: where
1221: %
1222: \bea S_R&=&\int d^6X \sqrt{-G}\frac{1}{\kappa^2}R, \\
1223:  S_F&=&-\frac{1}{4}\int d^6X \sqrt{-G}F^2, \\
1224: S_{\lambda}&=&\int d^6X\sqrt{-G}\left(-\lambda \right), \\
1225: S_{\phi}&=&\int d^6X \sqrt{-G}\left[-\left(\nabla_M\phi\right)^*\nabla^M\phi-V(\phi) \right].\label{actiondecomposition}
1226: \eea
1227: %
1228: In appendix \ref{SFSR} we prove that the contributions coming from
1229: $S_R$ and $S_F$ vanish, so only $S_{\phi}$ contributes to the spin-1
1230: masses up to $O(\eta^{3/2})$. The same low-energy spin-1 masses in
1231: (\ref{M0}) and (\ref{M1}) can be obtained also by using the general
1232: formula of \cite{RS}, which contains all the bilinear terms in the
1233: light cone gauge. The light cone gauge advantage is that the sectors
1234: with different spin decouple. However the derivation that we
1235: presented here shows that the unique contribution (at the leading
1236: order) to the spin-1 masses comes from $S_{\phi}$, like in the
1237: effective theory approach. This explains why the ratio
1238: $M^2_V/M^2_{V\pm}$ is correctly predicted by the 4D effective theory
1239: for every values of $c_i$.
1240: 
1241: \subsection{Spin-0 Spectrum} \label{spin-0}
1242: 
1243: We choose the light cone gauge \cite{RS, RSS2} in order to
1244: evaluate the spin-0 spectrum. In this gauge we have just two
1245: independent values for the indexes $\mu,\nu,...$ which label the
1246: 4-dimensional coordinates.
1247:   The bilinears for the
1248: fluctuations over the solution (\ref{solution2})
1249: can be simply computed
1250: with the general formula of \cite{RS}. For our model the helicity-0 $\mathcal{L}_0$ part is given by
1251: %
1252: \be \mathcal{L}_0=\mathcal{L}_0(\phi,\phi)+\mathcal{L}_0(h,h)+\mathcal{L}_0(\mathcal{V},\mathcal{V})+\mathcal{L}_0(\phi,h)
1253: +\mathcal{L}_0(\phi,\mathcal{V})+\mathcal{L}_0(h,\mathcal{V}),\ee
1254: %
1255: where
1256: %
1257: \bea \mathcal{L}_0(\phi,\phi)&=&\phi^*\partial^2\phi+\phi^*\nabla^2\phi-\left[m^2+(4\xi+e^2)|\Phi|^2
1258: +\kappa^2\left(\nabla_m\Phi\right)^*\nabla^m\Phi \right]|\phi|^2 \nonumber \\
1259: &&-\frac{1}{2}\left\{\left[(2\xi-e^2)\left(\Phi^*\right)^2+\kappa^2
1260: \left(\nabla_m\Phi \nabla^m\Phi \right)^*\right]\phi^2
1261: +c.\,c.\right\},\label{L01} \\
1262: \mathcal{L}_0(h,h)&=& \frac{1}{4\kappa^2}\left\{h_{mn}\partial^2
1263: h^{mn}+h_{mn}\nabla^2 h^{mn}
1264: +2R_{mn}^{\,\quad kl}h_l^m h_k^n \right. \nonumber \\
1265:  &&\left.+\kappa^2h_{ks}h_{mn}F^{km}F^{sn}-2\kappa^2h^l_mh_{ln}\left[\frac{1}{2}F^m_{\,\,\,\,k}F^{nk}+\left(\nabla^m\Phi\right)^*\nabla^n\Phi
1266:  \right]\right. \nonumber \\
1267: &&\left.+\frac{1}{2}
1268: h^i_i\partial^2h_j^j+\frac{1}{2}h_i^i\nabla^2h_j^j
1269: \right\},\label{L02}\\
1270: \mathcal{L}_0(\mv,\mv)&=& \frac{1}{2}\left\{\mv_m\partial^2\mv^m +\mv_m\nabla^2\mv^m-R_{mn}\mv^m\mv^n\right. \nonumber \\
1271: &&\left.-2e^2|\Phi|^2\mv^m\mv_m-\kappa^2\left(F_{ml}\mv^l\right)^2
1272: \right\}, \label{L03}\\
1273: \mathcal{L}_0(\phi,h)&=&\nabla_lh^{lm}\phi^*\nabla_{m}\Phi+h^{mn}\left(\nabla_m
1274: \phi \right)^*\nabla_n\Phi +c.\,c.\,,\label{L04} \\
1275: \mathcal{L}_0(\phi,\mathcal{V})&=&2ie\mv^m\phi^*\nabla_m\Phi-\kappa^2F^{lm}\mv_m\phi^*\nabla_l\Phi+c.\,c.\,,
1276: \label{L05}\\
1277: \mathcal{L}_0(h,\mathcal{V})&=&\mv^n\left(\nabla_mh_{ln}F^{lm}-h_l^m\nabla_mF^l_{\,\,\,\,n}\right),
1278: \label{L06}\eea
1279: %
1280: where $\Phi$ and $\phi$ are the background and the fluctuation of the 6-dimensional scalar.
1281: In this vanishing-helicity sector, it turns out that we have not only mixing terms of the order $\eta$
1282: but also mixing terms of the order $\eta^{1/2}$, coming from $\mathcal{L}_0(\phi,h)$ and $\mathcal{L}_0(\phi,\mathcal{V})$.
1283:  So now we can't neglect
1284: the mixing between the sectors with different values of $l$, as we
1285: did in the helicity $\pm 1$ sector. If we integrate these
1286: bilinears over the extra-dimensions we get an infinite dimensional
1287: squared mass matrix. However we are interested only in the light
1288: masses, therefore we can use the perturbation theory of quantum
1289: mechanics in order to extract the correction of the order $\eta$
1290: to the masses of the 6 real scalars which are massless for
1291: $\eta=0$. We already used this method for the computation of the
1292: mass spectrum in the scalar theories of section \ref{general}. We
1293: explain now how to use it in this framework.
1294: 
1295: Formally we can write the bilinears $\mathcal{L}_0$ of the scalar fields in this way
1296: %
1297: \be  \mathcal{L}_0 =\frac{1}{2}S^{\dag}\partial^2 S-\frac{1}{2}S^{\dagger}\mathcal{O} S,\label{formalbilinear}\ee
1298: %
1299: where $S$ is an array which includes all the scalar fluctuations; we choose
1300: %
1301: \be S=\left(\ba {c} \phi \\
1302:   \phi^*
1303:  \\
1304:   h_{++} \\
1305: h_{--}  \\
1306: h_{+-}  \\
1307:   \mathcal{V}_+  \\
1308: \mathcal{V}_-
1309: \ea\right). \label{Sdefinition}\ee
1310:  We have just to solve a 2-dimensional eigenvalue
1311: problem for the squared mass operator\footnote{The matrix elements of  $\mathcal{O}$ can be computed
1312: by comparing (\ref{formalbilinear}) with the explicit expression of $\mathcal{L}_0$.} $\mathcal{O}$:
1313: %
1314: \be \mathcal{O}S=M^2S. \ee
1315: %
1316: In particular we want to find the 6 values of $M^2$ which go to zero as $\eta $ goes to zero.
1317: Since we're working at the order $\eta$ we decompose $\mathcal{O}$ as follows
1318: %
1319: \be
1320: \mathcal{O}=\mathcal{O}_0+\mathcal{O}_1+\mathcal{O}_2, \label{Odecomposition}\ee
1321: %
1322: where $\mathcal{O}_0$ doesn't depend on $\eta$, $\mathcal{O}_1$ is
1323: proportional to $\eta^{1/2}$ and $\mathcal{O}_2$ is proportional to
1324: $\eta$. From the perturbation theory of quantum mechanics in the
1325: degenerate case we know that the 6 values of $M^2$ we are interested
1326: in are the eigenvalues of the following $6\times 6$
1327: matrix\footnote{Like in section \ref{general} we use the Dirac
1328: notation; for two states $|S_1>$ and $|S_2>$ and for an operator
1329: $A$, $<S_1|A|S_2>$ represents $\int S_1^{\dagger}AS_2$, where the
1330: integral is performed with the round $S^2$ metric.}:
1331: %
1332: \be M^2_{ij}=-\sum_{\tilde{i}}\frac{<i|\mathcal{O}_1|\tilde{i}><\tilde{i}|\mathcal{O}_1|j>}{M^2_{\tilde{i}}}
1333: +<i|\mathcal{O}_2|j>,\label{QMperturbation} \ee
1334: %
1335: where $|i>$, $i=1,...6$ represent the 6 orthonormal eigenfunctions of $\mathcal{O}_0$ with vanishing eigenvalue and
1336: they have the form
1337: %
1338: \be  |i>=\left(\ba {c} \phi \\
1339:   \phi^*
1340:  \\
1341:   0 \\
1342: .  \\
1343: .  \\
1344:   .  \\
1345: 0
1346: \ea\right). \ee
1347: %
1348: Moreover $|\tilde{i}>$ are all the remaining orthonormal eigenfunctions of $\mathcal{O}_0$
1349: and $M^2_{\tilde{i}}$ the corresponding eigenvalues. We note that the matrix elements $<i|\mathcal{O}_1|\tilde{i}>$ are
1350: non vanishing for
1351: %
1352: \be |\tilde{i}>=\left(\ba {c} 0 \\
1353:   0
1354:  \\
1355:   h_{++} \\
1356: h_{--}  \\
1357: h_{+-}  \\
1358:   \mathcal{V}_+  \\
1359: \mathcal{V}_-
1360: \ea\right). \label{tildei}\ee
1361: %
1362: Further the operator $\mathcal{O}_1$
1363: modifies the integration measure just by a factor proportional to
1364: the harmonics $\mathcal{D}^{(1)}$, therefore we need just
1365: a finite subset of ${|\tilde{i}>}$ for the evaluation of $M^2_{ij}$, namely those constructed through
1366: the harmonics with $l=0,1,2$, which are given in
1367: appendix \ref{GR}.
1368: An explicit form for $|i>$ and $|\tilde{i}>$, and the
1369: preliminary computations of the 6 eigenvalues we are interested in, are given in appendix \ref{spin-0calculation}.
1370: 
1371: We give here just the final result: we have two unphysical scalar
1372: fields (a real and a complex one) which form the helicity-0
1373: component of the massive vector fields; they have in fact the same
1374: squared masses given in (\ref{M0}) and (\ref{M1}), as it's
1375: required by Lorentz invariance, which is not manifest in the light cone gauge. Then we have a
1376: physical real scalar and a physical complex scalar, charged under
1377: the residual U(1) symmetry, with squared masses given respectively
1378: by (for $\mu^2<0$)
1379: 
1380: %
1381: \bea M^2_S&=&-2\mu^2, \nonumber \\
1382:  M^2_{S\pm}&=&-\mu^2\frac{\lambda_H+\lambda_G}{\lambda_H-\lambda_G}.\label{MS} \eea
1383: %
1384: For $\mu^2>0$, we get a negative value for $M^2_S$, therefore the
1385: corresponding solution is unstable. Note that the squared mass
1386: $M^2_S$ has exactly the same expression as in the 4D effective
1387: theory, for every $c_i$. But for $c_i=0$, which
1388: corresponds to neglecting the heavy modes contribution, the
1389: effective theory prediction for $M_{S\pm}^2$ in (\ref{MS+-}) is
1390: not equal to the correct value (\ref{MS}).
1391: We note that this is a physical inequivalence because the ratio $M^2_S/M^2_{S\pm}$, which is in principle
1392: a measurable quantity, is not correctly predicted by the 4D effective theory without the heavy modes contribution. More
1393: precisely the effective theory prediction for $M^2_S/M^2_{S\pm}$, in the case $c_i=0$, is always greater than
1394: the correct value.
1395: 
1396: 
1397: \subsection{Spin-1/2 Spectrum} \label{spin-1/2}
1398: 
1399: The spin-1/2 spectrum can be calculated by linearizing the equation of motion (\ref{fermioneq}): for $n=2$ we get
1400: %
1401: \bea &&\left(\partial^2 + 2 \nabla_+ \nabla_-
1402: -g_Y^2|\Phi|^2\right)\psi_{+L}
1403: =0, \nonumber \\
1404: &&\left(\partial^2 + 2 \nabla_- \nabla_+
1405: -g_Y^2|\Phi|^2\right)\psi_{-L}
1406: =0, \nonumber \\
1407: &&\left(\partial^2 + 2 \nabla_- \nabla_+
1408: -g_Y^2|\Phi|^2\right)\psi_{+R}
1409: +\sqrt2 g_{Y}\left(\nabla_+\Phi \right)^*\psi_{-R}=0, \nonumber \\
1410: &&\left(\partial^2 + 2 \nabla_+ \nabla_-
1411: -g_Y^2|\Phi|^2\right)\psi_{-R} +\sqrt2 g_{Y}\nabla_+\Phi
1412: \,\psi_{+R}=0, \label{linearizedfermion} \eea
1413: %
1414: where $\Phi$ represents again the background of the 6-dimensional scalar, namely the third line of
1415: (\ref{solution2}), and the covariant derivatives are evaluated with the background
1416: metric and background gauge field given by the first and the second line of (\ref{solution2}).
1417: These covariant derivatives are in the
1418: $\pm$ basis defined by (\ref{epmeps}) and it includes the modified spin connection when it acts on spinors:
1419: %
1420: \be \nabla_{\a} \psi_{\pm R} = e_{\a}^m(y,\eta)\left( \partial_m \pm\omega_m \frac{1}{2} +ie_{\pm} A_m \right)\psi_{\pm R},
1421: \ee
1422: %
1423: %
1424: \be \nabla_{\a} \psi_{\pm L} = e_{\a}^m(y,\eta)\left( \partial_m \mp\omega_m \frac{1}{2}  +ie_{\pm} A_m \right)\psi_{\pm L},
1425:  \ee
1426: %
1427: where $\omega_\theta = 0$, $\omega_{\varphi}\equiv \omega_{\varphi\,\,\, +} ^{\,\,+}$ is given in equation (\ref{spinconn})
1428: and the value of the charges $e_{\pm}$ and the
1429: iso-helicities\footnote{For $\eta\neq0$ we adopt the same harmonic
1430: expansion as in the $\eta=0$ case; this gives the correct result for
1431: the fermionic masses squared at the order $\eta$.} of the fermions
1432: are given at the end of subsection \ref{su2 x u1}. There we give
1433: also the fermionic massless spectrum for $\eta=0$: an $SU(2)$
1434: singlet from $\psi_{+L}$ and an $SU(2)$ triplet from $\psi_{-R}$.
1435: 
1436: From (\ref{linearizedfermion}) it's clear that the left handed
1437: sector doesn't present mixing terms of the order $\eta^{1/2}$ but
1438: only of the order $\eta$. Therefore the calculation of the squared
1439: mass $M_F^2$ of the light fermion coming from $\psi_{+L}$ is quite
1440: easy. The result is
1441: %
1442: \be M_F^2=\frac{3g_Y^2}{16\pi a^2}\frac{-\mu^2}{\lambda_H-\lambda_G}.\label{fermionmass}\ee
1443: %
1444: Instead the evaluation of the right-handed spectrum is complicated by the presence of mixing terms of the order $\eta^{1/2}$,
1445: as in the scalar sector. Therefore we use the perturbation theory of quantum mechanics also in the fermion right-handed sector.
1446: Formally we can write the eigenvalue equation for the mass squared operator $\mathcal{O}$
1447: acting in the right-handed sector as follows
1448: %
1449: \be \mathcal{O}F_R=M^2 F_R,\label{eigenfermion} \ee
1450: %
1451: where $F_R$ is an array which includes both the right-handed fermions; we choose
1452: %
1453: \be F_R=\left(\ba {c} \psi_{+R} \\
1454:   \psi_{-R}
1455: \ea\right). \label{FRdefinition}\ee
1456: %
1457: One can easily compute $\mathcal{O}$ acting on $F_R$ by performing
1458: the substitution $\partial^2\rightarrow M^2$ in the last two
1459: equations of (\ref{linearizedfermion}). Then we can proceed as in
1460: the scalar spectrum, performing the decomposition
1461: (\ref{Odecomposition}). However in this case the matrix $M^2_{ij}$
1462: in (\ref{QMperturbation}) is a 3$\times$3 matrix as the number of
1463: zero modes for $\eta=0$ in the right-handed sector is 3. Like in
1464: the scalar spectrum we need only those $|\tilde{i}>$ vectors made
1465: of harmonics with $l\leq 2$, because the operator $\mathcal{O}_1$
1466: modifies the integration measure just by a factor proportional to
1467: the harmonics $\mathcal{D}^{(1)}$.
1468:  In appendix \ref{spin-1/2calculation} we give an expression for the $|i>,i=1,-1,0,$
1469: vectors, for the $|\tilde{i}>$ vectors and the $M^2_{\tilde{i}}$ eigenvalues for the relevant values of $l$: $l=1,2$.
1470: Here we give the final result: the right-handed low energy spectrum has
1471: a pair of massless right-handed fermions as in the 4D effective theory, which have opposite charge
1472: under the residual $U(1)$ symmetry, and a massive right-handed fermion with the same squared mass given in (\ref{fermionmass}).
1473: This right-handed fermion together with the massive left-handed fermion form a massive Dirac spinor with mass $M_F$.
1474: 
1475: Also in the fermionic sector we note that the heavy modes
1476: contribution is needed in order that the effective theory
1477: reproduces the correct 6-dimensional result; this sentence is evident if one
1478: compares the effective theory prediction (\ref{efffermion}) with
1479: the correct result (\ref{fermionmass}).
1480: 
1481: 
1482: \section{Summary and Conclusions}\label{conclusions}\setcounter{equation}{0}
1483: 
1484: The principal result of this paper is that the contribution of the
1485: heavy KK modes to the effective 4-dimensional  action is necessary
1486: in order to reproduce the correct D-dimensional predictions
1487: concerning the light KK modes. We have calculated such a
1488: contribution for a class of scalar theories. However this result
1489: holds in a more general framework. In order to show this, we have
1490: studied a 6-dimensional gauge and gravitational theory which
1491: involves a complex scalar and, possibly, fermions. In particular we
1492: have considered the compactification over $S^2$, for a particular
1493: value of the monopole number ($n=2$), and the construction of a 4D
1494: $SU(2)\times U(1)$ effective theory. The latter contains a scalar
1495: triplet of  $SU(2)$ which, through an Higgs mechanism, gives masses
1496: to the vector, scalar and fermion fields. An explicit expressions
1497: for these masses and for the VEV of the scalar triplet was found at
1498: the leading order in the small mass ratio $\mu/M$, where $M$ is the lightest heavy mass. On the other
1499: hand, for $n=2$, we found a simple perturbative solution of the
1500: fundamental 6-dimensional EOMs with the same symmetry of the 4D
1501: effective theory in the broken phase. This solution presents a
1502: deformation of the internal space $S^2$ to an ellipsoid, which has
1503: isometry group $U(1)$ instead of $SU(2)$. Moreover we computed the
1504: corresponding vector, scalar and fermion spectrum with quantum
1505: mechanics perturbation theory technique. We have demonstrated by
1506: direct calculation that these quantities, computed in the
1507: 6-dimensional approach, are equal to the corresponding predictions
1508: of the 4D effective theory only if the contribution of the heavy KK
1509: modes  are taken into account. In table \ref{comparison} we give the
1510: spectrum predicted by the 4D effective theory for $c_i=0$, namely,
1511: without heavy KK modes contribution, and the low energy spectrum
1512: predicted by the 6-dimensional theory for the stable ($\mu^2<0$)
1513: solution, that we gave in the text.
1514: %
1515: \begin{table}[top]
1516: \begin{center}
1517: \begin{tabular}{|l|l|l|}
1518: \hline Squared Mass  & 4D Effective Theory &\rule{0.80cm}{0pt} 6D
1519: Theory \quad \quad \quad  \\ \hline
1520:  \raisebox{-0.30cm}{\rule{0pt}{0.80cm}} \quad $M^2_{V}$ & \quad \quad \quad $\frac{3e^2}{8\pi
1521: a^2}\frac{-\mu^2}{\lambda_H}$ & \quad \quad$\frac{3e^2}{8\pi
1522: a^2}\frac{-\mu^2}{\lambda_H-\lambda_G}$ \\ \hline
1523:  \raisebox{-0.30cm}{\rule{0pt}{0.80cm}} \quad $M^2_{V\pm}$ & \quad \quad \quad $\frac{9e^2}{16\pi a^2}\frac{-\mu^2}{\lambda_H}$ &
1524: \quad \quad $\frac{9e^2}{16\pi a^2}\frac{-\mu^2}{\lambda_H-\lambda_G}$ \\ \hline
1525: \raisebox{-0.30cm}{\rule{0pt}{0.80cm}} \quad $M^2_{S}$ & \quad \quad \quad $-2\mu^2$ & \quad \quad $-2\mu^2$ \\
1526: \hline      \raisebox{-0.30cm}{\rule{0pt}{0.80cm}} \quad $M^2_{S\pm}$ & \quad \quad \quad $-\mu^2$ &
1527: \quad \quad $-\mu^2\frac{\lambda_H+\lambda_G}{\lambda_H-\lambda_G}$ \\\hline
1528: \raisebox{-0.30cm}{\rule{0pt}{0.80cm}}
1529: \quad $M^2_{F}$ & \quad \quad \quad $\frac{3g_Y^2}{16\pi a^2}\frac{-\mu^2}{\lambda_H}$
1530: & \quad \quad $\frac{3g_Y^2}{16\pi a^2}\frac{-\mu^2}{\lambda_H-\lambda_G}$ \\
1531: \hline  \raisebox{-0.30cm}{\rule{0pt}{0.80cm}} \quad $M^2_{F\pm}$ & \quad \quad \quad $0$ &
1532: \quad \quad $0$ \\ \hline
1533: \end{tabular}
1534: \end{center}\caption{\footnotesize  The spectra predicted by the 4D effective theory without heavy modes
1535: contribution ($c_i=0$) and by the 6D theory.} \label{comparison}
1536: \end{table}
1537: %
1538: We observe that ratios of masses which involve only vector and
1539: fermion excitations are correctly predicted by the 4D effective
1540: theory even without the heavy KK modes contribution. But the ratios
1541: of masses which involve at least one scalar mode are not correctly
1542: predicted and the error is measured by $\lambda_G/\lambda_H$, where
1543: $\lambda_G$ and $\lambda_H$ are defined in equations (\ref{lGlH}). We can
1544: roughly estimate the magnitude of this disagreement: if we require
1545: $g_1$ and $g_2$ in (\ref{g12}) to be of the order of $1$ and we
1546: consider also the relation  between $\kappa$ and the 4-dimensional
1547: Planck length $\kappa_4$
1548: %
1549: \be \frac{4\pi a^2}{\kappa^2}=\frac{1}{\kappa_4^2}, \ee
1550: %
1551: we get that $\sqrt {\kappa}$, $e$ and $a$ are all of the order of $\kappa_4$.
1552: So roughly speaking the condition $\lambda_G/\lambda_H\ll 1$
1553: becomes $\lambda_H\gg 1$, which is a strong coupling regime. Therefore we can't probably
1554: neglect the heavy KK modes contribution and believe in the perturbation theory of quantum field theory at the same time.
1555: 
1556: Finally we note that there's a value of $c_1$ and $c_2$ ($c_1=-1/3$,
1557: $c_2=1$) such that the effective theory VEV and vector, scalar and
1558: fermion spectrum turn out to be correct, namely, they are equal to
1559: the corresponding quantities given in section \ref{u13}. This is a sign of the equivalence between the
1560: geometrical approach, which involves the deformed internal space
1561: geometry, to the spontaneous symmetry breaking and the Higgs
1562: mechanism in the 4D effective theory. In particular the heavy KK
1563: modes contribution can be interpreted in a geometrical way as the
1564: internal space deformation of the 6-dimensional solution: in fact if
1565: we put $\b=0$ but we keep $\a \neq 0$ in (\ref{solution2}), which
1566: corresponds to neglecting the $S^2$ deformation, we get exactly the
1567: VEV and the spectrum predicted by the 4D effective theory without
1568: heavy KK modes contribution.
1569: 
1570: Possible uses of our work can be its extension to the case which
1571: resembles more the standard electro-weak theory. The latter could be
1572: for instance the 6D gauge and gravitational theory of this paper,
1573: compactified over $S^2$ but with monopole number $n=1$; in this case
1574: we have in fact an Higgs doublet in the 4D effective theory. Other
1575: interesting applications could be models without fundamental
1576: scalars, which, in some sense, geometrize the Higgs mechanism or the
1577: context of supersymmetric version of 6D gauge and gravitational
1578: theories. Such supersymmetric theories have been recently
1579: investigated in connection with attempts to find a solution to the
1580:  cosmological dark energy  problem, a summary of which can be found in
1581: \cite{Burgess:2004ib}.
1582: 
1583: 
1584: 
1585: 
1586: 
1587: 
1588: \vspace{1cm}
1589: 
1590: {\bf Acknowledgments.} This work was supported in part by the Swiss
1591: Science Foundation. A.S. is appreciative of the hospitality at the
1592: IPT of Lausanne and the support by INFN.
1593: 
1594: 
1595: 
1596: \newpage
1597: 
1598: \appendix
1599: 
1600: {\Large \bf Appendix}
1601: 
1602: %\renewcommand{\thesection}{Appendix \Alph{section}}
1603: 
1604: \section{Conventions and Notations} \label{GR}\setcounter{equation}{0}
1605: 
1606: We choose the signature $-,+,+,+,...$ for the metric $G_{MN}$. The Riemann
1607: tensor is defined as follows
1608: %
1609: \be R_{MNS}^{R}=\partial_M \Gamma_{NS}^R -\partial_N \Gamma_{MS}^R +
1610: \Gamma_{MP}^R \Gamma_{NS}^P -\Gamma_{NP}^R \Gamma_{MS}^P,  \ee
1611: %
1612: where the $\Gamma 's$ are the Levi-Civita connection. While the Ricci tensor
1613: and the Ricci scalar
1614: %
1615: \be R_{MN}=R_{PMN}^{P}, \ \ \ \ R=G^{MN}R_{MN}.  \ee
1616: %
1617: 
1618: Our choice for the 6-dimensional gamma matrices is
1619: %
1620: \be \Gamma^{\mu}=\left(\bacc 0 & \gamma^{\mu} \\
1621: \gamma^{\mu} & 0 \ea \right), \quad  \Gamma^5=\left(\bacc 0 & \gamma^5 \\
1622: \gamma^5 & 0 \ea \right),\quad  \Gamma^6=\left(\bacc 0 & -i \\
1623: i & 0 \ea \right), \ee
1624: %
1625: where the $\gamma^{\mu}$ are the 4-dimensional gamma matrices and $\gamma^5$ the 4-dimensional
1626: chirality matrix.
1627: 
1628: 
1629: 
1630: We define the harmonics $\mathcal{D}^{(l)\lambda}_m$ as proportional to the matrix element
1631: %
1632: \be \left<l,\lambda \right|e^{i\varphi Q_3}e^{i(\pi-\theta)Q_2}e^{i\varphi Q_3}\left|l,m\right>, \ee
1633: %
1634: where the $Q_j$, $j=1,2,3$, are the generators of $SU(2)$:
1635: %
1636: \be \left[Q_j,Q_k\right]=i\epsilon_{jkl}Q_l, \ee
1637: %
1638: where $\epsilon_{jkl}$ is the totally antisymmetric Levi-Civita symbol with $\epsilon_{123}=1$. Moreover
1639: $\left|l,m\right>$ is the eigenvector of $\sum_j Q_j^2$ with eigenvalue $l(l+1)$ and the
1640: eigenvector of $Q_3$ with eigenvalue $m$.
1641: 
1642: We introduce also $\mathcal{D}^{(l)}_{\lambda, m}\equiv \mathcal{D}^{(l)-\lambda}_m $ and, for $l=1$,
1643: $\mathcal{D}_{\lambda, m}\equiv \mathcal{D}^{(1)}_{\lambda, m}$; our choice is
1644: %
1645: \be \mathcal{D}_{\hat{\alpha},\hat{\beta}}(\theta,\varphi)=
1646: \left(\ba {ccc} \frac{1}{2}(\cos\theta+1) &
1647: \frac{1}{2}(\cos\theta-1)e^{-2i\varphi} &
1648: -\frac{1}{\sqrt2}\sin\theta
1649: e^{-i\varphi} \\
1650:  \frac{1}{2}(\cos\theta-1)e^{2i\varphi}
1651:  & \frac{1}{2}(\cos\theta+1)
1652:  & -\frac{1}{\sqrt2}\sin\theta
1653: e^{i\varphi} \\
1654:  \frac{1}{\sqrt2}\sin\theta e^{i\varphi} &
1655: \frac{1}{\sqrt2}\sin\theta e^{-i\varphi} & \cos\theta
1656: \ea\right).\label{harmonics} \ee
1657: %
1658: In (\ref{harmonics}) the first, second and third rows correspond
1659: to $\hat{\alpha}=+,-,3$, the first, second and third columns to
1660: $\hat{\beta}=+,-,3$.
1661: 
1662: While our choice for $\mathcal{D}^{(2)}_{\lambda,m}$ is
1663: %
1664:  $$\mathcal{D}^{(2)}_{\lambda,2}(\theta,\varphi)=\left(\ba {c} \frac{1}{4}\left(1+\cos\theta \right)^2 \\
1665:   -\frac{1}{2}\sin\theta (1+\cos\theta)e^{i\varphi}
1666:  \\
1667:   \sqrt{\frac{3}{8}}\sin^2\theta e^{2 i\varphi} \\ -\frac{1}{2}\sin\theta (1-\cos\theta)e^{3 i\varphi}  \\
1668: \frac{1}{4}\left( 1-\cos\theta\right)^2 e^{4i\varphi}
1669: \ea\right),\, \mathcal{D}^{(2)}_{\lambda,1}(\theta,\varphi)=\left(\ba {c} -\frac{1}{2}\sin\theta (1+\cos\theta)e^{-i\varphi}  \\
1670: 
1671:   \frac{1}{2}(1-\cos\theta-2\cos^2\theta)  \\
1672: 
1673:  \sqrt{\frac{3}{2}}\sin\theta \cos\theta e^{i\varphi} \\
1674: 
1675: \frac{1}{4}(4\cos^2\theta-2\cos\theta -2) e^{2i\varphi}   \\
1676: 
1677: \frac{1}{2}\sin\theta (1-\cos\theta)e^{3i\varphi}
1678: \ea\right), $$
1679: %
1680: %
1681:  $$\mathcal{D}^{(2)}_{\lambda,0}(\theta,\varphi)=\left(\ba {c} \sqrt{\frac{3}{8}}\sin^2\theta e^{-2 i\varphi} \\
1682:   \sqrt{\frac{3}{2}}\sin\theta \cos\theta e^{-i\varphi}
1683:  \\
1684:   \frac{1}{2}(3\cos^2\theta -1)\\-\sqrt{\frac{3}{2}}\sin\theta \cos\theta e^{i\varphi}  \\
1685: \sqrt{\frac{3}{8}}\sin^2\theta e^{2 i\varphi}
1686: \ea\right),\, \mathcal{D}^{(2)}_{\lambda,-1}(\theta,\varphi)=\left(\ba {c} -\frac{1}{2}\sin\theta (1-\cos\theta)e^{-3i\varphi}  \\
1687: 
1688:   \frac{1}{4}(4\cos^2\theta-2\cos\theta -2) e^{-2i\varphi}  \\
1689: 
1690:  -\sqrt{\frac{3}{2}}\sin\theta \cos\theta e^{-i\varphi} \\
1691: 
1692: \frac{1}{2}(1-\cos\theta-2\cos^2\theta)   \\
1693: 
1694: \frac{1}{2}\sin\theta (1+\cos\theta)e^{i\varphi}
1695: \ea\right), $$
1696: %
1697: $$\mathcal{D}^{(2)}_{\lambda,-2}(\theta,\varphi)=\left(\ba {c} \frac{1}{4}\left( 1-\cos\theta\right)^2 e^{-4i\varphi} \\
1698:   \frac{1}{2}\sin\theta (1-\cos\theta) e^{-3i\varphi}
1699:  \\
1700:   \sqrt{\frac{3}{8}}\sin^2\theta e^{-2 i\varphi}\\\frac{1}{2}\sin\theta (1+\cos\theta)e^{-i\varphi}   \\
1701: \frac{1}{4}\left(1+\cos\theta \right)^2
1702: \ea\right),$$
1703: 
1704: where $\lambda $ is the row index.
1705: 
1706: 
1707: 
1708: 
1709: 
1710: 
1711: \section{Spin-1 Mass Terms from $S_F$ and $S_R$}\label{SFSR}\setcounter{equation}{0}
1712: 
1713: 
1714: 
1715: \subsection{ $S_F$ Contribution}\label{S_F}
1716: 
1717: In this subsection we write the contribution of
1718: %
1719: \be S_F\equiv -\frac{1}{4}\int d^6X \sqrt{-G}F^2 \ee
1720: %
1721: to the bilinear terms of $V$, $U$ and $W$.
1722: By direct computation we get kinetic terms for $V$ and $U$ and some mass terms for $U$ and $W$:
1723: %
1724: \bea &&-\frac{1}{4}\int d^2y \e F^2=
1725: -\frac{1}{4}V_{\mu\nu}V^{\mu\nu}K -\frac{1}{6}U_{\mu\nu}^{\hat{\alpha}}U^{\mu\nu \hb}K_{\ha \hb} \nonumber\\
1726:      && -\frac{2}{3}U_{\mu}^{\ha}U^{\mu \hb}M^{(1)}_{\ha \hb}+
1727: \frac{4}{3}U_{\mu}^{\ha}W^{\mu \hb}M^{(2)}_{\ha \hb}
1728: -\frac{2}{3}W_{\mu}^{\ha}W^{\mu \hb}M^{(3)}_{\ha \hb}+...,
1729: \eea
1730: %
1731: where the 4-dimensional curved indices $\mu$ and $\nu$ are contracted with the 4-dimensional metric $g_{\mu \nu}$,
1732: the dots are constant terms and interaction terms, moreover
1733: %
1734: \be V_{\mu\nu}=\partial_{\mu}V_{\nu}-\partial_{\nu}V_{\mu},\,\,\,
1735: U_{\mu\nu}^{\ha}=\partial_{\mu}U_{\nu}^{\ha}-\partial_{\nu}U_{\mu}^{\ha} \ee
1736: %
1737: and
1738: 
1739: %
1740: \bea K&=&\frac{1}{4\pi a^2}\int d^2y \,\e,
1741: \,\,\,\, K_{\ha \hb}=\frac{3}{4\pi}\left(\frac{\kappa}{\sqrt2 ea^2}\right)^2
1742: \int d^2 y \,\e\,\D^3_{\ha}\D^3_{\hb}, \nonumber \\
1743: M^{(1)}_{\ha \hb}&=&\frac{3}{8\pi}\left(\frac{\kappa}{\sqrt2 ea^2}\right)^2
1744: \int d^2 y \,\e\,g^{mn}\partial_m \D_{\ha}^3\partial_n\D_{\hb}^3 \nonumber \\
1745: M^{(2)}_{\ha \hb}&=&-\frac{3 \kappa^2}{16 \pi ea^3}\int d^2 y \,\e\,
1746: \partial_m\D_{\ha}^3\D_{\hb}^{\alpha}e_{\alpha}^ng^{mq}F_{nq} \nonumber \\
1747: M^{(3)}_{\ha \hb}&=&\frac{3\kappa^2}{16\pi a^2}
1748: \int d^2 y \,\e\,\D_{\ha}^{\alpha}e_{\alpha}^m\D_{\hb}^{\beta}e_{\beta}^pF_p^{\,\,n}F_{mn}.\label{F2bil}
1749: \eea
1750: %
1751: The results (\ref{F2bil}) are valid for all background $e^{\alpha}$ and $e^3$. We use the $SU(2)\times U(1)$ background
1752: in the subsection \ref{S^2simm}, the $U(1)_3$ background
1753: in the subsection \ref{S^2}.
1754: 
1755: 
1756: 
1757: \subsection{$S_R$ Contribution} \label{S_R}
1758: 
1759: In this subsection we write the contribution of
1760: %
1761: \be S_R\equiv \int d^6X \sqrt{-G}\frac{1}{\kappa^2}R \ee
1762: %
1763: to the bilinear terms of $W$. The complete contribution
1764: of $S_R$ to the 4-dimensional action is given in \cite{SS} in the case of non deformed background solutions. Here we
1765: need explicit expressions, at least for the bilinears, which are also valid for deformed solutions.
1766: We get a kinetic term and a mass term of $W$: up to a total derivative we have
1767: 
1768: %
1769: \bea &&\int d^2y \frac{1}{\kappa^2}\,\e\,R= -\frac{1}{6}W_{\mu\nu}^{\hat{\alpha}}W^{\mu\nu \hb}K'_{\ha \hb} \nonumber\\
1770:      && +W_{\mu}^{\ha}W^{\mu \hb}M^{(4)}_{\ha \hb}+...,
1771: \eea
1772: %
1773: where the dots include constant and interaction terms; moreover
1774: %
1775: \be W_{\mu\nu}^{\ha}=\partial_{\mu}W_{\nu}^{\ha}-\partial_{\nu}W_{\mu}^{\ha}, \ee
1776: %
1777: and
1778: %
1779: \bea K'_{\ha \hb}&=&\frac{3}{8\pi a^2}\int d^2 y \,\e\,\D^{\alpha}_{\ha}\D^{\beta}_{\hb}g_{\alpha \beta}, \nonumber \\
1780: M^{(4)}_{\ha \hb}&=&\frac{1}{4\pi a^2}\int d^2 y \,\e\left[\partial_n\D_{\ha}^{\a}\D_{\hb}^{\b}\left(-e^m_{\a}\o_{m\,\,\,
1781: \beta} ^{\,\,\,\c}e^n_{\c}-g_{\a\d}g^{nm}\o_{m\,\,\,
1782: \beta} ^{\,\,\,\d}+2e^n_{\a}e^m_{\c}\o_{m\,\,\,
1783: \beta} ^{\,\,\,\c} \right)\right.+ \nonumber \\
1784: &&+D_{\ha}^{\a}\D_{\hb}^{\b}\left(-\frac{1}{2}\o_{n\,\,\,
1785: \a} ^{\,\,\,\c}e^m_{\c}\o_{m\,\,\,
1786: \beta} ^{\,\,\,\d}e^n_{\d}-\frac{1}{2}\o_{n\,\,\,
1787: \a} ^{\,\,\,\d}g^{nm}\o_{m\d \beta}+\o_{n\,\,\,
1788: \a} ^{\,\,\,\d}e^n_{\d}\o_{m\,\,\,
1789: \beta} ^{\,\,\,\c}e^m_{\c}\right)\nonumber \\
1790: &&\left.+ \partial_n\D_{\ha}^{\a}\partial_m\D_{\hb}^{\b}\left(-\frac{1}{2}e^m_{\a}e^n_{\beta}+e^n_{\a}e^m_{\beta}
1791: -\frac{1}{2}g_{\a \b}g^{nm}\right)\right],
1792: \label{SRbil}\eea
1793: %
1794: where $\o_{n\,\,\,\beta} ^{\,\,\,\a}$ is the 2-dimensional spin connection for $e_n^{\a}$.
1795: The results (\ref{SRbil}) are also valid for every background $e^{\alpha}$ and $e^3$. We use the $SU(2)\times U(1)$ background
1796: in the subsection \ref{S^2simm}, the $U(1)_3$ background
1797: in the subsection \ref{S^2}.
1798: 
1799: 
1800: \subsection{The case of $SU(2)\times U(1)$ background} \label{S^2simm}
1801: 
1802: We use now the $SU(2)\times U(1)$ background, that is $\eta =0$. This computation is
1803: performed in \cite{RSS}. We have the following bilinear terms for $V$, $U$ and $W$:
1804: %
1805: \bea && -\frac{1}{4}V_{\mu\nu}V^{\mu\nu}
1806: -\frac{1}{6}U_{\mu\nu}^{\hat{\alpha}}U^{\mu\nu}_{ \ha}-\frac{1}{6}W_{\mu\nu}^{\hat{\alpha}}W^{\mu\nu}_{ \ha} \nonumber\\
1807: &&-\frac{2}{3a^2}\left(U_{\mu \hat{\alpha}}-W_{\mu
1808: \hat{\alpha}}\right)\left(U^{\mu \hat{\alpha}}-W^{\mu
1809: \hat{\alpha}}\right). \label{bilinearV}\eea
1810: %
1811: %
1812:  If we define
1813: %
1814: \bea \mathcal{A}&=&\sqrt{\frac{1}{3}}(W+U), \nonumber \\
1815: X&=&\sqrt{\frac{1}{3}}(W-U), \label{def2}\eea
1816: %
1817: we can write (\ref{bilinearV}) as follows
1818: %
1819: \bea
1820: &&-\frac{1}{4}V_{\mu\nu}V^{\mu\nu}-\frac{1}{4}\mathcal{A}_{\mu\nu}^{\hat{\alpha}}\mathcal{A}^{\mu\nu}_{ \ha} \nonumber \\
1821: &&-\frac{1}{4}X_{\mu\nu}^{\hat{\alpha}}X^{\mu\nu }_{\ha}
1822: -\frac{2}{a^2}X_{\mu\hat{\alpha}}X^{\mu\hat{\alpha}},\eea
1823: %
1824: So $\mathcal{A}$ is a massless field, in fact it's the $SU(2)$
1825: Yang-Mills field \cite{RSS}, while $X$ is a massive field which
1826: can be neglected in the low-energy limit.
1827: 
1828: 
1829: 
1830: 
1831: 
1832: 
1833: \subsection{The Case of $U(1)_3$ Background} \label{S^2}
1834: 
1835: Let's consider now the solution (\ref{solution2}).
1836:  First we note that $S_R$ and $S_F$ don't give mass terms
1837: for $V$; so the only source for the mass of $V$ is $S_{\phi}$.
1838: 
1839: We want to prove now that also the $SU(2)$
1840: Yang-Mills fields masses don't receive contributions from $S_R$ and $S_F$.
1841: First we give the bilinears for $U$ and $W$, which come from $S_R$ and $S_F$:
1842: %
1843: \bea &&-\frac{1}{6}U_{\mu\nu}^{\hat{\alpha}}U^{\mu\nu \hb}g_{\ha
1844: \hb}\left(1+|\eta| \beta k_{\ha}\right)
1845: -\frac{1}{6}W_{\mu\nu}^{\hat{\alpha}}W^{\mu\nu \hb}g_{\ha \hb}\left(1+|\eta| \beta k'_{\ha}\right) \nonumber\\
1846:  &&-\frac{2}{3}U_{\mu}^{\ha}U^{\mu \hb}g_{\ha \hb}\left(1+|\eta| \beta m^{(1)}_{\ha}\right)
1847: +\frac{4}{3}U_{\mu}^{\ha}W^{\mu \hb}g_{\ha \hb}\left(1+|\eta| \beta m^{(2)}_{\ha}\right) \nonumber\\
1848: &&-\frac{2}{3}W_{\mu}^{\ha}W^{\mu \hb}g_{\ha \hb}\left(1+|\eta| \beta m^{(3)}_{\ha}\right),\label{bilinearsu13} \nonumber\\
1849: \eea
1850: %
1851: where
1852: %
1853: \begin{displaymath} k_+=k_-=\frac{2}{5}, \,\,\, k_3=\frac{1}{5}, \,\,\,
1854: k'_+=k'_-=\frac{3}{10}, \,\,\, k'_3=\frac{2}{5}, \nonumber\ed
1855: %
1856: %
1857: \bd m^{(1)}_+=m^{(1)}_-=\frac{1}{5}, \,\,\,
1858: m^{(1)}_3=-\frac{2}{5}, \nonumber\ed
1859: %
1860: \bd m^{(2)}_+=m^{(2)}_-=-\frac{1}{20},\,\,\,
1861: m^{(2)}_3=-\frac{2}{5},\,\,\, m^{(3)}_+=m^{(3)}_-=-\frac{3}{10}, \,\,\,
1862: m^{(3)}_3=-\frac{2}{5}. \nonumber\ed
1863: %
1864: 
1865: In order to prove (\ref{bilinearsu13}) it's useful to use the following formula for the background
1866: spin connection:
1867:  %
1868: \be \omega_{\varphi\,\,\, +} ^{\,\,+}=-\omega_{\varphi\,\,\, -}
1869: ^{\,\,-}=\frac{i}{a}(\cos\theta-1 -\frac{1}{2}|\eta|\b\cos\theta \sin^{2}\theta) \label{spinconn}\ee
1870: %
1871: and $\omega_{\varphi\,\,\, +} ^{\,\,-}=\omega_{\varphi\,\,\, -} ^{\,\,+}=0$.
1872: 
1873: Now we define $X$ and $A$ as follows
1874: %
1875: \bea \left(1+\frac{|\eta| \beta}{2}k'_{\ha}\right)W^{\ha}=
1876: \sqrt{\frac{3}{2}}\left(\cos\theta_{\eta}^{\ha}X^{\ha}+\sin\theta_{\eta}^{\ha}\mathcal{A}^{\ha}\right), \nonumber\\
1877: \left(1+\frac{|\eta| \beta}{2}k_{\ha}\right)U^{\ha}=
1878: \sqrt{\frac{3}{2}}\left(-\sin\theta_{\eta}^{\ha}X^{\ha}+\cos\theta_{\eta}^{\ha}\mathcal{A}^{\ha}\right),
1879: \label{defeps} \eea
1880: %
1881: where the angle $\theta_{\eta}^{\ha}$ is defined by
1882: %
1883: \be \cos\theta_{\eta}^{\ha}=\frac{1+|\eta| \beta
1884: \delta^{\ha}}{\sqrt2},\,\,\,
1885: \sin\theta_{\eta}^{\ha}=\frac{1-|\eta| \beta
1886: \delta^{\ha}}{\sqrt2},\ee
1887: %
1888: and the quantities $\d^{\ha}$ are not still fixed. It's simple to check that the kinetic terms for $X$ and $\mathcal{A}$ are in the
1889: standard form for every $\d^{\ha}$ up to $O(\eta^{3/2})$.
1890: The definition (\ref{defeps}) reduce to (\ref{def2}) for $\eta=0$.
1891: 
1892: If we choose
1893: %
1894: \be \d^{\ha}=\frac{1}{8}\left(m^{(3)}_{\ha}-k'_{\ha}-m^{(1)}_{\ha}+k_{\ha}\right) \ee
1895: %
1896: we have no mass terms for $\mathcal{A}$ coming from $S_R+S_F$.
1897: 
1898: So the only source for the spin-1 low energy spectrum is $S_{\phi}$ and the
1899: result is given in equations (\ref{M0}) and (\ref{M1}).
1900: 
1901: 
1902: \section{Explicit Calculation of Spin-0 Spectrum for the 6D Theory}\label{spin-0calculation}\setcounter{equation}{0}
1903: 
1904: As we pointed out in the text, in order to find the spin-0 spectrum
1905: the expression of the $|i>$, $i=1,...,6$, vectors is needed; these
1906: are defined by $\mathcal{O}_0|i>=0$, which is equivalent to
1907: $\nabla^2\phi+\phi/a^2=0$, where $\nabla^2\phi$ is the laplacian
1908: over the charged scalar $\phi$, calculated with the round $S^2$
1909: metric. Our choice for the orthonormal vectors\footnote{We express a
1910: generic vector as in (\ref{Sdefinition}).}  $|i>$ is
1911: 
1912: %
1913:  $$|1>=\frac{1}{\sqrt{2}}\left(\ba {c} \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,1}\\
1914:   \sqrt{\frac{3}{4\pi}}\left(\mathcal{D}^{(1)}_{-1,1}\right)^*
1915:  \\
1916:   0  \\
1917: .\\
1918: . \\
1919:  .  \\
1920: 
1921:  0
1922: \ea\right),\,|2>=\frac{1}{\sqrt{2}}\left(\ba {c} i\sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,1}\\
1923:   -i\sqrt{\frac{3}{4\pi}}\left(\mathcal{D}^{(1)}_{-1,1}\right)^*
1924:  \\
1925:   0  \\
1926: .\\
1927: . \\
1928:  .  \\
1929:  0
1930: \ea\right), $$
1931: %
1932: %
1933:  $$|3>=\frac{1}{\sqrt{2}}\left(\ba {c} \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,0}\\
1934:   \sqrt{\frac{3}{4\pi}}\left(\mathcal{D}^{(1)}_{-1,0}\right)^*
1935:  \\
1936:   0  \\
1937: .\\
1938: . \\
1939:  .  \\
1940: 
1941:  0
1942: \ea\right), \,|4>=\frac{1}{\sqrt{2}}\left(\ba {c} i\sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,0}\\
1943:   -i\sqrt{\frac{3}{4\pi}}\left(\mathcal{D}^{(1)}_{-1,0}\right)^*
1944:  \\
1945:   0  \\
1946: .\\
1947: . \\
1948:  .  \\
1949:  0
1950: \ea\right), $$
1951: %
1952: %
1953: $$|5>=\frac{1}{\sqrt{2}}\left(\ba {c} \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,-1}\\
1954:   \sqrt{\frac{3}{4\pi}}\left(\mathcal{D}^{(1)}_{-1,-1}\right)^*
1955:  \\
1956:   0  \\
1957: .\\
1958: . \\
1959:  .  \\
1960: 
1961:  0
1962: \ea\right), \,|6>=\frac{1}{\sqrt{2}}\left(\ba {c} i\sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,-1}\\
1963:   -i\sqrt{\frac{3}{4\pi}}\left(\mathcal{D}^{(1)}_{-1,-1}\right)^*
1964:  \\
1965:   0  \\
1966: .\\
1967: . \\
1968:  .  \\
1969:  0
1970: \ea\right). $$
1971: %
1972: %
1973: 
1974: Another ingredient for the calculation of the spin-0 spectrum is an
1975: explicit expression of the vectors $|\tilde{i}>$ and of the
1976: eigenvalues $M^2_{\tilde{i}}$. As we explained in the text, only the
1977: $|\tilde{i}>$ like (\ref{tildei}) and made of $l=0,1,2$ harmonics
1978: are needed. The $|\tilde{i}> $ vectors must satisfy the following
1979: eigenvalue equations\footnote{We derive (\ref{eigen}) evaluating
1980: (\ref{L02}), (\ref{L03}) and (\ref{L06}) in the basis (\ref{epm})
1981: and performing the redefinition $h_{\pm \pm}\rightarrow
1982: \sqrt{2}\kappa h_{\pm \pm}$ and $h_{+-}\rightarrow
1983: h_{+-}\kappa/\sqrt{2}$, which normalizes the kinetic terms in the
1984: standard way.}:
1985: %
1986: \bea &&-\nabla^2h_{++}+2R_{+-+-}h_{++}-2\kappa^2F_{+-}^2h_{++}-\sqrt{2}\kappa\nabla_{+}\mv_{+}F_{-+}=M^2h_{++},\nonumber \\
1987: &&-\nabla^2h_{--}+2R_{+-+-}h_{--}-2\kappa^2F_{+-}^2h_{--}+\sqrt{2}\kappa\nabla_{-}\mv_{-}F_{-+}=M^2h_{--},\nonumber \\
1988: &&-\nabla^2h_{+-}-R_{+-+-}h_{+-}-\frac{\kappa}{\sqrt2}\nabla_+\mv_-F_{-+}+
1989: \frac{\kappa}{\sqrt2}\nabla_-\mv_+F_{-+}=M^2h_{+-},\nonumber \\
1990: &&-\nabla^2\mv_{+}+R_{+-}\mv_{+}-\kappa^2\mv_+ F^2_{+-}+
1991: \frac{\kappa}{\sqrt2}\nabla_+h_{+-}F_{-+}-\sqrt2 \kappa\nabla_-h_{++}F_{-+}=M^2\mv_{+},\nonumber \\
1992: &&-\nabla^2\mv_{-}+R_{+-}\mv_{-}-\kappa^2\mv_- F^2_{+-}-
1993: \frac{\kappa}{\sqrt2}\nabla_-h_{+-}F_{-+}+\sqrt2 \kappa\nabla_+h_{--}F_{-+}=M^2\mv_{-},\nonumber \\
1994: \label{eigen}\eea
1995: %
1996: where the background objects ($\nabla^2$, $R_{+-+-}$,...) correspond
1997: to the background (\ref{sphere}), (\ref{monopole}) and (\ref{phi0}).
1998: We can transform the differential problem (\ref{eigen}) into an
1999: algebraic one by using the expansion (\ref{lexpansion}). We get an
2000: eigenvalue problem for every value of $l$ and we give now an
2001: explicit expression for the $|\tilde{i}>$ vectors for the relevant
2002: value of $l$, namely $l=0,1,2$. For $l=0$ we get just one
2003: eigenvector $|\tilde{1}>$ with $M^2=1/a^2$:
2004: %
2005: \be |\tilde{1}>=\left(\ba {c} 0 \\
2006:   0
2007:  \\
2008:   0 \\
2009: 0  \\
2010: 1/\sqrt{4\pi}  \\
2011:   0  \\
2012: 0
2013: \ea\right). \ee
2014: %
2015: For $l=1$ we get three different eigenvalues: $M^2=2/a^2, 4/a^2, 5/a^2$ . The eigenvectors which
2016: correspond to $M^2=2/a^2$ are
2017: %
2018: \be |\tilde{2}_0>, \quad \frac{1}{\sqrt2}\left(|\tilde{2}_{1}>+|\tilde{2}_{-1}>\right),\quad
2019: \frac{1}{\sqrt2 i}\left(|\tilde{2}_{1}>-|\tilde{2}_{-1}>\right), \ee
2020: %
2021: where
2022: %
2023: \be |\tilde{2}_{m}>\equiv\frac{1}{\sqrt6}\left(\ba {c} 0 \\
2024:   0
2025:  \\
2026:   0 \\
2027: 0  \\
2028: 2\sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{0,m}  \\
2029:   -\sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{1,m}  \\
2030: -\sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,m}
2031: \ea\right). \ee
2032: %
2033: Instead the eigenvectors which
2034: correspond to $M^2=4/a^2$ are
2035: %
2036: \be i|\tilde{3}_0>, \quad \frac{1}{\sqrt2 i}\left(|\tilde{3}_{1}>+|\tilde{3}_{-1}>\right),\quad
2037: \frac{1}{\sqrt2 }\left(|\tilde{3}_{1}>-|\tilde{3}_{-1}>\right), \ee
2038: %
2039: where
2040: %
2041: \be |\tilde{3}_{m}>\equiv\frac{1}{\sqrt2}\left(\ba {c} 0 \\
2042:   0
2043:  \\
2044:   0 \\
2045: 0  \\
2046: 0  \\
2047:   -\sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{1,m}  \\
2048: \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,m}
2049: \ea\right). \ee
2050: %
2051: Moreover the eigenvectors which
2052: correspond to $M^2=5/a^2$ are
2053: %
2054: \be |\tilde{4}_0>, \quad \frac{1}{\sqrt2 }\left(|\tilde{4}_{1}>+|\tilde{4}_{-1}>\right),\quad
2055: \frac{1}{\sqrt2 i}\left(|\tilde{4}_{1}>-|\tilde{4}_{-1}>\right), \ee
2056: %
2057: where
2058: %
2059: \be |\tilde{4}_{m}>\equiv\frac{1}{\sqrt3}\left(\ba {c} 0 \\
2060:   0
2061:  \\
2062:   0 \\
2063: 0  \\
2064:  \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{0,m} \\
2065:   \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{1,m}  \\
2066: \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,m}
2067: \ea\right). \ee
2068: %
2069: Finally, for $l=2$ the values of $M^2$ are given by
2070: %
2071: \be a^2M^2=6,\,\,2(3-\sqrt3),\,\, 2(3+\sqrt3),\,\,\frac{1}{2}(13-\sqrt{73}) ,\,\,\frac{1}{2}(13+\sqrt{73}).\ee
2072: %
2073: The eigenvectors with $a^2M^2=6$ are
2074: %
2075: \bea &&|\tilde{5}_0>,\quad \frac{1}{\sqrt2}\left(|\tilde{5}_{1}>-|\tilde{5}_{-1}>\right),\quad
2076: \frac{1}{\sqrt2i}\left(|\tilde{5}_{1}>+|\tilde{5}_{-1}>\right),\nonumber \\
2077: &&\frac{1}{\sqrt2}\left(|\tilde{5}_{2}>+|\tilde{5}_{-2}>\right),\quad
2078: \frac{1}{\sqrt2i}\left(|\tilde{5}_{2}>-|\tilde{5}_{-2}>\right),
2079: \eea
2080: %
2081: where
2082: %
2083: \be |\tilde{5}_{m}>\equiv\frac{1}{3\sqrt2}\left(\ba {c} 0 \\
2084: 0 \\
2085: -\sqrt2 \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{2,m} \\
2086:   -\sqrt2 \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-2,m}
2087:  \\
2088:   -2\sqrt3 \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{0,m} \\
2089:   -\sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{1,m}  \\
2090: \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-1,m}
2091: \ea\right). \ee
2092: %
2093: For $a^2M^2=2(3-\sqrt3)$ we have the eigenvectors
2094:  %
2095: \bea &&i|\tilde{6}_0>,\quad \frac{1}{\sqrt2}\left(|\tilde{6}_{1}>+|\tilde{6}_{-1}>\right),\quad
2096: \frac{1}{\sqrt2i}\left(|\tilde{6}_{1}>-|\tilde{6}_{-1}>\right),\nonumber \\
2097: &&\frac{1}{\sqrt2}\left(|\tilde{6}_{2}>-|\tilde{6}_{-2}>\right),\quad
2098: \frac{1}{\sqrt2i}\left(|\tilde{6}_{2}>+|\tilde{6}_{-2}>\right),
2099: \eea
2100: %
2101: where
2102: %
2103: \be |\tilde{6}_{m}>\equiv\frac{1}{\sqrt{2(3+\sqrt3)}}\left(\ba {c} 0 \\
2104: 0 \\
2105: -\frac{1+\sqrt3}{\sqrt2} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{2,m} \\
2106:  \frac{1+\sqrt3}{\sqrt2} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-2,m}
2107:  \\
2108:  0 \\
2109:   \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{1,m}  \\
2110: \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-1,m}
2111: \ea\right). \ee
2112: %
2113: For $a^2M^2=2(3+\sqrt3)$ we have the eigenvectors
2114:  %
2115: \bea &&i|\tilde{7}_0>,\quad \frac{1}{\sqrt2}\left(|\tilde{7}_{1}>+|\tilde{7}_{-1}>\right),\quad
2116: \frac{1}{\sqrt2i}\left(|\tilde{7}_{1}>-|\tilde{7}_{-1}>\right),\nonumber \\
2117: &&\frac{1}{\sqrt2}\left(|\tilde{7}_{2}>-|\tilde{7}_{-2}>\right),\quad
2118: \frac{1}{\sqrt2i}\left(|\tilde{7}_{2}>+|\tilde{7}_{-2}>\right),
2119: \eea
2120: %
2121: where
2122: %
2123: \be |\tilde{7}_{m}>\equiv\frac{1}{\sqrt{2(3-\sqrt3)}}\left(\ba {c} 0 \\
2124: 0 \\
2125: -\frac{1-\sqrt3}{\sqrt2} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{2,m} \\
2126:  \frac{1-\sqrt3}{\sqrt2} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-2,m}
2127:  \\
2128:  0 \\
2129:   \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{1,m}  \\
2130: \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-1,m}
2131: \ea\right). \ee
2132: %
2133: Then for $a^2M^2=(13-\sqrt{73})/2$:
2134: %
2135: \bea &&|\tilde{8}_0>,\quad \frac{1}{\sqrt2}\left(|\tilde{8}_{1}>-|\tilde{8}_{-1}>\right),\quad
2136: \frac{1}{\sqrt2i}\left(|\tilde{8}_{1}>+|\tilde{8}_{-1}>\right),\nonumber \\
2137: &&\frac{1}{\sqrt2}\left(|\tilde{8}_{2}>+|\tilde{8}_{-2}>\right),\quad
2138: \frac{1}{\sqrt2i}\left(|\tilde{8}_{2}>-|\tilde{8}_{-2}>\right),
2139: \eea
2140: %
2141: where
2142: %
2143: \be |\tilde{8}_{m}>\equiv\frac{1+\sqrt{73}}{\sqrt{438+30\sqrt{73}}}\left(\ba {c} 0 \\
2144: 0 \\
2145: \frac{13\sqrt2+\sqrt{146}}{2(1+\sqrt{73})} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{2,m} \\
2146:   \frac{13\sqrt2+\sqrt{146}}{2(1+\sqrt{73})} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-2,m}
2147:  \\
2148:   -\frac{4\sqrt3}{1+\sqrt{73}} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{0,m} \\
2149:   -\sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{1,m}  \\
2150: \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-1,m}
2151: \ea\right). \ee
2152: %
2153: Finally for $a^2M^2=(13+\sqrt{73})/2$:
2154: %
2155: \bea &&|\tilde{9}_0>,\quad \frac{1}{\sqrt2}\left(|\tilde{9}_{1}>-|\tilde{9}_{-1}>\right),\quad
2156: \frac{1}{\sqrt2i}\left(|\tilde{9}_{1}>+|\tilde{9}_{-1}>\right),\nonumber \\
2157: &&\frac{1}{\sqrt2}\left(|\tilde{9}_{2}>+|\tilde{9}_{-2}>\right),\quad
2158: \frac{1}{\sqrt2i}\left(|\tilde{9}_{2}>-|\tilde{9}_{-2}>\right),
2159: \eea
2160: %
2161: where
2162: %
2163: \be |\tilde{9}_{m}>\equiv\frac{1-\sqrt{73}}{\sqrt{438-30\sqrt{73}}}\left(\ba {c} 0 \\
2164: 0 \\
2165: \frac{13\sqrt2-\sqrt{146}}{2(1-\sqrt{73})} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{2,m} \\
2166:   \frac{13\sqrt2-\sqrt{146}}{2(1-\sqrt{73})} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-2,m}
2167:  \\
2168:   -\frac{4\sqrt3}{1-\sqrt{73}} \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{0,m} \\
2169:   -\sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{1,m}  \\
2170: \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-1,m}
2171: \ea\right). \ee
2172: %
2173: 
2174: We can now calculate the $6\times6$ matrix $M^2_{ij}$ given in (\ref{QMperturbation}). In order to do that
2175: we need just the matrix elements $<i|\mathcal{O}_1|\tilde{i}>$ and $<i|\mathcal{O}_2|j>$, which can be
2176: computed by evaluating\footnote{For the background solution (\ref{sphere}), (\ref{monopole}) and (\ref{phi0})
2177: we have  $\mathcal{L}_0(\phi,\mv)=0$.} $\mathcal{L}_0(\phi,h)$ and $\mathcal{L}_0(\phi,\phi)$, which appears
2178: in (\ref{L01}) and (\ref{L04}), in the $\pm$ basis given in (\ref{epm}). After the redefinitions
2179: $h_{\pm \pm}\rightarrow \sqrt{2}k h_{\pm \pm}$ and $h_{+-}\rightarrow h_{+-}k/\sqrt{2}$,
2180: which normalize the kinetic terms in the standard way, we get (for $n=2$)
2181: %
2182: \bea
2183: \mathcal{L}_0(\phi,h)&=&\sqrt2 \kappa\nabla_{+}\Phi\nabla_+h_{--}\phi^{*}
2184: +\frac{\kappa}{\sqrt2}\nabla_{+}\Phi\nabla_-h_{+-}\phi^{*} \nonumber \\
2185: &&+\sqrt2 \kappa\nabla_{+}\Phi h_{--}\left(\nabla_-\phi\right)^{*}
2186: +\frac{\kappa}{\sqrt2}\nabla_{+}\Phi
2187: h_{+-}\left(\nabla_+\phi\right)^{*}+c.c. \, ,\nonumber \\
2188: \mathcal{L}_0(\phi,\phi)&=&\phi^*\partial^2\phi-\phi^*\left[-\nabla^2+m^2+(e^2+4\xi)\Phi^*
2189: \Phi
2190: +\kappa^2\nabla_+\Phi \left(\nabla_+\Phi\right)^*\right]\phi \nonumber \\
2191: &&-\frac{1}{2}\left[\phi(2\xi-e^2)\left(\Phi^*\right)^2\phi+c.c.
2192: \right].\eea
2193: %
2194: By using these expressions and the values of $|i>$ and $|\tilde{i}>$ given before, we find the following expression
2195: for $M^2_{ij}$:
2196: \be  \{M^2_{ij}\}=
2197: \left(\ba {cccccc} a_1 &
2198: 0 &
2199: 0 & 0 & a_4 & 0 \\
2200:  0
2201:  & a_1
2202:  & 0 & 0 & 0 & -a_4 \\
2203: 0 & 0 & a_2 & 0 & 0 & 0 \\
2204: 0 & 0 & 0 & a_3 & 0 & 0 \\
2205: a_4 &
2206: 0 &
2207: 0 & 0 & a_1 & 0 \\
2208: 0
2209:  & -a_4
2210:  & 0 & 0 & 0 & a_1 \ea\right),\label{M2ij} \ee
2211: %
2212: where
2213: %
2214: \bea a_1&=&\frac{|\eta|}{a^2}\left(-sign(\eta) +\frac{3}{10}\b + \frac{12}{5}\frac{\b \xi a^2}{\kappa^2}\right), \nonumber \\
2215:  a_2&=&\frac{|\eta|}{a^2}\left(-sign(\eta) -\frac{6}{5}\b + \frac{24}{5}\frac{\b \xi a^2}{\kappa^2}\right), \nonumber \\
2216: a_3&=&\frac{|\eta|}{a^2}\left(-sign(\eta) +\frac{4}{15}\b + \frac{8}{5}\frac{\b \xi a^2}{\kappa^2}\right), \nonumber \\
2217: a_4&=&\frac{|\eta|}{a^2}\b\left(\frac{3}{10}- \frac{4}{5}\frac{ \xi a^2}{\kappa^2}\right).\eea
2218: %
2219: By diagonalizing $M^2_{ij}$, we found exactly the spectrum that we discussed in the subsection \ref{spin-0}: the squared
2220: masses of the vector particles are reproduced\footnote{In order to see that we use the background
2221: constraints (\ref{constraint}).}, as required by the light cone gauge; moreover we get the two masses squared given in
2222: (\ref{MS}).
2223: 
2224: 
2225: \section{Explicit Calculation of Spin-1/2 Spectrum for the 6D Theory}\label{spin-1/2calculation}\setcounter{equation}{0}
2226: 
2227: Here we concentrate on the right-handed sector, which is the non trivial one
2228: because it presents $\eta^{1/2}$ mixing terms.
2229: 
2230: The eigenvalue equations for the unperturbed ($\eta=0$) mass squared operator $\mathcal{O}_0$,
2231: acting on the right-handed sector, are
2232: %
2233: \bea - 2 \nabla_- \nabla_+\psi_{+R}
2234: =M^2\psi_{+R}, \nonumber \\
2235: - 2\nabla_+ \nabla_- \psi_{-R}
2236: =M^2\psi_{-R}, \label{unperturbedfermion}
2237: \eea
2238: %
2239: The differential equation (\ref{unperturbedfermion}) can be
2240: transformed in an algebraic one through the harmonic expansion,
2241: remembering the iso-helicities of $\psi_{+R}$ and $\psi_{-R}$:
2242: $\lambda_{+R}=\lambda_{-R}=1$.
2243:  Therefore an explicit expression for the vectors $|i>$, which satisfies by definition $\mathcal{O}_0|i>=0$, is given by
2244: %
2245: \be |i>=\left(\ba {c} 0 \\
2246:   \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,i}
2247: \ea\right),\quad \quad i=1,-1,0. \ee
2248: %
2249: We give also an expression for the vectors $|\tilde{i}>$ and the corresponding non vanishing eigenvalues $M^2_{\tilde{i}}$.
2250: For $l=1$, we have just one eigenvalue $M^2=2/a^2$ and the corresponding eigenvectors are
2251: %
2252: \be |\tilde{1}_m>=\left(\ba {c}
2253:   \sqrt{\frac{3}{4\pi}}\mathcal{D}^{(1)}_{-1,m} \\
2254: 0
2255: \ea\right). \ee
2256: %
2257: For $l=2$ we have an eigenvalue $M^2=6/a^2$, which corresponds to the eigenvectors
2258: %
2259: \be |\tilde{2}_m>=\left(\ba {c}
2260:   \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(2)}_{-1,m} \\
2261: 0
2262: \ea\right), \ee
2263: %
2264: and an eigenvalue $M^2=4/a^2$, which corresponds to the eigenvectors
2265: %
2266: \be |\tilde{3}_m>=\left(\ba {c} 0 \\
2267:   \sqrt{\frac{5}{4\pi}}\mathcal{D}^{(1)}_{-1,m}
2268: \ea\right). \ee
2269: %%
2270: By inserting these eigenvectors and eigenvalues in the expression (\ref{QMperturbation}) we get
2271: %
2272: \be M^2_{ij}=diag\left(\,\,0,\,\,0,\,\,\frac{2}{3}|\eta|g^2_Y\frac{\b}{\kappa^2}\right), \ee
2273: %
2274: which corresponds to the spectrum we discussed at the end of section
2275: \ref{spin-1/2}.
2276: 
2277: 
2278: 
2279: \begin{thebibliography}{99}
2280: 
2281: %\cite{GUT}
2282: \bibitem{GUT}
2283:    J.~C.~Pati and A. Salam,
2284:   ``Unified Lepton - Hadron Symmetry, And A Gauge Theory Of The Basic
2285:   Interactions,''
2286:   Phys.\ Rev.\ D {\bf 8} (1973) 1240;
2287:   %%CITATION = PHRVA,D8,1240;%%
2288:   H.~Georgi and S.~L.~Glashow,
2289:   ``Unity Of All Elementary Particle Forces,''
2290:   Phys.\ Rev.\ Lett.\  {\bf 32} (1974) 438;
2291:   %%CITATION = PRLTA,32,438;%%
2292: 
2293: 
2294:   %\cite{Wilson:1974mb}
2295: \bibitem{Wilson}
2296:   K.~G.~Wilson,
2297:   %``The Renormalization Group: Critical Phenomena And The Kondo Problem,''
2298:   Rev.\ Mod.\ Phys.\  {\bf 47}, 773 (1975).
2299:   %%CITATION = RMPHA,47,773;%%
2300: 
2301: \bibitem{Weinberg:1980wa}
2302:   S.~Weinberg,
2303:   ``Effective Gauge Theories,''
2304:   Phys.\ Lett.\ B {\bf 91} (1980) 51.
2305:   %%CITATION = PHLTA,B91,51;%%
2306: 
2307: \bibitem{Appelquist:1974tg}
2308:   T.~Appelquist and J.~Carazzone,
2309:   ``Infrared Singularities And Massive Fields,''
2310:   Phys.\ Rev.\ D {\bf 11} (1975) 2856.
2311:   %%CITATION = PHRVA,D11,2856;%%
2312: 
2313: \bibitem{RSS}
2314:   S.~Randjbar-Daemi, A.~Salam and J.~A.~Strathdee,
2315:   ``Spontaneous Compactification In Six-Dimensional Einstein-Maxwell Theory,''
2316:   Nucl.\ Phys.\ B {\bf 214} (1983) 491.
2317:   %%CITATION = NUPHA,B214,491;%%
2318: 
2319: \bibitem{Gibbons:2003gp}
2320:   G.~W.~Gibbons and C.~N.~Pope,
2321:   ``Consistent $S^2$ Pauli reduction of six-dimensional chiral gauged
2322:   Einstein-Maxwell supergravity,''
2323:   Nucl.\ Phys.\ B {\bf 697} (2004) 225
2324:   [arXiv:hep-th/0307052].
2325:   %%CITATION = HEP-TH 0307052;%%
2326: 
2327: \bibitem{Salam:1984cj}
2328:   A.~Salam and E.~Sezgin,
2329:   ``Chiral Compactification On $(Minkowski)\times S^2$ Of N=2 Einstein-Maxwell
2330:   Supergravity In Six-Dimensions,''
2331:   Phys.\ Lett.\ B {\bf 147} (1984) 47.
2332:   %%CITATION = PHLTA,B147,47;%%
2333: 
2334: %\cite{Buras:1977yy}
2335: \bibitem{Buras:1977yy}
2336:   A.~J.~Buras, J.~R.~Ellis, M.~K.~Gaillard and D.~V.~Nanopoulos,
2337:   ``Aspects Of The Grand Unification Of Strong, Weak And Electromagnetic
2338:   Interactions,''
2339:   Nucl.\ Phys.\ B {\bf 135} (1978) 66.
2340:   %%CITATION = NUPHA,B135,66;%%
2341: 
2342: %\cite{Chan:1979ce}
2343: \bibitem{Chan:1979ce}
2344:   L.~H.~Chan, T.~Hagiwara and B.~A.~Ovrut,
2345:   ``The Effect Of Heavy Particles In Low-Energy Light Particle Processes,''
2346:   Phys.\ Rev.\ D {\bf 20} (1979) 1982.
2347:   %%CITATION = PHRVA,D20,1982;%%
2348: 
2349: \bibitem{Wigner} E. P. Wigner,
2350:    ``Group Theory and its Application to the Quantum Mechanics of Atomics Spectra,''
2351:    New York: Academic Press, 1959.
2352: 
2353: \bibitem{S}
2354:   J.~Sobczyk,
2355:   ``Symmetry Breaking In Kaluza-Klein Theories,''
2356:   Class.\ Quant.\ Grav.\  {\bf 4} (1987) 37.
2357:   %%CITATION = CQGRD,4,37;%%
2358: 
2359: \bibitem{RS} S.~Randjbar-Daemi and M.~Shaposhnikov,
2360:   ``A formalism to analyze the spectrum of brane world scenarios,''
2361:   Nucl.\ Phys.\ B {\bf 645} (2002) 188
2362:   [arXiv:hep-th/0206016].
2363:   %%CITATION = HEP-TH 0206016;%%
2364: 
2365: \bibitem{RSS2} S.~Randjbar-Daemi, A.~Salam and J.~A.~Strathdee,
2366:   ``Towards A Selfconsistent Computation Of Vacuum Energy In Eleven-Dimensional
2367:   Supergravity,''
2368:   Nuovo Cim.\ B {\bf 84} (1984) 167.
2369:   %%CITATION = NUCIA,B84,167;%%
2370: 
2371: %\cite{Burgess:2004ib}
2372: \bibitem{Burgess:2004ib}
2373:   C.~P.~Burgess,
2374:   %``Towards a natural theory of dark energy: Supersymmetric large extra
2375:   %dimensions,''
2376:   AIP Conf.\ Proc.\  {\bf 743} (2005) 417
2377:   [arXiv:hep-th/0411140];
2378:   %%CITATION = HEP-TH 0411140;%%
2379: C.~P.~Burgess,
2380:   ``Supersymmetric large extra dimensions and the cosmological constant
2381:   problem,''
2382:   arXiv:hep-th/0510123.
2383:   %%CITATION = HEP-TH 0510123;%%
2384: %\cite{SS}
2385: 
2386: \bibitem{SS}
2387:   A.~Salam and J.~A.~Strathdee,
2388:   ``On Kaluza-Klein Theory,''
2389:   Annals Phys.\  {\bf 141} (1982) 316.
2390:   %%CITATION = APNYA,141,316;%%
2391: 
2392: 
2393: \end{thebibliography}
2394: 
2395: 
2396: 
2397: \end{document}
2398: