hep-ph0605264/FD.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%% L a T e X  (no macros) %%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: \documentclass[12pt]{article}
3: 
4: \textwidth 16.25cm
5: \textheight 22.5cm
6: \hoffset -1.5cm
7: \voffset -1cm
8: 
9: \setlength{\parindent}{1cm}
10: \setlength{\parskip}{5pt plus 2pt minus 1pt}
11: \renewcommand{\baselinestretch}{1.2}
12: 
13: \usepackage{cite}
14: \usepackage{axodraw}
15: \usepackage[dvips]{graphicx}
16: \usepackage{epsfig}
17: \usepackage{amssymb}
18: \usepackage{rotating}
19: 
20: \def\theequation{\arabic{section}.\arabic{equation}}
21: \renewcommand{\textfraction}{0}
22: \renewcommand{\topfraction}{1}
23: \renewcommand{\bottomfraction}{1}
24: %\renewcommand{\thefootnote}{\fnsymbol{footnote}}
25: \renewcommand{\thefootnote}{\arabic{footnote}}
26: \renewcommand\figurename{\sc\small Figure}
27: \renewcommand\tablename{\sc\small Table}
28: 
29: 
30: %\def\tablename{\bf Table}
31: %\def\figurename{\bf Figure}
32: 
33: 
34: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
35: \begin{document}
36: \begin{flushright}
37: {\tt CERN-PH-TH/2006-100}\\[-2pt]
38: {\tt MAN/HEP/2006/17}\\[-2pt]
39: {\tt hep-ph/0605264}
40: \end{flushright}
41: \bigskip
42: 
43: \begin{center}
44: {\LARGE {\bf Anatomy of {\boldmath $F_D$}-Term Hybrid Inflation}}\\[1.5cm]
45: {\sc Bj\"orn Garbrecht$^{\, a}$, Constantinos Pallis$^{\, a}$ and
46: Apostolos Pilaftsis$^{\, a,b}$}\\[0.5cm]
47: {\em $^a$School of Physics and Astronomy, University of Manchester,}\\
48: {\em Manchester M13 9PL, United Kingdom}\\[0.3cm]
49: {\em $^b$CERN, Physics Department, Theory Division, CH-1211 Geneva 23,
50: Switzerland}
51: %{\em Email}: {\tt bjorn@hep.man.ac.uk, pallis@hep.man.ac.uk,
52: %pilaftsi@hep.man.ac.uk}
53: \end{center}
54: 
55: \vspace{1.cm} \centerline{\bf ABSTRACT}
56: 
57: \noindent
58: {\small
59: We analyze the cosmological  implications of $F$-term hybrid inflation
60: with   a  subdominant   Fayet--Iliopoulos   $D$-term  whose   presence
61: explicitly breaks a $D$-parity in the inflaton-waterfall sector.  This
62: scenario  of inflation, which  is called  $F_D$-term hybrid  model for
63: brevity,  can  naturally  predict   lepton  number  violation  at  the
64: electroweak  scale, by  tying the  $\mu$-parameter of  the MSSM  to an
65: SO(3)-symmetric Majorana mass $m_N$,  via the vacuum expectation value
66: of the inflaton field. We show how a negative Hubble-induced mass term
67: in a  next-to-minimal extension  of supergravity helps  to accommodate
68: the present CMB data and considerably weaken the strict constraints on
69: the theoretical  parameters, resulting  from cosmic string  effects on
70: the power  spectrum $P_{\cal  R}$.  The usual  gravitino overabundance
71: constraint  may  be significantly  relaxed  in  this  model, once  the
72: enormous  entropy  release from  the  late  decays  of the  ultraheavy
73: waterfall  gauge particles  is properly  considered.  As  the Universe
74: enters  a second  thermalization  phase involving  a  very low  reheat
75: temperature,  which  might  be   as  low  as  about  0.3~TeV,  thermal
76: electroweak-scale  resonant leptogenesis  provides a  viable mechanism
77: for  successful baryogenesis,  while  thermal right-handed  sneutrinos
78: emerge as  new possible  candidates for solving  the cold  dark matter
79: problem.  In~addition, we discuss grand unified theory realizations of
80: $F_D$-term hybrid  inflation devoid  of cosmic strings  and monopoles,
81: based  on  the  complete  breaking  of  an  SU(2)$_X$  subgroup.   The
82: $F_D$-term  hybrid model  offers rich  particle-physics phenomenology,
83: which  could  be  probed  at  high-energy colliders,  as  well  as  in
84: low-energy experiments of lepton flavour or number violation. }
85: 
86: 
87: 
88: \thispagestyle{empty}
89: 
90: 
91: \noindent
92: 
93: \medskip
94: \noindent
95: {\small PACS numbers: 98.80.Cq, 12.60.Jv, 11.30Pb}
96: 
97: \tableofcontents
98: \thispagestyle{empty}
99: 
100: \newpage
101: \pagestyle{plain}
102: \setcounter{page}{1}
103: 
104: 
105: \setcounter{equation}{0}
106: \section{Introduction}
107: 
108: 
109: Standard  big-bang cosmology faces  severe difficulties  in accounting
110: for  the observed  flatness  and  enormity of  the  causal horizon  of
111: today's Universe. It also leaves  unexplained the origin of the nearly
112: scale-invariant cosmic  microwave background (CMB), as was  found by a
113: number        of        observations        over       the        last
114: decade~\cite{COBE,WMAP,MT,WMAP3,Lyman}.   All these  pressing problems
115: can be successfully addressed  within the field-theoretic framework of
116: inflation~\cite{review}.   As a  source  of inflation,  it is  usually
117: considered to be a scalar field, the inflaton, which is displaced from
118: its  minimum and  whose  slow-roll dynamics  leads  to an  accelerated
119: expansion  of  the  early  Universe.   In this  phase  of  accelerated
120: expansion or inflation, the quantum fluctuations of the inflaton field
121: are  stretched on  large scales  and eventually  get frozen  when they
122: become much bigger than the Hubble radius.  These quantum fluctuations
123: get imprinted  in the form  of density perturbations, when  the former
124: are crossing  back inside the  Hubble radius long after  inflation has
125: ended. In this  way, inflation provides a causal  mechanism to explain
126: the observed nearly-scale invariant CMB spectrum.
127: 
128: A complete description  of the CMB spectrum involves  about a dozen of
129: cosmological parameters, such as  the power spectrum $P_{{\cal R}}$ of
130: curvature perturbations,  the spectral index $n_{\rm  s}$, the running
131: spectral index $dn_{\rm s}/d\ln  k$, the ratio $r$ of tensor-to-scalar
132: perturbations,   the  baryon-to-photon   ratio  of   number  densities
133: $\eta_B$, the fractions of  relic abundance $\Omega_{\rm DM}$ and dark
134: energy  $\Omega_{\rm   \Lambda}$  and  a  few   others.   Recent  WMAP
135: data~\cite{WMAP,WMAP3},     along      with     other     astronomical
136: observations~\cite{MT}, have improved upon the precision of almost all
137: of  the above  cosmological  observables. In  particular, the  precise
138: values of these cosmological  observables set stringent constraints on
139: the model-building  of successful models  of inflation. To  ensure the
140: slow-roll  dynamics of  the inflaton,  for example,  one would  need a
141: scalar potential, which  is almost flat.  Moreover, one  has to assure
142: that the  flatness of the inflaton  potential does not  get spoiled by
143: large quantum corrections that  depend quadratically on the cut-off of
144: the theory.   In this context, supersymmetry (SUSY),  softly broken at
145: the TeV scale,  emerges almost as a compelling  ingredient not only in
146: the model-building of inflationary  scenarios, but also for addressing
147: technically the so-called gauge-hierarchy problem.
148: 
149: One  of the  most  predictive and  potentially  testable scenarios  of
150: inflation   is  the  model   of  hybrid   inflation~\cite{Linde}.   An
151: advantageous feature  of this  model is that  the inflaton  $\phi$ may
152: start its  slow-roll from field  values well below the  reduced Planck
153: mass  $m_{\rm  Pl}  =   2.4\times  10^{18}$~GeV.   As  a  consequence,
154: cosmological observables,  such as $P_{{\cal R}}$ and  $n_{\rm s}$, do
155: not  generically  receive   significant  contributions  from  possible
156: higher-dimensional   non-renormalizable   operators,   as  these   are
157: suppressed  by inverse  powers of  $1/m_{\rm Pl}$.   Thus,  the hybrid
158: model becomes very predictive and possibly testable, in the sense that
159: the  inflaton dynamics  is  mainly governed  by  a few  renormalizable
160: operators  which  might have  observable  implications for  laboratory
161: experiments.  In  the hybrid  model, inflation terminates  through the
162: so-called waterfall mechanism.  This  mechanism is triggered, when the
163: inflaton field  $\phi$ passes below  some critical value  $\phi_c$. In
164: this case,  another field $X$  different from $\phi$, which  is called
165: the waterfall field and is  held fixed at origin initially, develops a
166: tachyonic  instability  and rolls  rapidly  down  to  its true  vacuum
167: expectation value~(VEV).
168: 
169: Hybrid  inflation can be  realized in  supersymmetric theories  in two
170: forms.   In the  first form,  the  hybrid potential  results from  the
171: $F$-terms of  a superpotential, where  the slope of the  potential may
172: come either from  supergravity (SUGRA) corrections~\cite{CLLSW} and/or
173: from   radiative   effects~\cite{DSS}.    The  second   supersymmetric
174: realization~\cite{Halyo}   of  hybrid   inflation   uses  a   dominant
175: Fayet--Iliopoulos~(FI) $D$-term~\cite{FI}, which may originate from an
176: anomalous  local  U(1)$_Q$  symmetry  within  the  context  of  string
177: theories.
178: 
179: All models  of inflation embedded in  SUGRA have to  address a serious
180: problem. This  is the  so-called gravitino overabundance  problem.  If
181: abundantly produced in the early Universe, gravitinos may disrupt, via
182: their late gravitationally-mediated decays, the nucleosynthesis of the
183: light elements.   In order to prevent this  from happening, gravitinos
184: $\widetilde{G}$   must   have    a   rather   low   abundance   today,
185: i.e.~$Y_{\widetilde{G}}  = n_{\widetilde{G}}/s\ \stackrel{<}{{}_\sim}\
186: 10^{-12}$--$10^{-15}$, where $n_{\widetilde{G}}$ is the number density
187: of  gravitinos and  $s$ is  the entropy  density. The  upper  bound on
188: $Y_{\widetilde{G}}$  depends on  the properties  of the  gravitino and
189: becomes tighter,  if gravitinos  decay appreciably to  hadronic modes.
190: These  considerations set a  strict upper  bound on  Universe's reheat
191: temperature  $T_{\rm  reh}$, generically  implying  that $T_{\rm  reh}
192: \stackrel{<}{{}_\sim}  10^{10}$--$10^7$~GeV~\cite{kohri,oliveg}.  This
193: upper  limit on  $T_{\rm  reh}$  severely restricts  the  size of  any
194: renormalizable superpotential coupling of the inflaton to particles of
195: the  Standard   Model~(SM).   All  these  couplings   must  be  rather
196: suppressed.    Typically,  they   have  to   be  smaller   than  about
197: $10^{-5}$~\cite{SS}.
198: 
199: The aforementioned  gravitino constraint may  be considerably relaxed,
200: if there is  a mechanism that could cause late  entropy release in the
201: evolution of the  early Universe.  Such a mechanism  could then dilute
202: the gravitinos to a level that would not upset the limits derived from
203: Big  Bang nucleosynthesis  (BBN).  This possibility  might arise  even
204: within the context  of $F$-term hybrid inflation, if  a subdominant FI
205: $D$-tadpole associated with the  gauge group U(1)$_X$ of the waterfall
206: sector  were  added  to  the  model.  Such  a  scenario  was  recently
207: discussed in~\cite{GP}.  It  has been observed that the  presence of a
208: FI $D$-term breaks explicitly an exact discrete symmetry acting on the
209: gauged waterfall  sector, i.e.~a kind of $D$-parity,  which would have
210: remained  otherwise  unbroken  even  after  the  spontaneous  symmetry
211: breaking  (SSB)  of  U(1)$_X$.    As  a  consequence,  the  ultraheavy
212: U(1)$_X$-gauge-sector  bosons  and  fermions,  which would  have  been
213: otherwise stable, can  now decay with rates controlled  by the size of
214: the FI  $D$-term. Since these  particles could be  abundantly produced
215: during the  preheating epoch, their late  decays could give  rise to a
216: second reheat phase in the evolution of the early Universe.  Depending
217: on the actual size of  the FI $D$-term, this second reheat temperature
218: may be as low as 0.5--1~TeV, resulting in an enormous entropy release.
219: This could be sufficient to render the gravitinos underabundant, which
220: might  be  copiously produced  during  the  first  reheating from  the
221: perturbative inflaton decays.
222: 
223: In  this paper  we  present  a detailed  analysis  of $F$-term  hybrid
224: inflation with  a subdominant FI $D$-tadpole. As  mentioned above, the
225: presence of the FI $D$-tadpole is essential for explicitly breaking an
226: exact discrete symmetry, a $D$-parity,  which was acting on the gauged
227: waterfall sector.  In~\cite{GP}, we termed this inflationary scenario,
228: in  short,  $F_D$-term  hybrid  inflation.   As  the  inflaton  chiral
229: superfield   $\widehat{S}$  couples   to   the  Higgs-doublet   chiral
230: superfields        $\widehat{H}_{u,d}$,       through       $\lambda\,
231: \widehat{S}\widehat{H}_u   \widehat{H}_d$,  the  model   generates  an
232: effective  $\mu$-parameter  for  the Minimal  Supersymmetric  Standard
233: Model~(MSSM), through the VEV $\langle S \rangle$~\cite{DLS}. The same
234: mechanism may also generate an  effective Majorana mass matrix for the
235: singlet  neutrino  superfields $\widehat{N}_{1,2,3}$~\cite{Francesca},
236: through   the    operator   $\frac{1}{2}\,   \rho_{ij}\,   \widehat{S}
237: \widehat{N}_i  \widehat{N}_j$~\cite{PU2,GP}.  Assuming that  this last
238: operator  is  SO(3)-symmetric or  very  close  to it,  i.e.~$\rho_{ij}
239: \approx  \rho\,  {\bf  1}_3$,  the  resulting  lepton-number-violating
240: Majorana mass, $m_N = \rho\,  \langle S \rangle$, will be closely tied
241: to  the $\mu$-parameter  of the  MSSM.   If $\lambda  \sim \rho$,  the
242: $F_D$-term hybrid  model will  then give rise  to 3  nearly degenerate
243: heavy  Majorana neutrinos $\nu_{1,2,3\,R}$,  as well  as to  3 complex
244: right-handed       sneutrinos       $\widetilde{N}_{1,2,3}$,      with
245: electroweak-scale  masses.    Such  a  mass  spectrum   opens  up  the
246: possibility  to explain  the baryon  asymmetry in  the  Universe (BAU)
247: $\eta_B$~\cite{FY,BAUpapers}  by  thermal  electroweak-scale  resonant
248: leptogenesis~\cite{APRD,APRL,PU2}, almost independently of the initial
249: baryon-number composition  of the primordial  plasma.  Moreover, since
250: the  $F_D$-term  hybrid   model  conserves  $R$-parity~\cite{GP},  the
251: lightest supersymmetric  particle (LSP)  is stable.  Here,  we examine
252: the  possibility  that   {\em  thermal}  right-handed  sneutrinos  are
253: responsible  for solving  the cold  dark matter  (CDM) problem  of the
254: Universe.
255: 
256: In  this   paper  we  also  improve   an  earlier  approach~\cite{GP},
257: concerning  the production of  the quasi-stable  U(1)$_X$ gauge-sector
258: particles  during the  preheating epoch.   In addition,  we  present a
259: numerical  analysis  that properly  takes  into  account the  combined
260: effect on the  reheat temperature $T_{\rm reh}$ from  the inflaton and
261: gauge-sector particle  decays and  their annihilations.  We  call this
262: two-states'  mechanism   of  reheating  the   Universe,  {\em  coupled
263: reheating}.   After   solving  numerically  a   network  of  Boltzmann
264: equations~(BEs) that appropriately  treat coupled reheating, we obtain
265: estimates for the present abundance of gravitinos in the Universe.  We
266: show  explicitly, how  a small  breaking  of $D$-parity  sourced by  a
267: subdominant  FI  $D$-tadpole  helps  to  relax  the  strict  gravitino
268: overproduction constraint.
269: 
270: In addition to gravitinos, one  might have to worry that topologically
271: stable cosmic strings do not contribute significantly to the CMB power
272: spectrum~$P_{\cal   R}$.   Cosmic  strings,   global  or   local,  are
273: topological defects and  usually form after the SSB  of some global or
274: local  U(1)  symmetry~\cite{NielsenOlesen,Vilenkin,HK}.  According  to
275: recent analyses~\cite{strings}, cosmic strings, if any, should make up
276: no more than about 10\% of the power spectrum $P_{\cal R}$.  This last
277: requirement  puts severe  limits  on the  allowed  parameter space  of
278: models    of    inflation.     There    have   already    been    some
279: suggestions~\cite{JKLS} on how to get  rid of cosmic strings, based on
280: modified versions  of hybrid inflation.   Here, we follow  a different
281: approach to solving  this problem.  We consider models,  for which the
282: waterfall sector  possesses an  SU(2)$_X$ gauge symmetry  which breaks
283: completely,  i.e.~SU(2)$_X  \to {\bf  I}$,  such  that neither  cosmic
284: strings nor monopoles  are produced at the end  of inflation.  In this
285: case,  gauge   invariance  forbids  the  existence   of  an  SU(2)$_X$
286: $D$-tadpole $D^a$.  However, Planck-mass suppressed non-renormalizable
287: operators  that  originate from  the  superpotential  or the  K\"ahler
288: potential  can give  rise  to explicit  breaking  of $D$-parity.   The
289: latter   may  manifest   itself   by  the   generation  of   effective
290: $D^a$-tadpole terms  that arise after  the SSB of~SU(2)$_X$.   In this
291: way, all the SU(2)$_X$ gauge-sector particles can be made unstable.
292: 
293: The organization of the paper is as follows: in Section~\ref{FDmodel},
294: we  describe the  $F_D$-term  hybrid model  and  calculate the  1-loop
295: effective potential  relevant to  inflation.  In addition,  we discuss
296: the  possible cosmological  consequences of  radiative effects  on the
297: flat directions  in the MSSM.   We conclude this section  by outlining
298: how the  $F_D$-term hybrid  model could generally  be embedded  into a
299: grand unified  theory~(GUT), including possible realizations  of a GUT
300: without  cosmic strings and  monopoles.  Technical  details concerning
301: mechanisms  of  explicit $D$-parity  breaking  in  SUGRA, e.g.~via  an
302: effective subdominant  $D$-tadpole or non-renormalizable  operators in
303: K\"ahler    potential,   are   given    in   Appendix~\ref{Dappendix}.
304: Section~\ref{inflation}  analyzes the  constraints on  the theoretical
305: parameters, which are mainly  derived from considerations of the power
306: spectrum $P_{\cal R}$ and a strongly red-tilted spectral index $n_{\rm
307: s}$,  with $n_{\rm  s} \approx  0.95$, as  observed most  recently by
308: WMAP~\cite{WMAP3,Lyman}.  We  show how a  negative Hubble-induced mass
309: term in  a next-to-minimal extension of supergravity  helps to account
310: for  the present  CMB data,  as well  as to  substantially  weaken the
311: strict constraints  on the  model parameters, originating  from cosmic
312: string effects on  $P_{\cal R}$, within a U(1)$_X$  realization of the
313: $F_D$-term hybrid model.
314: 
315: In  Section~\ref{Preheat},  we  analyze   the  mass  spectrum  of  the
316: inflaton-waterfall  sector in  the post-inflationary  era  and present
317: naive estimates  of the reheat  temperature $T_{\rm reh}$  as obtained
318: from perturbative  inflaton decays.  We  then make use of  an improved
319: approach  to  preheating  and   compute  the  energy  density  of  the
320: quasi-stable waterfall  gauge particles.  In  Section~\ref{reheat}, we
321: solve numerically  the BEs relevant  to coupled reheating  and present
322: estimates  for the gravitino  abundance in  the present  Universe.  In
323: Section~\ref{BAU},  we   demonstrate,  how  thermal  electroweak-scale
324: resonant  leptogenesis can  be realized  within the  $F_D$-term hybrid
325: model  and discuss  the possibility  of  solving the  CDM problem,  if
326: thermal right-handed sneutrinos  are considered to be the  LSPs in the
327: spectrum.  In  Section~\ref{conclusions}, we present  our conclusions,
328: including a  summary of possible particle-physics  implications of the
329: $F_D$-term hybrid  model for high-energy colliders  and for low-energy
330: experiments of lepton flavour and/or number violation.
331: 
332: 
333: \setcounter{equation}{0}
334: \section{General Setup}\label{FDmodel}
335: 
336: In this section  we first present the general  setup of the $F_D$-term
337: hybrid  model  within the  minimal  SUGRA  framework  and compute  the
338: renormalized  1-loop effective  potential relevant  to  inflation.  We
339: then discuss the cosmological implications of radiative effects on the
340: MSSM  flat directions  for $F_D$-term  hybrid inflation  and  for SUSY
341: inflationary models  in general. Finally, we analyze  the prospects of
342: embedding the $F_D$-term hybrid model into a GUT.
343: 
344: 
345: 
346: \subsection{The Model}\label{FD}
347: 
348: 
349: The renormalizable  superpotential of  the $F_D$-term hybrid  model is
350: given by
351: \begin{eqnarray}
352:   \label{Wmodel}
353:  W & =& \kappa\, \widehat{S}\, \Big( \widehat{X}_1
354: \widehat{X}_2\:  -\: M^2\Big)\ +\ \lambda\, \widehat{S} \widehat{H}_u
355: \widehat{H}_d\ +\ \frac{\rho_{ij}}{2}\, \widehat{S}\, \widehat{N}_i
356: \widehat{N}_j\ +\ h^{\nu}_{ij} \widehat{L}_i \widehat{H}_u
357: \widehat{N}_j\nonumber\\ &&+\ W_{\rm MSSM}^{(\mu = 0)}\; ,
358: \end{eqnarray}
359: where $W_{\rm  MSSM}^{(\mu =  0)}$ denotes the  MSSM superpotential
360: without the $\mu$-term:
361: \begin{equation} W_{\rm MSSM}^{(\mu = 0)}\ =\
362:   h^u_{ij}\,\widehat{Q}_i\widehat{H}_u\widehat{U}_j\: +\:
363: h^d_{ij}\,\widehat{H}_d\widehat{Q}_i\widehat{D}_j\: +\:
364:   h_l\, \widehat{H}_d\widehat{L}_l\widehat{E}_l \; .
365: \end{equation}
366: The first term in~(\ref{Wmodel}) describes the inflaton-waterfall (IW)
367: sector.   Specifically,  $\widehat{S}$   is  the  SM-singlet  inflaton
368: superfield, and $\widehat{X}_{1,2}$ is  a chiral multiplet pair of the
369: waterfall fields with opposite charges under the U(1)$_X$ gauge group,
370: i.e.~$Q (\widehat{X}_1) = - Q  (\widehat{X}_2) = 1$.  In addition, the
371: corresponding   inflationary   soft   SUSY-breaking  sector   obtained
372: from~(\ref{Wmodel}) reads:
373: \begin{equation}
374:   \label{Lsoft}
375: -\, {\cal L}_{\rm soft}\ =\ M^2_S S^*S\: +\: \Big(
376: \kappa A_\kappa\, S X_1X_2\: +\: \lambda A_\lambda S H_u H_d\: \: +\:
377: \frac{\rho}{2}\, A_\rho\, S \widetilde{N}_i\widetilde{N}_i\:
378: -\: \kappa a_S M^2 S \: \ +\ {\rm  H.c.}\,\Big)\,,
379: \end{equation}
380: where   $M_S$,   $A_{\kappa,\lambda,\rho}$    and   $a_S$   are   soft
381: SUSY-breaking mass parameters of order $M_{\rm SUSY} \sim 1$~TeV.
382: 
383: The    second   term    in~(\ref{Wmodel}),    $\lambda\,   \widehat{S}
384: \widehat{H}_u  \widehat{H}_d$, induces  an  effective $\mu$-parameter,
385: when the scalar component of $\widehat{S}$, $S$, acquires a VEV, i.e.
386: \begin{equation}
387:   \label{mu}
388: \mu\ =\  \lambda\, \langle S  \rangle\ \approx\ \frac{\lambda}{2\kappa}\,
389: |A_\kappa - a_S|\ .
390: \end{equation}
391: In obtaining the last approximate equality in~(\ref{mu}), we neglected
392: the VEVs  of $H_{u,d}$ and  considered the fact  that the VEVs  of the
393: waterfall fields $X_{1,2}$ after inflation are: $\langle X_{1,2}\rangle
394: =   M$~\cite{DLS}.    For  $\lambda   \sim   \kappa$,   the  size   of
395: $\mu$-parameter turns out to be of the order of the soft-SUSY breaking
396: scale $M_{\rm  SUSY}$, as required for a  successful electroweak Higgs
397: mechanism.    By   analogy,    the   third   term   in~(\ref{Wmodel}),
398: $\frac{1}{2}\,\rho_{ij}\,  \widehat{S}\, \widehat{N}_i \widehat{N}_j$,
399: gives  rise  to  an  effective lepton-number-violating  Majorana  mass
400: matrix, i.e.~$M_S  = \rho_{ij}\,  v_S$.  Assuming that  $\rho_{ij}$ is
401: approximately  SO(3) symmetric,  viz.~$\rho_{ij}  \approx \rho\;  {\bf
402: 1}_3$,  one   obtains  3  nearly   degenerate  right-handed  neutrinos
403: $\nu_{1,2,3\,R}$, with mass
404: \begin{equation}
405:   \label{mN}
406: m_N\ =\ \rho\, v_S\ .
407: \end{equation}
408: If  $\lambda$  and  $\rho$  are  comparable  in  magnitude,  then  the
409: $\mu$-parameter and  the SO(3)-symmetric Majorana mass  $m_N$ are tied
410: together, i.e.~$m_N  \sim \mu$, thus  leading to a scenario  where the
411: singlet   neutrinos  $\nu_{1,2,3\,R}$  can   naturally  have   TeV  or
412: electroweak-scale masses~\cite{PU2,GP}.
413: 
414: The renormalizable  superpotential~(\ref{Wmodel}) of the  model may be
415: uniquely determined by imposing the continuous $R$ symmetry:
416: \begin{equation}
417:   \label{RFD}
418: \widehat{S}\  \to\ e^{i\alpha}\,\widehat{S}\, ,\qquad
419: \widehat{L}\  \to\  e^{i\alpha}\, \widehat{L}\,, \qquad
420: \widehat{Q}\ \to\ e^{i\alpha}\, \widehat{Q}\; ,
421: \end{equation}
422: with $W \to e^{i\alpha} W$,  whereas all other fields remain invariant
423: under an $R$ transformation.  Notice that the $R$ symmetry~(\ref{RFD})
424: forbids  the  presence of  higher-dimensional  operators  of the  form
425: $\widehat{X}_1 \widehat{X}_2  \widehat{N}_i \widehat{N}_j/m_{\rm Pl}$.
426: This fact ensures that  the electroweak-scale Majorana mass $m_N$ does
427: not get destabilized by Planck-scale SUGRA effects.
428: 
429: One  may  now  observe   that  the  superpotential  (\ref{Wmodel})  is
430: symmetric   under   the   permutation   of   the   waterfall   fields,
431: i.e.~$\widehat{X}_1 \leftrightarrow  \widehat{X}_2$.  This permutation
432: symmetry persists,  even after the  SSB of U(1)$_X$, since  the ground
433: state, $\langle X_1  \rangle = \langle X_2 \rangle  = M$, is invariant
434: under the  same symmetry  as well. Hence,  there is an  exact discrete
435: symmetry acting on the gauged  waterfall sector, a kind of $D$-parity.
436: As a consequence of  $D$-parity conservation, the ultraheavy particles
437: of mass  $g M$, which  are related to  the U(1)$_X$ gauge  sector, are
438: stable.  Such a possibility is not very desirable, as these particles,
439: if abundantly produced, may overclose  the Universe at late times.  In
440: order to  break this unwanted  $D$-parity, a subdominant  FI $D$-term,
441: $-\frac{1}{2} g\, m^2_{\rm FI}\,  D$, is added to the model~\cite{GP},
442: giving rise  to the $D$-term potential~\footnote{The  $D$-parity is an
443: accidental discrete  symmetry and it  should not be confused  with the
444: ${\rm   U(1)}_X$   charge  conjugation   symmetry   realized  by   the
445: transformations: $X_1 \leftrightarrow  X_1^*$ and $X_2 \leftrightarrow
446: X_2^*$. Although  both discrete symmetries  have the same  effect when
447: acting on the  ${\rm U(1)}_X$ scalar current $j_X^\mu={\rm  i} ( X_1^*
448: \stackrel{\leftrightarrow}{\partial^\mu}       X_1       -       X^*_2
449: \stackrel{\leftrightarrow}{\partial^\mu}      X_2)$,     i.e.~$j_X^\mu
450: \leftrightarrow  -  j_X^\mu$,  they  crucially differ  when  they  are
451: applied on  the FI $D$-term:  $-\,\frac{g}{2} m^2_{\rm FI}  (|X_1|^2 -
452: |X_2|^2)$.  This term is even  under charge conjugation, but odd under
453: a $D$-parity conjugation.}
454: \begin{equation}
455:   \label{Dterm}
456: V_D\ =\ \frac{g^2}{8}\ \Big( |X_1|^2\, -\, |X_2|^2\, -\, m^2_{\rm
457:   FI}\,\Big)^2\; .
458: \end{equation}
459: The FI $D$-term will not  affect the inflationary dynamics, as long as
460: $g m_{\rm FI}  \ll \kappa M$. Technically, a  subdominant $D$-term can
461: be  generated  radiatively after  integrating  out Planck-scale  heavy
462: degrees    of    freedom.    Further    discussion    is   given    in
463: Section~\ref{postinfl} and in  Appendix~\ref{Dappendix}, where we also
464: discuss   the  possibility  of   breaking  explicitly   $D$-parity  by
465: non-renormalizable  K\"ahler potential  terms.   The post-inflationary
466: implications  of  the  FI  $D$-term,  $m_{\rm  FI}$,  for  the  reheat
467: temperature    $T_{\rm    reh}$    and   the    gravitino    abundance
468: $Y_{\widetilde{G}}$ will be analyzed in Section~\ref{reheat}.
469: 
470: The inflationary potential $V_{\rm inf}$ may be represented by the sum
471: \begin{equation}
472:   \label{Vinf}
473: V_{\rm inf}\ =\ V^{(0)}_{\inf}\: +\:  V^{(1)}_{\inf}\: +\: V_{\rm SUGRA}\ ,
474: \end{equation}
475: where  $V^{(0)}_{\inf}$   and  $V^{(1)}_{\inf}$  are   the  tree-level
476: potential and the 1-loop effective potential, respectively and $V_{\rm
477: SUGRA}$ contains the  SUGRA contribution. Including soft-SUSY breaking
478: terms related to $S$,  the tree-level contribution to the inflationary
479: potential is
480: \begin{equation}
481:   \label{V0inf}
482: V^{(0)}_{\inf}\ =\ {\cal Z}_S\, \kappa^2 M^4\: +\:
483: M^2_S\, S^* S\: -\: \Big( \kappa a_S M^2 S\: +\: {\rm H.c.}\Big)\ ,
484: \end{equation}
485: where ${\cal  Z}^{1/2}_S$ is the wave-function  renormalization of the
486: inflaton  field which is  needed to  renormalize the  1-loop effective
487: potential  given  below  in  the   SUSY  limit  of  the  theory.   The
488: counter-term,  $\delta  {\cal Z}_S  =  {\cal Z}_S  -  1$,  due to  $S$
489: wave-function  renormalization  may  be  obtained  from  the  inflaton
490: self-energy $\Pi_{SS}(p^2)$, through the relation
491: \begin{equation}
492: \delta  {\cal Z}_S\ =\ -\, \frac{d\,{\rm Re}\,
493:   \Pi_{SS}(p^2)}{dp^2}\Bigg|_{p^2 = 0}\ .
494: \end{equation}
495: Calculating the  UV part  of $\delta {\cal  Z}_S$ from this  very last
496: relation, we find
497: \begin{equation}
498:   \label{dZs}
499: \delta  {\cal   Z}_S\ =\ -\, \frac{1}{32 \pi^2}\,
500: \Bigg[\, 2{\cal N}\kappa^2\, \ln\Bigg(\frac{\kappa^2 M^2}{Q^2}\Bigg)\: +\:
501: 4\lambda^2\, \ln\Bigg(\frac{\lambda^2 M^2}{Q^2}\Bigg)\: +\:
502: 3\rho^2\,\ln\Bigg(\frac{\rho^2 M^2}{Q^2}\Bigg)\,\Bigg]\; ,
503: \end{equation}
504: where $Q^2$ is the renormalization  scale and the inflaton field value
505: $|S_R|  = M$  is taken  as a  common mass  renormalization  point.  In
506: addition,  the  parameter  ${\cal  N}$ in~(\ref{dZs})  represents  the
507: dimensionality of the waterfall sector.   For example, it is ${\cal N}
508: = 1$ for an U(1)$_X$ waterfall sector, whilst it is ${\cal N} = N$, if
509: $\widehat{X}_1$   ($\widehat{X}_2$)   belongs   to   the   fundamental
510: (anti-fundamental)  representation  of  an SU($N$)  theory.   Observe,
511: finally,  that  only  the  fermionic components  of  the  superfields,
512: $\widehat{X}_{1,2}$,    $\widehat{H}_{u,d}$,    $\widehat{N}_{1,2,3}$,
513: contribute to $\delta {\cal Z}_S$.
514: 
515: Ignoring  soft  SUSY-breaking terms,  the  1-loop effective  potential
516: relevant to inflation is calculated to be
517: \begin{eqnarray}
518:   \label{V1loop}
519: V^{(1)}_{\rm inf} \!\!&=&\!\! \frac{1}{32\pi^2}\, \Bigg\{
520: {\cal N}\kappa^4\, \Bigg[ |S^2 + M^2|^2\,
521: \ln\Bigg(\frac{\kappa^2 (|S|^2 +M^2)}{Q^2}\Bigg) +\,
522: |S^2 - M^2|^2\,
523: \ln\Bigg(\frac{\kappa^2 (|S|^2 - M^2)}{Q^2}\Bigg)\Bigg]\nonumber\\
524: &&\hspace{-0.17cm} +\, 2\lambda^4\, \Bigg[
525: |S^2 + {\textstyle \frac{\kappa}{\lambda}}\, M^2|^2\,
526: \ln\Bigg(\frac{\lambda^2 (|S|^2 + {\textstyle \frac{\kappa}{\lambda}}
527: M^2)}{Q^2}\Bigg)\, +\,
528: |S^2 - {\textstyle \frac{\kappa}{\lambda}} M^2|^2\,
529: \ln\Bigg(\frac{\lambda^2 (|S|^2 - {\textstyle \frac{\kappa}{\lambda}}
530: M^2)}{Q^2}\Bigg)\Bigg]\nonumber\\
531: &&\hspace{-0.17cm} +\, \frac{3\rho^4}{2}\, \Bigg[
532: |S^2 + {\textstyle \frac{\kappa}{\rho}}\, M^2|^2\,
533: \ln\Bigg(\frac{\rho^2 (|S|^2 + {\textstyle \frac{\kappa}{\rho}}
534: M^2)}{Q^2}\Bigg)\, +\,
535: |S^2 - {\textstyle \frac{\kappa}{\rho}} M^2|^2\,
536: \ln\Bigg(\frac{\rho^2 (|S|^2 - {\textstyle \frac{\kappa}{\rho}}
537: M^2)}{Q^2}\Bigg)\Bigg]\nonumber\\
538: &&\hspace{-0.17cm} -\, |S|^4\, \Bigg[\, 2{\cal N}\kappa^4\,
539: \ln\Bigg(\frac{\kappa^2\,|S|^2}{Q^2}\Bigg)\: +\: 4\lambda^4\,
540: \ln\Bigg(\frac{\lambda^2\,|S|^2}{Q^2}\Bigg)
541: \: +\:
542: 3\rho^4\,\ln\Bigg(\frac{\rho^2\, |S|^2}{Q^2}\Bigg)\,\Bigg]\,\Bigg\}\; .
543: \end{eqnarray}
544: Given~(\ref{dZs})  and~(\ref{V1loop}),  it  can  be checked  that  the
545: expression $V^{(0)}_{\rm  inf} + V^{(1)}_{\rm inf}$  is independent of
546: $\ln Q^2$, as it should be.
547: 
548: Finally,  the  SUGRA contribution  $V_{\rm  SUGRA}$  to $V_{\rm  inf}$
549: in~(\ref{Vinf}) is highly model-dependent.  In general, one expects an
550: infinite series  of non-renormalizable  operators to occur  in $V_{\rm
551: SUGRA}$, i.e.~\cite{CLLSW,CP,LR}
552: \begin{eqnarray}
553:   \label{VSUGRA}
554: V_{\rm SUGRA}\ =\ -\, c^2_H\, H^2\, |S|^2\: +\:
555: \kappa^2 M^4\, \frac{|S|^4}{2\,m^4_{\rm Pl}}\: +\: {\cal O}(|S|^6)\ .
556: \end{eqnarray}
557: where $H^2 = \kappa^2 M^4/(3 m^2_{\rm Pl})$ is the squared Hubble rate
558: during  inflation.   The  first  term in~(\ref{VSUGRA})  represents  a
559: Hubble-induced mass  term, which is preferably defined  to be negative
560: for observational reasons  to be discussed in Section~\ref{inflation}.
561: In  a model  with a  minimal K\"ahler  potential, the  parameter $c_H$
562: vanishes  identically.\footnote{Strictly  speaking, curvature  effects
563: related to  an expanding de  Sitter background will contribute  to the
564: potential a term given by $-\frac{3}{16\pi^2}\, (2{\cal N} \kappa^2 + 4
565: \lambda^2 + 3 \rho^2)\, H^2 |S|^2 \ln(|S|^2/Q^2)$, even in the minimal
566: K\"ahler potential case~\cite{BG}.  Such a term, however, turns out to
567: be  negligible to  affect  the inflation  dynamics  in the  $F_D$-term
568: hybrid model.  Finally,  this term may be partially  absorbed into the
569: RG running of $c_H^2 (Q^2)$.} In fact, if $|c_H| \stackrel{<}{{}_\sim}
570: 10^{-2}$, its  influence on the CMB  data~\cite{JP} gets marginalized.
571: In our analysis in Section~\ref{inflation}, we present results for two
572: representative  models:  (i)  the  scenario with  a  minimal  K\"ahler
573: potential  ($c_H  = 0$);  (ii)  a  next-to-minimal K\"ahler  potential
574: scenario with  $c_H \stackrel{<}{{}_\sim} 0.2$, where  only the effect
575: of  the  term  $(\widehat{S}^\dagger \widehat{S})^2/m^2_{\rm  Pl}$  is
576: considered  and  all  higher  order non-renormalizable  operators  are
577: ignored  in  the K\"ahler  manifold.   Moreover,  we neglect  possible
578: 1-loop       contributions      to       $V_{\rm       inf}$      from
579: $A_{\kappa,\lambda,\rho}$-terms, which are insignificant for values $M
580: \stackrel{>}{{}_\sim} 10^{15}$~GeV.  We  only include the tadpole term
581: $\kappa a_S M^2\, S$, which  may become relevant for values of $\kappa
582: \stackrel{<}{{}_\sim}   10^{-4}$,   but    ignore   all   other   soft
583: SUSY-breaking    terms,    since    they   are    negligible    during
584: inflation~\cite{SS}.
585: 
586: The stability  of the inflationary  trajectory in the presence  of the
587: Higgs  doublets  $H_{u,d}$   and  the  right-handed  scalar  neutrinos
588: $\widetilde{N}_{1,2,3}$ provides further restrictions on the couplings
589: $\lambda$  and  $\rho$.   In  order  to  successfully  trigger  hybrid
590: inflation,  the fields  at  the  start of  inflation  should obey  the
591: following conditions:
592: \begin{equation}
593:   \label{initial}
594: {\rm Re}\, S^{\rm in}\ =\ |S^{\rm in}|\ \stackrel{>}{{}_\sim}\ M\,,\qquad
595: X^{\rm in}_{1,2}\ =\ 0\,,\qquad
596: H^{\rm in}_{u,d}\ =\ 0\,,\qquad
597: \widetilde{N}^{\rm in}_{1,2,3}\ =\ 0\; .
598: \end{equation}
599: The precise start  values of the inflaton ${\rm  Re}\, S^{\rm in}$ are
600: determined by the number of $e$-folds ${\cal N}_e$, which is a measure
601: of Universe's  expansion during inflation (see also  our discussion in
602: Section~\ref{inflation}).    After   inflation   and   the   waterfall
603: transition mechanism  have been completed,  it is important  to ensure
604: that  the   waterfall  fields  acquire  a  high   VEV,  i.e.   $X^{\rm
605: end}_{1,2}\ =\ M$, while all other fields have small electroweak-scale
606: VEVs.  This  can be achieved  by requiring that the  Higgs-doublet and
607: the  sneutrino mass  matrices  stay positive  definite throughout  the
608: inflationary  trajectory up  to a  critical value  $|S_c|  \approx M$.
609: Instead, the corresponding mass matrix  of $X_{1,2}$ will be the first
610: to develop  a negative eigenvalue  and tachyonic instability  close to
611: $|S_c|$.  As a consequence, the  fields $X_{1,2}$ will be the first to
612: start  moving away  from 0  and set  in to  the `good'  vacuum $X^{\rm
613: end}_1\  =\  X^{\rm end}_2\  =\  M$,  well  before the  other  fields,
614: e.g.~$H^{\rm in}_{1,2}$ and  $\widetilde{N}^{\rm in}_{1,2,3}$, go to a
615: `bad' vacuum  where $X^{\rm end}_{1,2}\ =\ 0$,  $H^{\rm end}_{1,2}\ =\
616: \frac{\kappa}{\lambda}\,  M$  and  $\widetilde{N}^{\rm  in}_{1,2,3}  =
617: \frac{\kappa}{\rho}\,  M$.  To  better understand  this point,  let us
618: write down the mass matrix in the weak field basis $(H_d\,,\ H_u^* )$:
619: \begin{equation}
620:   \label{Mdoublet}
621: M^2_{\rm Higgs}\ =\ \left(\! \begin{array}{cc}
622: \lambda^2 |S|^2 & -\,\kappa \lambda (M^2 - X_1 X_2 ) \\
623: -\,\kappa\lambda (M^2- X^*_1 X^*_2) & \lambda^2 |S|^2 \end{array}\!\right)\ .
624: \end{equation}
625: Then,  positive definiteness  of $M^2_{\rm  Higgs}$ implies that
626: \begin{equation}
627:   \label{Scondition}
628: \lambda\, |S|^2\ \ge\ \kappa \, |M^2 - X_1 X_2 |\ .
629: \end{equation}
630: From~(\ref{Scondition}),  it is  evident that  the  condition $\lambda
631: \stackrel{>}{{}_\sim}   \kappa$  is   sufficient  for   ending  hybrid
632: inflation  to the  `good'  vacuum. Finally,  one  obtains a  condition
633: analogous to~(\ref{Scondition}) from  the sneutrino mass matrix, which
634: is  equivalent  to having  $\rho  \stackrel{>}{{}_\sim} \kappa$.   The
635: above two constraints on $\lambda$ and $\rho$, i.e.~$\lambda,\ \rho\ >
636: \kappa$,   will   be   imposed    in   the   analysis   presented   in
637: Section~\ref{inflation}.
638: 
639: 
640: 
641: \subsection{Radiative Lifting of MSSM Flat Directions}\label{RadLift}
642: 
643: Flat    directions   in    supersymmetric   theories,    e.g.~in   the
644: MSSM~\cite{GKM}, play an important role in cosmology~\cite{AD,DK}.  As
645: we  will demonstrate  in  this section,  however,  their influence  on
646: $F_D$-term  hybrid   inflation  is  minimal   under  rather  realistic
647: assumptions.
648: 
649: One possible consequence of flat directions could be the generation of
650: a   primordial  baryon   asymmetry  $\eta_B^{\rm   in}$   through  the
651: Affleck--Dine  mechanism~\cite{AD}.  However,  if this  initial baryon
652: asymmetry $\eta_B^{\rm in}$ is generated at temperatures $T > m_N$, it
653: will rapidly  be erased  by the strong  $(B-L)$-violating interactions
654: mediated  by electroweak-scale  heavy  Majorana neutrinos  at $T  \sim
655: m_N$. The BAU  will then reach the present observed  value by means of
656: the thermal  resonant leptogenesis mechanism  and will only  depend on
657: the   basic   theoretical   parameters   of  the   $F_D$-term   hybrid
658: model~\cite{APRL,PU2}.  More details are given in Section~\ref{BAU}.
659: 
660: In  addition,  one  might   argue  that  large  VEVs  associated  with
661: quasi-flat directions  in the  MSSM would make  all MSSM  particles so
662: heavy  after  inflation, such  that  all  perturbative  decays of  the
663: inflaton would  be kinematically blocked and hence  the Universe would
664: never  thermalize~\cite{AM}.  The  system  may fall  into a  false
665: vacuum with  a large VEV at  the start of inflation,  which could, for
666: example, be  triggered by a negative Hubble-induced  squared mass term
667: of order $H^2$~\cite{DRT}, along the flat direction. In the $F_D$-term
668: hybrid  model, however, spontaneous  SUSY breaking  due to  a non-zero
669: $\langle S  \rangle$ is communicated  radiatively to the  MSSM sector,
670: via  the renormalizable  operators $\lambda  \widehat{S} \widehat{H}_u
671: \widehat{H}_d$  and  $\rho  \widehat{S} \widehat{N}_i  \widehat{N}_i$.
672: Consequently, their effects  on the MSSM flat directions  can be large
673: and  so  affect  the  inflaton  decays  which  proceed  via  the  same
674: renormalizable operators. In the  following, we will present a careful
675: treatment  of this  radiative  lifting of  MSSM  flat directions,  and
676: examine  the  conditions,  under  which the  directions  would  remain
677: sufficiently  flat  so  as  to  prohibit  the  Universe  from  thermal
678: equilibration, shortly after inflation.
679: 
680: To  obtain a  flat direction  in supersymmetric  theories, one  has to
681: impose the conditions of $D$- and $F$-flatness on the scalar potential
682: $V$, namely  the vanishing  of all $F$-  and $D$-terms for  a specific
683: field configuration $\sigma$.  $D$-flatness  is automatic for any flat
684: direction associated with a  gauge-invariant operator, which is absent
685: in the MSSM,  e.g.~$\widehat{D}_i \widehat{D}_j \widehat{U}_k$.  Based
686: on   this   observation,   let   us  therefore   consider   here   the
687: gauge-covariant field configuration
688: \begin{equation}
689: \sigma\ =\ \frac{1}{\sqrt{3}}\,
690: \left(\,\
691: \frac{\tilde{u}^*_{R\,k}}{|\tilde{u}_{R\,k}|}\;\tilde{d}^*_{R\,i}
692: \: +\:
693: \frac{\tilde{u}^*_{R\,k}}{|\tilde{u}_{R\,k}|}\;\tilde{d}^*_{R\,j}
694: \: +\:
695: \tilde{u}_{R\,k}\, \right)\; ,
696: \end{equation}
697: where $i\not=j$.   It can be straightforwardly checked  that the field
698: configuration $\sigma$, with the constraint
699: \begin{equation}
700: \frac{\tilde{u}^*_{R\,k}}{|\tilde{u}_{R\,k}|}\;\tilde{d}^*_{R\,i}  \ =
701: \  \frac{\tilde{u}^*_{R\,k}}{|\tilde{u}_{R\,k}|}\;\tilde{d}^*_{R\,j} \
702: = \ \tilde{u}_{R\,k}\ \neq \ 0
703: \end{equation}
704: and all remaining fields being set  to zero, is a flat direction, with
705: vanishing  $F$- and $D$-terms.   It is  then easy  to verify  that the
706: scalar potential $V (\sigma )$ is truly flat, i.e.~$d V/d\sigma = 0$.
707: Although we  will consider  here the case  of $\sigma =  \tilde u_{R\,
708: k}$, the  discussion of  other squark and  slepton flat  directions is
709: completely  analogous.   For   notational  convenience,  we  drop  all
710: generation  indices from  the fields,  and denote  the  flat direction
711: simply by $\tilde u_R$.
712: 
713: 
714: Because of the  spontaneous SUSY breaking induced by  the non-zero VEV
715: of $S$, the flatness of the potential along the $\tilde u_R$-direction
716: gets lifted,  once radiative corrections  are taken into  account. The
717: non-renormalization  theorem  related to  theories  of  SUSY is  still
718: applicable and entails that this radiative lifting should be UV finite
719: and therefore calculable.  We start our calculation by considering the
720: pertinent  mass spectrum  in  the  background of  a  non-zero $S$  and
721: $\tilde{u}_R$.   The fermionic  sector  consists of  2 Dirac  higgsino
722: doublets, with  squared masses $m^2_{\tilde h}= \lambda^2  |S|^2 + h^2
723: |\tilde{u}_R|^2$, while the mass spectrum of the bosonic sector may be
724: deduced by the mass matrix
725: \begin{equation}
726:   \label{Mbosonic}
727: {\cal M}^2_H\ =\ \left(
728: \begin{array}{ccc}
729: \lambda^2 |S|^2 & - \kappa\lambda M^2 & h \lambda S \tilde u_R^* \\
730: - \kappa \lambda M^2 & \lambda^2 |S|^2 + h^2 |\tilde u_R|^2 & 0\\
731: h\lambda S \tilde u_R & 0 & h^2 |\tilde u_R|^2
732: \end{array} \right)\; ,
733: \end{equation}
734: which   is   defined   in   the   weak   basis   $(H_d\,,\   H_u^*\,,\
735: \widetilde{Q})$.   The coupling  $h$ in~(\ref{Mbosonic})  represents a
736: generic up-type quark Yukawa coupling.
737: 
738: In the  renormalization scheme  of dimensional reduction  with minimal
739: subtraction~$\overline{\rm  DR}$~\cite{Jones},  the  1-loop  effective
740: potential $V^{(1)}$ related to $\tilde{u}_R$ is given by
741: \begin{equation}
742:   \label{VuR}
743: V^{(1)} (\tilde{u}_R)\ =\ \frac{2\,Q^2}{16\pi^2}\; {\rm STr}\,{\cal M}^2\
744: +\ \frac{2}{32\pi^2}\; {\rm STr}\, \Bigg\{ {\cal M}^4\, \Bigg[\,\ln\Bigg(
745: \frac{ {\cal M}^2}{Q^2}\Bigg)\ -\ \frac{3}{2}\, \Bigg]\,\Bigg\}\ ,
746: \end{equation}
747: where ${\rm STr}$ denotes  the usual supertrace, e.g.~${\rm STr} {\cal
748: M}^2 = {\rm  Tr}\, {\cal M}^2_H - 2 m^2_{\tilde  h}$, ${\rm STr} {\cal
749: M}^4  = {\rm  Tr}\, {\cal  M}^4_H  - 2  m^4_{\tilde h}$  etc.  In  the
750: absence of soft  SUSY-breaking terms, one finds that  ${\rm STr} {\cal
751: M}^2 =  0$ and ${\rm STr}  {\cal M}^4 =  2\kappa^2\lambda^2 M^4$.  The
752: first  condition implies the  absence of  quadratic UV  divergences in
753: SUSY theories, whereas  the first together with the  second one ensure
754: the UV  finiteness along the $\tilde{u}_R$ direction,  namely the fact
755: that $d V^{(1)} (\tilde{u}_R)/d\tilde{u}_R$ is $Q^2$ independent.
756: 
757: It  would  be  more  illuminating  to  compute  the  1-loop  effective
758: potential in~(\ref{VuR}) in a  Taylor series expansion with respect to
759: $h^2|\tilde{u}_R|^2$.   To  order $h^4  |\tilde{u}_R|^4$,  the 3  mass
760: eigenvalues of ${\cal M}^2_H$ are approximately given by
761: \begin{eqnarray}
762: M^2_\pm & = & \lambda^2 |S|^2\: \pm\: \kappa\lambda M^2\: +\:
763: \frac{\kappa M^2 \pm 2 \lambda |S|^2}{2(\kappa M^2 \pm
764: \lambda |S|^2)}\; h^2|\tilde{u}_R|^2\: \pm\:
765: \frac{\kappa M^2(\kappa M^2 \pm 3 \lambda |S|^2)}{8\lambda(\kappa
766:   M^2 \pm \lambda |S|^2)^3}\;
767: h^4 |\tilde{u}_R|^4\; ,\nonumber\\
768: M_0^2 &=&
769: \frac{\kappa^2 M^4}{\kappa^2M^4 -\lambda^2 |S|^4}\; h^2 |\tilde{u}_R|^2\:
770: -\:
771: \frac{2\kappa^2\lambda^2 M^4 |S|^6}{(\kappa^2 M^4-\lambda^2 |S|^4)^3}\;
772: h^4 |\tilde{u}_R|^4\; .
773: \end{eqnarray}
774: Notice  that   in  the  limit   $\tilde{u}_R  \to  0$,   one  obtains:
775: $m^2_{\tilde{h}}  = \lambda^2  |S|^2$,  $M^2_\pm =  \lambda |S|^2  \pm
776: \kappa\lambda M^2$ and $M^2_0 =  0$, as expected.  Moreover, it is not
777: difficult  to  check  that ${\rm  STr}  {\cal  M}^2  = {\cal  O}  (h^6
778: |\tilde{u}_R|^6)$ and ${\rm STr} {\cal M}^4 = 2\kappa^2\lambda^2 M^4 +
779: {\cal  O} (h^6  |\tilde{u}_R|^6)$, in  accordance with  our discussion
780: given above.
781: 
782: Employing the fact that $|S|^2\gg \frac{\kappa}{\lambda}\, M^2$ at the
783: start   of  inflation,   the  1-loop   effective   potential  $V^{(1)}
784: (\tilde{u}_R)$ may further be approximated as follows:
785: \begin{eqnarray}
786: \label{VuRappr}
787: V^{(1)} (\tilde{u}_R) \!&=&\!
788: \frac{\kappa^2\lambda^2 M^4}{8\pi^2}\, \Bigg[
789: \ln\Bigg(\frac{\lambda^2 |S|^2}{Q^2}\Bigg)\: -\:  \frac{3}{2}\, \Bigg]\
790: -\ \frac{1}{48\pi^2}\frac{h^2 \kappa^4 M^8}{\lambda^2 |S|^6}\;
791: |\tilde{u}_R|^2\
792: +\
793: \frac{1}{16\pi^2}\frac{h^4\kappa^2 M^4}{\lambda^2 |S|^4}\; |\tilde{u}_R|^4
794: \nonumber\\
795: &&+\; \frac{1}{16\pi^2}
796: \left(\frac{h^2\kappa^2 M^4}{\lambda^2 |S|^4}\; |\tilde{u}_R|^2\, \right)^2
797: \ln\left(\frac{h^2\kappa^2 M^4}{\lambda^4 |S|^6}\; |\tilde{u}_R|^2
798: \right)\quad
799: +\quad {\cal O}(h^6 |\tilde{u}_R|^6)\,.
800: \end{eqnarray}
801: The   first  term   in~(\ref{VuRappr})  contributes   to   the  1-loop
802: inflationary    potential~(\ref{V1loop}),    while    the    remaining
803: $Q^2$-independent     terms    lift     the     flatness    of     the
804: $\tilde{u}_R$-direction.  Assuming  that $\kappa^2 \ll  \lambda^2$ and
805: $M\simeq |S|$ towards  the end of inflation, we  find the well-defined
806: minimum
807: \begin{equation}
808:   \label{VEVuR}
809: \langle\, \tilde{u}_R\, \rangle\ =\ \frac{\kappa}{\sqrt{6}\,h}\ M\; .
810: \end{equation}
811: We should remark  here that the above minimum  would remain unaltered,
812: even if the  flat direction were a squark or  slepton doublet. In this
813: case, only the overall  normalization of the $Q^2$-independent part of
814: $V^{(1)}  (\tilde{u}_R)$ would  have  changed by  a  factor 1/2.   The
815: loop-induced   VEV   of  $\tilde{u}_R$   generates   a  squared   mass
816: $M^2_{\tilde{u}_R}$ via the Higgs mechanism, which is given by
817: \begin{equation}
818:   \label{MuR}
819: M^2_{\tilde{u}_R}\ =\ \frac{1}{24\pi^2}\,
820: \frac{h^2\,\kappa^4}{\lambda^2}\; M^2\ .
821: \end{equation}
822: This squared mass $M^2_{\tilde{u}_R}$ should be compared with the size
823: of possible negative Hubble-induced squared mass terms of order $H^2 =
824: \kappa^2   M^4/    (3m_{\rm   Pl}^2)$,   e.g.~terms    of   the   form
825: $-c^2_{\tilde{u}}\,  H^2 |\tilde{u}_R|^2$ that  may occur  in $V^{(1)}
826: (\tilde{u}_R)$ and originate from  SUGRA effects. These terms may play
827: some   role   in   our   model,  unless   $c^2_{\tilde{u}}\,   H^2   <
828: M^2_{\tilde{u}_R}$.  The  latter condition may be  translated into the
829: inequality
830: \begin{equation}
831:   \label{cuR}
832: c_{\tilde{u}}\ <\ \frac{1}{2\sqrt{2}\pi}\, \frac{h\,\kappa}{\lambda}\;
833: \frac{m_{\rm Pl}}{M}\ .
834: \end{equation}
835: As a typical  example, let us consider an  inflationary scenario, with
836: $\lambda = 2\kappa$, $\kappa = 10^{-3}$ and $M = 10^{16}$~GeV. In this
837: case,  (\ref{cuR}) implies  that $c_{\tilde{u}}  < 0.87\,  h$.  Hence,
838: although  the required  tuning of  the coefficient  $c_{\tilde{u}}$ to
839: fulfill  this last  inequality may  not  be significant  for the  third
840: generation squarks  and sleptons, it  becomes excessive for  the first
841: generation, unless a minimal K\"ahler potential is assumed.  It should
842: be  stressed   here,  however,  that   the  deepest  and   hence  most
843: energetically favoured  minimum for all squark  and slepton directions
844: is the  one related  to $\tilde{t}_R$. In  other words,  given chaotic
845: initial conditions, the fields are  most likely to settle to minima of
846: quasi-flat directions involving large Yukawa couplings.  In this case,
847: radiative  effects play  an important  role  in the  dynamics of  flat
848: directions\footnote{We  should   note  that  the   evolution  of  flat
849: directions during the waterfall  and coherent oscillation periods is a
850: non-equilibrium  dynamics problem.   Moreover, no  theoretical methods
851: yet exist  that would  lead to a  practical solution to  this problem,
852: even though effective potential  corrections to the flat directions as
853: the ones considered here are  expected to be relevant during the above
854: cosmological periods.}.
855: 
856: 
857: Let  us  finally  assume  that   we  are  in  a  situation  where  the
858: Hubble-induced mass  terms can be neglected,  i.e.~$c_{\tilde{u}} = 0$
859: as is the case for  a minimal K\"ahler potential, for example. Suppose
860: that  the  loop-induced  VEV  of  the  quasi-flat  direction  persists
861: throughout  the  coherent  oscillatory  regime.   In  this  case,  the
862: VEV~(\ref{VEVuR}) gives rise to masses $h\langle \tilde{u}_R \rangle =
863: \kappa M/\sqrt{6}$ in the $\widehat{Q} \widehat{H}_u$-sector, which do
864: not   depend  on   the   Yukawa  coupling   $h$.   Consequently,   the
865: inflaton-related   fields   of  mass   $\sqrt{2}   \kappa\,  M$   (see
866: Table~\ref{spectrum}) will  have a large  decay rate to  those massive
867: particles, thus creating a non-thermal distribution.  This non-thermal
868: distribution will in turn induce $T$-dependent mass terms which can be
869: larger than  the expansion  rate $H(T)$ at  some temperature  $T$ soon
870: after inflation, such that  $\langle \tilde{u}_R \rangle$ will rapidly
871: relax  to   zero.   Of  course,  one  might   think  of  contemplating
872: configurations  where   multiple  flat  directions   have  VEVs  which
873: contribute constructively to the masses  of both $H_u$ and $H_d$, such
874: that all inflaton and waterfall particle decays would be kinematically
875: forbidden.   However,  we  consider   such  a  possibility  as  a  bit
876: contrived.    It    is   therefore   reasonable    to   assume   that,
877: provided~(\ref{cuR}) is fulfilled,  reheating and equilibration of all
878: MSSM degrees of freedom will take place in the $F_D$-term hybrid model
879: and  in  all  supersymmetric  models  of  inflation  that  include  an
880: unsuppressed   renormalizable  operator   of  the   form  $\widehat{S}
881: \widehat{H}_u \widehat{H}_d$.
882: 
883: 
884: \subsection{Topological Defects and GUT Embeddings}\label{TDGUT}
885: 
886: As  we mentioned  in the  Introduction, topological  defects,  such as
887: domain walls, cosmic  strings or monopoles, may be  created at the end
888: of inflation,  when a symmetry  group $G$, local, global  or discrete,
889: breaks  down into  a  subgroup $H$,  in  a way  such  that the  vacuum
890: manifold  $M =  G/H$ is  not trivial.   Specifically,  the topological
891: properties  of the  vacuum  manifold $M$  under  its homotopy  groups,
892: $\pi_n    (M)$,   determine    the   nature    of    the   topological
893: defects~\cite{Vilenkin,HK}.  Thus, one  generally has the formation of
894: domain walls for  $\pi_0 (M) \neq {\bf I}$,  cosmic strings for $\pi_1
895: (M) \neq {\bf I}$, monopoles if  $\pi_2 (M) \neq {\bf I}$, or textures
896: if $\pi_{n  > 2} (M)  \neq {\bf I}$~\cite{Vilenkin}. For  example, for
897: the SSB breaking pattern U(1)$_X \to {\bf I}$ in the waterfall sector,
898: the  first homotopy  group  of  the vacuum  manifold  is not  trivial,
899: i.e.~$\pi_1({\rm  U}(1)/{\bf I})  = {\bf  Z}$.  In  this  case, cosmic
900: strings will  be produced  at the end  of inflation.  In  general, the
901: non-observation  of  any  cosmic  string  contribution  to  the  power
902: spectrum $P_{\cal R}$ at the 10\% level introduces serious constraints
903: on the theoretical parameters of hybrid inflation models.
904: 
905: A  potentially   interesting  inflationary  scenario   arises  if  the
906: waterfall sector possesses an SU(2)$_X$ gauge symmetry.  In this case,
907: the SSB  breaking pattern  is: SU(2)$_X \to  {\bf I}$,  i.e.~the group
908: SU(2)$_X$ breaks completely. It is worth stressing here that this is a
909: unique  property of  the SU(2)  group,  since the  breaking of  higher
910: SU($N$)  groups,  with $N  >  2$, into  the  identity~{\bf  I} is  not
911: possible.    Moreover,   an  homotopy   group   analysis  gives   that
912: $\pi_{0,1,2} ({\rm SU}(2)_X/{\bf I}) = {\bf I}$, implying the complete
913: absence  of domain  walls,  cosmic strings  and  monopoles.  The  only
914: non-trivial homotopy  group is $\pi_3  ({\rm SU}(2)_X/{\bf I})  = {\bf
915: Z}$, thus signifying the formation  of textures, in case the SU$(2)_X$
916: group is global.  If the SU(2)$_X$ group is local, however, observable
917: textures do not occur.  Since their corresponding field configurations
918: never leave the  vacuum manifold, the would-be textures  can always be
919: compensated by  local SU(2)$_X$ gauge transformations~\cite{Vilenkin}.
920: It  is therefore  essential  that the  $X$-symmetry  of the  waterfall
921: sector is local in the $F_D$-term hybrid model.
922: 
923: It is now interesting to  explore whether generic scenarios exist, for
924: which the waterfall gauge groups  ${\rm U}(1)_X$ or ${\rm SU}(2)_X$ of
925: the $F_D$-term hybrid model  may, partially or completely, be embedded
926: into a  GUT. As a key  element for such a  model-building, we identify
927: the maintenance of $D$-parity  conservation in the $X$-gauged waterfall
928: sector, which  is discussed  in detail in  Section~\ref{postinfl}.  In
929: order to  preserve $D$-parity, the waterfall sector  should be somehow
930: `hidden'  from the  perspective of  the SM  gauge group  $G_{\rm SM}$.
931: This  means that  the SM  fields must  be neutral  under $X$  and vice
932: versa, the $X$-gauge and waterfall sector fields should not be charged
933: under  $G_{\rm SM}$.   Consequently,  we  have to  require,  as a  GUT
934: breaking   route,  that   the  waterfall   $X$-gauge  group   and  the
935: GUT-subgroup that contains  $G_{\rm SM}$ factor out into  a product of
936: two independent groups without overlapping charges.
937: 
938: It is reasonable to assume  that the GUT-subgroup is broken to $G_{\rm
939: SM}$ before  or while inflation takes place.   Then, possible unwanted
940: topological  defects due to  the various  stages of  symmetry breaking
941: from the  GUT-subgroup down to the  SM will be  inflated away.  Notice
942: that we do not have to require that the GUT-subgroup breaking scale is
943: higher than the respective  $X$-symmetry breaking scale, but only that
944: the  reheat temperature  $T_{\rm  reh}$  is low  enough  such that  no
945: symmetries  of  the GUT-subgroup  are  restored  during reheating.   A
946: related  discussion  within  the   context  of  SO(10)  may  be  found
947: in~\cite{SO10inflation}.
948: 
949: 
950: Let us first investigate whether a `hidden' gauge group ${\rm U}(1)_X$
951: related to the waterfall sector  can be embedded into a GUT.  Although
952: `hidden'  ${\rm   U}(1)$'s  naturally   arise  in  models   of  string
953: compactification~\cite{ExtraU1fromStrings},  our interest  here  is to
954: identify possible ${\rm  U}(1)_X$ factors that can be  embedded into a
955: simple GUT.   Given the above criterion, the  frequently discussed GUT
956: based on ${\rm SO}(10)$ should  be excluded, since it does not contain
957: `hidden'  U(1)$_X$ groups~\cite{Slansky:1981}.   As  a next  candidate
958: theory,  we may  consider the  exceptional  group E(6),  with the  SSB
959: breaking  path ${\rm E}(6)\to  {\rm U}(1)  \times {\rm  SO}(10)$.  The
960: fundamental representation of ${\rm  E}(6)$ is the chiral ${\bf 27}_F$
961: representation, which branches under  ${\rm U}(1) \times {\rm SO}(10)$
962: as follows:
963: \begin{equation}
964: {\bf 27}_F\  =\ (4,{\bf 1})\: +\: (-2, {\bf 10})\:  +\: (1,{\bf 16})\; .
965: \end{equation}
966: Although  the SM  particles  may fit  into  ${\bf 16}$,  they are  not
967: neutral  under the extra  U(1). Higher  representations, such  as $(0,
968: {\bf 45})$ stemming from ${\bf 78}$ of ${\rm E}(6)$, are neutral under
969: the  ${\rm  U}(1)$ factor,  but  they  are  not suitable  to  properly
970: accommodate all the SM particles.
971: 
972: We  therefore turn  our  attention to  possible  breaking patterns  of
973: maximal  groups that  contain a  `hidden' ${\rm  SU}(2)_X$  factor.  A
974: promising  example is  ${\rm  E}(6)\supset {\rm  SU}(2)_X \times  {\rm
975: SU}(6)$, where the fundamental representation ${\bf 27}_F$ follows the
976: branching:
977: \begin{equation}
978:   \label{27F}
979: {\bf 27}_F\ =\ ({\bf 2} ,{\bf \overline 6})\: +\: ({\bf 1},{\bf 15})\; .
980: \end{equation}
981: Under ${\rm SU}(6) \supset {\rm  SU}(5) \times {\rm U}(1)$, ${\bf 15}$
982: is an antisymmetric  representation of SU(6) and one  of its branching
983: rules is
984: \begin{equation}
985:   \label{15SM}
986: {\bf 15}\ =\ ({\bf 5},-4)\: +\: ({\bf 10},2)\; .
987: \end{equation}
988: However, we need a ${\bf \overline 5}$ of ${\rm SU}(5)$, together with
989: ${\bf 10}$ in~(\ref{15SM}), in  order to appropriately describe all SM
990: fermions.   This shortcoming may  be circumvented  by adding  an extra
991: ${\bf  \overline{27}}_F$ of ${\rm  E}(6)$ to  the spectrum,  where the
992: missing ${\bf \overline 5}$ may be obtained from the complex conjugate
993: branching of~(\ref{15SM}).  Such an extension of the particle spectrum
994: may even be welcome to resolve the proton stability problem, through a
995: kind of split multiplet mechanism~\cite{proton}.  Within the framework
996: of SUSY, the quark and lepton Yukawa interactions may be generated via
997: the  introduction of  a pair  of  the multiplets  ${\bf 27}_H$,  ${\bf
998: \overline{27}}_H$.  Finally,  in such an E(6) unified  scenario, the 3
999: right-handed neutrinos can only appear as singlets.
1000: 
1001: Another possible  GUT scenario that  complies with our criterion  of a
1002: hidden  SU(2)$_X$  is ${\rm  E}(7)\supset  {\rm  SU}(2)_X \times  {\rm
1003: SO}(12)$. The fundamental representation  is ${\bf 56}_F$ and branches
1004: under ${\rm SU}(2)_X \times {\rm SO}(12)$ as follows:
1005: \begin{equation}
1006:   \label{56E7}
1007: {\bf 56}_F\ =\ ({\bf 2} , {\bf \overline{12}})\: +\:
1008:                                                 ({\bf 1},{\bf 32})\; .
1009: \end{equation}
1010: Subsequently,  SO(12) breaks  spontaneously  into ${\rm  SO}(10)\times
1011: {\rm  U}(1)$,  where  ${\bf  32}  = ({\bf  16},1)  +  ({\bf  \overline
1012: {16}},-1)$  is a  vector-like representation.   However, one  may well
1013: envisage   a    string-theoretic   framework,   in    which   orbifold
1014: compactification  projects  out  the undesirable  anti-chiral  states.
1015: Then,  all SM  particles,  including right-handed  neutrinos, will  be
1016: contained in one of the ${\bf 16}$'s of ${\bf 32}$. Related discussion
1017: of missing  or incomplete multiplets due  to orbifold compactification
1018: may be found in~\cite{orbifold}.
1019: 
1020: Building  a realistic  GUT model  from  the blocks  stated above  lies
1021: beyond the scope  of this paper.  We have  demonstrated here, however,
1022: that the  embedding of an SU(2)$_X$  gauge group into a  GUT, which is
1023: hidden but nevertheless takes  part in the gauge coupling unification,
1024: appears feasible within E(6) and E(7) unified theories.
1025: 
1026: We conclude this section by observing that the presence of the singlet
1027: inflaton  field~$S$ offers  alternative options,  for  suppressing the
1028: heavy Majorana  neutrino masses within  SUSY GUTs.  As an  example, we
1029: mention the  breaking scenario, where  ${\rm SO}(10) \to  {\rm SU}(5)$
1030: via  the VEV of  a ${\bf  126}_H$ Higgs  representation and  the usual
1031: superpotential  term  ${\bf  16}_F  \langle  {\bf  126}_H\rangle\,{\bf
1032: 16}_F$ induces heavy  Majorana masses of the GUT  scale $M_{\rm GUT}$.
1033: Given  that the  above renormalizable  operator is  forbidden  by some
1034: $R$-symmetry,  the presence of  an $R$-charged  inflaton $S$  may give
1035: rise to a drastic suppression  of the GUT-scale Majorana mass, through
1036: a superpotential  term of the  form $\widehat{S}\, {\bf  16}_F \langle
1037: {\bf 126}_H\rangle\,{\bf 16}_F/m_{\rm Pl}$.   Since $S$ receives a VEV
1038: of  order  $M_{\rm SUSY}/\kappa$  in  general  $F$-term hybrid  models
1039: [cf.~(\ref{mu})], one naturally obtains heavy Majorana neutrino masses
1040: of order $M_{\rm SUSY}$,  if $\kappa \sim \langle {\bf 126}_H\rangle\,
1041: /m_{\rm  Pl} \sim  10^{-3}$. Such  values of  $\kappa$ do  satisfy the
1042: current inflationary constraints which we discuss in the next section.
1043: 
1044: 
1045: 
1046: 
1047: 
1048: \setcounter{equation}{0}
1049: \section{Inflation}\label{inflation}
1050: 
1051: 
1052: Here,  we  first  briefly  review  in  Section~\ref{intro}  the  basic
1053: formalism   of   inflation,  including   the   constraints  from   the
1054: non-observation of  cosmic strings in the power  spectrum $P_{\cal R}$
1055: of  the  CMB data.   Then,  in  Section~\ref{numinf},  we present  our
1056: numerical results  for two scenarios:  (i) the minimal  SUGRA (mSUGRA)
1057: scenario and  (ii) the  next-to-minimal SUGRA (nmSUGRA)  scenario.  In
1058: particular, we  exhibit numerical  predictions for the  spectral index
1059: $n_{\rm  s}$  and  discuss  its  possible  reduction  in  the  nmSUGRA
1060: scenario.  Finally,  we  analyze   the  combined  constraints  on  the
1061: fundamental  theoretical parameters  $\kappa$, $\lambda$,  $\rho$, and
1062: $M$, which result from the recent CMB observations and inflation.
1063: 
1064: \subsection{Basic Formalism}\label{intro}
1065: 
1066: According to the  inflationary paradigm~\cite{review}, the horizon and
1067: flatness  problems   of  the   standard  Big-Bang  Cosmology   can  be
1068: technically  addressed, if  our observable  Universe has  undergone an
1069: accelerated  expansion   of  a  number  50--60   of  $e$-folds  during
1070: inflation.  In  the slow-roll approximation, the  number of $e$-folds,
1071: ${\cal N}_e$, is related to the inflationary potential through:
1072: \begin{equation}
1073:   \label{Nefold}
1074: {\cal N}_e\ =\ \frac{1}{m^2_{\rm Pl}}\; \int_{\phi_{\rm
1075:     end}}^{\phi_{\rm exit}}\, d\phi\: \frac{V_{\rm inf}}{V'_{\rm inf}}\
1076:     \simeq\ 55\; .
1077: \end{equation}
1078: Hereafter, a  prime on $V_{\rm inf}$ will  denote differentiation with
1079: respect  to  the inflaton  field  $\phi=\sqrt{2}\,  {\rm Re}\,S$.   In
1080: addition, $\phi_{\rm exit}$  is the value of $\phi$,  when our present
1081: horizon scale crossed outside inflation's horizon and $\phi_{\rm end}$
1082: is the  value of  $\phi$ at  the end of  inflation.  In  the slow-roll
1083: approximation, the field value $\phi_{\rm end}$ is determined from the
1084: condition:
1085: \begin{equation}
1086:   \label{slow}
1087: {\sf max}\{\epsilon(\phi_{\rm end}),|\eta(\phi_{\rm
1088: end})|\}\ =\ 1\, ,
1089: \end{equation}
1090: where
1091: \begin{equation}
1092:   \label{epseta}
1093: \epsilon\ =\ \frac{m_{\rm Pl}^2}{2}\ \left(
1094: \frac{V'_{\rm inf}}{V_{\rm inf}}\right)^2\,,\qquad
1095: \eta\ =\  m_{\rm Pl}^2\ \frac{V''_{\rm inf}}{V_{\rm inf}}\ .
1096: \end{equation}
1097: We  have checked  that  the slow-roll  condition~(\ref{slow}) is  well
1098: satisfied up to the critical point $\phi_{\rm end}=\sqrt{2} M$, beyond
1099: which  the waterfall  mechanism takes  place.  We  also find  that the
1100: slow-roll condition  remains valid,  even within the  nmSUGRA scenario
1101: with $c_H\neq0$ and with appreciable non-renormalizable SUGRA effects.
1102: Finally,  we note that  the assumed  value of  ${\cal N}_e  \simeq 55$
1103: in~(\ref{Nefold})   is   slightly  higher   than   the  one   computed
1104: consistently  from~(\ref{Ng}), which  is about  50 for  our low-reheat
1105: cosmological  scenario.   However,  our numerical  results  concerning
1106: $P_{\cal R}$ and $n_{\rm s}$ do not depend on such a 10\% variation of
1107: ${\cal N}_e$ in any essential way.
1108: 
1109: The power  spectrum $P_{\cal R}$  is a cosmological observable  of the
1110: curvature perturbations, which  sensitively depends on the theoretical
1111: parameters of the inflationary potential. The square root of the power
1112: spectrum, $P^{1/2}_{\cal R}$, may be conveniently written down as
1113: \begin{equation}
1114:     \label{PR}
1115: P^{1/2}_{\cal R}\ =\ \frac{1}{2\sqrt{3}\, \pi m^3_{\rm Pl}}\;
1116: \frac{V_{\rm inf}^{3/2}(\phi_{\rm exit})}{|V'_{\rm inf}(\phi_{\rm
1117: exit})|}\ .
1118: \end{equation}
1119: The recent  WMAP~\cite{WMAP,WMAP3} results, which  are compatible with
1120: the  ones suggested  for the  COBE  normalization~\cite{COBE}, require
1121: that
1122: \begin{equation}
1123:   \label{Pr}
1124: P^{1/2}_{\cal R}\ \simeq\ 4.86\times 10^{-5}\, .
1125: \end{equation}
1126: 
1127: In addition  to scalar  curvature perturbations, tensor  gravity waves
1128: and cosmic string effects may  also contribute to $P_{\cal R}$. In the
1129: $F_D$-term  hybrid model  with an  Abelian $U(1)_X$  waterfall sector,
1130: cosmic strings arise after the SSB of the gauge symmetry (see also our
1131: discussion   in  Section~\ref{TDGUT}).    In  this   case,  additional
1132: constraints  are obtained  from the  non-observation of  cosmic string
1133: effects on~$P_{\cal  R}$~\cite{cstrings1,cstrings}.  The evaluation of
1134: such effects involves a certain degree of uncertainty in the numerical
1135: simulations  of string  networks \cite{cstrings2}.   Nevertheless, the
1136: common approach  taken to cosmic string  effects~\cite{mairi,JP} is to
1137: require that  their contribution $(P_{\cal R})_{\rm cs}$  to the power
1138: spectrum $P_{\cal  R}$ does not exceed the  10\% level, i.e.~$(P_{\cal
1139: R})_{\rm  cs}/P_{\cal R}  \stackrel{<}{{}_\sim}  0.1$.  In~detail,  we
1140: require that
1141: \begin{equation}
1142:   \label{Prcs}
1143: (P^{1/2}_{\cal R})_{\rm cs}\ \leq\ 1.54 \times 10^{-5}\; .
1144: \end{equation}
1145: The cosmic string contribution $(P_{\cal  R})_{\rm cs}$ to the power
1146: spectrum may be computed by
1147: \begin{equation}
1148: (P^{1/2}_{\cal R})_{\rm cs}\ = \
1149: {\sqrt{15}\over 4\pi}\: {\mu_{\rm cs}\over m^2_{\rm Pl}}\ y_{\rm cs}\; ,
1150: \end{equation}
1151: where the tension of the cosmic strings, $\mu_{\rm cs}$, is calculated
1152: using the formulae:
1153: \begin{equation}
1154:   \label{mucs}
1155: \mu_{\rm cs}\ =\ 2\pi M^2\epsilon_{\rm cs}(\beta)\,,\qquad
1156: \epsilon_{\rm cs}(\beta)\ \simeq\ \left\{\matrix{
1157: %\begin{array}{rl}
1158: 1.04\ \beta^{0.195}\hfill ,  & \mbox{for}~~\beta>10^{-2},
1159: \hfill \cr
1160: %
1161: 2.4\,/ \ln(2/\beta) \hfill ,  &\mbox{for}~~\beta\leq10^{-2}\; .
1162: \hfill \cr}
1163: %\end{array}
1164: \right.
1165: \end{equation}
1166: In~(\ref{mucs})   the  argument   $\beta$   is  given   by  $\beta   =
1167: \kappa^2/(2g^2)$, while  the U(1)$_X$  gauge coupling constant  $g$ is
1168: considered to  assume the value $g \simeq  0.7$ as is the  case in GUT
1169: models.  The  central value of the  parameter $y_{\rm cs}$  is 8.9 and
1170: its error  margin lies  in the interval  [6.7,11.6], according  to the
1171: analysis in~\cite{cstrings}.
1172: 
1173: 
1174: The  recently  announced   three-years  results  of  WMAP~\cite{WMAP3}
1175: improved  upon  the  precision  of  a  number  of  other  cosmological
1176: observables.  The  merits of  an inflationary model  can be  judged by
1177: comparing its predictions for  the scalar spectral index, $n_{\rm s}$,
1178: the  tensor to  scalar ratio,  $r$, and  the running  of  $n_{\rm s}$,
1179: $dn_{\rm s}/d\ln\kappa$, with the  CMB data.  In the $F_D$-term hybrid
1180: model,  $r=16\epsilon(\phi_{\rm exit})$  is much  lower than  the WMAP
1181: bound,  i.e.~well  below~$10^{-2}$,  and  $dn_{\rm  s}/d\ln\kappa$  is
1182: always smaller  than $10^{-3}$ and  so unobservable. In  addition, the
1183: spectral index  $n_{\rm s}$ in our  model may well  be approximated as
1184: follows:~\cite{review}
1185: \begin{equation}
1186:   \label{nS}
1187: n_{\rm s}\ =\ 1-6\epsilon(\phi_{\rm exit})\ +\ 2\eta(\phi_{\rm exit})\
1188: \simeq\ 1\ +\ 2\eta(\phi_{\rm exit}),
1189: \end{equation}
1190: since  $\epsilon$  is negligible.  The  predicted  value  needs to  be
1191: compared with the recent WMAP results~\cite{WMAP3}:
1192: \begin{equation}
1193:   \label{nswmap}
1194: n_{\rm  s}\ =\ 0.951_{-0.019}^{+0.015}\ .
1195: \end{equation}
1196: The latter is translated into the double inequality,
1197: \begin{equation}
1198:   \label{ns95}
1199: 0.913\ \lesssim\  n_{\rm s}\ \lesssim\ 0.981\; ,
1200: \end{equation}
1201: at the 95\% confidence level (CL).
1202: 
1203: 
1204: The  result~(\ref{ns95})  brings  under  considerable  stress  minimal
1205: $F$-term hybrid  inflation models~\cite{DSS}. This is due  to the fact
1206: that these models predict $n_{\rm s}$ extremely close to unity without
1207: much running.  To be  precise, when the radiative corrections dominate
1208: the slope of the potential, we obtain
1209: \begin{equation}
1210:   \label{nSrc}
1211: n_{\rm s}\ \simeq\ 1\ -\ {1/{\cal N}_e}\ \simeq\ 0.98\; ,
1212: \end{equation}
1213: for ${\cal  N}_e = 55$. On  the other hand,  if the non-renormalizable
1214: operator  $|S|^4$ in $V_{\rm  SUGRA}$ of~(\ref{VSUGRA})  dominates the
1215: slope of  the potential of  a mSUGRA model with  $c_H=0$~\cite{SS}, we
1216: obtain a blue-tilted spectrum, with
1217: \begin{equation}
1218:   \label{nSrc2}
1219: n_{\rm s}\ \simeq\ 1\ +\ {6M^2\over m_{\rm Pl}^2-2M^2{\cal N}_e}\
1220: \stackrel{>}{{}_\sim}\ 1\; .
1221: \end{equation}
1222: A possible Hubble-induced positive  term $+c_H^2 H^2 |S|^2$ in $V_{\rm
1223: SUGRA}$~\cite{CP, JP}  implies an  even more pronounced  blue spectrum
1224: and is therefore excluded by the current WMAP data.
1225: 
1226: As  noticed  earlier   in~\cite{hilltop}  and  elaborated  further  in
1227: Ref.~\cite{king},  agreement of  theory's  prediction for~$n_{\rm  s}$
1228: with  observation  strongly  suggests   the  presence  of  a  negative
1229: Hubble-induced  mass  term  $-c^2_H  H^2 |S|^2$  in  $V_{\rm  SUGRA}$,
1230: thereby  clearly disfavouring the  minimal K\"ahler  potential. In~our
1231: analysis, we therefore consider the following next-to-minimal form for
1232: the K\"{a}hler manifold~\cite{CP}:
1233: \begin{equation}
1234:   \label{qK}
1235: K_S\ =\ |S|^2\ +\ k_S\;{|S|^4\over 4\,m^2_{\rm Pl}}\ ,
1236: \end{equation}
1237: where  the  constant  $k_S$   can  be  either  positive  or  negative.
1238: Substituting (\ref{qK}) into the general formula for the $F$-term type
1239: contributions to the SUGRA potential~(see, e.g.~\cite{CLLSW}),
1240: \begin{equation}
1241:   \label{VF}
1242: V_F\ =\ e^{K_S/m^2_{\rm Pl}}\, \Bigg[\, F^i (K^{-1}_S)_i^j F_j\ -\ 3\,
1243: \frac{|W|^2}{m^2_{\rm Pl}}\,\Bigg]\ ,
1244: \end{equation}
1245: we  arrive at  the result~(\ref{VSUGRA})  with $c^2_H  =  3k_S$, after
1246: neglecting   higher-order   terms    that   are   small   for   $|c_H|
1247: \stackrel{<}{{}_\sim}   0.2$.    In    (\ref{VF}),   $F^i$   are   the
1248: SUGRA-generalized  $F$-terms  and  $(K^{-1}_S)_i^j$ is  the  so-called
1249: inverse  metric   of  the  K\"ahler  manifold,   where  a  superscript
1250: (subscript)  index $i$ or  $j$ on  $K_S$ denotes  differentiation with
1251: respect to $S$ ($S^*$).
1252: 
1253: The  aforementioned nmSUGRA  inflationary potential,  with  a negative
1254: Hubble-induced mass term,  reaches a local minimum and  maximum at the
1255: points  $\phi_{\rm min}$  and $\phi_{\rm  max}$,  respectively.  These
1256: points can be estimated by
1257: \begin{equation}
1258:  \label{phimax}
1259: \phi_{\rm max}\ \simeq\  {m_{\rm Pl}\over 4\pi c_H}\
1260: \Big(\, 6\kappa^2{\cal N}\: +\: 12\lambda^2\: +\:
1261: 9\rho^2\,\Big)^{1/2}\,, \qquad
1262: \phi_{\rm min}\ \simeq\ \sqrt{2\over3}\; c_H m_{\rm Pl}\ .
1263: \end{equation}
1264: For relevant  parameter values,  for which $\phi_{\rm  max} <\phi_{\rm
1265: min}$, and under convenient  initial conditions, the so-called hilltop
1266: inflation~\cite{hilltop}  can  take  place,  where $\phi$  rolls  from
1267: $\phi_{\rm max}$ down to smaller  values, such that $\phi_{\rm exit} <
1268: \phi_{\rm max}$.  In  this nmSUGRA scenario, the value  of $n_{\rm s}$
1269: can be significantly lowered and can be approximately given by
1270: \begin{equation}
1271:   \label{nShilltop}
1272: n_{\rm s}\ \simeq\ 1\: -\ \frac{1}{{\cal N}_e}\ -\ c_H^2\; .
1273: \end{equation}
1274: As  we will show  more explicitly  in the  next section,  the spectral
1275: index $n_{\rm s}$ can be easily driven into the range of~(\ref{ns95}),
1276: for  values   of  $c_H  \sim   0.1$.   The  presence  of   the  second
1277: next-to-minimal term proportional  to $k_S$ in~(\ref{qK}) modifies the
1278: analytic    expressions    of~(\ref{Nefold}),    (\ref{epseta})    and
1279: (\ref{Pr})~\cite{king}.  However,  these modifications turn  out to be
1280: numerically insignificant for the predicted values of ${\cal N}_e$ and
1281: $P_{\cal   R}$,    if   $c_H$    is   not   very    large,   e.g.~$c_H
1282: \stackrel{<}{{}_\sim} 0.2$.
1283: 
1284: 
1285: 
1286: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1287: \begin{figure}[t]
1288: \centerline{\epsfig{file=mSUGRAPR.eps,angle=-90,width=13.cm}} \hfill
1289: \vspace*{-.15in} \hfill
1290: \centerline{\epsfig{file=mSUGRAns.eps,angle=-90,width=13.cm}}\hfill
1291: \caption{\sl\small The values of the inflationary scale $M$ allowed
1292: by~(\ref{Nefold}) and (\ref{Pr}) {\sf (a)} and the predicted values of
1293: the spectral index $n_{\rm s}$ {\sf (b)} as a function of $\kappa$ for
1294: ${\cal N}=1$ and $\rho=\lambda=\kappa$ (light grey lines) or
1295: $\rho=\lambda=4\kappa$ (grey lines), including the one-loop radiative
1296: corrections (dashed lines) or the mSUGRA ($c_H=0$) contributions with
1297: a$_S=1~{\rm TeV}$ (solid lines). The upper bound of~(\ref{Prcs}) for
1298: $y_{\rm cs} =6.7,~8.9,~11.6$ (from top to bottom) [cf.~(\ref{nswmap})]
1299: is also shown by thin lines {\sf (a)} [{\sf (b)}].}
1300: \label{fig:mSUGRAPR}
1301: \end{figure}
1302: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1303: 
1304: 
1305: \subsection{Numerical results}\label{numinf}
1306: 
1307: In  our  numerical estimates,  we  use  the  full expression  for  the
1308: inflationary  potential  $V_{\rm  inf}$ given  in~(\ref{Vinf}),  which
1309: consists  of the  tree-level,  1-loop and  SUGRA contributions,  given
1310: in~(\ref{V0inf}), (\ref{V1loop}) and (\ref{VSUGRA}), respectively.  We
1311: will ignore all soft SUSY-breaking  terms, but the tadpole term $a_S$.
1312: To facilitate  our numerical analysis,  we introduce the  real tadpole
1313: parameter $\mbox{a}_S$,  which is defined, in terms  of the Lagrangian
1314: parameter $a_S$, by the relation:
1315: \begin{equation}
1316:   \label{aS}
1317: \mbox{a}_S\  =\ -\, 2  |a_S|\, \cos{(\arg{a_S}  + \arg{S})}\; .
1318: \end{equation}
1319: For any  given value of  $\kappa,~\lambda,~\rho$, a$_S$ and  $c_H$, we
1320: determine    $\phi_{\rm   exit}$    and   $M$,    by    imposing   the
1321: conditions~(\ref{Nefold}) and  (\ref{Pr}) for the  number ${\cal N}_e$
1322: of $e$-folds and the  power spectrum $P^{1/2}_{\cal R}$, respectively.
1323: In  addition, we  compute  $n_{\rm s}$  by  means of~(\ref{nS}).   Our
1324: results  are  presented  in  Fig.~\ref{fig:mSUGRAPR}  for  the  mSUGRA
1325: scenario  and in  Fig.~\ref{fig:nmSUGRAPR} for  the  nmSUGRA scenario.
1326: They will be analyzed in more detail in the following two subsections.
1327: 
1328: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1329: \begin{figure}[!th]
1330: \centerline{\epsfig{file=mSUGRAlr.eps,angle=-90,width=13.cm}} \hfill
1331: \caption{\sl\small The allowed values of $\lambda/\kappa$ versus
1332: $\rho/\kappa$ for the mSUGRA scenario with $M=2\times10^{16}~{\rm
1333: GeV}$ and $\kappa=0.05$ (dark grey line), $\kappa=0.01$ (grey line) or
1334: $\kappa=0.005$ (light grey line).}\label{fig:mSUGRAlr}
1335: \end{figure}
1336: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1337: 
1338: 
1339: 
1340: 
1341: 
1342: 
1343: \subsubsection{The minimal SUGRA scenario}\label{msugra}
1344: 
1345: Here,  we present  numerical  results for  the  mSUGRA scenario.   The
1346: values  of the  inflationary scale  $M$ allowed  by~(\ref{Nefold}) and
1347: (\ref{Pr}) and  the predicted values  of $n_{\rm s}$, as  functions of
1348: $\kappa$,   for   $\rho=\lambda=\kappa$   (light   grey   lines)   and
1349: $\rho=\lambda=4\kappa$     (grey    lines),    are     displayed    in
1350: Fig.~\ref{fig:mSUGRAPR}(a)   and~\ref{fig:mSUGRAPR}(b),  respectively.
1351: Dashed  lines   indicate  results  obtained,  when   only  the  1-loop
1352: contribution to $V_{\rm inf}$ is considered and $\mbox{a}_S$ is set to
1353: zero, whilst  solid lines represent numerical values  obtained, if the
1354: remaining   contributions   are    included,   namely   those   coming
1355: from~(\ref{V0inf})  with  a$_S=1~{\rm  TeV}$  and~(\ref{VSUGRA})  with
1356: $c_H=0$.  In  Fig.~\ref{fig:mSUGRAPR}, we  observe that as  the common
1357: value for  $\rho$, $\lambda$ and  $\kappa$ increases, $M$  and $n_{\rm
1358: s}$ increase as well. In  particular, $M$ gets closer to the GUT-scale
1359: value $2  \times 10^{16}~{\rm GeV}$ for $\kappa  \sim 10^{-3}$, unlike
1360: the case $\lambda=\rho=0$,  where $M$ takes on much  smaller values at
1361: this  point~\cite{DSS,SS,JP}.
1362: 
1363: It  is now  not difficult  to identify  in~Fig.~\ref{fig:mSUGRAPR} the
1364: regimes,  in  which  the  different  contributions  to  $V_{\rm  inf}$
1365: dominate.   More explicitly, for  $\kappa \gtrsim  4\times10^{-3}$ and
1366: $\rho=\lambda=\kappa$     or    $\kappa    \gtrsim     10^{-3}$    and
1367: $\rho=\lambda=4\kappa$,    the     non-renormalizable    SUGRA    term
1368: of~(\ref{VSUGRA}) dominates and drives  $n_{\rm s}$ to values close to
1369: or larger than~1 [cf.~Fig.~\ref{fig:mSUGRAPR}(b)].  On the other hand,
1370: for  $4\times10^{-4}  \lesssim \kappa  \lesssim  4\times 10^{-3}$  and
1371: $\rho = \lambda =  \kappa$ or $4\times10^{-4} \lesssim \kappa \lesssim
1372: 10^{-3}$   and    $\rho   =    \lambda   =   4\kappa$,    the   1-loop
1373: corrections~(\ref{V1loop}) dominate, in  which case the spectral index
1374: $n_{\rm  s}$   takes  on  the   predicted  value  $\sim   0.98$  given
1375: in~(\ref{nSrc}).  Finally,  for $\kappa \lesssim  4\times 10^{-3}$ and
1376: $\rho=\lambda=\kappa$  or  $\rho=\lambda=4\kappa$,  the  tadpole  term
1377: in~(\ref{V0inf}) starts playing an  important role.  As $M$ increases,
1378: the  non-renormalizable  SUGRA  term of~(\ref{VSUGRA})  becomes  again
1379: important~\cite{SS,JP}.    In   this    case,   the   prediction   for
1380: $P^{1/2}_{\cal R}$ and $n_{\rm s}$ is almost independent of $\rho$ and
1381: $\lambda$, as expected.  For lower values of a$_S$, the solid lines in
1382: the   latter    regime   would   eventually    approach   the   dashed
1383: lines~\cite{JP}.  In Fig.~\ref{fig:mSUGRAPR}(b), we also indicate with
1384: a  thin  line  the  $95\%$  CL  upper  limit  on  $n_{\rm  s}$  stated
1385: in~(\ref{ns95}). Clearly,  a mSUGRA  version of the  $F_D$-term hybrid
1386: model appears to be disfavoured by the most recent WMAP results.
1387: 
1388: 
1389: In addition, we show in Fig.~\ref{fig:mSUGRAPR}(a) upper limits due to
1390: cosmic   string  effects  based   on~(\ref{Prcs}),  for   $y_{\rm  cs}
1391: =6.7,~8.9,~11.6$  (from top  to  bottom).  Such  constraints are  only
1392: relevant for an Abelian realization of the waterfall-gauge sector.  We
1393: observe  that the  presence of  cosmic strings  severely  restrict the
1394: available parameter space of the U(1)$_X$ $F_D$-term hybrid model.  As
1395: we discussed  in Section~\ref{TDGUT}, however, these  constraints do no
1396: longer  apply, if  the  waterfall-gauge sector  realizes an  SU(2)$_X$
1397: gauge  symmetry.  Since  the dimensionality  of the  representation is
1398: ${\cal N}=2$ in  this case, the allowed range of $M$  as a function of
1399: $\kappa$ slightly changes.  In fact,  the allowed values of $M$ become
1400: marginally     larger     than      the     ones     already     shown
1401: in~Fig.~\ref{fig:mSUGRAPR}(a)      by     up     to      12\%,     for
1402: $\rho=\lambda=\kappa$,  while   they  stay  at  the   1\%  level,  for
1403: $\rho=\lambda=4\kappa$. Likewise, the  predicted values of $n_{\rm s}$
1404: remain  almost  unaffected at  the  2\%  level,  from those  presented
1405: in~Fig.~\ref{fig:mSUGRAPR}(b). Obviously, as $\rho$ and $\lambda$ gets
1406: larger  than $\kappa$,  the difference  between the  ${\cal  N}=1$ and
1407: ${\cal N}=2$ cases becomes practically unobservable.
1408: 
1409: Finally,  in Fig.~\ref{fig:mSUGRAlr}  we  plot the  allowed values  of
1410: $\lambda/\kappa$     versus    $\rho/\kappa$,    subject     to    the
1411: constraints~(\ref{Nefold})   and  (\ref{Pr}),  for   $M  =   2  \times
1412: 10^{16}$~GeV  (close to  the GUT  scale) and  for different  values of
1413: $\kappa$: $\kappa=0.005$ (dark  grey line), $\kappa=0.001$ (grey line)
1414: or  $\kappa=0.0005$ (light  grey  line).  Along  these contour  lines,
1415: $\phi_{\rm exit}$  and $n_{\rm  s}$ remain constant  and equal  to the
1416: values  presented in  Table~\ref{tab1}.  We  observe that  as $\kappa$
1417: increases, $\phi_{\rm exit}$ and $n_{\rm s}$ increase as well.
1418: 
1419: 
1420: 
1421: 
1422: \begin{table}[!t]
1423: \begin{center}
1424: \begin{tabular}{|l|c|c||c|c|c|c|c|}\hline
1425: %
1426: \multicolumn{3}{|c||}{Fig.~\ref{fig:mSUGRAlr}}&\multicolumn{5}{|c|}{Fig.~\ref{fig:nmSUGRAlr}}\\\hline
1427: %
1428: \multicolumn{3}{|c||}{$M=2\times10^{16}~{\rm GeV}$}&\multicolumn{5}{|c|}{
1429: $\kappa=0.005,~M=10^{16}~{\rm GeV}$}\\\hline
1430: %
1431: {$~~~\kappa$}&{$\phi_{\rm
1432: exit}$}&{$n_{\rm s}$}&{$c_H$}
1433: &{$\phi_{\rm min}$}&{$\phi_{\rm max}$}&{$\phi_{\rm exit}$}&{$n_{\rm s}$}\\
1434: \hline\hline
1435: %
1436: $0.005$&{$6.28$}&{$1.017$}&{$0.07$}&{$8.4$}&{$-$}&{$5.1$}&{$0.978$}\\
1437: %
1438: $0.001$&{$2.14$}&{$0.99$}&{$0.14$}&{$16.5$}&{$10.5$}&{$8.1$}&{$0.955$}\\
1439: %
1440: $0.0005$&{$1.51$}&{$0.99$}&{$0.18$}&{$21.7$}&{$12.8$}&{$11.2$}&{$0.941$}\\ \hline
1441: %
1442: \end{tabular}
1443: \end{center}
1444: \caption{\sl\small The values of $\phi_{\rm exit}$ (in units
1445: $\sqrt{2}M$) and $n_{\rm s}$ for several $\kappa$'s along the curves
1446: in Fig.~\ref{fig:mSUGRAlr} and the values of $\phi_{\rm min}$, $\phi_{\rm
1447: max}$, $\phi_{\rm exit}$ (in units $\sqrt{2}M$) and $n_{\rm s}$ for
1448: several $c_H$'s along the curves in Fig.~\ref{fig:nmSUGRAlr}.} \label{tab1}
1449: \end{table}
1450: 
1451: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1452: \begin{figure}[t]
1453: \centerline{\epsfig{file=nmSUGRAPR.eps,angle=-90,width=13.cm}} \hfill
1454: \vspace*{-.15in} \hfill
1455: \centerline{\epsfig{file=nmSUGRAns.eps,angle=-90,width=13.cm}}\hfill
1456: \caption{\sl\small The values of the inflationary scale $M$ allowed
1457: by~(\ref{Nefold}) and (\ref{Pr}) {\sf (a)} and the predicted values of
1458: the spectral index $n_{\rm s}$ {\sf (b)} as a function of $\kappa$ for
1459: ${\cal N}=1$ and $\rho=\lambda=\kappa$ (light grey lines) or
1460: $\rho=\lambda=4\kappa$ (grey lines), for the nmSUGRA scenario with
1461: $c_H=0.07$ (dashed lines) or $c_H=0.14$ (solid lines). In both cases
1462: we take a$_S=1~{\rm TeV}$.  The upper bound given in~(\ref{Prcs}) (for
1463: $y_{\rm cs}=6.7,~8.9,~11.6$ from top to bottom) [cf.~(\ref{nswmap})]
1464: is also depicted by thin lines {\sf (a)} [{\sf (b)}].}
1465: \label{fig:nmSUGRAPR}
1466: \end{figure}
1467: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1468: 
1469: 
1470: \subsubsection{The next-to-minimal SUGRA scenario}\label{qsugra}
1471: 
1472: We now turn  our attention to the nmSUGRA  scenario.  Although we
1473: take the tadpole  term to be a$_S=1~{\rm  TeV}$, its impact  on
1474: our results turns out  to be  insignificant for the  whole range
1475: of  parameters we have  scanned.   The values  of  the
1476: inflationary  scale $M$  allowed by~(\ref{Nefold}) and  (\ref{Pr})
1477: and the  predicted $n_{\rm s}$  as a function  of  $\kappa$  are
1478: presented  in  Fig.~\ref{fig:nmSUGRAPR}(a)
1479: and~\ref{fig:nmSUGRAPR}(b),  respectively,  for
1480: $\rho=\lambda=\kappa$ (light  grey  lines)  and
1481: $\rho=\lambda=4\kappa$  (grey  lines).   We consider  the  two
1482: cases:  $c_H=0.07$  (dashed  lines) and  $c_H=0.14$ (solid lines).
1483: As in the case of mSUGRA, $M$ and $n_{\rm s}$ increase, with
1484: increasing  $\rho, \lambda$  and $\kappa$. Moreover,  as $\kappa$
1485: decreases, the non-renormalizable SUGRA contribution in
1486: (\ref{VSUGRA}) becomes  subdominant  and $n_{\rm  s}$  decreases.
1487: Such a  reduction becomes  even more  drastic with  increasing
1488: $c_H$,  as can  be easily inferred  from
1489: Fig.~\ref{fig:nmSUGRAPR}(a),  where  the 95\%~CL  upper bound  on
1490: $n_{\rm  s}$  [cf.~(\ref{ns95})] is  indicated  with a  thin
1491: horizontal line  on the  same plot.  In  stark contrast to  the
1492: mSUGRA scenario, we  observe that our  model can become  perfectly
1493: consistent with  the recent  WMAP result  for $0.04  \lesssim c_H
1494: \lesssim 0.22$. Note that the various lines terminate at large
1495: values of $\kappa$, for which  the two restrictions~(\ref{Nefold})
1496: and~(\ref{Pr}) cannot  be simultaneously met.
1497: 
1498: 
1499: 
1500: \begin{sidewaystable}
1501: \begin{center}
1502: \begin{tabular}{|l||c|c|c|c|c|c||c|c|c|c|c|c||c|c|c|c|c|c|}\hline
1503: %
1504: {$~~\kappa$}&{$c_H$}&{$M$}&{$\phi_{\rm min}$}&{$\phi_{\rm max}$}&{$\phi_{\rm
1505: exit}$}&{$\Delta_{\rm exit}$}
1506: &{$c_H$}&{$M$}&{$\phi_{\rm min}$}&{$\phi_{\rm max}$}&{$\phi_{\rm
1507: exit}$}&{$\Delta_{\rm exit}$}&{$c_H$}&
1508: {$M$}&{$\phi_{\rm min}$}&{$\phi_{\rm max}$}&{$\phi_{\rm exit}$}&{$\Delta_{\rm
1509: exit}$}\\
1510: \cline{2-19}
1511: %
1512: &\multicolumn{6}{|c||}{$n_{\rm s}=0.913$}&\multicolumn{6}{|c||}{$n_{\rm
1513: s}=0.951$}
1514: &\multicolumn{6}{|c|}{$n_{\rm s}=0.981$}\\ \hline\hline
1515: %
1516: \multicolumn{19}{|c|}{$\lambda=\rho=\kappa$}\\ \hline
1517: %
1518: $0.01$&{$0.179$}&{$0.34$}&{$73.6$}&{$11.9$}&{$11.3$}&{$0.050$}
1519:       &{$0.130$}&{$0.53$}&{$32.0$}&{$10.8$}&{$8.75$}&{$0.19$}&
1520:       {$0.065$}&{$0.78$}&{$16.7$}&{$-$}&{$7.50$}&{$-$}\\
1521: %
1522: $0.005$&{$0.176$}&{$0.34$}&{$73.1$}&{$6.0$}&{$5.7$}&{$0.053$}
1523:        &{$0.120$}&{$0.53$}&{$32.2$}&{$6.2$}&{$4.48$}&{$0.18$}
1524:        &{$0.040$}&{$0.78$}&{$8.20$}&{$-$}&{$3.90$}&{$-$}\\
1525: %
1526: $0.001$&{$0.173$}&{$0.25$}&{$95.6$}&{$1.64$}&{$1.6$}&{$0.028$}
1527:         &{$0.120$}&{$0.38$}&{$45.0$}&{$1.55$}&{$1.42$}&{$0.09$}
1528:         &{$0.060$}&{$0.58$}&{$19.0$}&{$2.10$}&{$1.36$}&{$0.34$}\\
1529: %
1530: $0.0005$&{$0.165$}&{$0.19$}&{$121$}&{$1.23$}&{$1.21$}&{$0.014$}
1531:         &{$0.116$}&{$0.28$}&{$58.8$}&{$1.19$}&{$1.15$}&{$0.04$}
1532:         &{$0.060$}&{$0.43$}&{$20.0$}&{$1.37$}&{$1.13$}&{$0.17$}\\\hline
1533: %
1534: \multicolumn{19}{|c|}{$\lambda=\rho=4\kappa$}\\ \hline
1535: %
1536: $0.01$&{$0.216$}&{$0.56$}&{$49$}&{$23.0$}&{$22.0$}&{$0.046$}
1537:       &{$0.190$}&{$0.83$}&{$23.0$}&{$21.9$}&{$17.0$}&{$0.22$}
1538:       &{$0.169$}&{$1.12$}&{$26.0$}&{$-$}&{$14.3$}&{$-$}\\
1539: %
1540: $0.005$&{$0.188$}&{$0.61$}&{$41$}&{$11.4$}&{$10.8$}&{$0.050$}
1541:        &{$0.146$}&{$0.96$}&{$26.0$}&{$9.1$}&{$8.30$}&{$0.19$}
1542:        &{$0.103$}&{$1.36$}&{$8.6$}&{$-$}&{$7.05$}&{$-$}\\
1543: %
1544: $0.001$&{$0.177$}&{$0.57$}&{$43$}&{$2.48$}&{$2.38$}&{$0.043$}
1545:        &{$0.125$}&{$0.89$}&{$24.6$}&{$2.28$}&{$1.96$}&{$0.14$}
1546:         &{$0.058$}&{$1.30$}&{$4.7$}&{$-$}&{$1.82$}&{$-$}\\
1547: %
1548: $0.0005$&{$0.178$}&{$0.46$}&{$54$}&{$1.53$}&{$1.49$}&{$0.028$}
1549:         &{$0.129$}&{$0.68$}&{$26$}&{$1.45$}&{$1.33$}&{$0.08$}
1550:         &{$0.070$}&{$1.00$}&{$9.6$}&{$1.83$}&{$1.30$}&{$0.29$}\\ \hline
1551: \end{tabular}
1552: \end{center}
1553: \caption{\sl\small The  values of  $c_H$, $M$ (in  units $10^{16}~{\rm
1554: GeV}$) $\phi_{\rm min}$, $\phi_{\rm max}$, $\phi_{\rm exit}$ (in units
1555: $\sqrt{2}M$)  and $\Delta_{\rm  exit}  = (\phi_{\rm  max} -  \phi_{\rm
1556: exit})/\phi_{\rm max}$, for selected values of $\kappa$, $\lambda$ and
1557: $\rho$, and for fixed values of the spectral index $n_{\rm s}$.}\label{tab2}
1558: \end{sidewaystable}
1559: 
1560: 
1561: It  is interesting  to further  investigate the  inflationary dynamics
1562: described   by  $V_{\rm   inf}$  in   the  presence   of   a  negative
1563: Hubble-induced mass term.  To this end, we exhibit in Table~\ref{tab2}
1564: the values of $c_{H}$,  $\phi_{\rm min}$, $\phi_{\rm max}$, $\phi_{\rm
1565: exit}$ (in units $\sqrt{2}M$) and the inflationary scale $M$ (in units
1566: of $10^{16}$~GeV) which are obtained for different values of $\kappa$,
1567: assuming that $\lambda = \rho = \kappa$ or $\lambda = \rho = 4\kappa$,
1568: and for fixed values of $n_{\rm s}$, i.e.~$n_{\rm s} = 0.913,\ 0.951,\
1569: 0.981$, compatible  with the 95\% CL limits  given in~(\ref{ns95}). In
1570: addition,  we present values  for the  parameter $\Delta_{\rm  exit} =
1571: (\phi_{\max}  -  \phi_{\rm   exit})/\phi_{\rm  exit}$,  which  somehow
1572: quantifies the degree of tuning  required in the initial conditions of
1573: inflation.  The entries without a value assigned (in Tables~\ref{tab1}
1574: and~\ref{tab2})  mean  that   the  respective  inflationary  potential
1575: $V_{\rm inf}$  has no distinguishable nearby  local maximum $\phi_{\rm
1576: max}$. We notice from  Table~\ref{tab2}, that as $n_{\rm s}$ decreases
1577: with  fixed  values  of  $\kappa$,  $c_{H}$ increases  while  $M$  and
1578: $\Delta_{\rm exit}$  decrease. Moreover,  for fixed values  of $n_{\rm
1579: s}$  and decreasing $\kappa$,  $c_H$ and  $M$ decrease  and $\phi_{\rm
1580: exit}$  approaches $\phi_{\max}$.   On the  contrary,  with increasing
1581: $\kappa$, $\lambda$  and $\rho$, the inflationary  scale $M$ increases
1582: and the parameter $\Delta_{\rm  exit}$ becomes larger. We have checked
1583: that the  inequality $\phi_{\max}>\phi_{\rm exit}$  is fulfilled along
1584: the lines presented in  Fig.~\ref{fig:nmSUGRAPR}.  In this respect, we
1585: also  note  that $\phi_{\rm  min}$  is  in  general much  larger  than
1586: $\phi_{\max}$ especially for low values of $n_{\rm s}$.
1587: 
1588: 
1589: It  is important  to observe  from  Table~\ref{tab2} that  there is  a
1590: degree  of  tuning required  for  the  values  $\phi_{\rm exit}$  with
1591: respect to $\phi_{\max}$.  For values of $\kappa \stackrel{>}{{}_\sim}
1592: 10^{-3}$,  we find  that the  degree of  tuning required  is  not very
1593: serious,   i.e.~$\Delta_{\rm    exit}   \stackrel{>}{{}_\sim}   10\%$.
1594: However,  the  situation  becomes  rather delicate  as  $\kappa$  gets
1595: smaller  than $10^{-3}$, for  $n_{\rm s}  \stackrel{<}{{}_\sim} 0.97$.
1596: In  this case,  we find  that $\phi_{\max}  \approx  \phi_{\rm exit}$,
1597: leading  to a  substantial tuning  at the  few per  cent level  in the
1598: initial conditions of inflation.
1599: 
1600: 
1601: As in the mSUGRA case, we also show in Fig.~\ref{fig:nmSUGRAPR}(a) the
1602: upper bounds resulting  from cosmic-string effects [cf.~(\ref{Prcs})],
1603: for $y_{\rm cs}  = 6.7,~8.9,~11.6$ (from top to  bottom). As mentioned
1604: above,   these  constraints   are   only  relevant   for  an   Abelian
1605: waterfall-gauge sector  with dimensionality  ${\cal N} =  1$. However,
1606: unlike  in the mSUGRA  case, these  restrictions appear  less harmful,
1607: since  the  inflationary scale  $M$  assumes  smaller  values
1608: (see also Table~\ref{tab2}) and  the tadpole term becomes
1609: unimportant. Thus, larger values of $\kappa$ up to order $10^{-2}$ can be
1610: tolerated in this case.  For the non-Abelian SU(2)$_X$   $F_D$-term    hybrid
1611: model,   the    restrictions   from considerations  of cosmic-string
1612: effects are  totally lifted  and the
1613: lines depicted  in Fig.~\ref{fig:nmSUGRAPR}  only vary within  the few
1614: per  cent level.  Such  a variation  becomes even  smaller if  $\rho >
1615: \kappa$ and/or $\lambda > \kappa$.
1616: 
1617: 
1618: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1619: \begin{figure}[!t]
1620: \centerline{\epsfig{file=cHk.eps,angle=-90,width=13.cm}} \hfill
1621: \caption{\sl\small The parameter values $(\kappa,c_H)$ allowed by
1622: (\ref{Nefold}), (\ref{Pr}) and (\ref{nswmap}) in the nmSUGRA scenario,
1623: for $\rho=\lambda=\kappa$ (light grey hatched area) and
1624: $\rho=\lambda=4\kappa$ (grey hatched area). The grey (light grey) line
1625: has been obtained by fixing $n_{\rm s}$ to its central value given in
1626: (\ref{nswmap}), for $\rho=\lambda=4\kappa$
1627: ($\rho=\lambda=\kappa$).}\label{fig:cHk}
1628: \end{figure}
1629: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1630: 
1631: In Fig.~\ref{fig:cHk},  we present the  parameter space
1632: $(\kappa,c_H)$ which  is  allowed by  the
1633: conditions~(\ref{Nefold}), (\ref{Pr})  and (\ref{nswmap}) in the
1634: nmSUGRA  scenario. The light grey (grey) hatched area   indicates
1635: the   allowed   region  for   $\rho=\lambda=4\kappa$
1636: ($\rho=\lambda=\kappa$). The lower (upper) boundaries of the
1637: allowed regions correspond to the upper (lower) bound on $n_{\rm
1638: s}$, cf.~(\ref{ns95}), while the solid lines correspond to the
1639: central value of $n_{\rm s}$, cf.~(\ref{nswmap}). We find that
1640: values of $c_H \sim  0.2$ and $\kappa \sim  0.05$ are still
1641: possible  in a nmSUGRA  extension of the $F_D$-term hybrid model.
1642: 
1643: 
1644: 
1645: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1646: \begin{figure}[!t]
1647: \centerline{\epsfig{file=nmSUGRAlr.eps,angle=-90,width=13.cm}} \hfill
1648: \caption{\sl\small The allowed values of $\rho/\kappa$ versus $\lambda/\kappa$
1649: for the nmSUGRA scenario with $\kappa=0.005,~M=10^{16}~{\rm GeV}$ and
1650: $c_H=0.18$ (dark grey line), $c_H=0.14$ (grey line) or
1651: $c_H=0.07$ (light grey line).}\label{fig:nmSUGRAlr}
1652: \end{figure}
1653: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1654: 
1655: 
1656: Finally,  we plot  in Fig.~\ref{fig:nmSUGRAlr}  the allowed  values of
1657: $\lambda/\kappa$ versus $\rho/\kappa$,  on account of the inflationary
1658: constraints~(\ref{Nefold})  and (\ref{Pr}),  for $\kappa=0.005$,  $M =
1659: 10^{16}$~GeV, and  for $c_H=0.18$  (dark grey line),  $c_H=0.14$ (grey
1660: line) and  $c_H=0.07$ (light grey  line). We have selected  a slightly
1661: lower value for $M$, because no viable nmSUGRA scenarios seem to exist
1662: with acceptable values  for $n_{\rm s}$, if $M  = 2\times 10^{16}$~GeV
1663: and    $c_H    \ge    0.07$.     Along   the    contour    lines    in
1664: Fig.~\ref{fig:nmSUGRAlr},   $\phi_{\rm    min}$,   $\phi_{\rm   max}$,
1665: $\phi_{\rm exit}$  and $n_{\rm  s}$ remain constant  and equal  to the
1666: values  presented  in  Table~\ref{tab1}.    We  observe  that  as  $c_H$
1667: increases, $\phi_{\rm  exit}$ approaches $\phi_{\rm  max}$, $\phi_{\rm
1668: min}$  increases,   while  $n_{\rm  s}$   decreases.   This  kinematic
1669: behaviour    is   in   agreement    with   our    discussion   related
1670: to~(\ref{phimax}) and~(\ref{nShilltop}).
1671: 
1672: 
1673: 
1674: \setcounter{equation}{0}
1675: \section{Preheating}\label{Preheat}
1676: 
1677: As  stated  in the  Introduction,  gravitinos,  if thermally  produced
1678: during the early  stages of the evolution of  the Universe, will spoil
1679: the  successful predictions  of  BBN~\cite{Sarkar}.  Their  disastrous
1680: consequences may  be avoided, if the reheat  temperature $T_{\rm reh}$
1681: of the  Universe is  not very  high. In fact,  depending on  the decay
1682: properties   of   the   gravitino,   it   should   be   $T_{\rm   reh}
1683: \stackrel{<}{{}_\sim}  10^7$--$10^{10}$~GeV~\cite{kohri,oliveg}.  This
1684: fact leads to a tension between the allowed range of $T_{\rm reh}$ and
1685: the natural  scale of  hybrid inflation $M$,  which is of  order $\sim
1686: 10^{16}$~GeV. The traditional way taken  to get around this problem is
1687: to  consider scenarios  where the  decay rate  of the  inflaton  to SM
1688: particles  is extremely suppressed,  e.g.~by suppressing  all possible
1689: couplings of the inflaton to the SM fields.
1690: 
1691: 
1692: In  this  and next  sections,  we  present  in detail  an  alternative
1693: solution to the  above gravitino overabundance problem~\cite{GP}.  Our
1694: solution  relies on  the huge  entropy  release caused  from the  late
1695: out-of-equilibrium decays of the supermassive waterfall particles. The
1696: entropy produced  through this mechanism  is sufficient to  reduce the
1697: gravitino abundance $Y_{\widetilde{G}}$  to levels compatible with BBN
1698: limits     discussed     in     detail    in     Section~\ref{reheat}.
1699: Figure~\ref{figure:cartoon}  gives a  schematic representation  of the
1700: post-inflationary dynamics  of the early Universe, as  is predicted by
1701: the $F_D$-term hybrid model.  Shortly after inflation ends, the energy
1702: density  $\rho_\kappa$  of the  Universe  is  predominantly stored  to
1703: coherently oscillating  inflaton condensates which  scale as $a^{-3}$,
1704: where  $a$  is the  usual  cosmological  scale  factor describing  the
1705: expansion  of   the  Universe.   The  coherent   oscillations  of  the
1706: inflaton-related    condensates   also    give    rise   to    another
1707: non-perturbative  mechanism  called  preheating.   During  preheating,
1708: waterfall  gauge particles  of  energy density  $\rho_g$ are  produced
1709: almost instantaneously,  which are absolutely stable  if a $D$-parity,
1710: an analogue of the usual  $R$-parity in the MSSM, is conserved.  Then,
1711: the   following  scenario   visualized   in  Fig.~\ref{figure:cartoon}
1712: emerges.   First,  $\rho_g/\rho_\kappa$  remains constant  during  the
1713: epoch of coherent oscillations,  since both $\rho_g$ and $\rho_\kappa$
1714: behave as  matter energy densities  and scale as $a^{-3}$  during this
1715: period. The constancy of $\rho_g/\rho_\kappa$ ceases to hold, when the
1716: coherently  oscillating inflaton  condensates decay  and  their energy
1717: density  $\rho_\kappa$  gets   distributed  among  light  relativistic
1718: degrees  of freedom.  As  a consequence  of the  latter, $\rho_\kappa$
1719: will  be $\propto  a^{-4}$,  whilst $\rho_g$  will  still be  $\propto
1720: a^{-3}$.   If the initial  value of  $\rho_g/\rho_\kappa$ is  not very
1721: suppressed,  e.g.~it is of  order $10^{-4}$--$10^{-5}$,  the waterfall
1722: gauge  particles will eventually  dominate the  energy density  of the
1723: Universe, leading to  a second matter dominated epoch  which will last
1724: until these particles decay  via $D$-parity violating couplings.  This
1725: is expected to  produce an enormous entropy release  and so reduce the
1726: gravitino-to-entropy  ratio $Y_{\widetilde{G}}$  to  values compatible
1727: with BBN constraints.
1728: 
1729: 
1730: \begin{figure}[t]
1731: \begin{center}
1732: \epsfig{file=cartoon.eps, height=3.5in,width=4in}
1733: \end{center}
1734: \caption{\sl\small Schematic representation of the thermal history of
1735: the Universe in the $F_D$-term hybrid model.}\label{figure:cartoon}
1736: \end{figure}
1737: 
1738: The  discussion   in  this  section   is  organized  as   follows:  in
1739: Section~\ref{postinfl},  we pay  special attention  to  $D$-parity and
1740: derive the particle spectrum of the combined inflaton-waterfall sector
1741: in the  supersymmetric limit of  the theory.  In addition,  we compute
1742: the decay rates of all inflaton-related and waterfall gauge particles.
1743: Finally, in Section~\ref{preheat}, we  discuss how the waterfall gauge
1744: particles   are  instantaneously   produced  through   preheating  and
1745: calculate  the  resulting energy  density  $\rho_g$  carried by  these
1746: particles.
1747: 
1748: 
1749: 
1750: \subsection{{\boldmath $D$}-Parities and the
1751: Inflaton-Waterfall Sector}\label{postinfl}
1752: 
1753: Let  us  first  consider  a  model  with  a  U(1)$_X$  gauge-symmetric
1754: waterfall  sector.   The  case  of  a  waterfall  sector  realizing  a
1755: non-Abelian  SU(2)$_X$  gauge  symmetry   is  analogous  and  will  be
1756: discussed later.  In terms  of superfields, the  minimal gauge-kinetic
1757: Lagrangian of the U(1)$_X$ model reads:
1758: \begin{equation}
1759:   \label{U1kin}
1760: {\cal L}_{\rm kin}\ =\
1761: \int d^4 \theta\, \left(
1762: \frac{1}{2}\, W^\alpha W_\alpha\,\delta^{(2)}(\bar{\theta} )\ +\
1763: \frac{1}{2}\, \overline{W}_{\dot{\alpha}}
1764: \overline{W}^{\dot{\alpha}}\, \delta^{(2)}(\theta )\  +\
1765: \widehat X_1^\dagger {\rm e}^{2 g \widehat V _X} \widehat X_1\
1766: +\ \widehat X_2^\dagger
1767: {\rm e}^{-2 g \widehat{V}_X} \widehat X_2 \right)\; ,
1768: \end{equation}
1769: where $\widehat{V}_X$ is the U(1)$_X$ vector superfield and $W_\alpha$
1770: ($\overline{W}_{\dot{\alpha}}$)    are    their   respective    chiral
1771: (anti-chiral) field strengths. The latter are given by
1772: \begin{equation}
1773: W_\alpha\ =\ -\, \frac{1}{8g}\, \bar{D}^2\, ( e^{-2g\widehat{V}_X} D_\alpha\,
1774: e^{2g\widehat{V}_X} )\,,\qquad
1775: \overline{W}_{\dot{\alpha}}\ =\
1776: \frac{1}{8g}\, D^2\, ( e^{2g\widehat{V}_X}
1777: \bar{D}_{\dot{\alpha}}\, e^{-2g\widehat{V}_X} )\,,
1778: \end{equation}
1779: where   $D_\alpha$   and   $\bar{D}_{\dot{\alpha}}$  are   the   usual
1780: SUSY-covariant  derivatives which  are irrelevant  for  our discussion
1781: here.   The minimal  gauge-kinetic  Lagrangian~(\ref{U1kin}) possesses
1782: the discrete symmetry
1783: \begin{equation}
1784:   \label{DsuperU1}
1785: D:\qquad \widehat{X}_1\ \leftrightarrow\ \widehat{X}_2\,,\qquad
1786: \widehat{V}_X\ \to\ -\,\widehat{V}_X\; ,
1787: \end{equation}
1788: whereas all other superfields do not transform. It is not difficult to
1789: verify  that  the  complete  $F_D$-term hybrid  model,  including  the
1790: superpotential~(\ref{Wmodel})  and its  associated  soft SUSY-breaking
1791: sector, is  invariant under the  discrete symmetry~(\ref{DsuperU1}) in
1792: the  unbroken phase  of the  theory. After  the SSB  of  U(1)$_X$, the
1793: waterfall fields receive the VEVs:  $\langle X_1 \rangle = \langle X_2
1794: \rangle = M$.  Thus, the  above discrete symmetry survives even in the
1795: spontaneously broken phase of the theory.  Since the discrete symmetry
1796: acts on  a gauged waterfall sector,  it manifests itself as  a kind of
1797: parity, which we call $D$-parity.
1798: 
1799: It therefore proves convenient to choose a weak basis where the fields
1800: are  eigenstates of  $D$-parity.  To  this end,  we define  the linear
1801: combinations in terms of the waterfall superfields
1802: \begin{equation}
1803:   \label{Xparity}
1804: \widehat{X}_\pm \ =\ \frac{1}{\sqrt{2}}\, \bigg(\, \widehat{X}_1\
1805: \pm\ \widehat{X}_2\,\bigg)\; .
1806: \end{equation}
1807: Evidently, the  superfield $\widehat{X}_+$ ($\widehat{X}_-$)  has even
1808: (odd) $D$-parity;  its $D$-parity quantum number is  $+1$ ($-1$).  The
1809: vector  superfield  $\widehat{V}_X$,  which  is already  a  $D$-parity
1810: eigenstate, has  odd $D$-parity.  All remaining  fields, including the
1811: inflaton superfield $\widehat{S}$ and the other MSSM superfields, have
1812: positive $D$-parity.
1813: 
1814: As  a consequence  of $D$-parity  conservation, all  $D$-odd particles
1815: will  be stable,  in as  much  the same  way as  the usual  $R$-parity
1816: guarantees  that the  LSP of  the MSSM  is stable.   As  we explicitly
1817: mentioned in Section~\ref{FD}, the simplest way to break $D$-parity is
1818: to  add  a FI  $D$-term  to  the  model, e.g.
1819: \begin{equation}
1820:   \label{LFI}
1821: {\cal L}_{\rm FI}\ =\ -\, \frac{g}{2}\, m^2_{\rm FI}\,
1822:                        \int d^4 \theta\; \widehat{V}_X\ =\ -\,
1823:                        \frac{g}{2}\, m^2_{\rm FI}\, D\; ,
1824: \end{equation}
1825: where  $D$  is  the   auxiliary  component  of  the  vector  superfield
1826: $\widehat{V}_X$.  It  is obvious that  ${\cal L}_{\rm FI}$  flips sign
1827: under  the  discrete  symmetry~(\ref{DsuperU1}). Other  mechanisms  of
1828: explicitly      breaking     $D$-parity      are      discussed     in
1829: Appendix~\ref{Dappendix}.
1830: 
1831: We now calculate the particle spectrum of the inflaton-waterfall sector
1832: in the presence  of a subdominant FI $D$-term $m_{\rm  FI}$ and in the
1833: supersymmetric  limit of  the theory.   With this  aim, we  expand the
1834: scalar $D$-parity eigenstates $X_\pm$ about their VEVs:
1835: \begin{equation}
1836:   \label{Xpm}
1837: X_\pm \ =\ \langle X_\pm \rangle\: +\:  \frac{1}{\sqrt{2}}\,
1838: \Big(\, R_\pm\: +\: {\rm i}I_\pm\,\Big)\; .
1839: \end{equation}
1840: The VEVs $\langle X_\pm \rangle$ are determined from the minimization
1841: conditions of the combined $F$- and $D$-term scalar potential
1842: \begin{eqnarray}
1843:   \label{VFD}
1844: V_{FD}\ =\  F_S^* F_S\: +\: \frac 12\, D^2\,,
1845: \end{eqnarray}
1846: where
1847: \begin{equation}
1848: F_S \ =\ \frac{\kappa}{2}\, \Big(\, X^2_+\: -\: X^2_-\:
1849: -\: 2\,M^2\,\Big)\,,\qquad
1850: D \ =\ \frac{g}{2}\, \Big(\,
1851: X^*_+ X_-\: +\: X^*_- X_+\: -\: m^2_{\rm FI}\, \Big)\; .
1852: \end{equation}
1853: Since  SUSY  is  preserved  after  the SSB  of  U(1)$_X$,  the  scalar
1854: potential  $V_{FD}$ will  vanish  at its  ground state,  i.e.~$\langle
1855: V_{\rm FD}  \rangle =  0$. Consequently, to  leading order  in $m_{\rm
1856: FI}/M$, the VEVs of the scalar inflaton-waterfall fields are
1857: \begin{equation}
1858:   \label{Xdec}
1859: \langle S\rangle\ =\ 0\,,\qquad \langle X_+\rangle\ =\ \sqrt{2}\,M \,,\qquad
1860: \langle X_-\rangle \ =\ \frac{v}{\sqrt{2}} \ ,
1861: \end{equation}
1862: where $v  = m^2_{\rm FI}/(2M)$. Notice  that the VEVs of  the $F$- and
1863: $D$-terms   vanish   through    order   $m_{\rm   FI}/M$   considered,
1864: i.e.~$\langle  D \rangle  = 0$  and $\langle  F_S \rangle  =  {\cal O}
1865: (m^4_{\rm FI}/M^2)$.
1866: 
1867: To derive the mass spectrum,  we expand the potential about its ground
1868: state  up to terms  quadratic in  all the  fields involved.   We first
1869: consider the $F$-terms.  To order $v/M\ (= m^2_{\rm FI}/M^2)$, we find
1870: the approximate mass eigenstates:
1871: \begin{equation}
1872: S\ =\ \frac{1}{\sqrt{2}}\; \Big(\, \phi\: +\: {\rm i} a\, \Big)\,,\qquad
1873: R_+\: -\: \frac{v}{2M}\,R_-\,,\qquad I_+\: -\: \frac{v}{2M}\, I_-\ .
1874: \end{equation}
1875: All the  above fields, consisting of  4 bosonic degrees  of freedom in
1876: total, share the common mass
1877: \begin{equation}
1878:   \label{Mkappa}
1879: m_\kappa\ =\ \sqrt 2\, \kappa M\; .
1880: \end{equation}
1881: As a  consequence of  SUSY, the corresponding  4 fermionic  degrees of
1882: freedom form  a Dirac  spinor $\psi_\kappa$, which  also has  the same
1883: mass~(\ref{Mkappa}).  We refer  to these particles as inflaton-related
1884: or $\kappa$-sector particles.
1885: 
1886: 
1887: \begin{table}[t]
1888: \begin{center}
1889: \begin{tabular}{|c|c|c|c|}
1890: \hline
1891: % & & & \\
1892: Sector & Boson & Fermion & Mass\\
1893: % & & & \\
1894: \hline\hline
1895: % & & & \\
1896: \begin{tabular}{l}
1897: Inflaton\\
1898: ($\kappa$-sector)\\
1899: $D$-parity: $+1$
1900: \end{tabular}
1901: &
1902: \begin{tabular}{l}
1903: $S\,$,\\
1904: $R_+-\frac{v}{2M}R_-\,$,\\
1905: $I_+-\frac{v}{2M}I_-$
1906: \end{tabular}
1907: &
1908: $\psi_\kappa=
1909: \left(
1910: \begin{array}{c}
1911: \psi_{X_+} - \frac{v}{2M}\,\psi_{X_-}\\
1912: \psi_S^\dagger
1913: \end{array}
1914: \right)^{\phantom{X}}
1915: $
1916: &
1917: $\sqrt 2 \kappa M$
1918: \\
1919: \hline
1920: \begin{tabular}{l}
1921: U(1)$_X$\\ Waterfall Gauge\\
1922: ($g$-sector)\\
1923: $D$-parity: $-1$
1924: \end{tabular}
1925: &
1926: \begin{tabular}{l}
1927: $V_\mu\, [I_-+\frac v{2M}I_+]\,$,\\
1928: $R_-+\frac v{2M}R_+$
1929: \end{tabular}
1930: &
1931: $\psi_g=
1932: \left(
1933: \begin{array}{c}
1934: \psi_{X_-} + \frac{v}{2M}\,\psi_{X_+}\\
1935: -{\rm i}\lambda^\dagger
1936: \end{array}
1937: \right)^{\phantom{X}}
1938: $
1939: &
1940: $g M$
1941: \\
1942: \hline
1943: \end{tabular}
1944: \end{center}
1945: \caption{\sl\small Particle spectrum of the inflaton and the ${\rm
1946: U(1)}_X$ waterfall-gauge sectors after inflation, where the
1947: approximate $D$-parity for each sector is displayed.  The field
1948: $V_\mu$ denotes the ${\rm U(1)}_X$ gauge boson and $\lambda$~its
1949: associate gaugino.  The would-be Goldstone boson related to the
1950: longitudinal degree of $V_\mu$ appears in the square
1951: brackets.}\label{spectrum}
1952: \end{table}
1953: 
1954: 
1955: The remaining scalar fields receive  their masses from the $D$-term of
1956: the scalar potential  $V_{FD}$ in~(\ref{VFD}). Performing an analogous
1957: calculation as  outlined above,  we obtain to  order $v/M$  the scalar
1958: mass eigenstates:
1959: \begin{equation}
1960: I_-\: +\: \frac{v}{2M}\, I_+\,,\qquad R_-\: +\: \frac{v}{2M}\, R_+\ .
1961: \end{equation}
1962: The  first field  is absorbed  by  the longitudinal  component of  the
1963: U(1)$_X$  gauge  field  $V_\mu$,  via  the Higgs  mechanism.   In  the
1964: supersymmetric  limit,  all  these  fields, which  mediate  4  bosonic
1965: degrees  of freedom, are  degenerate and  characterized by  the common
1966: mass
1967: \begin{equation}
1968:   \label{Mg}
1969: m_g\ =\ g\, M\ .
1970: \end{equation}
1971: Like in  the $\kappa$-sector case, the respective  4 fermionic degrees
1972: of freedom will  make up a 4-component Dirac spinor  of mass $m_g$. We
1973: refer  to this  group of  particles as  waterfall gauge  or $g$-sector
1974: particles.  In  Table~\ref{spectrum}, we present a summary  of all the
1975: inflaton-related  ($\kappa$-sector)  and waterfall-gauge  ($g$-sector)
1976: particles. As  can also been seen from  the same Table~\ref{spectrum},
1977: $\kappa$-sector  particles  are  predominantly $D$-even,  whereas  the
1978: $g$-sector ones have approximately $D$-odd parity.
1979: 
1980: 
1981: It is  now interesting to calculate  the decay rates  of the $\kappa$-
1982: and $g$-sector particles and analyze their implications for the reheat
1983: temperature of the Universe.  Starting  with the singlet field $S$, it
1984: decays  predominantly into  pairs  of charged  and neutral  higgsinos,
1985: $\tilde{h}^\pm_{u,d}$, $\tilde{h}^0_{u,d}$, $\tilde{\bar{h}}^0_{u,d}$,
1986: and  into pairs of  right-handed Majorana  neutrinos $\nu_{1,2,3\,R}$.
1987: On  the  other  hand, the  scalars  $R_+$  and  $I_+$ decay  into  the
1988: SUSY-conjugate partners of the aforementioned fields at the same rate.
1989: In fact, we  find a common decay rate for  each of the $\kappa$-sector
1990: particles:
1991: \begin{equation}
1992:   \label{infldecay}
1993: \Gamma_\kappa\ =\ \frac{1}{32\pi}\:  \Big(\, 4\lambda^2\: +\: 3 \rho^2\,
1994: \Big)\: m_\kappa\; .
1995: \end{equation}
1996: The  reheat temperature $T_\kappa$  resulting from  these perturbative
1997: decays  of the $\kappa$-sector  particles may  be estimated  using the
1998: relation $\Gamma_\kappa  = H (T_\kappa )$, where  the Hubble parameter
1999: $H(T)$ is  given in the radiation  dominated era of  the Universe.  In
2000: this way, we obtain
2001: \begin{equation}
2002:   \label{Tkappa}
2003: T_\kappa\ =\ \left( \frac{90}{\pi^2\, g_*}\right)^{1/4}\,
2004: \sqrt{\Gamma_\kappa\: m_{\rm Pl} }\ ,
2005: \end{equation}
2006: where $g_* = 240$ is the number of the relativistic degrees of freedom
2007: in   the  $F_D$-term  hybrid   model.   Substituting~(\ref{infldecay})
2008: and~(\ref{Mkappa}) into~(\ref{Tkappa}), we arrive at the expression:
2009: \begin{equation}
2010: T_\kappa\ =\ 8.1 \cdot
2011: 10^{15}~{\rm GeV} \times
2012: \Big[\kappa (4\lambda^2+3 \rho^2)\Big]^{1/2}
2013: \left(\frac M{10^{16}{\rm GeV}}\right)^{1/2} \,.
2014: \end{equation}
2015: Assuming that  no relevant  amount of entropy  is released  during the
2016: subsequent thermal  history of the Universe,  the gravitino constraint
2017: on  the reheat  temperature $T_\kappa  \stackrel{<}{{}_\sim} 10^9$~GeV
2018: requires that each individual  coupling $\kappa$, $\lambda$ and $\rho$
2019: must be smaller than about $10^{-5}$, if $M \sim 10^{16}$~GeV. Further
2020: details are given in Section~\ref{reheat}.
2021: 
2022: The above unnatural tuning of  all inflaton couplings to SM fields may
2023: be avoided, if  the large entropy release from the  late decays of the
2024: $g$-sector particles is  properly considered.  An extensive discussion
2025: of  this  issue is  given  in  Section~\ref{reheat}.  Here, we  simply
2026: compute the decay rates of  the $g$-sector particles which are induced
2027: by  a  non-vanishing FI  term  $m_{\rm  FI}$.  In~fact,  the  relevant
2028: interaction Lagrangian is given by
2029: \begin{equation}
2030:   \label{Lint}
2031: {\cal L}_{\rm int}\  =\ \frac{g^2 m^2_{\rm FI}}{8 M}\;  R_-\, (R_+^2 +
2032: I_+^2)\; .
2033: \end{equation}
2034: As  mentioned  above, this  induces  a  decay  width for  the  $D$-odd
2035: particle $R_-$, which is easily calculated to be
2036: \begin{equation}
2037:   \label{GammaR}
2038: \Gamma_g\ =\ \frac{g^3}{128 \pi}\, \frac{m^4_{\rm FI}}{M^3}\,.
2039: \end{equation}
2040: In close  analogy with  the $\kappa$-sector, each  $g$-sector particle
2041: decay rate is equal to $\Gamma_g$.
2042: 
2043: 
2044: Let  us now  consider a  model with  a waterfall  sector based  on the
2045: SU(2)$_X$  gauge  group. As  was  mentioned  in Section~\ref{FD},  the
2046: waterfall superfields  $\widehat{X}_1$ and $\widehat{X}_2$  are chosen
2047: to   belong  in  this   case  to   the  2-component   fundamental  and
2048: anti-fundamental representations of SU(2)$_X$, respectively.  Although
2049: the  two  representations  are   equivalent  for  the  SU(2)  case,  we
2050: nevertheless  use this  convention,  such that  its generalization  to
2051: ${\rm   SU}(N)$  theories,  with   $N>2$,  is   straightforward.   The
2052: superpotential is almost identical to the one given in~(\ref{Wmodel}),
2053: with  the obvious  substitution: $\widehat{X}_{1}  \widehat{X}_{2} \to
2054: \widehat{X}_{1}^T  \widehat{X}_{2}$.   Extending~(\ref{U1kin}) to  the
2055: SU(2)$_X$ case, the minimal gauge-kinetic Lagrangian is written down
2056: \begin{eqnarray}
2057:   \label{SU2kin}
2058: {\cal L}_{\rm kin} &=&
2059: \int d^4 \theta\ \Bigg[\,
2060: \frac{1}{2}\, {\rm Tr}\,(W^\alpha W_\alpha)\,\delta^{(2)}(\bar{\theta} )\ +\
2061: \frac{1}{2}\, {\rm Tr}\,(\overline{W}_{\dot{\alpha}}
2062: \overline{W}^{\dot{\alpha}})\, \delta^{(2)}(\theta )\nonumber\\
2063: &&  +\
2064: \widehat X_1^\dagger {\rm e}^{2 g \widehat V _X} \widehat X_1\
2065: +\ \widehat X_2^\dagger
2066: {\rm e}^{-2 g \widehat{V}^T_X} \widehat X_2\, \Bigg]\; .
2067: \end{eqnarray}
2068: In the above, $\widehat{V}_X = \widehat{V}^a_X\, T^a$ is the SU(2)$_X$
2069: vector    superfield    and     $W_\alpha    =    W^a_\alpha\,    T^a$
2070: ($\overline{W}_{\dot{\alpha}}  = \overline{W}^a_{\dot{\alpha}}\, T^a$)
2071: are the corresponding non-Abelian chiral (anti-chiral) field strengths
2072: in the  so-called Wess--Zumino~(WZ)  gauge.  The superscript  `$T$' on
2073: $\widehat{V}_X$, i.e.~$\widehat{V}^T_X$,  indicates transposition that
2074: acts on  the generators $T^a  = \frac 12  \tau^a$ of the  SU(2) group,
2075: where $\tau^{1,2,3}$ are the usual Pauli matrices.  Finally, the trace
2076: in~(\ref{SU2kin}) is understood to be taken over the group space.
2077: 
2078: 
2079: The minimal SU(2)$_X$ gauge-kinetic Lagrangian is invariant under the
2080: discrete transformations
2081: \begin{equation}
2082:   \label{D1}
2083: D_1:\qquad    \widehat{X}_1\    \leftrightarrow\    \widehat{X}_2\,,\qquad
2084: \widehat{V}_X\ \to\ -\widehat{V}^T_X\; .
2085: \end{equation}
2086: Notice  that  under  the  action  of $D_1$  in~(\ref{D1}),  the  field
2087: strengths   transform   as:   $W_\alpha   \to  -   (W_\alpha)^T$   and
2088: $\overline{W}_{\dot{\alpha}} \to - (\overline{W}_{\dot{\alpha}})^T$ in
2089: {\em  any}  SUSY  gauge,  including   the  WZ  gauge.   If  all  other
2090: superfields  do   not  transform,  the  complete   Lagrangian  of  the
2091: non-Abelian  $F_D$-term  hybrid  model  will be  invariant  under  the
2092: discrete  transformation~(\ref{D1})  in  the  unbroken  phase  of  the
2093: theory.
2094: 
2095: Our discussion so far has made no reference to the specific properties
2096: of  SU(2)$_X$ and  so applies  equally well  to any  SU($N>2$) theory.
2097: However, in  the SU(2)$_X$ case, the $F_D$-term  hybrid model exhibits
2098: an  additional Abelian  or diagonal  discrete symmetry.   This  may be
2099: defined by
2100: \begin{equation}
2101:   \label{D2}
2102: D_2:\qquad \widehat{X}_1\ \to\ \tau^3 \widehat{X}_1\,,\qquad
2103: \widehat{X}_2\ \to\ \tau^3 \widehat{X}_2\,,\qquad
2104: \widehat{V}_X\ \to\ \tau^3\,\widehat{V}_X\,\tau^3\; ,
2105: \end{equation}
2106: whereas all other superfields do not transform. It is then easy to see
2107: that  (\ref{D2}) implies:  $W_\alpha \to  \tau^3\,W_\alpha\,\tau^3$ in
2108: any SUSY gauge  and likewise for $\overline{W}_{\dot{\alpha}}$.  Since
2109: $\tau^3 = (\tau^3)^T$ and $(\tau^3)^2  = {\bf 1}_2$, the invariance of
2110: the Lagrangian~(\ref{SU2kin}) and of  the whole model under the action
2111: of~$D_2$ is evident.\footnote{In general, for a waterfall-gauge sector
2112: based  on  an  SU($N>2$)   group,  there  are  $N$  distinct  discrete
2113: symmetries.   The first  is given  by~(\ref{D1}), while  the remaining
2114: $N-1$  symmetries result  from  replacing $\tau^3$  with  $D_n =  {\rm
2115: diag}\, (1,1,\dots,1,  -1,1,\dots, 1)$. The  entry $-1$ occurs  at the
2116: $n$ position  of the $N$-dimensional  diagonal matrix $D_n$,  with the
2117: restriction $1< n \leq N$.  Obviously,  it is $D_n = D^T_n$ and $D^2_n
2118: =  {\bf  1}_N$.  These  discrete  symmetries  are  non-Abelian in  the
2119: adjoint group space, in the sense that the eigenvalue matrix~$c^{ab}$,
2120: determined by means of the relation $D_n T^a D_n = c^{ab} T^b$, is not
2121: diagonal.}
2122: 
2123: We now  proceed to  compute the particle  spectrum of  the non-Abelian
2124: $F_D$-term hybrid model after the  SSB of SU(2)$_X$. For this purpose,
2125: it is useful to introduce the notation
2126: \begin{equation}
2127: Z\ =\
2128: \left(\begin{array}{c}
2129: ^+\!Z\\
2130: ^-\!Z
2131: \end{array}\right)
2132: \,,
2133: \end{equation}
2134: where  $Z$  is  a  generic  ${\rm  SU}(2)_X$-doublet  or  anti-doublet
2135: (conjugate)  field.  The left  superscripts~$\pm$  on  $Z$ denote  the
2136: eigenvalues of  the discrete  symmetry transformation operator  $D_2 =
2137: \tau^3$ defined  in~(\ref{D2}), and they  should not be  confused with
2138: the  corresponding eigenvalues of  the isospin  operator $T^3$  of the
2139: SU(2)$_X$  group.  In  the unitary  gauge, the  minimum of  the scalar
2140: potential occurs for the field values
2141: \begin{equation}
2142: \label{SU2:VEV}
2143: ^+\!X_1\ =\ ^+\!X_2\ =\ M\,,\qquad ^-\!X_1\ =\ ^-\!X_2\ =\ 0\;.
2144: \end{equation}
2145: Consequently,   the  discrete   symmetries  $D_1$   and   $D_2$  given
2146: in~(\ref{D1})  and~(\ref{D2})  remain  intact  after the  SSB  of  the
2147: SU(2)$_X$ gauge  group. Since they  act on a gauged  waterfall sector,
2148: they  are  actually  parities.   We   refer  to  them  as  $D_1$-  and
2149: $D_2$-parities, or collectively as $D$-parities.
2150: 
2151: Analogously to  the U(1)$_X$ case,  we express the  SU(2)$_X$ doublets
2152: $X_{1,2}$   in  terms   of  eigenstates   of   the  $D_{1,2}$-parities
2153: [cf.~(\ref{Xparity})]. In terms of  their components, these fields may
2154: be conveniently expressed as follows:
2155: \begin{equation}
2156: ^\pm\!X_\pm \ =\
2157: \langle X_\pm \rangle\: +\: \frac{1}{\sqrt{2}}\,
2158: \Big(\, ^\pm\!R_\pm\ +\ {\rm i}\,^\pm\!I_\pm\, \Big)\; ,
2159: \end{equation}
2160: with $\langle X_+  \rangle = \sqrt{2}\, M$ and  $\langle X_- \rangle =
2161: 0$ in the absence of  any $D$-parity violating coupling in the theory.
2162: Moreover, the SU(2)$_X$ $D$-terms are given by
2163: \begin{equation}
2164: D^a\ =\ \frac g2 \left(
2165: X_1^\dagger \tau^a X_1\: -\: X_2^T \tau^a X_2^* \right)\,.
2166: \end{equation}
2167: In the $D$-parity eigenbasis~(\ref{Xparity}), they take on the form
2168: \begin{equation}
2169: D^a\ =\ \frac g2 \times
2170: \left\{
2171: \begin{array}{l}
2172: X_+^\dagger \tau^a  X_- + X_-^\dagger \tau^a X_+\,, \quad
2173:                 \textnormal{for }\tau^a \textnormal{ symmetric } (a=1,3)\\
2174: X_+^\dagger \tau^a X_+ + X_-^\dagger \tau^a X_-\,, \quad
2175:                 \textnormal{for }\tau^a \textnormal{ antisymmetric } (a=2)
2176: \end{array}
2177: \right.
2178: \;.
2179: \end{equation}
2180: Exactly as in the U(1)$_X$ case,  we find that there are two groups of
2181: mass-degenerate   fields,  $\kappa$-   and  $g$-sector,   with  masses
2182: $m_\kappa$   and  $m_g$  given   in  (\ref{Mkappa})   and  (\ref{Mg}),
2183: respectively.   The complete  inflaton-waterfall spectrum,  along with
2184: their    $D_1$     and    $D_2$    parities,     is    exhibited    in
2185: Table~\ref{spectrum:SU2}.
2186: 
2187: 
2188: \begin{table}
2189: \begin{center}
2190: \begin{tabular}{|c|c|c|c|c|c|}
2191: \hline
2192: % & & & \\
2193: Sector & Boson & Fermion & Mass & $\,D_1\mbox{-parity}\!\!$
2194: & $\,D_2 \textnormal{-parity}\!\!$
2195: \\
2196: % & & & \\
2197: \hline\hline
2198: % & & & \\
2199: \begin{tabular}{l}
2200: Inflaton\\
2201: ($\kappa$-sector)
2202: \end{tabular}
2203: &
2204: $S\,$,
2205: $\!^+\!R_+$,
2206: $\!^+I_+$
2207: &
2208: $\psi_\kappa=
2209: \left(
2210: \begin{array}{c}
2211: \psi_{^+\!X_+}
2212: \\
2213: \psi_S^\dagger
2214: \end{array}
2215: \!\!\!\right)^{\phantom{X}}_{\phantom{X}}
2216: $
2217: &
2218: $\,\sqrt 2 \kappa M$
2219: &
2220: $+$
2221: &
2222: $+$
2223: \\
2224: %& & & \\
2225: % & & & \\
2226: \hline
2227: % & & & \\
2228: 
2229: &
2230: \begin{tabular}{l}
2231: $V_\mu^1 [^-\!I_-]\,,$\\
2232: ${^-\!R_-}\,$;
2233: \end{tabular}
2234: &
2235: $\psi_g^1=
2236: \left(
2237: \begin{array}{c}
2238: \psi_{^-\!X_-}
2239: \\
2240: -{\rm i} {\lambda^1}^\dagger
2241: \end{array}
2242: \!\!\!\right)_{\phantom{X}}^{\phantom{X}}
2243: $
2244: &
2245: $g M$
2246: &
2247: $-$
2248: &
2249: $-$
2250: \\
2251: \begin{tabular}{c}
2252: SU(2)$_X$\\ Waterfall Gauge\\
2253: ($g$-sector)
2254: \end{tabular}
2255: &
2256: \begin{tabular}{l}
2257: $V_\mu^2 [^-\!R_+]\,,$\\
2258: ${^-\!I_+}\,$;
2259: \end{tabular}
2260: &
2261: $\psi_g^2=
2262: \left(
2263: \begin{array}{c}
2264: {\rm i}\,\psi_{^-\!X_+}
2265: \\
2266: -{\rm i} {\lambda^2}^\dagger
2267: \end{array}
2268: \!\!\!\right)_{\phantom{X}}^{\phantom{X}}
2269: $
2270: &
2271: $g M$
2272: &
2273: $+$
2274: &
2275: $-$
2276: \\
2277: &
2278: \begin{tabular}{l}
2279: $V_\mu^3 [^+\!I_-]\,,$\\
2280: ${^+\!R_-}$
2281: \end{tabular}
2282: &
2283: $\psi_g^3=
2284: \left(
2285: \begin{array}{c}
2286: \psi_{^+\!X_-}
2287: \\
2288: -{\rm i} {\lambda^3}^\dagger
2289: \end{array}
2290: \!\!\!\right)^{\phantom{X}}
2291: $
2292: &
2293: $g M$
2294: &
2295: $-$
2296: &
2297: $+$
2298: \\
2299: %& & & \\
2300: % & & & \\
2301: \hline
2302: \end{tabular}
2303: \end{center}
2304: \caption{\sl\small Particle  spectrum of  the inflaton
2305: and  an  SU(2)$_X$-gauged   waterfall  sectors  after  inflation.  The
2306: would-be Goldstone bosons of the respective SU(2)$_X$ gauge fields are
2307: given in the square brackets \label{spectrum:SU2}}
2308: \end{table}
2309: 
2310: The conservation of both  $D_{1,2}$-parities enforces the stability of
2311: all $g$-sector  particles. Instead, if only the  $D_1$-parity, but not
2312: $D_2$,  is   conserved,  then   only  the  $D_1$-odd   particles  from
2313: Table~\ref{spectrum:SU2}  will  be   stable,  and  {\it  vice  versa}.
2314: Obviously, both $D_1$-  and $D_2$-parities need be broken  to make all
2315: $g$-sector   particles  unstable.   In   Appendix~\ref{Dappendix},  we
2316: discuss  possible mechanisms  of explicit  $D$-parity breaking  for an
2317: SU(2)$_X$   waterfall-gauge  sector.   In   general,  there   are  two
2318: mechanisms  for  breaking  $D$-parity.   The  first  one  consists  of
2319: including  higher-order non-renormalizable  operators in  the K\"ahler
2320: potential  whose  presence explicitly  breaks  $D$-parity, whilst  the
2321: second one is  very analogous to the ${\rm  U}(1)_X$ case.  Although a
2322: bare FI  $D$-term is not  possible in non-Abelian  theories, effective
2323: $D^a$-tadpole  terms  may appear  after  the  SSB  of SU(2)$_X$.   The
2324: effective $D^a$-tadpole  terms do not break SUSY.   They get generated
2325: either  from a  non-renormalizable K\"ahler  potential or  are induced
2326: radiatively,  after integrating out  Planck-scale degrees  of freedom.
2327: Thus, without excessive tuning,  the effective $D^a$-tadpole terms can
2328: in general  be small of  the size required  to obtain a  second reheat
2329: phase in the evolution of the Universe.
2330: 
2331: Independently of  the mechanism which is invoked  to break $D$-parity,
2332: we  may in general  parameterize the  $g$-sector particle  decay rates
2333: through the  $D$-parity-violating mass $m_{\rm FI}$,  which enters the
2334: relation~(\ref{GammaR}).   In the  next section,  we will  discuss how
2335: these  relatively  long-lived $g$-sector  particles  are produced  via
2336: preheating.
2337: 
2338: 
2339: 
2340: 
2341: \subsection{Preheating and Thermalization}\label{preheat}
2342: 
2343: 
2344: After the inflaton field $\phi$  passes below a certain critical value
2345: $\phi_c \approx M$, the  so-called waterfall mechanism gets triggered.
2346: In this case, the inflaton $\phi$ and all other $\kappa$-sector fields
2347: (see  Tables~\ref{spectrum}  and~\ref{spectrum:SU2})  oscillate  about
2348: their true supersymmetric minima:  $\langle S\rangle = 0$ and $\langle
2349: X_+ \rangle  = \sqrt{2}\,  M$.  In this  waterfall epoch, most  of the
2350: energy  density  of  the   Universe  is  stored  to  these  coherently
2351: oscillating $\kappa$-sector  field condensates and  is given initially
2352: by  $\rho_\kappa  =  \kappa^2  M^4$.   During  the  waterfall  regime,
2353: however,  there is  an  additional mechanism  for particle  production
2354: called {\em preheating}.
2355: 
2356: In  general, there  are two  phenomena associated  with the  notion of
2357: preheating:
2358: \begin{itemize}
2359: 
2360: \item  The  first  effect  of  preheating  arises  from  the  negative
2361:   curvature  of  the potential  with  respect  to the  $\kappa$-sector
2362:   fields.   Such  a  negative  curvature  corresponds  to  a  negative
2363:   tachyonic  mass  term  in  the  potential.  As  a  consequence,  the
2364:   particle  number within infrared  modes of  momentum less  than this
2365:   tachyonic mass grows exponentially.  This phenomenon is known as the
2366:   {\em    negative    coupling    instability}   or    \emph{tachyonic
2367:   preheating}~\cite{TACHYPREH}.  Numerical simulations have shown that
2368:   the  field  amplitudes  suffer   strong  damping  during  the  first
2369:   oscillation, due to  the energy transfer to the  infrared modes.  In
2370:   the  $F_D$-term  hybrid model,  only  $\kappa$-sector particles  are
2371:   produced  by tachyonic preheating.   A full  study of  this process,
2372:   including thermal equilibration of the $\kappa$-sector particles, is
2373:   a highly  nontrivial matter  and has so  far only been  achieved for
2374:   very  particular models of  preheating.  Since  the fraction  of the
2375:   energy   density  transferred  instantaneously   to  $\kappa$-sector
2376:   particles through tachyonic preheating  is rather small, compared to
2377:   their  initial energy  density $\rho_\kappa$,  these model-dependent
2378:   details fortunately have no dramatic impact on the expansion and the
2379:   thermal history of the Universe.   Therefore, we do not consider the
2380:   phenomenon of tachyonic preheating in the $F_D$-term hybrid model.
2381: 
2382: \item  Particle   production  may  also  occur   during  the  coherent
2383:   oscillation  regime,  because  both  the  $\kappa$-  and  $g$-sector
2384:   particles have masses  that can vary very strongly  with time.  This
2385:   effect is called {\em preheating  via a time-varying mass} or simply
2386:   {\em preheating}~\cite{PREHEATING,GBRM}.   As we will  show below, a
2387:   small but  significant fraction of  the total energy density  of the
2388:   Universe $\rho_\kappa$  can be transferred,  almost instantaneously,
2389:   to the $g$-sector particles, e.g.~$\rho_g \sim 10^{-4} \rho_\kappa$,
2390:   for $\kappa  \sim 10^{-2}$.  As we illustrated  in the  beginning of
2391:   this section and will  show more explicitly in Section~\ref{reheat},
2392:   this small  fraction of the $g$-sector energy  density is sufficient
2393:   to alter dramatically the thermal history of the Universe.
2394: 
2395: 
2396: \end{itemize}
2397: 
2398: 
2399: Our interest lies therefore in computing the production energy density
2400: $\rho_g$ of the $g$-sector particles  via preheating. A key element in
2401: such  a computation is  the profile  of the  time-varying mass  of the
2402: $g$-sector  particles, $m_g (t)  = g\,  X_+ (t)/\sqrt{2}$.   The exact
2403: time  dependence of  $m_g (t)$  depends crucially  on the  dynamics of
2404: tachyonic preheating.  Comparative  numerical studies strongly suggest
2405: that a sufficiently accurate description  of the time evolution of the
2406: $g$-sector  mass is  obtained by~\cite{GBRM}\footnote{To  be specific,
2407: the mass-term  time-variation studied  in~\cite{GBRM} was for  a model
2408: with  a single field  rolling from  the top  of a  local maximum  of a
2409: quartic potential.  It was  found that a $\tanh$-functional dependence
2410: accurately  captures the  evolution  of the  time-varying mass.   Even
2411: though the  model considered~\cite{GBRM}  is still different  from our
2412: hybrid inflationary potential,  the derived $\tanh$-functional profile
2413: for  the  time-varying  mass  should  be  regarded  as  a  substantial
2414: improvement over the one assumed in~\cite{GP}.}
2415: \begin{equation}
2416: \label{tanhmass}
2417: m_g(t)\ =\ \frac{g M}{2}\; \Big[\, \tanh ( \kappa M t)\: +\: 1\,
2418:   \Big]\; .
2419: \end{equation}
2420: Notice that the time-dependent function $m_g(t)$ properly interpolates
2421: between the values $m_g (t \to -\infty) = 0$ and $m_g (t \to \infty) =
2422: g M$ that  occur in the beginning and the end  of the waterfall epoch,
2423: respectively.
2424: 
2425: 
2426: Given  the time-dependent  mass~(\ref{tanhmass}), we  may  compute the
2427: occupation  number of the  fermionic $g$-sector  modes by  solving the
2428: Dirac equation
2429: \begin{equation}
2430:   \label{uheq}
2431: \Big[\, {\rm i}\, \gamma^0\, \partial_t\: -\:
2432: \mbox{\boldmath $\gamma$}\cdot\mathbf{k}\: -\: m_g(t)\,\Big]\; u_h(t)\ =\ 0\, .
2433: \end{equation}
2434: The  solution  to   the  above  equation  may  be   expressed  by  the
2435: time-dependent Dirac spinor $u_h (t)$ in the chiral representation:
2436: \begin{equation}
2437: u_h (t)\ =\ \left( \begin{array}{c} L_h (t) \\ R_h (t) \end{array} \right)\,
2438: \otimes\, \xi_h\; ,
2439: \end{equation}
2440: where $\xi_h$  is the helicity two-component  eigenspinor for helicity
2441: $h=\pm$.   The  occupation  number  of  Dirac  fermions  produced  via
2442: preheating in the true supersymmetric vacuum at $t\to \infty$ is given
2443: by
2444: \begin{equation}
2445:   \label{nFh}
2446: n^{\rm F}_h(k)\ =\ \frac 1{2\omega(k)}\, \Big[\,
2447: hk (|R_h|^2\, -\, |L_h|^2)\: -\: m_g (L_hR_h^*\, +\, L_h^*R_h)\,\Big]
2448: \ +\ \frac 12\ ,
2449: \end{equation}
2450: where $k  = |\mathbf{k}|$ is the  modulus of the  3-momentum. With the
2451: help  of~(\ref{nFh}), the  $k$-mode  energy density  is calculated  by
2452: $\rho(k)    =    \sum_h    \omega(k)\,    n^{\rm    F}_h(k)$,    where
2453: $\omega(k)=\sqrt{k^2+m_g^2(t\to   \infty)}$.   To   obtain   a  unique
2454: solution to  the linear differential  equation~(\ref{uheq}), we impose
2455: initial  conditions  that  correspond  to a  zero  occupation  number,
2456: i.e.~$n^{\rm F}_h(k)=0$.  These are given at $t\to -\infty$ by
2457: \begin{equation}
2458:   \label{initF}
2459: L_h\ =\ \sqrt\frac{\omega(k) + hk}{2\omega}\ ,\qquad
2460: R_h\ =\ \sqrt\frac{\omega(k) - hk}{2\omega}\ .
2461: \end{equation}
2462: 
2463: By analogy, the occupation number  of the bosonic $g$-sector modes are
2464: determined by solving the Klein--Gordon equation of motion
2465: \begin{equation}
2466:   \label{phieq}
2467: \Big[\, \partial_t^2\: +\: \mathbf{k}^2\: +\: m_g^2(t)\,\Big]\,
2468: \varphi(t) \ =\ 0\; ,
2469: \end{equation}
2470: and imposing the initial conditions at $t\to -\infty$,
2471: \begin{equation}
2472: \varphi\ =\ \frac{1}{2\sqrt{\omega(k)}}\ ,\qquad
2473: \frac{\partial \varphi}{\partial t}\ =\ -\,{\rm i}\,\sqrt{\frac\omega{2}}\ .
2474: \end{equation}
2475: As in the Dirac case, these initial conditions correspond to vanishing
2476: occupation  numbers.  The occupation  number of  the bosonic  modes at
2477: $t\to \infty$ is given by
2478: \begin{equation}
2479:   \label{initB}
2480: n^{\rm B}(k)\ =\ \frac 12\; \omega(k)\, |\varphi|^2\ +\
2481: \frac 1{2\omega(k)}\; \left|\frac{d \varphi}{dt}\right|^2\ -\ \frac 12\ .
2482: \end{equation}
2483: 
2484: 
2485: Using  the time-dependent  mass-term~(\ref{tanhmass}), along  with the
2486: initial  conditions~(\ref{initF}) and  (\ref{initB}),  one may  obtain
2487: analytical     expressions      in     terms     of     hypergeometric
2488: functions~\cite{GBRM},  for  the  particle  production  between  $t\to
2489: -\infty$ and $t\to \infty$.  For $\kappa\ll  g$ and $k \ll g M$, these
2490: analytical expressions reduce to
2491: \begin{equation}
2492: n(k)\ =\  \frac{2}{\exp{\left(\frac{\displaystyle \pi k}
2493: {\displaystyle \kappa M}\right)}\: \pm\:  1}\ \ ,
2494: \end{equation}
2495: where the sign $+$ applies  for $n(k)=n^{\rm F}_h(k)$ and the sign $-$
2496: for $n(k)=n^{\rm  B}(k)$.  Recalling that there are  2 helicity states
2497: for  a  $g$-sector  fermion  and  4  real degrees  of  freedom  for  a
2498: $g$-sector  boson,  we may  calculate  the  occupation  number of  all
2499: $g$-sector modes as follows:
2500: \begin{equation}
2501: n_g(k)\ =\ N_{\rm b}\, \left(\, \sum\limits_{h=\pm}\, n_h^{\rm F}(k)\: +\:
2502: 4 n^{\rm B}(k)\,\right)\,,
2503: \end{equation}
2504: where $N_{\rm b}$ is the  number of broken generators of the waterfall
2505: gauge symmetry.  In particular, it is $N_{\rm b}=1$ for ${\rm U}(1)_X$
2506: and  $N_{\rm  b}=3$  for  ${\rm  SU}(2)_X$  [cf.~Tables~\ref{spectrum}
2507: and~\ref{spectrum:SU2}].     Since   the   produced    particles   are
2508: non-relativistic,   i.e.~$k   \ll   gM$,   their   occupation   number
2509: distribution  $n_g (k)$  can easily  be integrated  to give  the total
2510: energy density carried by the $g$-sector fields, i.e.
2511: \begin{equation}
2512:   \label{rhog}
2513: \frac{\rho_g}{\rho_\kappa}\ \approx\
2514: \frac{gM}{\rho_\kappa\,2\pi^2}\, \int\limits_0^\infty k^2 dk\, n_g(k)\
2515: \approx\ 2.1 \times 10^{-2}\, N_{\rm b}\; \kappa g\; .
2516: \end{equation}
2517: Here  $\rho_{\kappa}=\kappa^2  M^4$  is  the  energy  density  of  the
2518: $\kappa$-sector  particles shortly  before  the waterfall  transition.
2519: Equation~(\ref{rhog}) will be a valuable input for the next section to
2520: compute  the true reheat  temperature $T_{\rm  reh}$ of  the Universe,
2521: which arises from the combined  effect of the $\kappa$- and $g$-sector
2522: particle decays.
2523: 
2524: 
2525: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2526: \setcounter{equation}{0}
2527: \section{Coupled Reheating and Gravitino Abundance}\label{reheat}
2528: 
2529: In the previous  section, we have seen that  the $g$-sector particles,
2530: e.g.~$\psi_g$, $R_-$  and $V_\mu$,  can be abundantly  produced during
2531: the  preheating epoch.  Assuming  that they  dominate the  Universe at
2532: some  later  time,  their  decays  induced  by  the  small  $D$-parity
2533: violating  couplings will give  rise to  a second  reheat temperature,
2534: which we  denote here by $T_g$. As  we will show in  this section, the
2535: large entropy, which is released  by the late decays of the $g$-sector
2536: particles,  will be  sufficient  to dilute  the  gravitinos to  levels
2537: compatible  with BBN  limits.
2538: 
2539: More  explicitly, we  present  a detailed  numerical  analysis of  the
2540: gravitino abundance~$Y_{\widetilde{G}}$, where  the combined effect of
2541: the $\kappa$-  and $g$-sector particle decays is  carefully taken into
2542: account.   As  we  mentioned  in  the Introduction,  we  call  such  a
2543: two-states'   mechanism  of  reheating   the  Universe   {\em  coupled
2544: reheating}.  In Section~\ref{Beqs}, we set the BEs relevant to coupled
2545: reheating  and give  numerical  estimates of  the gravitino  abundance
2546: $Y_{\widetilde{G}}$ and  the energy densities  $\rho_\kappa$, $\rho_g$
2547: and $\rho_{\rm rad}$ related to the $\kappa$- and $g$-sector particles
2548: and their radiation,  respectively.  In Section~\ref{Seqs}, we present
2549: a semi-analytic approach to  BEs, where useful approximate expressions
2550: for~$Y_{\widetilde{G}}$ are  obtained.  Finally, in Section~\ref{para}
2551: we   derive  gravitino  abundance   constraints  on   the  theoretical
2552: parameters.
2553: 
2554: 
2555: 
2556: \subsection{Boltzmann Equations}\label{Beqs}
2557: 
2558: The number density $n_{{\tilde G}}$ of gravitinos, the energy density
2559: $\rho_\kappa$ ($\rho_g$) of the $\kappa$~($g$)-sector particles and
2560: the energy density $\rho_{\rm rad}$ of the radiation produced by their
2561: decays satisfy the following system of BEs~\cite{kolb}:
2562: \begin{eqnarray}
2563:   \label{ng}
2564: \dot n_{{\widetilde G}}\: +\: 3Hn_{{\widetilde G}} &=&
2565:                                          C_{\widetilde G}\, T^6\,,\nonumber\\
2566:   \label{nf}
2567: \dot \rho_\kappa\: +\: 3H\rho_\kappa &=& -\,\Gamma_\kappa\,
2568: \rho_\kappa\,,\nonumber\\
2569:   \label{nfb}
2570: \dot\rho_g\: +\: 3H\rho_g &=& -\,\Gamma_g\,\rho_g\,,\nonumber\\
2571:   \label{rR}
2572: \dot\rho_{\rm rad}\: +\: 4H\rho_{\rm rad} &=&
2573: \Gamma_\kappa\,\rho_\kappa\: +\: \Gamma_g\,\rho_g\; ,
2574: \end{eqnarray}
2575: where a dot on  $n_{\widetilde{G}}$, $\rho_{\kappa, g}$ and $\rho_{\rm
2576: rad}$  indicates  differentiation  with  respect to  the  cosmic  time
2577: $t$.  The quantity  $C_{{\widetilde G}}(T)$  is a  collision  term for
2578: gravitino production  calculated in~\cite{brand,kohri} and  the Hubble
2579: parameter $H$ is given by
2580: \begin{equation}
2581:   \label{Hini}
2582: H\ =\ \frac{1}{\sqrt{3}\,m_{\rm Pl}}\;
2583: \bigg(\, m_{{\widetilde G}}\,n_{{\widetilde G}}\:
2584: +\: \rho_\kappa\: +\: \rho_g\: +\: \rho_{\rm rad}\, \bigg)^{1/2}\; ,
2585: \end{equation}
2586: where $m_{{\widetilde  G}}$ is the  mass of the  gravitino $\widetilde
2587: G$. In addition,  the temperature $T$ and the  entropy density $s$ may
2588: be determined through the relations:
2589: \begin{equation}
2590:   \label{rs}
2591: \rho_{\rm rad}\ =\ \frac{\pi^2}{30}\; g_*\, T^4\,,\qquad
2592: s\ =\ \frac{2\pi^2}{45}\; g_*\ T^3,
2593: \end{equation}
2594: where  $g_* (T)$  is the  effective number  of degrees  of  freedom at
2595: temperature  $T$. Since  the initial  temperature is  $T_{\rm  in} \ll
2596: \kappa M$, it is $g_* = 240$ for all $T > M_{\rm SUSY}$.
2597: 
2598: Here  we should  note that  in  BEs~(\ref{rR}) we  have neglected  the
2599: collision  terms  related   to  the  self-annihilation  of  $g$-sector
2600: particles.   Their thermally  averaged cross  section  times velocity,
2601: $\langle \sigma_{\rm ann} v \rangle$, is estimated to be
2602: \begin{equation}
2603: \langle \sigma_{\rm ann}\, v \rangle\ \stackrel{<}{{}_\sim}\
2604: 10^{-35}~{\rm GeV}^{-2}\; ,
2605: \end{equation}
2606: which is numerically negligible.
2607: 
2608: The  numerical  analysis of  the  BEs  (\ref{rR})  gets simplified  by
2609: absorbing the Hubble expansion terms into new variables.  To this end,
2610: we define the following dimensionless quantities \cite{riotto}:
2611: \begin{equation}
2612:   \label{fdef}
2613: f_{{\widetilde G}}\ =\ n_{{\widetilde G}} a^3\,,\quad f_\kappa\ =\ \rho_\kappa
2614: a^3\,,\quad f_g\ =\ \rho_g a^3\,, \quad f_{\rm rad}\ =\ \rho_{\rm rad} a^4.
2615: \end{equation}
2616: where $a$  is the  usual expansion scale  factor of the  Universe.  We
2617: also convert the  time derivatives to derivatives with  respect to the
2618: logarithmic   time  $\ln\left(a/a_{\rm   I}\right)$~\cite{qui},  where
2619: $a_{\rm I}$  is some initial or  reference value for  the scale factor
2620: $a$.   With  the  above   substitutions,  the  BEs~(\ref{rR})  may  be
2621: re-written as
2622: \begin{eqnarray}
2623:  \label{fg}
2624: H f^\prime_{\widetilde G} &=& C_{{\widetilde G}}\, T^6 a^3\,,\nonumber\\
2625:  \label{ff}
2626: H f^\prime_\kappa &=& -\,\Gamma_\kappa f_\kappa\,,\nonumber\\
2627:  \label{ffb}
2628: H f^\prime_g &=& -\,\Gamma_g\, f_g\,,\nonumber\\
2629:  \label{fR}
2630: H f^\prime_{\rm rad} &=& \Gamma_\phi f_\phi a\: +\: \Gamma_g f_g a\,,
2631: \end{eqnarray}
2632: where   the   prime   now   denotes   differentiation   with   respect
2633: to~$\ln\left(a/a_{\rm I}\right)$. Correspondingly, the Hubble parameter
2634: $H$ and  temperature $T$ may  now be expressed  in terms of  the newly
2635: introduced variables~(\ref{fdef}) as follows:
2636: \begin{equation}
2637:   \label{H2exp}
2638: H\ =\ \frac{1}{\sqrt{3}\, a^{3/2}\, m_{\rm Pl}}\;
2639: \bigg(\, m_{\widetilde{G}} f_{\widetilde{G}}\: +\:
2640: f_\kappa\: +\: f_g\:  +\: a^{-1} f_{\rm rad}\,\bigg)^{1/2}\,,\quad
2641: T\ =\ \left(\frac{30\, f_{\rm rad}}{\pi^2 g_{\ast} a^4}\right)^{1/4}\; .
2642: \end{equation}
2643: The transformed system of  BEs~(\ref{fR}) can be numerically solved by
2644: imposing the following initial conditions:
2645: \begin{equation}
2646:   \label{init}
2647: f_{\kappa,{\rm I}}\, a_{\rm I}^3\ =\ \kappa^2 M^4\,,\quad
2648: f_{g,{\rm I}}\, a_{\rm I}^3\ =\ 2.1\times 10^{-2} g \kappa^3\, M^4\,,
2649: \quad f_{{\rm rad},{\rm I}}\ =\ 0\; ,
2650: \end{equation}
2651: where    the   subscript~I   refers    to   quantities    defined   at
2652: $\ln\left(a/a_{\rm  I}\right)=0$.   Notice   that  the  initial  value
2653: $f_{g,{\rm  I}}\,  a_{\rm  I}^3$   is  equal  to  the  energy  density
2654: $\rho_{g,{\rm  I}}$  of   the  $g$-sector  particles  produced  during
2655: preheating and is given in~(\ref{rhog}) .
2656: 
2657: 
2658: In  Fig.~\ref{fig:rY}(a),  we   present  numerical  estimates  of  the
2659: cosmological evolution of energy densities $\rho_{\kappa,g,{\rm rad}}$
2660: as functions  of the temperature  $T$ in a double  $x$-$y$ logarithmic
2661: plot, where $\rho_\kappa$ is represented by a dark grey line, $\rho_g$
2662: by  a grey line  and $\rho_{\rm  rad}$ by  a light  grey line.   As an
2663: example,   we    use   $M    =   0.7   \times    10^{16}~{\rm   GeV}$,
2664: $\rho=\lambda=\kappa=10^{-3}$ and  $m_{\rm FI}/M =  4.3\times 10^{-7}$
2665: (bold  lines)  and  $m_{\rm  FI}/M  = 10^{-3}$  (thin  lines).   Since
2666: $\rho_{\rm  rad}$ is  affected very  little  for the  larger value  of
2667: $m_{\rm FI}/M$, it  has not been added to  the plot.  The intersection
2668: point  of  the $T$-dependent  functions  $\rho_\kappa$ and  $\rho_{\rm
2669: rad}$ signals  the completion of the  $\kappa$-sector particle decays.
2670: For  the   specific  example,  this   point  occurs  at   $T_\kappa  =
2671: 3.2\times10^{11}~{\rm   GeV}$.   For  $m_{\rm   FI}/M  =   4.3  \times
2672: 10^{-7}~{\rm  GeV}$,  we  obtain   two  more  intersections:  one  for
2673: $T=T_{\rm eq} \simeq 3.9  \times 10^6~{\rm GeV}$ where $\rho_g (T_{\rm
2674: eq}) = \rho_{\rm rad} (T_{\rm  eq})$ and another one for $T=T_g \simeq
2675: 200~{\rm  GeV}$,  where  the  $g$-sector  particles  have  practically
2676: decayed away and $\rho_g (T_g) = \rho_{\rm rad} (T_g)$.  Thanks to the
2677: huge   entropy  release   in  this   case,  the   gravitino  abundance
2678: $Y_{\widetilde  G}\ =\  n_{{\widetilde G}}/s$  gets  sharply decreased
2679: from  about $2.2  \times  10^{-11}$ to  $2.4  \times 10^{-15}$.   This
2680: dramatic    reduction   of    $Y_{\widetilde   G}$    is    shown   in
2681: Fig.~\ref{fig:rY}(b).  On  the contrary, if $m_{\rm  FI}/M = 10^{-3}$,
2682: no intersection of $\rho_g$ with  $\rho_{\rm rad}$ takes place and, in
2683: consequence,  no  phase of  second  reheating  occurs.   This is  also
2684: illustrated in Fig.~\ref{fig:rY}(a),  where the dependence of $\rho_g$
2685: is   displayed   by   a   thin    line.    As   can   be   seen   from
2686: Fig.~\ref{fig:rY}(b),  the   gravitino  abundance  $Y_{\widetilde  G}$
2687: remains  unsuppressed  in   this  case,  i.e.~$Y_{\widetilde  G}  \sim
2688: 10^{-10}$.   As we will  see below  in Section~\ref{para},  such large
2689: values  of   $Y_{\widetilde  G}$  are  in  gross   conflict  with  BBN
2690: constraints.
2691: 
2692: 
2693: 
2694: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2695: \begin{figure}[t]
2696: \begin{center}
2697: \epsfig{file=rT.eps,height=5.2in,angle=-90} \\[3mm]
2698: \epsfig{file=YgT.eps,height=5.2in,angle=-90}
2699: \end{center}
2700: \caption{\sl\small The evolution as a function of $\log
2701: T$ of the quantities: {\sf (a)} $\log\rho_i$ with $i=\kappa$ (dark
2702: grey line), $i=g$ (grey line), $i={\rm rad}$ (light grey line)
2703: {\sf (b)} $\widetilde G$ yield, $Y_{\widetilde G}$. In both cases, we take
2704: $M=0.7\times10^{16}~{\rm GeV}$, $\rho=\lambda=\kappa=0.001$ and
2705: $m_{\rm FI}/M=4.3\times10^{-7}~{\rm GeV}$ (bold lines) and $m_{\rm
2706: FI}/M=1\times10^{-3}$ (thin lines).} \label{fig:rY}
2707: \end{figure}
2708: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2709: 
2710: 
2711: 
2712: \subsection{Semi-analytic Approach} \label{Seqs}
2713: 
2714: We now  present a more intuitive  and rather accurate  approach to the
2715: dynamics  of  coupled   reheating,  and  find  approximate  analytical
2716: expressions    that   describe   the    evolution   of    the   energy
2717: densities~$\rho_{\kappa,g,{\rm  rad}}$.  In  addition,  we derive  the
2718: conditions that ensure  the existence of a second  reheat phase in the
2719: evolution  of  the  Universe.   Finally,  we  estimate  the  gravitino
2720: abundance $Y_{\widetilde G}$ due to coupled reheating.
2721: 
2722: Shortly after inflation ends, the energy of our observable Universe is
2723: dominated by the inflaton $S$ and the other $\kappa$-sector particles,
2724: with an initial energy density  $\rho_{\kappa, \rm I} = \kappa^2 M^4$.
2725: As   we  schematically   illustrated  in   Section~\ref{preheat},  the
2726: $\kappa$-sector particles decay  into relativistic degrees of freedom,
2727: producing  an energy  density  $\rho_{\rm rad}$.   The energy  density
2728: $\rho_g$  of the $g$-sector  particles is  subdominant at  these early
2729: stages after  the first reheating due to  the $\kappa$-sector particle
2730: decays.  In  fact, for  temperatures $T >  T_{\rm eq}$,  where $T_{\rm
2731: eq}$ is  the first intersection  point of the  $T$-dependent functions
2732: $\rho_{\rm rad}$ and $\rho_g$  [see~(\ref{trg})], the evolution of all
2733: relevant energy densities may be approximately described as follows:
2734: \begin{equation}
2735:   \label{rRfq}
2736: \rho_\kappa\ =\ \rho_{\kappa,{\rm I}}\, \left(a/a_{\rm I}\right)^{-3}\,,\quad
2737: \rho_g\ =\ \rho_{g,{\rm I}}\, \left(a/a_{\rm I}\right)^{-3}\,,\quad
2738: \rho_{\rm rad}\ =\ \rho_{\rm rad}(T_\kappa)\,
2739: \left(T/T_\kappa \right)^4\; .
2740: \end{equation}
2741: As mentioned  above, the  $T$-dependent function $\rho_{\rm  rad}$ may
2742: first cross the corresponding $\rho_g$ at $T=T_{\rm eq}$, where
2743: \begin{equation}
2744:   \label{trg}
2745: \rho_{\rm rad} (T_{\rm eq })\ =\ \rho_g (T_{\rm eq })\; .
2746: \end{equation}
2747: Using  the   fact  that  $\rho_{\rm  rad}   (T_\kappa)  =  \rho_\kappa
2748: (T_\kappa)$ and assuming that the Universe expands isentropically with
2749: $a  \propto T^{-1}$ when  $T_{\rm eq}\leq  T\leq T_\kappa$,  we obtain
2750: from~(\ref{trg}) the approximate relation
2751: \begin{equation}
2752:   \label{Teq}
2753: T_{\rm eq}\ \simeq\ T_\kappa\
2754: \frac{\rho_{g, {\rm I}}}{\rho_{\kappa,{\rm I }}}\ .
2755: \end{equation}
2756: In deriving~(\ref{Teq}), we have also made use of~(\ref{rRfq}).
2757: 
2758: A second  reheat phase in the  evolution of the  Universe takes place,
2759: only  if  $T_g  <  T_{\rm  eq}$,  where  $T_g$  is  the  naive  reheat
2760: temperature  due to  the $g$-sector  particles  decays [see~(\ref{Tg})
2761: below].  To better explore the consequences of this last condition, we
2762: use  the  abbreviation  $g$-DAD  ($g$-DBD)  to  indicate  whether  the
2763: $g$-sector  particles Decay  After  (Before) the  Domination of  their
2764: energy  density.  With  the  aid  of~(\ref{Teq}),  the  following  two
2765: conditions for $g$-DAD and $g$-DBD may be deduced:
2766: \begin{equation}
2767:   \label{cond}
2768: \frac{T_g}{T_\kappa}\ <\
2769:   \frac{\rho_{g,\rm I}}{\rho_{\kappa, \rm I}}\quad (g\mbox{-DAD}),\qquad
2770: \frac{T_g}{T_\kappa}\ \geq\ \frac{\rho_{g,\rm I}}{\rho_{\kappa,\rm
2771:   I}}\quad (g\mbox{-DBD})\; .
2772: \end{equation}
2773: These two possible scenarios  are illustrated in Fig.~\ref{fig:rY} for
2774: $m_{\rm FI}/M  = 4.3\times 10^{-7}$ ($m_{\rm FI}/M  = 10^{-3}$), where
2775: the bold (thin) lines correspond to $g$-DAD ($g$-DBD).
2776: 
2777: The gravitino abundance $Y_{\widetilde  G}$ can be calculated by simply
2778: integrating  $f^\prime_{\widetilde G}$  that  occurs in  the first  BE
2779: of~(\ref{fg})   and  using   the   fact  that   $Y_{\widetilde  G}   =
2780: f_{\widetilde G}/sa^3$.   It turns out  that the main  contribution to
2781: $Y_{\widetilde G}$  comes from the integration  after the commencement
2782: of  the  radiation  dominated  era, i.e.~for  $T\leq  T_\kappa$.   The
2783: so-derived  formula  reproduces rather  accurately  the one  presented
2784: in~\cite{kohri} in the massless gluino limit, where
2785: \begin{equation}
2786:   \label{Yk}
2787: Y^\kappa_{\widetilde G}\ =\ 1.6 \times 10^{-12}
2788: \left(\frac{T_\kappa}{10^{10}~{\rm GeV}}\right)\ .
2789: \end{equation}
2790: Note that (\ref{Yk}) is only valid for the $g$-DBD case.
2791: 
2792: The  situation is  different for  the  $g$-DAD case,  where a  drastic
2793: reduction  of the  gravitino  abundance, caused  by  the huge  entropy
2794: release  from the $g$-sector  particle decays,  takes place.   In this
2795: case, the gravitino abundance $Y^g_{\widetilde G}$ may be estimated in
2796: the following way. We first notice that
2797: \begin{equation}
2798:   \label{Yg1}
2799: Y^g_{\widetilde G}\ =\ Y^\kappa_{\widetilde G}\
2800: \frac{s(T_{\rm eq})\, a^3 (T_{\rm eq})}{s(T_g)\, a^3(T_g)}\ .
2801: \end{equation}
2802: Then, with the help of (\ref{rs}) and (\ref{rRfq}), we may obtain the
2803: relation
2804: \begin{equation}
2805:   \label{dilution}
2806: \frac{s(T_{\rm eq})\, a^3 (T_{\rm eq})}{s(T_g)\, a^3(T_g)}\ =\
2807: \left(\frac{T_{\rm eq}}{T_g}\right)^3\;
2808: \left(\frac{\rho_g(T_{\rm    eq})}{\rho_g(T_g)}\right)^{-1}\
2809: =\ \frac{T_g}{T_{\rm eq}}\ .
2810: \end{equation}
2811: Substituting    the   respective    expressions   of~(\ref{dilution}),
2812: (\ref{Yk}) and~(\ref{Teq}) into~(\ref{Yg1}) yields
2813: \begin{equation}
2814:   \label{Yg}
2815: Y^g_{\widetilde G}\ =\ 1.6\times10^{-12}\, \left(
2816: \frac{T_g}{10^{10}~{\rm GeV}}\right)\; \frac{\rho_{\kappa , \rm
2817: I}}{\rho_{g, \rm I}}\ \simeq\ \frac{7.6\times 10^{-11}}{\kappa
2818: g}\; \left(\frac{T_g}{10^{10}~{\rm GeV}}\right)\ ,
2819: \end{equation}
2820: where  we  have  used  (\ref{rhog})  to derive  the  last  approximate
2821: equality. We have checked that the semi-analytic formula (\ref{Yg}) is
2822: in  remarkable  agreement  with  numerical estimates  in  the  $g$-DAD
2823: regime.
2824: 
2825: Finally, we should  comment on the fact that  the number of $e$-folds,
2826: ${\cal  N}_e$  gets modified  in  the  $g$-DAD  case, because  of  the
2827: occurrence  of  a  $g$-sector-matter  dominated era.   Making  use  of
2828: standard methods~\cite{CLLSW,review}, we  are able to determine ${\cal
2829: N}_e$  at the  WMAP pivotal  point $k_0=0.002~{\rm  Mpc}^{-1}$  by the
2830: following relation:
2831: \begin{equation}
2832:   \label{Ng}
2833: {\cal N}_e\ =\ 22.6\ +\ {1\over 6}\ln (\kappa^2M^4)\
2834: +\ {1\over3}\ln T_g\ +\ {1\over3}\ln{
2835: \rho_{\kappa,\rm I}\over\rho_{g,\rm I}}\ .
2836: \end{equation}
2837: This result,  however, does  not crucially alter  the value  of ${\cal
2838: N}_e$,  which  remains  close  to  $55-60$  in  the  $g$-DAD  case  as
2839: well. Interestingly  enough, $Y^g_{\widetilde G}$ and  ${\cal N}_e$ do
2840: not directly  depend on $T_\kappa$ given  in~(\ref{Tkappa}).  In fact,
2841: in the $g$-DAD  case, $Y^g_{\widetilde G}$ and ${\cal  N}_e$ are fully
2842: independent of the superpotential  couplings $\lambda$ and $\rho$, and
2843: only  have  a mild  linear  and  logarithmic  dependence on  $\kappa$,
2844: respectively.  As we will discuss below, it is this last property that
2845: leads to a significant  relaxation of the strict gravitino constraints
2846: on these couplings, when compared to the $g$-DBD case.
2847: 
2848: 
2849: 
2850: \subsection{Gravitino Abundance Constraints}\label{para}
2851: 
2852: In order to avoid destroying the apparent success between the standard
2853: theory  for BBN  and observation,  gravitinos must  have  an abundance
2854: $Y_{\widetilde G}$ below certain  upper limits, which crucially depend
2855: on  their decay  properties~\cite{oliveg,kohri}.   Some representative
2856: upper  bounds  on~$Y_{\widetilde  G}$,   obtained  in  a  very  recent
2857: analysis~\cite{kohri}, are
2858: \begin{equation}
2859:   \label{bYg}
2860: Y_{\widetilde G}\ \stackrel{<}{{}_\sim}\ \left\{\matrix{
2861: %\begin{array}{rl}
2862: 10^{-15}\hfill ,  & \mbox{for}~~m_{\widetilde G}\simeq360~{\rm GeV},
2863: \hfill \cr
2864: %
2865: 10^{-14}\hfill ,  &\mbox{for}~~ m_{\widetilde G}\simeq600~{\rm GeV},
2866: \hfill \cr
2867: %
2868: 10^{-13}\hfill ,  &\mbox{for}~~ m_{\widetilde G}\simeq7.5~{\rm TeV},
2869: \hfill \cr
2870: %
2871: 10^{-12}\hfill ,  &\mbox{for}~~m_{\widetilde G}\simeq9.3~{\rm TeV}\; .
2872: \hfill \cr}
2873: %\end{array}
2874: \right.
2875: \end{equation}
2876: The above bounds  pertain to the less restrictive  case of a gravitino
2877: that  decays with a  small branching  ratio $B_h=0.001$  into hadronic
2878: modes.    For   the   $g$-DBD   case  discussed   above,   the   upper
2879: limits~(\ref{bYg}) imply the corresponding stringent bounds on $T_{\rm
2880: reh}$:
2881: \begin{equation}
2882:   \label{bTr}
2883: T_{\rm reh}\ \stackrel{<}{{}_\sim}\ \left\{\matrix{
2884: %\begin{array}{rl}
2885: 9\times10^{6}~{\rm GeV}\hfill ,  & \mbox{for}~~m_{\widetilde
2886: G}\simeq360~{\rm GeV}\,, \hfill \cr
2887: %
2888: 7\times10^{7}~{\rm GeV}\hfill ,  &\mbox{for}~~ m_{\widetilde
2889: G}\simeq600~{\rm GeV}\,, \hfill \cr
2890: %
2891: 7\times10^{8}~{\rm GeV}\hfill ,  &\mbox{for}~~ m_{\widetilde
2892: G}\simeq7.5~{\rm TeV}\,, \hfill \cr
2893: %
2894: 7\times10^{9}~{\rm GeV}\hfill ,  &\mbox{for}~~m_{\widetilde
2895: G}\simeq9.3~{\rm TeV}\; . \hfill \cr}
2896: %\end{array}
2897: \right.
2898: \end{equation}
2899: 
2900: 
2901: The  aforementioned upper limits  lead to  serious constraints  on the
2902: basic couplings  $\kappa$, $\lambda$ and $\rho$,  usually forcing them
2903: to  acquire  very  small  values,  i.e.~$\kappa.\,  \lambda,  \,  \rho
2904: \stackrel{<}{{}_\sim}  10^{-5}$.   For  the standard  $F$-term  hybrid
2905: model within  mSUGRA and with  a soft SUSY-breaking  tadpole parameter
2906: $\mbox{a}_S =  1$~TeV, the requirement of accounting  for the observed
2907: power spectrum  $P_{\cal R}$, with  a number of $e$-folds  ${\cal N}_e
2908: =50$--60, implies  that $\kappa> 10^{-4}$ and $T_\kappa  = T_{\rm reh}
2909: \stackrel{>}{{}_\sim} 9  \times 10^{9}$~GeV.  Such a  high lower bound
2910: on $T_{\rm reh}$ invalidates  all the limits presented in~(\ref{bTr}),
2911: thereby ruling out the above $F$-term hybrid model.
2912: 
2913: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2914: \begin{figure}[t]
2915: \begin{center}
2916: \epsfig{file=Ygm.eps,height=6.4in,angle=-90} \\[3mm]
2917: \end{center}
2918: \caption{\sl\small  The dependence  of $\log  Y_{\widetilde  G}$ on
2919: $m_{\rm    FI}/M$,   for    $\kappa=10^{-2}$    (dark   grey    line),
2920: $\kappa=10^{-3}$   (grey  line)   and  $\kappa=10^{-4}$   (light  grey
2921: line). }\label{fig:Ygm}
2922: \end{figure}
2923: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2924: 
2925: 
2926: The above  situation, however,  changes drastically in  the $F_D$-term
2927: hybrid model with small $D$-parity violation, e.g.~due to the presence
2928: of a  subdominant FI $D$-term.   This corresponds to the  $g$-DAD case
2929: described    in   the   previous    subsection,   where    the   upper
2930: bounds~(\ref{bYg})  translate,  by  means  of~(\ref{Yg}),  into  upper
2931: bounds on $m_{\rm FI}/M$ for $\kappa > 8\times 10^{-5}$.  The required
2932: size of the $D$-parity violating parameter $m_{\rm FI}$ may naively be
2933: estimated using a relation very analogous to (\ref{Tkappa}), viz.\
2934: \begin{equation}
2935:   \label{Tg}
2936: T_g\ =\ \left( \frac{90}{\pi^2\, g_*}\right)^{1/4}\,
2937: \sqrt{\Gamma_g\: m_{\rm Pl} }\ ,
2938: \end{equation}
2939: where $\Gamma_g$  is the decay width  of a $g$-sector  particle and is
2940: given in~(\ref{GammaR}).  If we solve~(\ref{Tg}) for the ratio $m_{\rm
2941: FI}/M$, we obtain
2942: \begin{equation}
2943:   \label{ratio:mFI:M}
2944: \frac{m_{\rm FI}}{M} \approx \ 8.4 \cdot 10^{-4}\times \left(
2945: \frac{0.5}{g}\right)^{3/4} \left(\frac{T_g}{10^9~{\rm
2946: GeV}}\right)^{1/2}\, \left( \frac{10^{16}~{\rm
2947: GeV}}{M}\right)^{1/4}\; .
2948: \end{equation}
2949: For  second  reheat   temperatures  $T_g$  of  cosmological  interest,
2950: i.e.~$0.2~{\rm  TeV}  \stackrel{<}{{}_\sim} T_g  \stackrel{<}{{}_\sim}
2951: 10^9~{\rm GeV}$, the following double inequality for $M = 10^{16}$~GeV
2952: may be derived:
2953: \begin{equation}
2954:   \label{FIcombined}
2955: 4\times 10^{-7}\ \stackrel{<}{{}_\sim}\ \frac{m_{\rm FI}}{M}\
2956: \stackrel{<}{{}_\sim}\ 10^{-3}\; .
2957: \end{equation}
2958: The  lower  bounds  on  $T_g$  and  $m_{\rm  FI}/M$  result  from  the
2959: requirement  that thermal  electroweak-scale resonant  leptogenesis be
2960: successfully realized. More details are given in Section~\ref{BAU}.
2961: 
2962: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2963: \begin{figure}[t]
2964: \begin{center}
2965: \epsfig{file=km.eps,height=6.4in,angle=-90}
2966: \end{center}
2967: \caption{\sl\small The allowed region on the $(m_{\rm FI}/M,\kappa)$
2968: plane for $Y_{\widetilde G} < 10^{-15}$ (black area), $Y_{\widetilde
2969: G}<10^{-14}$ (light grey area), $Y_{\widetilde G}<10^{-13}$ (grey
2970: area) and $Y_{\widetilde G} < 10^{-12}$ (dark grey area).}\label{fig:mFI}
2971: \end{figure}
2972: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2973: 
2974: 
2975: A  numerical  analysis  of  the gravitino  abundance  predictions  and
2976: constraints  related to  the $g$-DAD  scenario has  been  performed in
2977: Figs.~\ref{fig:Ygm}  and \ref{fig:mFI},  respectively.   Our numerical
2978: results apply equally  well to both mSUGRA and  nmSUGRA scenarios.  In
2979: detail, Fig.~\ref{fig:Ygm} shows $\log Y_{\widetilde G}$ as a function
2980: of $m_{\rm  FI}/M$ and  $T_g$, for the  different values of  $\kappa =
2981: 10^{-4},\,  10^{-3}\,,  10^{-2}$, while  $M$  is  fixed  by the  usual
2982: inflationary constraints on ${\cal N}_e$ and $P_{\cal R}$, for $\kappa
2983: =  \lambda   =  \rho$  and  $c_H   =  0$.   The   different  lines  in
2984: Fig.~\ref{fig:Ygm} terminate  at high values of  $m_{\rm FI}/M$, since
2985: the inequality $T_g < T_\kappa$ does no longer hold.  The lowest value
2986: of  $m_{\rm FI}/M$  is determined  by the  condition $T_g  > 200$~GeV,
2987: which  results  from   the  aforementioned  requirement  that  thermal
2988: electroweak-scale     resonant     leptogenesis    is     successfully
2989: realized~\cite{APRD,APRL,PU2}.
2990: 
2991: 
2992: In Fig.~\ref{fig:Ygm},  we also observe  the two regimes:  $g$-DBD and
2993: $g$-DAD.   In  the $g$-DBD  regime,  $\log  Y_{\widetilde G}$  remains
2994: constant  for given  $\kappa$ up  to some  value $m_{\rm  FI}/M$.  For
2995: example, for  $\kappa = 10^{-3}$, $\log Y_{\widetilde  G}$ is constant
2996: for  $m_{\rm  FI}/M  \stackrel{>}{{}_\sim}  10^{-4}$. This  result  is
2997: consistent with~(\ref{Yk}). For smaller  values of $m_{\rm FI}/M$, one
2998: enters  the $g$-DAD  regime.  In  this case,  $\log  Y_{\widetilde G}$
2999: decreases   rapidly,  as  $m_{\rm   FI}/M$,  or   equivalently  $T_g$,
3000: decreases.   This  behaviour  of  $Y_{\widetilde G}$  is  expected  on
3001: account  of~(\ref{Yg}).   Also,   in  agreement  with~(\ref{Yg}),  the
3002: reduction of~$Y_{\widetilde G}$ becomes more drastic for larger values
3003: of $\kappa$.
3004: 
3005: In  Fig.~\ref{fig:mFI}  we  delineate   the  allowed  regions  on  the
3006: $(\kappa,   m_{\rm  FI}/M)$   plane   for  the   discrete  values   of
3007: $Y_{\widetilde  G}  =  10^{-15},  10^{-14}, 10^{-13},  10^{-12}$,  for
3008: $\kappa \geq  8 \times10^{-5}$.  The  upper boundaries of  the various
3009: areas are obtained using (\ref{Yg}).  For $\kappa < 8 \times 10^{-5}$,
3010: we  are   in  the  $g$-DBD   region,  where  we  obtain   $10^{-13}  <
3011: Y_{\widetilde  G}   <  10^{-12}$,  almost   independently  of  $m_{\rm
3012: FI}/M$~[cf.~(\ref{Yk})]. Therefore, we only display values for $m_{\rm
3013: FI}/M$, for which $g$-DAD becomes  relevant.  We observe that the most
3014: stringent  limit on  $Y_{\widetilde  G}$ can  still  be fulfilled  for
3015: $\kappa    \stackrel{>}{{}_\sim}    10^{-2}$    and   $m_{\rm    FI}/M
3016: \stackrel{<}{{}_\sim}  10^{-6}$. Such large  values of  $\kappa$ would
3017: have been excluded from  naive estimates of the $\kappa$-sector reheat
3018: temperature  $T_\kappa$ due  to the  $\kappa$-sector  particle decays.
3019: According to our analysis in this section, however, these large values
3020: of $\kappa$,  $\lambda$ and  $\rho$ of order  $10^{-2}$--$10^{-1}$ are
3021: allowed within  the $F_D$-term hybrid inflationary model.   As we will
3022: see  in  the  next section,  this  is  a  distinctive feature  of  the
3023: $F_D$-term hybrid  model that opens up novel  possibilities in solving
3024: the CDM problem.
3025: 
3026: At  the  end  of this  section,  we  wish  to  comment on  a  possible
3027: $F_D$-term  hybrid scenario, where  the $\kappa$-sector  particles can
3028: decay  directly  into  the  $g$-sector  ones.  This  can  happen,  for
3029: example, if $m_\kappa > 2 m_g$ or equivalently when $\kappa > \sqrt{2}
3030: g$.   Since the gauge  coupling $g$  of the  waterfall sector  must be
3031: smaller than 0.1  in this case, it would be difficult  to embed such a
3032: $F_D$-term hybrid scenario into  a GUT. The energy density transferred
3033: from  the $\kappa$-sector particles  into the  $g$-sector ones  may be
3034: calculated by
3035: \begin{equation}
3036: \frac{\rho_g}{\rho_\kappa}\  =\ \frac{g}{\sqrt{2}\, \kappa}\ B_{\kappa
3037: \to g}\ .
3038: \end{equation}
3039: Here $B_{\kappa \to  g}$ denotes the branching ratio  of the decays of
3040: the  $\kappa$- to $g$-sector  particles. Assuming  conservatively that
3041: $B_{\kappa  \to g}  \sim 10^{-2}$  and $\kappa  = 2  g$, we  obtain an
3042: estimate for the gravitino abundance $Y_{\widetilde{G}} \sim 10^{-18}$
3043: for  $m_{\rm FI}/M  \stackrel{<}{{}_\sim} 10^{-6}$,  thereby rendering
3044: gravitinos quite harmless.
3045: 
3046: 
3047: 
3048: \setcounter{equation}{0}
3049: \section{Baryon Asymmetry and Cold Dark Matter}\label{BAU}
3050: 
3051: In this  section we briefly discuss  further cosmological implications
3052: of the $F_D$-term hybrid model for the BAU and the CDM.
3053: 
3054: 
3055: \subsection{Resonant Flavour-Leptogenesis at the Electroweak Scale}
3056: 
3057: Earlier  studies  of  the  BAU  in  supersymmetric  models  of  hybrid
3058: inflation  have  mainly  been  focused  on  scenarios  of  non-thermal
3059: leptogenesis~\cite{LS}, with  an hierarchical heavy  Majorana neutrino
3060: spectrum, e.g.~$m_{N_1} < m_{N_2}  \ll m_{N_3}$. The simplest model of
3061: this type  is obtained by  identifying the waterfall gauge  group with
3062: U(1)$_{B-L}$, which allows the presence of the operator $\gamma_{ij}\,
3063: \widehat{X}_2 \widehat{X}_2 \widehat{N}_i \widehat{N_j}/m_{\rm Pl}$ in
3064: the  superpotential. Notice  that  such  a term  is  forbidden in  the
3065: $F_D$-term hybrid model by virtue of the $R$ symmetry~(\ref{RFD}).  In
3066: the non-thermal leptogenesis  model, the reheat temperature consistent
3067: with  the observed BAU  $\eta_B =  6.1\times 10^{-10}$  and low-energy
3068: neutrino data is estimated to be~\cite{SS}
3069: \begin{equation}
3070:   \label{Tnonth}
3071: T_{\rm reh}\ \stackrel{>}{{}_\sim}\  2.5\cdot 10^7~{\rm GeV} \times
3072: \Bigg( \frac{10^{16}~{\rm GeV}}{M}\Bigg)^{1/2}\,
3073: \Bigg(\frac{\kappa}{10^{-5}}\Bigg)^{3/4}\ ,
3074: \end{equation}
3075: where the superpotential couplings  $\lambda,\, \rho$ are set to zero.
3076: If  $\lambda =  \kappa$,  the lower  bound~(\ref{Tnonth}) gets  larger
3077: roughly  by  a  factor  20.   It  is  obvious  that  in  this  generic
3078: non-thermal leptogenesis scenario, the gravitino constraint on $T_{\rm
3079: reh}$  favours   rather  small  values  of   $\kappa$  and  $\lambda$,
3080: e.g.~$\kappa,\,  \lambda  \stackrel{<}{{}_\sim}  10^{-5}$ for  $T_{\rm
3081: reh}   \stackrel{<}{{}_\sim}   10^8$~GeV.    As   was   discussed   in
3082: Section~\ref{inflation}.2,  however,  such  small values  of  $\kappa$
3083: introduce strong tuning  at a less than 1\% level  to the horizon exit
3084: values of the  inflaton field $\phi_{\rm exit}$ in  a nmSUGRA scenario
3085: that accounts  for the recently  observed value of the  spectral index
3086: $n_{\rm  s}$ given  by~(\ref{nswmap}).   Moreover, the  success of  this
3087: scenario  relies heavily  on the  assumption  that there  is no  other
3088: source of baryogenesis,  e.g.~through the Affleck--Dine mechanism, nor
3089: of entropy release, e.g.~from possible late decays of moduli or flaton
3090: fields~\cite{DLES},  between the energy  scales $m_{N_1}\  (\gg T_{\rm
3091: reh})$ and the electroweak phase transition.
3092: 
3093: In  the  $F_D$-term  hybrid  model, non-thermal  leptogenesis  is  not
3094: possible for one  of the reasons mentioned above.   The late decays of
3095: the  $g$-sector  ($D$-odd) particles  generally  lead  to an  enormous
3096: entropy release,  so that  not only gravitinos,  but also  any initial
3097: lepton-number excess will be  diluted to unobservable values. However,
3098: as has  already been discussed in~\cite{GP}, the  $F_D$-term model can
3099: realize  electroweak-scale  resonant  leptogenesis~\cite{PU2}, if  the
3100: coupling of  the inflaton  superfield $\widehat{S}$ to  the respective
3101: right-handed   neutrinos   $\widehat{N}_i$  is   very   close  to   an
3102: SO(3)-symmetric form, i.e.~$\rho_{ij} \approx \rho\, {\bf 1}_3$.  This
3103: will give rise to 3 nearly  heavy Majorana neutrinos of mass $m_N$ and
3104: so would enable a  successful realization of the resonant leptogenesis
3105: mechanism at the electroweak  scale.  The required SO(3)-breaking may,
3106: for  example, originate from  renormalization-group (RG)~\cite{Branco}
3107: or possible GUT threshold effects~\cite{PU2,Lindner}.
3108: 
3109: 
3110: 
3111: \begin{figure}[t]
3112: \begin{center}
3113: \begin{picture}(600,320)
3114: \put(50,0){\includegraphics[scale=0.55]{initialconds-250.eps}}
3115: \end{picture}
3116: \end{center}
3117: \caption{\sl\small Numerical estimates of  the BAU for a scenario with
3118:   $m_N = 250$~GeV and  for different initial lepton- and baryon-number
3119:   abundances, $\eta^{\rm in}_{L_l}$ and $\eta^{\rm in}_B$, assuming an
3120:   initial  thermal  distribution  for  the heavy  Majorana  neutrinos,
3121:   i.e.~$\eta^{\rm  in}_{N_{1,2,3}}  = 1$.   The  horizontal grey  line
3122:   shows   the   BAU   needed    to   agree   with   today's   observed
3123:   value.}\label{mN250GeV}
3124: \end{figure}
3125: 
3126: 
3127: An order  of magnitude estimate  of the final BAU  $\eta_B$, including
3128: single lepton flavour effects, may be obtained as~\cite{APRL,PU2}
3129: \begin{equation}
3130:   \label{Bestimate}
3131: \eta_B\ \sim\ -\, 10^{-2}\,\times\, r(T_g/m_N)\
3132:  \sum_{l=1}^3\, \sum_{N_i}\:
3133: \delta^l_{N_i}\: \frac{K^l_{N_i}}{K_l\,K_{N_i}}\ ,
3134: \end{equation}
3135: where
3136: \begin{equation}
3137:   \label{KlN}
3138: K^l_{N_i}\ =\ \frac{\Gamma (N_i \to L_l \Phi)\: +\:
3139: \Gamma(N_i \to L^C_l \Phi^\dagger)}{H (T=m_N)}
3140: \end{equation}
3141: is a  lepton-flavour dependent wash-out factor, which  quantifies in a
3142: way the degree of in- or  out-of-equilibrium of the decay rates of the
3143: heavy Majorana neutrino mass eigenstates  $N_i$ ($i = 1,2,3$) into the
3144: SM-like Higgs  doublet $\Phi$ and  the lepton doublet  $L_l$ ($l=e,\mu
3145: ,\tau$).  The  remaining $K$-factors in~(\ref{Bestimate})  are defined
3146: with the help of $K^{l}_{N_i}$ as follows:
3147: \begin{eqnarray}
3148:   \label{Kfactors}
3149: K_{N_i} \!&=&\!  \sum_{l = 1}^3\ K^{l}_{N_i}\;,\qquad
3150: K_l \ =\ \sum_{N_i}\, K^{l}_{N_i}\; .
3151: \end{eqnarray}
3152: The  parameters $\delta^l_{N_i}$  denote the  different lepton-flavour
3153: asymmetries related to  the decays $N_i \to L_l  \Phi$ and are defined
3154: by
3155: \begin{equation}
3156: \delta^l_{N_i}\ =\ \frac{\Gamma (N_i \to L_l \Phi)\: -\:
3157: \Gamma(N_i \to L^C_l \Phi^\dagger)}{\Gamma (N_i \to L_l \Phi)\: +\:
3158: \Gamma(N_i \to L^C_l \Phi^\dagger)} \ .
3159: \end{equation}
3160: Finally, the prefactor $r(T_g/m_N)$ in~(\ref{Bestimate}) takes care of
3161: a possible  dilution effect  on the  BAU that might  be caused  by the
3162: entropy  release of  late  $g$-sector particle  decays. This  dilution
3163: effect is  only relevant,  if the second  reheat temperature  $T_g$ is
3164: smaller  than  the   leptogenesis  scale  $m_N$.   Employing  standard
3165: arguments of thermodynamics, one may estimate that
3166: \begin{equation}
3167: r (T_g / m_N)\ \sim \ \Bigg(\frac{T_g}{m_N}\Bigg)^5\; .
3168: \end{equation}
3169: Instead,  if   $T_g  \gg   m_N$,  the  dilution   factor  $r(T_g/m_N)$
3170: approaches~1 and can therefore be omitted.
3171: 
3172: In Fig.~\ref{mN250GeV}, we display  numerical estimates of the BAU for
3173: a resonant leptogenesis scenario with  $m_N = 250$~GeV and an inverted
3174: hierarchical  light-neutrino spectrum.  For  a detailed  discussion of
3175: the  heavy and light  neutrino spectra  of this  model, the  reader is
3176: referred to~\cite{PU2}.  As can  be seen from Fig.~\ref{mN250GeV}, one
3177: advantageous feature of resonant leptogenesis is that the final baryon
3178: asymmetry~$\eta_B$  does not  sensitively depend  on  any pre-existing
3179: lepton-   or  baryon-number   abundance,   $\eta^{\rm  in}_{L_l}$   or
3180: $\eta^{\rm  in}_B$.    For  instance,  assuming   an  initial  thermal
3181: distribution  for   the  heavy  Majorana   neutrinos,  i.e.~$\eta^{\rm
3182: in}_{N_{1,2,3}}  = 1$,  and primordial  baryon  asymmetries $\eta^{\rm
3183: in}_B  \stackrel{<}{{}_\sim}  10^{-2}$,  we  observe  that  the  final
3184: $\eta_B$ is  practically independent  of the initial  conditions, once
3185: the  relevant particle-physics  model  parameters, such  as the  heavy
3186: Majorana masses and their respective Yukawa couplings, are fixed.
3187: 
3188: It is important to comment here on the fact that the above property of
3189: the  independence  of the  BAU  on  the  initial conditions  does  not
3190: necessarily  get  spoiled,  if  the second  reheat  temperature  $T_g$
3191: happens to be smaller than  the resonant leptogenesis scale $m_N$.  In
3192: this  case, one  only needs  to make  sure that  the  entropy dilution
3193: suppression factor  $\sim (T_g/m_N)^5$ does not lead  to a significant
3194: reduction  of  the  BAU.   Therefore, we  have  rather  conservatively
3195: assumed throughout our numerical analysis in Section~\ref{reheat} that
3196: $T_g \stackrel{>}{{}_\sim} m_N \sim  250$~GeV, even though $T_g$ could
3197: still be somewhat smaller than the resonant leptogenesis scale $m_N$.
3198: 
3199: Another  point that  deserves to  be  clarified here  is the  physical
3200: significance of lepton-flavour effects  on the BAU.  In general, there
3201: are  two sources  of lepton  flavour:  (i) the  charged lepton  Yukawa
3202: couplings $h_l$  and (ii) the neutrino  Yukawa couplings $h^\nu_{ij}$.
3203: The    former    has     been    extensively    discussed    in    the
3204: literature~\cite{BAULFV} and may affect the predictions for the BAU by
3205: up to one order of  magnitude, depending on the scale of leptogenesis.
3206: For our electroweak-scale leptogenesis scenario, these effects are not
3207: significant,  since  all   charged  lepton  Yukawa  couplings  mediate
3208: interactions that  are in thermal  equilibrium.  The second  source of
3209: flavour effects  is due to neutrino-Yukawa  couplings $h^\nu_{ij}$ and
3210: has  been  studied only  very  recently in~\cite{APRL,PU2,Bari}.   The
3211: effect on the  BAU is most relevant when  the heavy Majorana neutrinos
3212: get   closer   in  mass.    In   models   of  resonant   leptogenesis,
3213: neutrino-Yukawa  coupling effects can  have a  dramatic impact  on the
3214: predictions  for  the BAU,  enhancing  its  value  by many  orders  of
3215: magnitude~\cite{APRL,PU2}.
3216: 
3217: This last fact opens up  new vistas in the model-building of scenarios
3218: that   can  be  phenomenologically   more  accessible   to  laboratory
3219: experiments.   For   instance,  if  a  certain   hierarchy  among  the
3220: Yukawa-neutrino couplings  $h^\nu_{ij}$ is assumed, e.g.~$h^{\nu}_{i2}
3221: =  i  h^{\nu}_{i3}  \sim  10^{-2}  \sim  h_\tau$  and~$h^{\nu}_{i1}  =
3222: 10^{-6}$--$10^{-7} \sim h_e$,  resulting from the approximate breaking
3223: of  some global  U(1)$_l$  symmetry,  the required  BAU  can still  be
3224: generated  successfully from  an individual  lepton  number asymmetry,
3225: namely $L_\tau$ in  this case.  For this particular  model of resonant
3226: $\tau$-leptogenesis,  the   values  of  the   $K$-factors  defined  in
3227: (\ref{KlN}) are:
3228: \begin{equation}
3229: K^\tau_{N_{1,2,3}}\ \sim\ 10\,,\
3230: \qquad K^{e,\mu}_{N_3}\ \sim\ 30\,,
3231: \qquad K^{e,\mu}_{N_{1,2}}\ \sim\ 10^{10}\; .
3232: \end{equation}
3233: Given that the leptonic asymmetry is $\delta^\tau_{N_3} \sim 10^{-6}$,
3234: one  can  estimate from~(\ref{Bestimate})  that  the  right amount  of
3235: baryon asymmetry is produced, with~$\eta_B \sim 10^{-9}$. This is also
3236: shown in  Fig.~\ref{mN250GeV}.  Instead, older approaches  to BEs that
3237: do  not appropriately treat  lepton flavour  effects via  the neutrino
3238: Yukawa couplings $h^\nu_{ij}$ would  have predicted a value that would
3239: have been short of a huge factor $\sim 10^{-6}$~\cite{APRL,PU2}.
3240: 
3241: As can be  seen from the above example,  the lepton-flavour directions
3242: $L_{e,\mu}$ orthogonal  to $L_\tau$ can involve  large neutrino Yukawa
3243: couplings  of  order  $10^{-2}$.   Such  couplings can  give  rise  to
3244: distinctive    signatures   in   the    production   and    decay   of
3245: electroweak-scale heavy  Majorana neutrinos at  high-energy colliders,
3246: such  as the  LHC~\cite{NprodLHC}, the  International  Linear $e^+e^-$
3247: Collider~(ILC)~\cite{NprodILC}         and         other        future
3248: colliders~\cite{NprodEG}.  Moreover,  electroweak-scale heavy Majorana
3249: neutrinos can give  rise to phenomena of lepton  flavour and/or number
3250: violation,    such    as    the   neutrinoless    double-beta    decay
3251: ($0\nu\beta\beta$), the  decays $\mu \to  e\gamma$~\cite{CL}, $\mu \to
3252: eee$,      $\mu       \to      e$      conversion       in      nuclei
3253: etc~\cite{IP,LFVN,LFVrev,LLfit}.    A  detailed   discussion   of  the
3254: low-energy phenomenology of resonant  leptogenesis models may be found
3255: in~\cite{PU2}.
3256: 
3257: 
3258: \subsection{Thermal Right-Handed Sneutrinos as CDM}
3259: 
3260: 
3261: An  interesting  feature  of  the  $F_D$-term  hybrid  model  is  that
3262: $R$-parity is conserved, even though the lepton number $L$, as well as
3263: $B-L$,    are   explicitly   broken    by   the    Majorana   operator
3264: $\frac{1}{2}\,\rho\,  \widehat{S}  \widehat{N}_i  \widehat{N}_i$.   In
3265: fact, in  our model, all superpotential couplings  either conserve the
3266: $B-L$ number  or break it by  even number of units.   For example, the
3267: coupling $\rho$  breaks explicitly $L$, along with  $B-L$, by 2~units.
3268: Since the $R$-parity of  each superpotential operator is determined to
3269: be $R  = (-1)^{3(B-L)}  = +1$, the  $F_D$-term hybrid  model conserves
3270: $R$-parity.  As a  consequence, the LSP of the  spectrum is stable and
3271: so  becomes a  viable  candidate to  address  the CDM  problem of  the
3272: Universe.
3273: 
3274: 
3275: In addition to the standard  CDM candidates of the MSSM, e.g.~a stable
3276: neutralino,  it  would  be  interesting  to  explore  whether  thermal
3277: right-handed sneutrinos as LSPs could solve the CDM problem. Before we
3278: estimate  their   relic  abundance,   we  first  observe   that  light
3279: right-handed sneutrinos  may easily appear in  the spectrum.  Ignoring
3280: the small  neutrino-Yukawa coupling terms,  the right-handed sneutrino
3281: mass matrix  ${\cal M}^2_{\widetilde N}$  is written down in  the weak
3282: basis $(\widetilde{N}_{1,2,3}, \widetilde{N}^*_{1,2,3})$:
3283: \begin{equation}
3284: {\cal   M}^2_{\widetilde    N}\ =\ \frac{1}{2}\, \left(\! \begin{array}{cc}
3285: \rho^2 v^2_S\: +\: M^2_{\widetilde  N} & \rho A_\rho
3286: v_S\: +\: \rho\lambda v_u v_d\\
3287: \rho A^*_\rho v_S\: +\: \rho\lambda v_u v_d &
3288: \rho^2 v^2_S\: +\: M^2_{\widetilde  N}
3289: \end{array} \!\right)\; ,
3290: \end{equation}
3291: where $v_S =  \langle S \rangle$, $v_{u,d} =  \langle H_{u,d} \rangle$
3292: and  $M^2_{\widetilde N}$  is the  soft SUSY-breaking  mass parameters
3293: associated with the sneutrino fields. The sneutrino spectrum will then
3294: consist of 3 heavy (light) right-handed sneutrinos of mass
3295: $$\rho^2 v^2_S\:  +\: M^2_{\widetilde N}\  +\, (-)\ \Big(  \rho A_\rho
3296: v_S\: +\: \rho\lambda v_u v_d\Big)\, .$$ Hence, the 3 light sneutrinos
3297: can   act   as   LSPs,   which   we  collectively   denote   them   by
3298: $\widetilde{N}_{\rm LSP}$.
3299: 
3300: Recently, the possibility that right-handed sneutrinos are the CDM was
3301: considered  in~\cite{GGP}.  This recent  analysis showed  that thermal
3302: right-handed  sneutrinos have  rather high  relic abundances  and will
3303: generally overclose the Universe  in a supersymmetric extension of the
3304: MSSM with  right-handed neutrino superfields  $\widehat{N}_i$ and bare
3305: Majorana   masses  $(m_M)_{ij}   \widehat{N}_i   \widehat{N}_j$.   The
3306: underlying  reason  is  that  because  of  the  small  Yukawa-neutrino
3307: couplings $h^\nu_{ij}$, the  self- and co-annihilation interactions of
3308: the  sneutrino LSP  with itself  and other  MSSM particles  are rather
3309: weak.  These  weak processes  do not allow  the sneutrino LSP  to stay
3310: long enough in thermal  equilibrium before its freeze-out temperature,
3311: such that its  number density gets reduced to  a level compatible with
3312: the CMB  data, i.e.~$\Omega_{\rm DM} h^2 \approx  0.15$.  Instead, the
3313: predicted values turn  out to be many orders  of magnitude larger than
3314: 1.
3315: 
3316: In the  $F_D$-term hybrid model,  however, there is a  new interaction
3317: that can make the right-handed sneutrinos annihilate more efficiently.
3318: This is  the quartic coupling~\footnote{The implications  of a generic
3319: quartic coupling of the same  form for the CDM abundance and detection
3320: was studied earlier  in~\cite{mcdonald,pospelov} within the context of
3321: a simple non-SUSY model. These studies will not be directly applicable
3322: to  our  more  elaborate   case  of  a  supersymmetric  scenario  with
3323: right-handed sneutrinos.  However, we have used their results to check
3324: our qualitative estimates for the CDM abundance.}
3325: \begin{equation}
3326:   \label{Llsp}
3327: {\cal L}^{\rm LSP}_{\rm int}\ =\
3328: \frac{1}{2}\,\lambda \rho\,  \widetilde{N}^*_i \widetilde{N}^*_i H_u
3329:   H_d\quad  +\quad  {\rm H.c.}
3330: \end{equation}
3331: It  results  from  the  $F$-term  of the  inflaton  field:  $F_S  \sim
3332: \frac{1}{2}\,\rho    \widehat{N}_i    \widehat{N}_i\:   +\:    \lambda
3333: \widehat{H}_u  \widehat{H}_d$.   To  assess  the significance  of  the
3334: interaction~(\ref{Llsp}),   we   estimate   the   relic   density   of
3335: $\widetilde{N}_{\rm LSP}$ in different kinematic regions.
3336: 
3337: We   first  consider   the   self-annihilation  off-resonant   process
3338: $\widetilde{N}_{\rm   LSP}   \widetilde{N}_{\rm   LSP}   \to   \langle
3339: H_u\rangle\, H_d\to W^+W^-$,  which occurs when $m_{\widetilde{N}_{\rm
3340: LSP}} > M_W$. A simple estimate yields
3341: \begin{equation}
3342: \Omega_{\rm DM}\, h^2\ \sim\
3343: \Bigg(\frac{10^{-4}}{\rho^2\lambda^2}\Bigg)\;
3344: \Bigg(\frac{\tan\beta\, M_H}{g_w\, M_W}\Bigg)^2\, .
3345: \end{equation}
3346: To obtain an acceptable relic density, we need relatively large $\rho$
3347: and  $\lambda$ couplings, i.e.~$\rho,\,  \lambda \stackrel{>}{{}_\sim}
3348: 0.1$\footnote{An  upper bound on  the product  $\rho\lambda$, although
3349: somewhat model-dependent,  can be derived from  experimental limits on
3350: the  flux  of  energetic  upward  muons that  occur  in  the  possible
3351: detection  of  CDM  using  neutrino telescopes  \cite{pospelov}.   Our
3352: initial estimates indicate that it should be $\rho\lambda\lesssim0.03$
3353: for $m_{\widetilde N_{\rm LSP}}\sim 50~{\rm GeV}$, which is not a very
3354: rectrictive bound.}.  Such values  go in opposite direction with those
3355: obtained   by  requiring  successful   inflation  with   a  red-tilted
3356: spectrum. Therefore,  as far as  inflation is concerned,  they signify
3357: the necessity of going well beyond the minimal K\"ahler potential.
3358: 
3359: The  above situation may  slightly improve  for $m_{\widetilde{N}_{\rm
3360: LSP}} <  M_W$, in large $\tan\beta$ scenarios  with light Higgs bosons
3361: that couple appreciably to $b$-quarks~\cite{Sabine}. In particular, in
3362: the kinematic region  $M_{H_d} \approx 2 m_{\widetilde{N}_{\rm LSP}}$,
3363: the     self-annihilation     process     $\widetilde{N}_{\rm     LSP}
3364: \widetilde{N}_{\rm  LSP}  \to  \langle H_u\rangle\,  H_d\to  b\bar{b}$
3365: becomes resonant, and the above estimate modifies to
3366: \begin{equation}
3367: \Omega_{\rm   DM}\,   h^2\    \sim\
3368: 10^{-4}\times   B^{-1}(H_d   \to \widetilde{N}_{\rm LSP}
3369:  \widetilde{N}_{\rm LSP})\, \times\,
3370: \Bigg(\frac{M_H}{100~{\rm GeV}}\Bigg)^2\, .
3371: \end{equation}
3372: Consequently, if the  couplings $\lambda ,\, \rho$ are  not too small,
3373: e.g.~$\lambda,\,   \rho   \stackrel{>}{{}_\sim}   10^{-2}$,  the   LSP
3374: right-handed  sneutrinos  $\widetilde{N}_{\rm  LSP}$  can  efficiently
3375: annihilate via  a Higgs  resonance $H_d$ into  pairs of  $b$-quarks in
3376: this kinematic region, thus  obtaining a relic density compatible with
3377: the CMB data.  A~detailed study of the  thermal right-handed sneutrino
3378: as CDM could be given elsewhere.
3379: 
3380: 
3381: 
3382: \setcounter{equation}{0}
3383: \section{Conclusions}\label{conclusions}
3384: 
3385: We  have analyzed the  cosmological implications  of a  novel $F$-term
3386: hybrid  inflationary  model, in  which  the  inflaton  and the  gauged
3387: waterfall sectors  respect an  approximate discrete symmetry  which we
3388: called here $D$-parity.  The approximate breaking of $D$-parity occurs
3389: explicitly either through the presence of a subdominant FI $D$-term or
3390: through  non-renormalizable operators in  the K\"ahler  potential. For
3391: brevity,  this  scenario of  inflation  was  termed $F_D$-term  hybrid
3392: inflation.  One of the most  interesting features of the model is that
3393: the VEV of  the inflaton field closely relates  the $\mu$-parameter of
3394: the MSSM to an SO(3)  symmetric Majorana mass $m_N$.  If $\lambda \sim
3395: \rho$,  this implies  that $\mu  \sim m_N$,  so the  $F_D$-term hybrid
3396: model may naturally predict lepton-number violation at the electroweak
3397: scale.
3398: 
3399: Before summarizing the  cosmological and particle-physics implications
3400: of the  $F_D$-term hybrid model, it  might be interesting  to list our
3401: basic assumptions pertinent to inflation and to the model itself:
3402: 
3403: \begin{itemize}
3404: 
3405: \item[ (i)] The standard assumption for successful hybrid inflation is
3406:   that the  inflaton field  $\phi$ should be  displaced from  its true
3407:   minimum at the  start of inflation, whereas all  other scalar fields
3408:   in the  spectrum must have  zero VEVs [c.f.~(\ref{initial})].   In a
3409:   nmSUGRA scenario of hybrid  inflation, however, additional tuning is
3410:   required  beyond the  above standard  assumption.  The  horizon exit
3411:   values of the  inflaton field $\phi_{\rm exit}$ have  to be close to
3412:   the value $\phi_{\rm max}$,  at which the inflationary potential has
3413:   a  maximum.    Nevertheless,  such  a  tuning  is   not  so  strong,
3414:   i.e.~$(\phi_{\max}     -     \phi_{\rm    exit})/\phi_{\rm     exit}
3415:   \stackrel{>}{{}_\sim}     10\%$,      as     long     as     $\kappa
3416:   \stackrel{>}{{}_\sim} 10^{-3}$.
3417: 
3418: \item[(ii)]  Although  there  may  exist  several  ways  of  breaking
3419:   $D$-parity explicitly, we have  considered here two possibilities to
3420:   motivate  the required  small  amount of  $D$-parity violation.   As
3421:   discussed  in  Appendix  A,  we  first  considered  the  case  where
3422:   $D$-parity is broken by a  subdominant FI $D$-term, which is induced
3423:   radiatively after heavy degrees of freedom have been integrated out.
3424:   Another  minimal  way   would  be  to  introduce  non-renormalizable
3425:   operators   in  the   K\"ahler  potential   that   break  $D$-parity
3426:   explicitly.
3427: 
3428: \item[  (iii)]  In  order  to  be  able  to  realize  thermal  resonant
3429:   leptogenesis  at a  low scale,  the coupling  matrix  $\rho_{ij}$ is
3430:   assumed  to be  close  to SO(3)  symmetric, i.e.~$\rho_{ij}  \approx
3431:   \rho\, {\bf 1}_3$.
3432: 
3433: \end{itemize}
3434: 
3435: 
3436: The $F_D$-term hybrid model has several cosmological implications that
3437: may be summarized as follows:
3438: 
3439: \begin{itemize}
3440: 
3441: \item  The  model  can   accommodate  the  currently  favoured  strong
3442:   red-tilted    spectrum    with    $n_{\rm    s}    -    1    \approx
3443:   -0.05$~\cite{WMAP3,Lyman}, if the  radiative corrections dominate the
3444:   slope of  the inflationary potential and  a next-to-minimal K\"ahler
3445:   potential  is assumed,  where the  parameter $c_H$  is in  the range
3446:   0.05--0.2.   The radiative  corrections  dominate the  slope of  the
3447:   potential,  if  the  superpotential couplings,  $\kappa,\  \lambda,\
3448:   \rho$,  lie in  a certain  interval:  $10^{-4} \stackrel{<}{{}_\sim}
3449:   \kappa,\   \lambda,\   \rho   \stackrel{<}{{}_\sim}  10^{-2}$.    In
3450:   addition, the  actual value of  the power spectrum $P_{\cal  R}$ and
3451:   the required  number of $e$-folds, ${\cal N}_e  \approx 55$, provide
3452:   further constraints on these couplings  and the SSB scale $M$ of the
3453:   waterfall gauge symmetry.   For example, for $M\approx 10^{16}$~GeV,
3454:   one finds the allowed parameter space: $\kappa \stackrel{<}{{}_\sim}
3455:   \lambda,\  \rho  \stackrel{<}{{}_\sim}  4\kappa$, for  $\kappa  \sim
3456:   10^{-3}$--$10^{-2}$    and     $0.05    \stackrel{<}{{}_\sim}    c_H
3457:   \stackrel{<}{{}_\sim} 0.1$.
3458: 
3459: \item For $F_D$-term hybrid  models with spontaneously broken U(1)$_X$
3460:   gauge symmetry, the non-observation  of a cosmic string contribution
3461:   to the  power spectrum at the  10\% level implies an  upper bound on
3462:   the      superpotential     coupling      $\kappa$,     i.e.~$\kappa
3463:   \stackrel{<}{{}_\sim} 10^{-3}$. This  strict upper bound on $\kappa$
3464:   can be  weakened by  one order  of magnitude in  a nmSUGRA  model of
3465:   $F_D$-term hybrid inflation, with $\kappa = \lambda = \rho$ and $c_H
3466:   =  0.14$.  On the  other hand,  this upper  limit can  be completely
3467:   evaded,  if  the watefall  sector  of  the  $F_D$-term hybrid  model
3468:   realizes an  SU(2)$_X$ local symmetry that breaks  completely to the
3469:   identity ${\bf I}$,  i.e.~SU(2)$_X \to {\bf I}$.  In  this case, not
3470:   only  cosmic  strings  but  any  other topological  defects  can  be
3471:   avoided,  such  as  monopoles  and  textures.   As  we  outlined  in
3472:   Section~\ref{FDmodel}, GUTs, such as  those based on the exceptional
3473:   groups E(6) and E(7),  have breaking patterns that contain SU(2)$_X$
3474:   subgroups  uncharged under the  SM gauge  group and  so are  able to
3475:   realize $F_D$-term  hybrid inflation devoid of  monopoles and cosmic
3476:   strings.
3477: 
3478: 
3479: \item To  avoid overproduction  of gravitinos, one  needs to  impose a
3480:   strict upper limit on  the reheat temperature $T_{\rm reh}$ obtained
3481:   from   the   perturbative   inflaton   decays,   i.e.~$T_{\rm   reh}
3482:   \stackrel{<}{{}_\sim} 10^{10}$--$10^7$~GeV. This upper bound depends
3483:   on the  decay properties  of the gravitino  and gives rise  to tight
3484:   constraints on the  basic theoretical parameters $\kappa$, $\lambda$
3485:   and $\rho$,  i.e.~$\kappa,\, \lambda,\, \rho\ \stackrel{<}{{}_\sim}\
3486:   10^{-5}$. However,  these tight limits may  be significantly relaxed
3487:   by considering the late decays of the so-called $g$-sector particles
3488:   which are  induced by small $D$-parity violating  couplings that may
3489:   result from  either a subdominant FI  $D$-term or non-renormalizable
3490:   K\"ahler potential  terms.  These $g$-sector  particles are produced
3491:   during the  preheating epoch,  and if they  are abundant,  they will
3492:   lead  to a  second reheating  phase in  the evolution  of  the early
3493:   Universe, giving  rise to a  rather low reheat temperature,  even as
3494:   low as $0.3$~TeV.   In this case, the enormous  entropy release from
3495:   the  $g$-sector   particles  may  reduce   the  gravitino  abundance
3496:   $Y_{\widetilde{G}}$    below   the    BBN   limits    discussed   in
3497:   Section~\ref{reheat}.
3498: 
3499: 
3500: \item After the inflaton $S$ receives a VEV, one ends up with 3 nearly
3501:   degenerate heavy  Majorana neutrinos with masses  at the electroweak
3502:   scale.   As we  discussed in  Section~\ref{BAU}, this  opens  up the
3503:   possibility  to  successfully address  the  BAU  within the  thermal
3504:   electroweak-scale   resonant  leptogenesis   framework,  in   a  way
3505:   independent of any pre-existing lepton- or baryon-number abundance.
3506: 
3507: 
3508: \item The  $F_D$-term hybrid model  conserves $R$-parity, in  spite of
3509:   the fact that the lepton number is explicitly broken by the Majorana
3510:   operator    $\frac{1}{2}\,    \rho\,    \widehat{S}    \widehat{N}_i
3511:   \widehat{N}_i$.   This is so,  because all  superpotential couplings
3512:   either  conserve the  $B-L$ number  or break  it by  even  number of
3513:   units. The  aforementioned Majorana operator  breaks explicitly $L$,
3514:   as well as $B-L$, by 2~units.  Consequently, the LSP of the spectrum
3515:   is stable and so qualifies  as candidate to address the CDM problem.
3516:   The  new aspect  of  the  $F_D$-term hybrid  model  is that  thermal
3517:   right-handed  sneutrinos  emerge as  new  candidates  to solve  this
3518:   problem,  by virtue of  the quartic  coupling: $\frac{1}{2}\,\lambda
3519:   \rho\, \widetilde{N}^*_i \widetilde{N}^*_i  H_u H_d\ +\ {\rm H.c.}$.
3520:   This new  quartic coupling results  in the Higgs potential  from the
3521:   $F$-terms of the  inflaton field, and it is not  present in the more
3522:   often-discussed extension  of the MSSM,  where right-handed neutrino
3523:   superfields have bare Majorana  masses.  Provided that the couplings
3524:   $\lambda$  and  $\rho$  are  not too  small,  e.g.~$\lambda,\,  \rho
3525:   \stackrel{>}{{}_\sim}  10^{-2}$,  the  LSP  right-handed  sneutrinos
3526:   $\widetilde{N}_{\rm  LSP}$ can  efficiently annihilate  via  a Higgs
3527:   resonance $H_d$  into pairs of  $b$-quarks, in the  kinematic region
3528:   $M_{H_d} \approx 2  m_{\widetilde{N}_{\rm LSP}}$, and so drastically
3529:   reduce its relic density to values compatible with the CMB data.
3530: 
3531: 
3532: \end{itemize}
3533: 
3534: 
3535: In  addition to  the above  cosmological implications,  the $F_D$-term
3536: hybrid  model has  a  rich particle-physics  phenomenology. The  main
3537: phenomenological characteristics of the model are:
3538: 
3539: 
3540: \begin{itemize}
3541: 
3542: \item[(a)] It is straightforward  to embed the $F_D$-term hybrid model
3543:   into minimal or next-to-minimal  SUGRA, where the soft SUSY-breaking
3544:   parameters are  constrained at the gauge  coupling unification point
3545:   $M_X$.  Instead, electroweak baryogenesis is not viable in a minimal
3546:   SUGRA scenario  of the MSSM.  It requires an  unconventionally large
3547:   hierarchy   between    the   left-handed   and    right-handed   top
3548:   squarks~\cite{EWBAU},  which  is  difficult  to  obtain  within  the
3549:   framework  of minimal SUGRA.   In addition,  the CP-odd  soft phases
3550:   required  for   successful  electroweak  baryogenesis   face  severe
3551:   constraints from  the absence of observable  2-loop contributions to
3552:   the electron and neutron electric dipole moments~\cite{CKP}.
3553: 
3554: \item[(b)] As has been  discussed in Section~\ref{BAU}, if one assumes
3555:   that  the  neutrino-Yukawa  couplings  $h^\nu_{ij}$ have  a  certain
3556:   hierarchical  structure controlled  by the  approximate  breaking of
3557:   global  flavour  symmetries, the  model  can  have further  testable
3558:   implications  for low-energy  observables of  lepton  flavour and/or
3559:   number  violation, e.g.~$0\nu\beta\beta$  decay, $\mu  \to e\gamma$,
3560:   $\mu \to eee$,  $\mu \to e$ conversion in  nuclei etc.  In addition,
3561:   electroweak-scale heavy Majorana neutrinos may be copiously produced
3562:   at high-energy colliders,  such as the LHC, the  ILC and $e^-\gamma$
3563:   colliders,  whose  decays give  rise  to  distinctive signatures  of
3564:   lepton-number  violation which are  usually manifested  by like-sign
3565:   dileptons accompanied by hadron jets.
3566: 
3567: 
3568: \item[(c)]  Since  successful   inflation  requires  small  couplings,
3569:   i.e.~$\kappa,\,  \lambda,\,  \rho\ \stackrel{<}{{}_\sim}\  10^{-2}$,
3570:   the  inflaton  field   decouples  effectively  from  the  low-energy
3571:   spectrum and the Higgs-sector of  the model becomes identical to the
3572:   one of the  MSSM.  In spite of the  aforementioned decoupling of the
3573:   inflaton,  however, the  $F_D$-term hybrid  model could  still point
3574:   towards particular benchmark scenarios of the MSSM.  For example, if
3575:   $\lambda \gg  \kappa$, the $F_D$-term  hybrid model may  explain the
3576:   origin   of  a   possible  large   value  of   the  $\mu$-parameter.
3577:   Specifically, if  $\lambda = 2\kappa$,  $A_\kappa = -a_S  = 2
3578:   M_{\rm SUSY}$, one gets from~(\ref{mu}) the hierarchy $\mu \approx 4
3579:   M_{\rm SUSY}$,  where $M_{\rm SUSY}$ is a  common soft SUSY-breaking
3580:   scale  of  all   scalar  fermion  fields  in  the   model.   If  one
3581:   additionally requires $A_t  = A_b = 2 M_{\rm  SUSY}$, the low-energy
3582:   limit  of  the $F_D$-term  hybrid  model  becomes  identical to  the
3583:   so-called  CPX benchmark  scenario~\cite{CPX} describing  maximal CP
3584:   violation in  the MSSM  Higgs sector at  low and moderate  values of
3585:   $\tan \beta$.  In the CPX scenario, the lightest neutral Higgs boson
3586:   weighing less than 60~GeV might have escaped detection at LEP. There
3587:   have  been several  strategies to  unravel the  existence of  such a
3588:   light CP-violating Higgs boson~\cite{DP}.
3589: 
3590: \item[(d)] The possible CDM  scenario with the right-handed sneutrinos
3591:   as  LSPs requires large  $\lambda$ and  $\rho$ couplings  that could
3592:   make  Higgs bosons decay  invisibly, e.g.~$H  \to \widetilde{N}_{\rm
3593:   LSP}\, \widetilde{N}_{\rm LSP}$. Also, right-handed sneutrinos could
3594:   be  present in  the  cascade decays  of  the heavier  supersymmetric
3595:   particles.  The  collider phenomenology of such a  CDM scenario lies
3596:   beyond the scope of the present article.
3597: 
3598: 
3599: 
3600: \end{itemize}
3601: 
3602: 
3603: The $F_D$-term hybrid  model studied in this paper  should be regarded
3604: as   a  first   attempt   towards  the   formulation   of  a   minimal
3605: Particle-Physics and Cosmology Standard  Model, which does not involve
3606: excessive fine-tuning in the  fundamental parameters of the theory. As
3607: we outlined  above, it might be  possible to test the  validity of our
3608: model by  a number of laboratory experiments  and further substantiate
3609: it by  future astronomical observations.  The  $F_D$-term hybrid model
3610: is   not   plagued  with   the   usual   gauge-hierarchy  problem   of
3611: non-supersymmetric theories and can,  in principle, be embedded within
3612: an  E(6) or  E(7) GUT,  within the  framework of  SUGRA where  SUSY is
3613: softly broken  at the  TeV scale.   In~the same vein,  we note  that a
3614: possible natural solution to  the famous cosmological constant problem
3615: will   shed  valuable   light   on  the   model-building  aspects   of
3616: cosmologically viable  models.  It  will also open  up new  avenues in
3617: quantitatively  addressing  the   major  energy-density  component  of
3618: today's Universe, the  so-called Dark Energy.  We hope  that all these
3619: insights,  along with  new observational  and experimental  data, will
3620: help  us  to  improve   further  our  present  bottom-up  approach  to
3621: formulating  a more  complete minimal  model of  particle  physics and
3622: cosmology.
3623: 
3624: 
3625: 
3626: 
3627: \subsection*{Acknowledgements}
3628: 
3629: We thank Richard Battye,  Zurab Tavartkiladze and Thomas Underwood for
3630: illuminating discussions.  This work is supported in part by the PPARC
3631: research grants: PP/D000157/1 and PP/C504286/1.
3632: 
3633: 
3634: 
3635: \newpage
3636: 
3637: \def\theequation{\Alph{section}.\arabic{equation}}
3638: \begin{appendix}
3639: 
3640: \setcounter{equation}{0}
3641: \section{Mechanisms of Explicit {\boldmath $D$}-Parity
3642: Breaking}\label{Dappendix}
3643: 
3644: Here  we will  present mechanisms  for explicitly  breaking $D$-parity
3645: within the  SUGRA framework, pointing out  their possible implications
3646: for  the  decay rates  of  the  $g$-sector  particles.  We  separately
3647: discuss  the breaking  of $D$-parity  for an  Abelian U(1)$_X$  and an
3648: non-Abelian SU(2)$_X$ waterfall-gauge sector.
3649: 
3650: 
3651: \subsection{{\boldmath $D$}-Parity
3652: Breaking in the U(1)$_X$ Waterfall-Gauge Sector}
3653: 
3654: 
3655: As already  discussed in  Section~\ref{postinfl}, the simplest  way of
3656: breaking $D$-parity  is to add  a subdominant bare FI  $D$-term ${\cal
3657: L}_{\rm  FI}$  to  the  Lagrangian~[cf.~(\ref{LFI})].   As  was  shown
3658: in~\cite{GP},  however, even  if  such  a term  were  absent from  the
3659: tree-level    Lagrangian,   it   could    still   be    generated   by
3660: quantum-mechanical effects  in an effective  manner, after integrating
3661: out Planck-scale degrees  of freedom. It should be  stressed here that
3662: the radiative generation of an {\em effective} FI $D$-term occurs only
3663: {\em after} the SSB of the U(1)$_X$ gauge symmetry.
3664: 
3665: To elucidate  this point,  let us consider  a simple extension  of the
3666: $F_D$-term  hybrid  model,  which   includes  a  pair  of  superfields
3667: $\widehat{\overline{X}}_{1,2}$    of    opposite   U(1)$_X$    charge,
3668: i.e.~$Q(\widehat{\overline{X}}_2)  = - Q(\widehat{\overline{X}}_1  ) =
3669: Q(\widehat{X}_1  ) =  - Q(\widehat{X}_2  ) =  1$.  In  this  case, the
3670: superpotential $W_{\rm IW}$ pertinent to the inflaton-waterfall sector
3671: may be extended as follows:
3672: \begin{equation}
3673:   \label{Wdterm}
3674:  W_{\rm IW} \ =\ \kappa\, \widehat{S}\, \Big( \widehat{X}_1
3675: \widehat{X}_2\:  -\: M^2\Big)\ +\ \xi\, m_{\rm Pl}\,
3676: \widehat{\overline{X}}_1\,\widehat{\overline{X}}_2\ +\
3677: \xi_1\, \frac{ ( \widehat{\overline{X}}_1\widehat{X}_1 )^2}{2\,m_{\rm
3678:     Pl}}\ +\ \xi'_1\,
3679:    \frac{ ( \widehat{\overline{X}}_2\widehat{X}_2 )^2}{2\,m_{\rm Pl}}\ \dots
3680: \end{equation}
3681: where the  dots stand for  subleading terms that multiply  the leading
3682: operators      by       extra      powers      of      $(\widehat{X}_1
3683: \widehat{X}_2)^n/m^{2n}_{\rm Pl}$,  with $n \ge  1$.  These subleading
3684: operators are irrelevant  for our discussion here and  can be ignored,
3685: within a  perturbative framework of SUGRA.  The  leading operator form
3686: of  the superpotential~(\ref{Wdterm})  may  be reinforced  by the  $R$
3687: symmetry:
3688: \begin{equation}
3689:   \label{Rsym}
3690: \widehat{S}\  \to\ e^{i\alpha}\,  \widehat{S}\,,\quad
3691: \widehat{\overline{X}}_{1,2}\ \to\ e^{i\alpha/2}\,
3692: \widehat{\overline{X}}_{1,2}\,,\quad
3693: (\widehat{L},\ \widehat{Q})\   \to\   e^{i\alpha}\, (\widehat{L},\
3694: \widehat{Q})\,,
3695: \end{equation}
3696: with $W  \to e^{i\alpha} W$.  As  before, all remaining  fields do not
3697: transform   under   the   $R$  symmetry.\footnote{Observe   that   the
3698: $R$-symmetry~(\ref{Rsym}) allows for  the subleading operator $\kappa'
3699: S (\widehat{X}_1 \widehat{X}_2 )^2/m^2_{\rm Pl}$.  This superpotential
3700: term  can break  the U(1)$_X$  gauge symmetry  along  the inflationary
3701: trajectory,    thereby    inflating    away    unwanted    topological
3702: defects~\cite{JKLS}.   This  scenario   is  known  as  shifted  hybrid
3703: inflation.}
3704: 
3705: We will now show that a  FI $D$-term, $-\frac{1}{2} g m^2_{\rm FI} D$,
3706: will   be  generated  if   the  ultraheavy   Planck-scale  superfields
3707: $\widehat{\overline{X}}_{1,2}$  are  integrated  out.  As  a  starting
3708: point, we consider the U(1)$_X$ $D$-term Lagrangian
3709: \begin{equation}
3710:   \label{LDX}
3711: {\cal L}_{D}\ =\ \frac{1}{2}\, D^2\: +\:
3712: \frac{g}{2}\, D\, \Big(\, X^*_1 X_1\: -\: X^*_2 X_2\: -\:
3713: \overline{X}^*_1 \overline{X}_1\: +\:
3714: \overline{X}^*_2\overline{X}_2\,\Big)\; .
3715: \end{equation}
3716: In  order to integrate  out the  fields $\overline{X}_{1,2}$,  we need
3717: their  mass  spectrum in  the  post-inflationary  era, where  $\langle
3718: X_{1,2}\rangle = M$ and  $\langle \overline{X}_{1,2} \rangle = \langle
3719: S   \rangle   =  0$.    In   the   weak   basis  $\overline{X}_\pm   =
3720: \frac{1}{\sqrt{2}}\, (  \overline{X}_1 \pm \overline{X}_2  )$, this is
3721: approximately given by the Lagrangian
3722: \begin{equation}
3723:   \label{LXmass}
3724: -\, {\cal L}^{\overline{X}_\pm}_{\rm mass}\ \approx\
3725: ( \overline{X}^*_+,\ \overline{X}^*_- )\,
3726: \left(\! \begin{array}{cc}
3727: \xi^2 m^2_{\rm Pl} + \xi\, (\xi_1 + \xi'_1 )\, M^2 & (\xi^2_1 -
3728: \xi'^2_1 )\, \frac{\displaystyle M^4}{\displaystyle 2\, m^2_{\rm Pl}}
3729: \\ (\xi^2_1 - \xi'^2_1 )\,
3730: \frac{\displaystyle M^4}{\displaystyle 2\, m^2_{\rm Pl}} &
3731: \xi^2 m^2_{\rm Pl} - \xi\, (\xi_1 + \xi'_1 )\, M^2
3732: \end{array}\!\right)\,
3733: \left(\! \begin{array}{c} \overline{X}_+ \\
3734: \overline{X}_- \end{array} \!\right)\, .
3735: \end{equation}
3736: The approximate mass eigenstates derived from~(\ref{LXmass}) are
3737: \begin{equation}
3738: \widetilde{\overline{X}}_+\ =\ \overline{X}_+\ +\ s_\theta\,
3739: \overline{X}_-\,,\qquad
3740: \widetilde{\overline{X}}_-\ =\ \overline{X}_-\ -\ s_\theta\,
3741: \overline{X}_+\,,
3742: \end{equation}
3743: where $s_\theta \approx (\xi_1 - \xi'_1 ) M^2/ (4\xi m^2_{\rm Pl})$ is
3744: a mixing  angle which is typically  much smaller than~1.   In terms of
3745: the mass-eigenstates $\widetilde{\overline{X}}_{\pm}$, the part of the
3746: Lagrangian~(\ref{LDX})  linear in  the $D$-terms  associated  with the
3747: Planck-scale degrees of freedom reads:
3748: \begin{eqnarray}
3749:   \label{LoXD}
3750: {\cal L}^{\overline{X}_\pm}_{D} & = &
3751: -\, \frac{g}{2}\, D\, \Big(\,
3752: \overline{X}^*_+ \overline{X}_-\: +\:
3753: \overline{X}^*_- \overline{X}_+\,\Big)\nonumber\\
3754: & = &
3755: -\, \frac{g}{2}\, D\, \Big[\,
3756: \widetilde{\overline{X}}^*_+\, \widetilde{\overline{X}}_-\: +\:
3757: \widetilde{\overline{X}}^*_-\, \widetilde{\overline{X}}_+\ +\
3758: 2s_\theta\,
3759: \Big( \widetilde{\overline{X}}_+^*\, \widetilde{\overline{X}}_+\: -\:
3760: \widetilde{\overline{X}}^*_-\, \widetilde{\overline{X}}_- \Big)\: +\:
3761: {\cal O}(s^2_\theta)\, \Big]\; .
3762: \end{eqnarray}
3763: 
3764: %******************************************************************
3765: %%% Figure 1
3766: %******************************************************************
3767: \begin{figure}[t]
3768: \begin{center}
3769: \begin{picture}(200,100)(0,0)
3770: \SetWidth{0.8}
3771: 
3772: 
3773: \Line(100,20)(100,50)\DashArrowArc(100,75)(25,-90,270){3}
3774: \Text(100,10)[]{$D$}\Text(130,75)[l]{$\widetilde{\overline{X}}_{+},\
3775: \widetilde{\overline{X}}_{-}$}
3776: 
3777: \end{picture}
3778: \end{center}
3779: \caption{\sl\small  Radiative  generation   of  an  effective  FI  $D$-term,
3780:   $-\frac{g}{2}\,m^2_{\rm FI}\, D$.}\label{DXtad}
3781: \end{figure}
3782: 
3783: 
3784: A FI $D$-tadpole can only be generated from terms linear in $s_\theta$
3785: in  the Lagrangian~(\ref{LoXD}).   This result  should be  expected on
3786: symmetry grounds,  since terms  linear in $s_\theta$  explicitly break
3787: the  $D$-symmetry.   The
3788: $D$-tadpole  $m^2_{\rm FI}$,  calculated  from the  one-loop graph  of
3789: Fig,~\ref{DXtad}, is found to be
3790: \begin{equation}
3791:   \label{FIdterm}
3792: m^2_{\rm FI}\ \approx\ \frac{ \xi^2_1 - \xi'^2_1 }{8\pi^2}\
3793: \frac{M^4}{m^2_{\rm Pl}}\ \ln\left(\frac{m_{\rm Pl}}{M}\right)\ .
3794: \end{equation}
3795: Typically, one gets  $m_{\rm FI}/M \stackrel{<}{{}_\sim} 10^{-6}$, for
3796: $M = 10^{16}$~GeV and $\xi_1,\ \xi'_1 \stackrel{<}{{}_\sim} 10^{-3}$.
3797: 
3798: For later use, it is  interesting to outline a short-cut derivation of
3799: the  result~(\ref{FIdterm}),  using the  original  weak  basis of  the
3800: fields,  i.e.~$X_{1,2}$  and  $\overline{X}_{1,2}$.  We  notice  that,
3801: after the  SSB of U(1)$_X$, the $F$-terms  of $\overline{X}_{1,2}$ give
3802: rise to the $D$-odd operator,
3803: \begin{equation}
3804:   \label{FU1}
3805: {\cal F}\ =\ (\xi_1^2-\xi'^2_1)\ \frac{M^4}{2\, m_{\rm Pl}^2}\
3806: \Big(\,\overline X_1^* \overline X_1\:  -\:
3807: \overline X_2^* \overline X_2\, \Big)\; ,
3808: \end{equation}
3809: in the scalar potential of the extended $F_D$-term hybrid model.  This
3810: operator induces,  via the  diagram shown in  Fig.~\ref{DXtad:alt}, an
3811: effective FI $D$-tadpole. Since the scalar fields $\overline{X}_{1,2}$
3812: are degenerate in  mass to leading order, i.e.~$M_{\overline{X}_{1,2}}
3813: \approx \xi  m_{\rm Pl}$, the graph  in~Fig.~\ref{DXtad:alt} is easily
3814: calculated   using   standard    field-theoretic   methods.    It   is
3815: logarithmically divergent,  and in an  effective cut-off theory  it is
3816: given by~(\ref{FIdterm}).   We will use this  short-cut approach below
3817: to calculate  effective $D$-tadpoles  in more elaborate  extensions of
3818: the inflation-waterfall sector.
3819: 
3820: \begin{figure}[t]
3821: \begin{center}
3822: \begin{picture}(200,100)(0,0)
3823: \SetWidth{0.8}
3824: 
3825: 
3826: \Line(100,20)(100,50)
3827: \DashArrowArc(100,75)(25,-90,90){3}
3828: \DashArrowArc(100,75)(25,90,270){3}
3829: \SetWidth{1}
3830: \Line(96,104)(104,96)
3831: \Line(104,104)(96,96)
3832: \SetWidth{0.5}
3833: \Text(100,10)[]{$D$}
3834: \Text(130,75)[l]{$\overline{X}_{1}$ ($\overline{X}_{2}$)}
3835: \Text(70,75)[r]{$\overline{X}_{1}$ ($\overline{X}_{2}$)}
3836: 
3837: \Text(100,113)[]{$\cal F$}
3838: 
3839: \end{picture}
3840: \end{center}
3841: \caption{\sl\small  Diagram pertinent to a short-cut derivation of
3842: the  effective  FI  $D$-term.}\label{DXtad:alt}
3843: \end{figure}
3844: 
3845: The  size  of  the FI  $D$-term  may  be  further suppressed,  if  the
3846: Planck-mass chiral  superfields $\widehat{\overline{X}}_{1,2}$ possess
3847: higher U(1)$_X$ charges.  In general, one may assume that the U(1)$_X$
3848: charges         of         $\widehat{\overline{X}}_{1,2}$         are:
3849: $Q(\widehat{\overline{X}}_2)  = -  Q(\widehat{\overline{X}}_1 )  = n$,
3850: where  $n\ge  1$. In  this  case, the  leading  operator  form of  the
3851: inflaton-waterfall superpotential reads:
3852: \begin{equation}
3853:   \label{Wdtermn}
3854: W_{\rm IW} \ =\ \kappa\, \widehat{S}\, \Big( \widehat{X}_1
3855: \widehat{X}_2\: -\: M^2\Big)\ +\ \xi\, m_{\rm Pl}\,
3856: \widehat{\overline{X}}_1\,\widehat{\overline{X}}_2\ +\ \xi_n\, \frac{
3857: (\widehat{\overline{X}}_1)^2\, (\widehat{X}_1)^{n+1}}{2\,m^n_{\rm
3858: Pl}}\ +\ \xi'_n\, \frac{ (
3859: \widehat{\overline{X}}_2)^2\,(\widehat{X}_2)^{n+1}}{2\,m^n_{\rm Pl}}\ .
3860: \end{equation}
3861: Employing the  short-cut method outlined above,  it is straightforward
3862: to compute the loop-induced $D$-term, which is given by
3863: \begin{equation}
3864:   \label{FIdtermn}
3865: m^2_{\rm FI}\ \approx\ \frac{\xi^2_n - \xi'^2_n}{8\pi^2}\
3866: \frac{M^{2(n+1)}}{m^{2n}_{\rm Pl}}\ \ln\left(\frac{m_{\rm Pl}}{M}\right)\ .
3867: \end{equation}
3868: To  obtain a  small ratio  $m_{\rm FI}/M  \sim 10^{-6}$  with $\xi_n,\
3869: \xi'_n \sim 1$, one would simply need $n = 5,\ 6$.  Finally, we should
3870: remark  that the loop-induced  $D$-term does  not lead  to spontaneous
3871: breakdown of global supersymmetry.
3872: 
3873: 
3874: \subsection{{\boldmath $D$}-Parity Breaking in the SU(2)$_X$
3875:                                               Waterfall-Gauge Sector}
3876: 
3877: Here we  outline two possible  mechanisms for explicitly  breaking the
3878: $D$-parities,  $D_1$ and $D_2$  defined in~(\ref{D1})  and (\ref{D2}),
3879: which  govern  the  minimal  inflaton-waterfall  sector  based  on  an
3880: SU(2)$_X$ gauge group.
3881: 
3882: The  first mechanism utilizes  a non-minimal  K\"ahler waterfall-gauge
3883: sector, where  the two  $D$-parities are broken  by non-renormalizable
3884: operators.  To  be  specific,  a  minimal  $D_{1,2}$-parity  violating
3885: K\"ahler potential of the waterfall-gauge  sector may be cast into the
3886: form:
3887: \begin{eqnarray}
3888:   \label{Kwf}
3889: K_{\rm WF} &=&   \int   d^4\theta\     \Bigg(\,
3890: \widehat{X}^\dagger_1\, {\rm e}^{2g\widehat{V}_X} \widehat{X}_1\
3891: +\  \widehat{X}^T_2\, {\rm e}^{-2g\widehat{V}_X} \widehat{X}_2^*\
3892: +\  \kappa_1\,   \frac{(\widehat{X}^\dagger_1 \, {\rm e}^{2g\widehat{V}_X}
3893: \widehat{X}_1)^2}{2\,m^2_{\rm Pl}} \nonumber\\
3894: &&+\    \kappa_2\, \frac{(\widehat{X}^T_2 \, {\rm e}^{-2g\widehat{V}_X}
3895: \widehat{X}_2^*)^2}{2\,m^2_{\rm   Pl}}\
3896: +\ \frac{
3897: \kappa_1^\prime
3898: (\widehat X_1 ^\dagger {\rm e}^{2 g \widehat V_X} {\rm i} \tau^2 \widehat X_2 )
3899: (\widehat X_1 ^\dagger {\rm e}^{2 g \widehat V_X} \widehat X_1)
3900: +{\rm H.c.}
3901: }{2\,m_{\rm Pl}^2}\nonumber\\
3902: &&
3903: +\ \frac{
3904: \kappa_2^\prime
3905: (\widehat X_1 ^\dagger {\rm e}^{2 g \widehat V_X} {\rm i} \tau^2 \widehat X_2 )
3906: (\widehat X_2 ^T {\rm e}^{-2 g \widehat V_X} \widehat X_2^*)
3907: +{\rm H.c.}
3908: }{2\,m_{\rm Pl}^2}\  +\ \dots\Bigg)\, ,
3909: \end{eqnarray}
3910: where  the ellipses  denote  possible higher-order  non-renormalizable
3911: operators.    The   couplings   $\kappa_{1,2}$   are   real,   whereas
3912: $\kappa^\prime_{1,2}$  can  in  general  be  complex.   Moreover,  the
3913: difference  $\kappa_- =  \kappa_1 -  \kappa_2$  signifies $D_1$-parity
3914: violation, whilst $\kappa_-^\prime=\kappa_1^\prime-\kappa_2^\prime$ is
3915: a parameter of $D_2$-parity  violation.  Hence, non-zero values of the
3916: parameters $\kappa_-$  and $\kappa^\prime_-$ will give  rise to $D_1$-
3917: and $D_2$-parity  violation in the  waterfall-gauge K\"ahler potential
3918: $K_{\rm  WF}$.   Notice that,  as  far  as  $D_1$-parity violation  is
3919: concerned, the present mechanism applies to the Abelian case as well.
3920: 
3921: There are several sources of $D$-parity violation contained in $K_{\rm
3922: WF}$.  More explicitly, $D$-parity violation will first originate from
3923: the terms  linear in  $D^a$, where $D^a$  are the  auxiliary SU(2)$_X$
3924: components of  the gauge-vector  superfield $\widehat V_X$.   In fact,
3925: these are  the lowest dimensional $D_{1,2}$-odd  operators that emerge
3926: after  the  SSB  of the  SU(2)$_X$  and  are  given by  the  effective
3927: Lagrangian
3928: \begin{equation}
3929: {\cal L}^{D^a-{\rm tad}}_{\rm eff} \ =\
3930:                            \frac{g}{2}\ \frac{M^4}{m_{\rm Pl}^2}\
3931: \Big(\, {\rm Re}\kappa^\prime_-\, D^1\: -\:
3932: {\rm Im}\kappa^\prime_-\, D^2\: +\: \kappa_-\, D^3 \Big)\; .
3933: \end{equation}
3934: These  effective FI  $D$-terms can  be included  in the  Lagrangian by
3935: adding the constants $\frac{g}{2}\,  (m^a_{\rm FI})^2$ to the on-shell
3936: constrained $D^a$ terms, according to the scheme: $D^a \to D^a + \frac
3937: g2 ({m_{\rm FI}^a})^2$, where
3938: \begin{equation}
3939:   \label{mFI123}
3940: (m^1_{\rm FI})^2\ =\ \frac{M^4}{m_{\rm Pl}^2}\
3941:                          {\rm Re}\kappa^\prime_-\,,\qquad
3942: (m^2_{\rm FI})^2\ =\ -\ \frac{M^4}{m_{\rm Pl}^2}\
3943:                          {\rm Im}\kappa^\prime_-\,,\qquad
3944: (m^3_{\rm FI})^2\ =\ \frac{M^4}{m_{\rm Pl}^2}\ \kappa_-\; .
3945: \end{equation}
3946: One may obtain a fair estimate of the $g$-sector particle decay rates,
3947: using  the formula~(\ref{GammaR})  and identifying  $m_{\rm  FI}$ with
3948: $m^a_{\rm FI}$. In this way, we obtain
3949: \begin{equation}
3950:   \label{GgD}
3951: \Gamma [^-\!R_-\,,\ ^-\!I_+\,,\ ^+\!R_- ] \ \sim\
3952: [\,\kappa_-^2\,,\ {\rm Re}^2(\kappa_-^\prime)\,,\ {\rm Im}^2
3953: (\kappa_-^\prime )\,]\
3954: \frac {g^3}{128 \pi}\;  \frac{M^5}{m_{\rm Pl}^4}\ .
3955: \end{equation}
3956: In   addition  to   the  effective   $D$-tadpoles,  higher-dimensional
3957: operators  will  also  break   the  $D$-parities  and  so  render  the
3958: $g$-sector  particles  unstable.   For  example, after  expanding  the
3959: superfields  $\widehat{X}_{1,2}$  about  their  VEVs in  the  K\"ahler
3960: potential~(\ref{Kwf}),  we   find  the  non-renormalizable  $D$-parity
3961: violating interactions described by the Lagrangian
3962: \begin{equation}
3963:   \label{Lnonren}
3964: {\cal L}_{\rm non-ren} \ =\
3965: -\; \frac{M}{2 m_{\rm Pl}^2} \kappa_-\, ^+\!R_-\,
3966: |\partial_\mu {}^+\!X_+|^2\ +\ \frac{M}{4 \sqrt{2} m_{\rm Pl}^2}\;
3967: \Big(\, \kappa^\prime_-\, ^-\!X_-\ +\ {\rm H.c.}\,\Big)\,
3968: |\partial_\mu {}^+\!X_+|^2\; .
3969: \end{equation}
3970: With the aid of~(\ref{Lnonren}), an order of magnitude estimate of the
3971: $g$-sector  particle decay  rates gives:  $\Gamma_g  \sim (\kappa_-^2,
3972: |\kappa_-^\prime|^2)\,  g^3 M^5/m_{Pl}^4$.   These  are of  comparable
3973: order  to the  ones obtained  earlier in~(\ref{GgD}).   For  a typical
3974: inflationary scale, $M  = 10^{16}$~GeV (with $g \sim  1$), we find the
3975: decay   width  $\Gamma_g   \sim   (\kappa_-^2,  |\kappa_-^\prime|^2)\,
3976: 10^7$~GeV.  The latter should be compared with the bounds: $3.8 \times
3977: 10^{-13}~{\rm        GeV}        \stackrel{<}{{}_\sim}        \Gamma_g
3978: \stackrel{<}{{}_\sim}  4.3~{\rm GeV}$, corresponding  to an  upper and
3979: lower  limit on the  second reheat  temperature $T_g$  of cosmological
3980: interest:      $0.3~{\rm      TeV}     \stackrel{<}{{}_\sim}      T_g\
3981: \stackrel{<}{{}_\sim}  10^9~{\rm GeV}$.  Consequently,  values ranging
3982: from  $10^{-9}$  to  $10^{-2}$  for the  couplings  $\kappa_-$  and/or
3983: $\kappa^\prime_-$ are required  for successful coupled reheating.  The
3984: lower end values  of order $10^{-9}$ may possibly be  seen as a strong
3985: tuning of the  parameters.  One way to explain  the smallness of these
3986: parameters  is  to  contemplate  K\"ahler  manifolds  that  break  the
3987: $D$-parities   by  even  higher-order   non-renormalizable  operators,
3988: e.g.~of  order $\sim  1/m^4_{\rm Pl}$.   In this  case,  the couplings
3989: $\kappa_-$  and $\kappa'_-$  will be  multiplied by  extra  factors of
3990: $M^2/m^2_{\rm Pl} \sim 10^{-4}$, so these couplings can have values of
3991: order~1  and  still predict  a  second  reheat  temperature $T_g  \sim
3992: 0.3$~TeV.
3993: 
3994: We now describe a second mechanism of $D$-parity violation which might
3995: be useful  to obtain small $D$-parity violating  interactions.  Let us
3996: therefore   assume   that   the   K\"ahler  potential   respects   the
3997: $D$-parities. In this  case, we may invoke a  mechanism very analogous
3998: to  the   Abelian  case.    We  extend  the   field  content   of  the
3999: inflaton-waterfall  sector  by adding  a  pair  of Planck-mass  chiral
4000: superfields $\widehat{\overline{X}}_1$ and $\widehat{\overline{X}}_2$,
4001: which belong  to the anti-fundamental  and fundamental representations
4002: of SU(2)$_X$,  respectively.  As in the U(1)$_X$  case, the superheavy
4003: superfields  $\widehat{\overline{X}}_{1,2}$   are  charged  under  the
4004: continuous $R$-symmetry given in~(\ref{Rsym}).  With this restriction,
4005: the leading operator form  of the inflaton-waterfall superpotential is
4006: given by
4007: \begin{eqnarray}
4008:   \label{WIFSU2}
4009: W_{\rm IW} &=& \kappa\, \widehat{S}\, \Big( \widehat{X}_1{}\!^T
4010: \widehat{X}_2\:  -\: M^2\Big)\ +\
4011: \xi\, m_{\rm Pl}\,
4012: \widehat{\overline{X}}_1{}\!^T \,\widehat{\overline{X}}_2\ +\
4013: \theta_1\; \frac{ ( \widehat{\overline{X}}_1{}\!^T
4014: \widehat{X}_1 )\, ( \widehat{\overline{X}}_2{}\!^T \widehat{X}_2)}
4015: {m_{\rm Pl}}\nonumber\\
4016: &&+\; \theta_2\; \frac{ ( \widehat{\overline{X}}_1{}\!^T {\rm i}\tau^2
4017: \widehat{X}_2)\, ( \widehat{\overline{X}}_2{}\!^T {\rm i} \tau^2
4018: \widehat{X}_1)}{m_{\rm    Pl}}\ +\ \zeta_1\;
4019: \frac{(\widehat{\overline{X}}_1{}\!^T {\rm i} \tau^2 \widehat{X}_2)\,
4020: (\widehat{\overline{X}}_2{}\!^T \widehat{X}_2)}{m_{\rm Pl}}\nonumber\\
4021: &&+\ \zeta_2\;
4022: \frac{(\widehat{\overline{X}}_1{}\!^T  \widehat{X}_1)\,
4023: (\widehat{\overline{X}}_2{}\!^T {\rm i} \tau^2 \widehat{X}_1)}
4024: {m_{\rm Pl}}\quad + \quad \dots\; ,
4025: \end{eqnarray}
4026: where  the dots stand  for additional  operators that  turn out  to be
4027: irrelevant  for  the   generation  of  effective  $D^a$-tadpoles,  and
4028: especially  for those  related  to the  $D^1$-  and $D^2$-terms.   The
4029: presence of  these operators is  only important to lift  an accidental
4030: global  U(1)$_{\overline{X}}$ symmetry  that  governs this  restricted
4031: part of  the superpotential  $W_{\rm IW}$ under  consideration.  Here,
4032: all non-renormalizable couplings  $\theta_{1,2}$ and $\zeta_{1,2}$ can
4033: in general be complex.  Extending  the notion of $D_{1,2}$ parities to
4034: the Planck-mass superfields $\widehat{\overline{X}}_{1,2}$, we observe
4035: that  the  operators related  to  the  couplings  $\kappa$, $\xi$  and
4036: $\theta_{1,2}$ are  even under $D_1$ and $D_2$,  whereas those related
4037: to   the  couplings  $\zeta_{1,2}$   are  $D_2$-odd.    Moreover,  the
4038: superpotential operators proportional to the couplings $\zeta_{+(-)} =
4039: \zeta_1 +(-)\ \zeta_2$ are $D_1$-even ($D_1$-odd).
4040: 
4041: 
4042: %******************************************************************
4043: %%% Effective D-term for SU(2)_X
4044: %******************************************************************
4045: \begin{figure}[t]
4046: \begin{center}
4047: \begin{picture}(600,100)(0,0)
4048: \SetWidth{0.8}
4049: 
4050: 
4051: \Line(100,20)(100,50)
4052: \DashArrowArc(100,75)(25,-90,90){3}
4053: \DashArrowArc(100,75)(25,90,270){3}
4054: \SetWidth{1}
4055: \Line(96,104)(104,96)
4056: \Line(104,104)(96,96)
4057: \SetWidth{0.5}
4058: \Text(100,10)[]{$D^1_{(-,-)}$}
4059: \Text(130,75)[l]{$\overline{X}_{1\, (2)}$}
4060: \Text(70,75)[r]{$\overline{X}_{1\, (2)}$}
4061: \Text(100,113)[]{${\cal F}^1_{(-,-)}$}
4062: 
4063: \Line(230,20)(230,50)
4064: \DashArrowArc(230,75)(25,-90,90){3}
4065: \DashArrowArc(230,75)(25,90,270){3}
4066: \SetWidth{1}
4067: \Line(226,104)(234,96)
4068: \Line(234,104)(226,96)
4069: \SetWidth{0.5}
4070: \Text(230,10)[]{$D^2_{(+,-)}$}
4071: \Text(260,75)[l]{$\overline{X}_{1\, (2)}$}
4072: \Text(200,75)[r]{$\overline{X}_{1\, (2)}$}
4073: \Text(230,113)[]{${\cal F}^2_{(+,-)}$}
4074: 
4075: \Line(360,20)(360,50)
4076: \DashArrowArc(360,75)(25,-90,90){3}
4077: \DashArrowArc(360,75)(25,90,270){3}
4078: \SetWidth{1}
4079: \Line(356,104)(364,96)
4080: \Line(364,104)(356,96)
4081: \SetWidth{0.5}
4082: \Text(360,10)[]{$D^3_{(-,+)}$}
4083: \Text(390,75)[l]{$\overline{X}_{1\, (2)}$}
4084: \Text(330,75)[r]{$\overline{X}_{1\, (2)}$}
4085: \Text(360,113)[]{${\cal F}^3_{(-,+)}$}
4086: 
4087: \end{picture}
4088: \end{center}
4089: \caption{\sl\small   Diagrams  responsible for the  generation of  effective
4090: $D^{1,2,3}$-tadpoles  for  the ${\rm  SU}(2)_X$  case,  in the  single
4091: insertion approximation  of the $D$-odd  operators ${\cal F}^{1,2,3}$.
4092: The subscripts  in parentheses label  the $(D_1,D_2)$ parities  of the
4093: respective operator.}\label{DXtad:SU2}
4094: \end{figure}
4095: 
4096: To calculate  the effective $D^{1,2,3}$-tadpoles after the  SSB of the
4097: SU(2)$_X$ gauge  group, we use the short-cut  approach described above
4098: in     Section~\ref{Dappendix}.1.     Thus,    the     $F$-terms    of
4099: $\widehat{\overline{X}}_{1,2}$  give  rise  to the  following  $D$-odd
4100: contributions to the scalar potential:
4101: \begin{eqnarray}
4102:   \label{F1}
4103: {\cal F}^1_{(-,-)} &=& \theta^*_1\, \zeta_-\
4104: \frac{M^2}{2\,m^2_{\rm Pl}}\
4105: \bigg[\,
4106: \Big( \overline{X}^\dagger_1 \langle X_1^*\rangle \Big)\,
4107: \Big( \overline{X}^T_1\, {\rm i}\tau^2 \langle X_2\rangle \Big)\
4108: -\ (1\leftrightarrow 2)\, \bigg]\nonumber\\
4109: && -\, \theta^*_2\, \zeta_-\
4110: \frac{M^2}{2\,m^2_{\rm Pl}}\
4111: \bigg[\,
4112: \Big( \overline{X}^\dagger_1\, {\rm i}\tau^2 \langle X^*_2\rangle \Big)\,
4113: \Big( \overline{X}^T_1  \langle X_1 \rangle \Big)\
4114: -\ (1\leftrightarrow 2)\, \bigg]\quad  +\quad {\rm H.c.}\; ,\\[4mm]
4115:   \label{F2}
4116: {\cal F}^2_{(+,-)} &=& \theta^*_1\, \zeta_+\
4117: \frac{M^2}{2\,m^2_{\rm Pl}}\
4118: \bigg[\,
4119: \Big( \overline{X}^\dagger_1 \langle X_1^*\rangle \Big)\,
4120: \Big( \overline{X}^T_1\, {\rm i}\tau^2 \langle X_2\rangle \Big)\
4121: +\ (1\leftrightarrow 2)\, \bigg]\nonumber\\
4122: && +\, \theta^*_2\, \zeta_+\
4123: \frac{M^2}{2\,m^2_{\rm Pl}}\
4124: \bigg[\,
4125: \Big( \overline{X}^\dagger_1\, {\rm i}\tau^2 \langle X^*_2\rangle \Big)\,
4126: \Big( \overline{X}^T_1  \langle X_1 \rangle \Big)\
4127: +\ (1\leftrightarrow 2)\, \bigg]\quad  +\quad {\rm H.c.}\; ,\\[4mm]
4128:   \label{F3}
4129: {\cal F}^3_{(-,+)} & =&
4130: -\,{\rm Re}\,(\zeta_+\zeta^*_-)\ \frac{M^2}{2\,m^2_{\rm Pl}}\
4131: \bigg[\,  \Big( \overline{X}^\dagger_1\, i\tau^2 \langle X_2^*\rangle \Big)\,
4132: \Big( \langle X^T_2 \rangle\, i\tau^2 \overline{X}_1 \Big)\ +\
4133: \Big( \overline{X}^\dagger_1 \langle X^*_1 \rangle \Big)\,
4134: \Big( \langle X^T_1 \rangle \overline{X}_1 \Big)\nonumber\\
4135: &&-\quad (1 \leftrightarrow 2)\, \bigg]\; ,
4136: \end{eqnarray}
4137: where the subscripts in  parentheses indicate the $(D_1,D_2)$ parities
4138: of  the  above  operators.    Note  that  possible  $D$-odd  operators
4139: proportional to  $\xi \theta_{1,2}$  and $\xi\zeta_\pm$ have  not been
4140: displayed, since they do not contribute to the generation of effective
4141: $D^a$-tadpoles.   To  be  specific,  the  effective  $D$-tadpoles  are
4142: induced radiatively via the graphs shown in~Fig.~\ref{DXtad:SU2}, once
4143: the operators ${\cal F}^{1,2,3}$  are individually contracted with the
4144: $D$-term  operator ${\cal  D}^{1,2,3}_{\overline{X}}$  related to  the
4145: $\overline{X}_{1,2}$ fields:
4146: \begin{equation}
4147: {\cal D}^a_{\overline{X}}\ =\ -\; \frac{g}{2}\ \Big(\,
4148: \overline{X}^T_1\, \tau^a\, \overline{X}^*_1\: -\:
4149: \overline{X}^\dagger_2\, \tau^a\, \overline{X}_2\, \Big)\; .
4150: \end{equation}
4151: These  loop-induced effective  FI  $D$-terms can  be  included in  the
4152: effective Lagrangian by shifting  the on-shell constrained $D^a$ terms
4153: by constants, according to the  above described scheme: $D^a \to D^a +
4154: \frac  g2 ({m_{\rm FI}^a})^2$.   In this  scheme, the  mass parameters
4155: $(m^a_{\rm FI})^2$ are found to be
4156: \begin{eqnarray}
4157:   \label{mFISU2}
4158: \big(m_{\rm FI}^1\big)^2 & = &  -\,
4159:              \frac{{\rm Re}\,(\theta^*_- \zeta_-)}{4\pi^2}\
4160: \frac{M^4}{m_{\rm Pl}^2}\ \ln \left(\frac{m_{\rm Pl}}{M}\right)\; ,\nonumber\\
4161: \big(m_{\rm FI}^2\big)^2 & = &
4162:              \frac{{\rm Im}\,(\theta^*_-\zeta_+)}{4\pi^2}\
4163: \frac{M^4}{m_{\rm Pl}^2}\ \ln \left(\frac{m_{\rm Pl}}{M}\right)\; ,\\
4164: \big( m_{\rm FI}^3 \big)^2 & = & -\,
4165:              \frac{{\rm Re}\,(\zeta_+\zeta^*_-)}{4\pi^2}\
4166: \frac{M^4}{m_{\rm Pl}^2}\ \ln \left(\frac{m_{\rm Pl}}{M}\right)\; ,\nonumber
4167: \end{eqnarray}
4168: where  $\theta_\pm =  \theta_1  \pm \theta_2$.   It  can be  estimated
4169: from~(\ref{mFISU2})  that  for  values $\theta_\pm\,,  \zeta_\pm  \sim
4170: 10^{-4}$,   one  gets   $m^{1,2,3}_{\rm   FI}/M  \stackrel{<}{{}_\sim}
4171: 10^{-6}$,  leading to  a low  second reheat  temperature  $T_g$, below
4172: 1~TeV.   In this  context,  one should  notice  that the  size of  the
4173: effective $D$-tadpoles  is very sensitive to the  cut-off scale, which
4174: we have  chosen here to be  the reduced Planck mass  $m_{\rm Pl}$. For
4175: instance, if a cut-off larger  by one order of magnitude were adopted,
4176: then   values   of   order   $10^{-2}$  for   the   non-renormalizable
4177: superpotential couplings would be sufficient to generate the effective
4178: $D^a$-tadpoles at the required size.
4179: 
4180: 
4181: The violation  of $D$-parities will also affect  the particle spectrum
4182: of the  SU(2)$_X$ inflaton-waterfall sector.  This will  depend on the
4183: particular choice of the non-renormalizable part of the superpotential
4184: and K\"ahler potential.  Our intention is not to pursue this issue any
4185: further  here,  by  putting   forward  a  specific  non-minimal  SUGRA
4186: scenario.  Instead,  our goal has  been to explicitly  demonstrate the
4187: existence   of   at   least   two  mechanisms,   which   utilize   the
4188: non-renormalizable part of the K\"ahler potential or superpotential to
4189: break the $D$-parities at the required order of magnitude.
4190: 
4191: 
4192: \end{appendix}
4193: 
4194: 
4195: 
4196: 
4197: \newpage
4198: \begin{thebibliography}{99}
4199: 
4200: 
4201: \bibitem{COBE}  G.F. Smoot  {\it et  al.}, Astrophys.\  J.\  {\bf 396}
4202:   (1992) L1.
4203: 
4204: \bibitem{WMAP} D.N. Spergel {\it et al.}, Astrophys.\ J.\ Suppl.\ {\bf
4205: 148} (2003) 175.
4206: 
4207: \bibitem{MT} M. Tegmark {\it et al.}, Phys.\ Rev.\ D~{\bf 69} (2004) 103501.
4208: 
4209: \bibitem{WMAP3} D.N. Spergel {\it et al.}, astro-ph/0603449.
4210: 
4211: \bibitem{Lyman} For a recent global analysis that includes the Ly-$a$
4212:     forest power spectrum, see,\\ U. Seljak, A. Slosar and
4213:     P. McDonald, astro-ph/0604335.
4214: 
4215: \bibitem{review} For reviews, see,\\
4216: D.~H.~Lyth and A.~Riotto,
4217: %``Particle physics models of inflation and the cosmological density
4218: %perturbation,''
4219: Phys.\ Rept.\  {\bf 314} (1999) 1;\\
4220: B.A. Bassett, S. Tsujikawa and D. Wands, astro-ph/0507632.
4221: 
4222: \bibitem{Linde} A.~D.~Linde,
4223: %``Axions in inflationary cosmology,''
4224: Phys.\ Lett.\ B {\bf 259} (1991) 38.
4225: 
4226: \bibitem{CLLSW} E.~J.~Copeland, A.~R.~Liddle, D.~H.~Lyth,
4227:   E.~D.~Stewart and D.~Wands,
4228: %``False vacuum inflation with Einstein gravity,''
4229:   Phys.\ Rev.\ D {\bf 49} (1994) 6410.
4230: 
4231: \bibitem{DSS}  G.~R.~Dvali, Q.~Shafi and R.~K.~Schaefer,
4232: %``Large scale structure and supersymmetric inflation without fine tuning,''
4233:   Phys.\ Rev.\ Lett.\  {\bf 73} (1994) 1886.
4234: 
4235: \bibitem{Halyo} E. Halyo, Phys.\ Lett.\ B~{\bf 387} (1996) 43;
4236: Phys.\ Lett.\ B~{\bf 454} (1999) 223;\\
4237: P. Bin\'etruy and G. Dvali, Phys.\ Lett.\ B~{\bf 388} (1996) 241.
4238: 
4239: \bibitem{FI} P.~Fayet and J.~Iliopoulos,
4240:   %``Spontaneously Broken Supergauge Symmetries And Goldstone Spinors,''
4241:   Phys.\ Lett.\ B {\bf 51} (1974) 461.
4242: 
4243: \bibitem{SS} V.N. Senoguz and Q. Shafi, Phys.\ Rev.\ D~{\bf 71} (2005) 043514.
4244: 
4245: \bibitem{GP}
4246:   B.~Garbrecht and A.~Pilaftsis,
4247:   %``F(D)-term hybrid inflation with electroweak-scale lepton number
4248:   %violation,''
4249:   Phys. Lett. B {\bf 636} (2006) 154 [hep-ph/0601080].
4250:   %%CITATION = HEP-PH 0601080;%%
4251: 
4252: \bibitem{DLS} For a related observation in other variants of hybrid
4253:   inflation, see,\\
4254: G.~R.~Dvali, G.~Lazarides and Q.~Shafi,
4255: %``mu problem and hybrid inflation in supersymmetric SU(2)L x SU(2)R x
4256: %U(1)B-L,''
4257: Phys.\ Lett.\ B {\bf 424} (1998) 259;\\
4258: S.~F.~King and Q.~Shafi,
4259: %``Minimal supersymmetric SU(4) x SU(2)L x SU(2)R,''
4260: Phys.\ Lett.\ B {\bf 422} (1998) 135.
4261: 
4262: 
4263: \bibitem{Francesca} For alternative suggestions based on
4264:   non-renormalizable operators that involve higher powers of $\phi$,
4265:   see,\\
4266: F.~Borzumati and Y.~Nomura,
4267: %``Low-scale see-saw mechanisms for light neutrinos,''
4268: Phys.\ Rev.\ D {\bf 64} (2001) 053005;\\
4269: N.~Arkani-Hamed, L.~J.~Hall, H.~Murayama, D.~R.~Smith and N.~Weiner,
4270: %``Small neutrino masses from supersymmetry breaking,''
4271: Phys.\ Rev.\ D {\bf 64} (2001) 115011;\\
4272: S. Dar, S. Huber, V.N. Senoguz and Q. Shafi, Phys.\ Rev.\ D~{\bf 69}
4273: (2004) 077701.
4274: 
4275: \bibitem{PU2} A. Pilaftsis and T.E.J.~Underwood, Phys.\ Rev.\ D~{\bf
4276: 72} (2005) 113001.
4277: 
4278: \bibitem{FY} M. Fukugita and T. Yanagida, Phys.\ Lett.\ B {\bf 174}
4279: (1986) 45.
4280: 
4281: \bibitem{BAUpapers}  S.~Davidson and A.~Ibarra,
4282: %``A lower bound on the right-handed neutrino mass from leptogenesis,''
4283:   Phys.\ Lett.\ B {\bf 535} (2002) 25;\\
4284: W.~Buchmuller, P.~Di Bari and M.~Plumacher,
4285: % ``Cosmic microwave background, matter-antimatter asymmetry and neutrino
4286: % masses,''
4287:   Nucl.\ Phys.\ B {\bf 643} (2002) 367;\\
4288: G.~C.~Branco, R.~Gonzalez Felipe, F.~R.~Joaquim, I.~Masina,
4289: M.~N.~Rebelo and C.~A.~Savoy,
4290: %``Minimal scenarios for leptogenesis and CP violation,''
4291:   Phys.\ Rev.\ D {\bf 67} (2003) 073035;\\
4292: G.~F.~Giudice, A.~Notari, M.~Raidal, A.~Riotto and A.~Strumia,
4293: %``Towards a complete theory of thermal leptogenesis in the SM and MSSM,''
4294:   Nucl.\ Phys.\ B~{\bf 685} (2004) 89.
4295: 
4296: 
4297: \bibitem{APRD} A. Pilaftsis, Phys.\ Rev.\ D~{\bf 56} (1997) 5431;\\
4298: A. Pilaftsis and T.E.J.~Underwood, Nucl.\ Phys.\ B~{\bf 692} (2004) 303;\\
4299: T.~Hambye, J.~March-Russell and S.~M.~West,
4300: %``TeV scale resonant leptogenesis from supersymmetry breaking,''
4301:   JHEP {\bf 0407} (2004) 070;\\
4302: E.~J.~Chun,
4303: %``TeV leptogenesis in Z-prime models and its collider probe,''
4304:   Phys.\ Rev.\ D {\bf 72} (2005) 095010.
4305: 
4306: \bibitem{APRL} A. Pilaftsis, Phys.\ Rev.\ Lett.\ {\bf 95} (2005) 081602.
4307: 
4308: \bibitem{NielsenOlesen}
4309:   H.~B.~Nielsen and P.~Olesen,
4310:   %``Vortex-Line Models For Dual Strings,''
4311:   Nucl.\ Phys.\ B {\bf 61} (1973) 45.
4312:   %%CITATION = NUPHA,B61,45;%%
4313: 
4314: \bibitem{Vilenkin}
4315: A~Vilenkin and E.~P.~S.~Shellard, "Cosmological Strings and Other
4316:     Defects," Cambridge University Press, Cambridge UK (1994).
4317: 
4318: \bibitem{HK} For a pedagogical review, see,\\
4319:   M.B. Hindmarsh and T.W.B. Kibble, Rept.\ Prog.\ Phys.\ {\bf 58} (1995) 477.
4320: 
4321: 
4322: \bibitem{strings}
4323:   L.~Pogosian and T.~Vachaspati,
4324:   Phys.\ Rev.\ D {\bf 60} (1999) 083504;\\
4325:   L.~Pogosian, I.~Wasserman and M.~Wyman,
4326:   arXiv:astro-ph/0604141.
4327: 
4328: 
4329: \bibitem{JKLS} R. Jeannerot, S. Khalil, G. Lazarides and Q. Shafi,
4330: JHEP {\bf 10} (2000) 012.
4331: 
4332: 
4333: 
4334: \bibitem{LR} A.~D. Linde and A. Riotto, Phys.\ Rev.\ D~{\bf 56} (1997)
4335:   1841.
4336: 
4337: \bibitem{CP} C. Panagiotakopoulos, Phys.\ Rev.\ D~{\bf 55} (1997)
4338:   7335; Phys.\ Rev.\ D~{\bf 71} (2005) 063516.
4339: 
4340: \bibitem{BG} B. Garbrecht, hep-th/0604166.
4341: 
4342: \bibitem{JP} R.~Jeannerot and M.~Postma,
4343: %``Confronting hybrid inflation in supergravity with CMB data,''
4344:   JHEP {\bf 0505} (2005) 071.
4345: 
4346: \bibitem{GKM}
4347:   T.~Gherghetta, C.~F.~Kolda and S.~P.~Martin,
4348:   %``Flat directions in the scalar potential of the supersymmetric standard
4349:   %model,''
4350:   Nucl.\ Phys.\ B {\bf 468} (1996) 37
4351: 
4352: \bibitem{AD}
4353:   I.~Affleck and M.~Dine,
4354:   %``A New Mechanism For Baryogenesis,''
4355:   Nucl.\ Phys.\ B {\bf 249} (1985) 361.
4356: 
4357: \bibitem{DK}
4358:   M.~Dine and A.~Kusenko,
4359:   %``The origin of the matter-antimatter asymmetry,''
4360:   Rev.\ Mod.\ Phys.\  {\bf 76} (2004) 1.
4361:   %%CITATION = HEP-PH 0303065;%%
4362:   %%Cited 56 times in SPIRES-HEP
4363: 
4364: %\cite{Allahverdi:2005fq}
4365: \bibitem{AM}
4366:   R.~Allahverdi and A.~Mazumdar,
4367:   %``Quasi-thermal universe: From cosmology to colliders,''
4368:   hep-ph/0505050;
4369: %\cite{Allahverdi:2006wh}
4370: %\bibitem{AM2}
4371: %  R.~Allahverdi and A.~Mazumdar,
4372:   %``Towards a successful reheating within supersymmetry,''
4373:   hep-ph/0603244.
4374:   %%CITATION = HEP-PH 0603244;%%
4375: 
4376: %\cite{Dine:1995kz}
4377: \bibitem{DRT}
4378:   M.~Dine, L.~Randall and S.~D.~Thomas,
4379:   %``Baryogenesis from flat directions of the supersymmetric standard model,''
4380:   Nucl.\ Phys.\ B {\bf 458} (1996) 291.
4381:   %%CITATION = HEP-PH 9507453;%%
4382: 
4383: \bibitem{Jones}  D.~M.~Capper, D.~R.~T.~Jones and P.~van Nieuwenhuizen,
4384:   %``Regularization By Dimensional Reduction Of Supersymmetric And
4385:   %Nonsupersymmetric Gauge Theories,''
4386:   Nucl.\ Phys.\ B {\bf 167} (1980) 479.
4387: 
4388: \bibitem{SO10inflation}
4389:   B.~Kyae and Q.~Shafi,
4390:   %``Inflation with realistic supersymmetric SO(10),''
4391:   Phys.\ Rev.\ D {\bf 72} (2005) 063515;\\
4392:   %%CITATION = HEP-PH 0504044;%%
4393: %\bibitem{Garbrecht:2005rr}
4394:   B.~Garbrecht, T.~Prokopec and M.~G.~Schmidt,
4395:   %``SO(10) - GUT coherent baryogenesis,''
4396:   Nucl.\ Phys.\ B {\bf 736} (2006) 133.
4397:   %%CITATION = HEP-PH 0509190;%%
4398: 
4399: 
4400: \bibitem{ExtraU1fromStrings}
4401: %\bibitem{Cvetic:1995rj}
4402:   M.~Cvetic and P.~Langacker,
4403:   %``Implications of Abelian Extended Gauge Structures From String Models,''
4404:   Phys.\ Rev.\ D {\bf 54} (1996) 3570.
4405:   %%CITATION = HEP-PH 9511378;%%
4406: 
4407: \bibitem{Slansky:1981}
4408:   R.~Slansky,
4409:   %``Group Theory For Unified Model Building,''
4410:   Phys.\ Rept.\  {\bf 79} (1981) 1.
4411:   %%CITATION = PRPLC,79,1;%%
4412: 
4413: \bibitem{proton} For related suggestions in 4- and 5-dimensional
4414: theories, see,\\
4415: Q.~Shafi and Z.~Tavartkiladze,
4416:   %``Bi-maximal neutrino mixings and proton decay in SO(10) with anomalous
4417:   %flavor U(1),''
4418:   Phys.\ Lett.\ B {\bf 459} (1999) 563;
4419: Phys.\ Lett.\ B {\bf 487} (2000) 145;\\
4420: G.~Altarelli and F.~Feruglio,
4421: %``SU(5) grand unification in extra dimensions and proton decay,''
4422: Phys.\ Lett.\ B {\bf 511} (2001) 257;\\
4423: A.~B.~Kobakhidze,
4424:   %``Proton stability in TeV-scale GUTs,''
4425:   Phys.\ Lett.\ B {\bf 514} (2001) 131.
4426: 
4427: \bibitem{orbifold} S.~Forste, H.~P.~Nilles, P.~K.~S.~Vaudrevange and
4428:   A.~Wingerter,
4429: %``Heterotic brane world,''
4430:   Phys.\ Rev.\ D {\bf 70} (2004) 106008.
4431: 
4432: \bibitem{cstrings1}
4433:   B.~Allen, R.R.~Caldwell, E.P.S.~Shellard, A.~Stebbins and S.~Veeraraghavan,
4434:   Phys.\ Rev.\ Lett.\  {\bf 77} (1996) 3061.
4435: 
4436: \bibitem{cstrings} M.~Landriau and E.P.~S.~Shellard,
4437: Phys.\ Rev.\ D {\bf 69} (2004) 023003.
4438:   %%CITATION = ASTRO-PH 0302166;%%
4439: 
4440: \bibitem{cstrings2} D. Austin, E.J. Copeland and T.W.B Kibble,
4441: Phys.\ Rev.\ D {\bf 48} (1993) 5594.
4442: 
4443: \bibitem{mairi} J.~Rocher and M.~Sakellariadou,
4444:   %``Supersymmetric grand unified theories and cosmology,''
4445:   JCAP {\bf 0503} (2005) 004.
4446:   %%CITATION = HEP-PH 0406120;%%
4447: 
4448: \bibitem{hilltop} L.~Boubekeur and D.H.~Lyth, JCAP {\bf 0507} (2005) 010.
4449: 
4450: \bibitem{king} M. Bastero-Gil, S.F. King and Q. Shafi, hep-ph/0604198.
4451: 
4452: \bibitem{Sarkar} J.~R.~Ellis, D.~V.~Nanopoulos and S.~Sarkar,
4453:   %``The Cosmology Of Decaying Gravitinos,''
4454:   Nucl.\ Phys.\ B {\bf 259} (1985) 175;\\
4455: J.~R.~Ellis, G.~B.~Gelmini, J.~L.~Lopez, D.~V.~Nanopoulos and S.~Sarkar,
4456: %``Astrophysical Constraints On Massive Unstable Neutral Relic Particles,''
4457:   Nucl.\ Phys.\ B {\bf 373} (1992) 399.
4458: 
4459: \bibitem{TACHYPREH}
4460:   T.~Prokopec and T.~G.~Roos,
4461:   %``Lattice study of classical inflaton decay,''
4462:   Phys.\ Rev.\ D {\bf 55} (1997) 3768;
4463: %  [arXiv:hep-ph/9610400];
4464: %%CITATION = HEP-PH 9610400;%%
4465: \\
4466:   T.~Prokopec,
4467:   %``Negative coupling instability and grand unified baryogenesis,''
4468:   arXiv:hep-ph/9708428;
4469:   %%CITATION = HEP-PH 9708428;%%
4470: \\
4471:   G.~N.~Felder, J.~Garcia-Bellido, P.~B.~Greene, L.~Kofman,
4472:   A.~D.~Linde and I.~Tkachev,
4473:   %``Dynamics of symmetry breaking and tachyonic preheating,''
4474:   Phys.\ Rev.\ Lett.\  {\bf 87} (2001) 011601.
4475: %[arXiv:hep-ph/0012142];
4476: %%CITATION = HEP-PH 0012142;%%
4477: 
4478: \bibitem{PREHEATING}
4479:   J.~H.~Traschen and R.~H.~Brandenberger,
4480:   %``Particle Production During Out-Of-Equilibrium Phase Transitions,''
4481:   Phys.\ Rev.\ D {\bf 42} (1990) 2491;
4482: \\
4483:   L.~Kofman, A.~D.~Linde and A.~A.~Starobinsky,
4484:   %``Towards the theory of reheating after inflation,''
4485:   Phys.\ Rev.\ D {\bf 56} (1997) 3258;
4486: %  [arXiv:hep-ph/9704452];
4487:   %%CITATION = HEP-PH 9704452;%%
4488: \\
4489:   D.~J.~H.~Chung, E.~W.~Kolb, A.~Riotto and I.~I.~Tkachev,
4490:   %``Probing Planckian physics: Resonant production of particles during
4491:   %inflation and features in the primordial power spectrum,''
4492:   Phys.\ Rev.\ D {\bf 62} (2000) 043508;
4493: %  [arXiv:hep-ph/9910437];
4494: \\
4495:   M.~Peloso and L.~Sorbo,
4496:   %``Preheating of massive fermions after inflation: Analytical results,''
4497:   JHEP {\bf 0005} (2000) 016;
4498: %  [arXiv:hep-ph/0003045];
4499: \\
4500:   B.~Garbrecht, T.~Prokopec and M.~G.~Schmidt,
4501:   %``Particle number in kinetic theory,''
4502:   Eur.\ Phys.\ J.\ C {\bf 38} (2004) 135;
4503: %  [arXiv:hep-th/0211219];
4504: \\
4505:   J.~Berges and J.~Serreau,
4506:   %``Parametric resonance in quantum field theory,''
4507:   Phys.\ Rev.\ Lett.\  {\bf 91} (2003) 111601.
4508: %  [arXiv:hep-ph/0208070].
4509: 
4510: \bibitem{GBRM}
4511:   J.~Garcia-Bellido and E.~Ruiz Morales,
4512:   %``Particle production from symmetry breaking after inflation,''
4513:   Phys.\ Lett.\ B {\bf 536} (2002) 193.
4514: %[arXiv:hep-ph/0109230].
4515: %%CITATION = HEP-PH 0109230;%%
4516: 
4517: \bibitem{kohri} M.~Kawasaki, K.~Kohri and T.~Moroi,  Phys.\ Lett.\ B {\bf
4518:   625} (2005) 7;\\
4519:   M. Kawasaki, K. Kohri and T. Moroi,   Phys.\ Rev.\ D {\bf 71}
4520:   (2005) 083502.
4521: 
4522: \bibitem{brand} M. Bolz, A. Brandenburg and W. Buchmuller,
4523: Nucl.\ Phys.\ B {\bf 606} (2001) 518.
4524: 
4525: \bibitem{kolb} E.W.~Kolb and M.S.~Turner,
4526: %``The Early Universe,''
4527: {\it  Redwood City, USA: Addison-Wesley (1990)}.
4528: 
4529: \bibitem{riotto} G.F. Giudice, E.W. Kolb, A. Riotto, Phys.\ Rev.\ D {\bf
4530:   64} (2001) 023508.
4531: 
4532: \bibitem{qui} C. Pallis, Astropart.\ Phys.\ {\bf 21} (2004) 689;
4533:   hep-ph/0510234.
4534: 
4535: \bibitem{oliveg} R.H.~Cyburt, J.R.~Ellis, B.D.~Fields and K.A.~Olive,
4536:   Phys.\ Rev.\ D {\bf 67} (2003) 103521;\\
4537:   J.R.~Ellis, K.A.~Olive and E.~Vangioni, Phys.\ Lett.\ B {\bf 619} (2005) 30.
4538: 
4539: \bibitem{LS} G. Lazarides and Q. Shafi, Phys.\ Lett.\ {\bf B258}
4540:   (1991) 305;\\
4541:   H. Murayama, H. Suzuki, T. Yanagida and J. Yokoyama,
4542:   Phys.\ Rev.\ Lett.\ {\bf 70} (1993) 1912; Phys.\ Rev.\ {\bf D50}
4543:   (1994) 2356;\\
4544:   T. Asaka, K. Hamaguchi, M. Kawasaki and T. Yanagida,
4545:   Phys.\ Lett.\ {\bf B464} (1999) 12.
4546: 
4547: \bibitem{DLES} D.H. Lyth and E.D. Stewart, Phys.\ Rev.\ {\bf D53}
4548:   (1996) 1784.
4549: 
4550: \bibitem{Branco} R.~Gonzalez Felipe, F.~R.~Joaquim and B.~M.~Nobre,
4551:   %``Radiatively induced leptogenesis in a minimal seesaw model,''
4552:   Phys.\ Rev.\ D {\bf 70} (2004) 085009;\\
4553:   G.~C.~Branco, R.~Gonzalez Felipe, F.~R.~Joaquim and B.~M.~Nobre,
4554:   %``Enlarging the window for radiative leptogenesis,''
4555:   hep-ph/0507092.
4556: 
4557: \bibitem{Lindner} For recent works on GUT flavour symmetries, see,\\
4558:   C.~Hagedorn, M.~Lindner and R.~N.~Mohapatra,
4559:   %``S(4) flavor symmetry and fermion masses: Towards a grand unified theory of
4560:   %flavor,''
4561:   hep-ph/0602244;\\
4562:   C.~Hagedorn, M.~Lindner and F.~Plentinger,
4563:   %``The discrete flavor symmetry D(5),''
4564:   hep-ph/0604265.
4565: 
4566: \bibitem{BAULFV} R.~Barbieri, P.~Creminelli, A.~Strumia and N.~Tetradis,
4567:   %``Baryogenesis through leptogenesis,''
4568:   Nucl.\ Phys.\ B {\bf 575} (2000) 61;\\
4569: E.~Nardi, Y.~Nir, J.~Racker and E.~Roulet,
4570:   %``On Higgs and sphaleron effects during the leptogenesis era,''
4571:   JHEP {\bf 0601} (2006) 068;\\
4572: A.~Abada, S.~Davidson, F.~X.~Josse-Michaux, M.~Losada and A.~Riotto,
4573:   %``Flavour issues in leptogenesis,''
4574:   JCAP {\bf 0604} (2006) 004;\\
4575: E.~Nardi, Y.~Nir, E.~Roulet and J.~Racker,
4576:   %``The importance of flavor in leptogenesis,''
4577:   JHEP {\bf 0601} (2006) 164.
4578: 
4579: \bibitem{Bari} P.~Di Bari,
4580:   %``Seesaw geometry and leptogenesis,''
4581:   Nucl.\ Phys.\ B {\bf 727} (2005) 318;\\
4582: O.~Vives,
4583:   %``Flavoured leptogenesis: A successful thermal leptogenesis with N(1) mass
4584:   %below 10**8-GeV,''
4585:   Phys.\ Rev.\ D {\bf 73} (2006) 073006.
4586: 
4587: \bibitem{NprodLHC} A.~Pilaftsis,
4588:   %``Radiatively induced neutrino masses and large Higgs neutrino couplings in
4589:   %the standard model with Majorana fields,''
4590:   Z.\ Phys.\ C {\bf 55} (1992) 275;\\
4591: A.~Datta, M.~Guchait and A.~Pilaftsis,
4592:   %``Probing lepton number violation via majorana neutrinos at hadron
4593:   %supercolliders,''
4594:   Phys.\ Rev.\ D {\bf 50} (1994) 3195;\\
4595: F.~M.~L.~Almeida, Y.~A.~Coutinho, J.~A.~Martins Simoes and M.~A.~B.~do Vale,
4596:   %``On a signature for heavy Majorana neutrinos in hadronic collisions,''
4597:   Phys.\ Rev.\ D {\bf 62} (2000) 075004;\\
4598:  T.~Han and B.~Zhang,
4599:   %``Signatures for Majorana neutrinos at hadron colliders,''
4600:   hep-ph/0604064.
4601: 
4602: 
4603: \bibitem{NprodILC} W.~Buchmuller and C.~Greub,
4604: %``Heavy Majorana neutrinos in electron - positron and electron - proton
4605: % collisions,''
4606:   Nucl.\ Phys.\ B {\bf 363}  (1991) 345;\\
4607: %%CITATION = NUPHA,B363,345;%%
4608: G.~Cvetic, C.~S.~Kim and C.~W.~Kim,
4609: % ``Heavy Majorana neutrinos at e+ e- colliders,''
4610:   Phys.\ Rev.\ Lett.\  {\bf 82} (1999) 4761;\\
4611: F.~del Aguila, J.~A.~Aguilar-Saavedra, A.~Martinez de la Ossa and D.~Meloni,
4612: %``Flavour and polarisation in heavy neutrino production at e+ e- colliders,''
4613:   Phys.\ Lett.\ B~{\bf 613} (2005) 170;\\
4614: F.~del Aguila and J.~A.~Aguilar-Saavedra,
4615: JHEP {\bf 0505} (2005) 026.
4616: 
4617: \bibitem{NprodEG}
4618: S.~Bray, J.~S.~Lee and A.~Pilaftsis,
4619:   %``Heavy Majorana neutrino production at e- gamma colliders,''
4620:   Phys.\ Lett.\ B {\bf 628} (2005) 250.
4621: 
4622: \bibitem{CL} T.P.~Cheng and L.F.~Li, Phys.\ Rev.\ Lett.\ {\bf 45}
4623:   (1980) 1908.
4624: 
4625: \bibitem{IP} A. Ilakovac and A. Pilaftsis, Nucl.\ Phys.\ B {\bf 437}
4626:   (1995) 491.
4627: 
4628: \bibitem{LFVrev}
4629:   G.  Bhattacharya, P. Kalyniak and I. Mello, Phys.\
4630:   Rev.\ D {\bf 51} (1995) 3569;\\
4631:   A. Ilakovac, B.A.  Kniehl, and A. Pilaftsis, Phys.\ Rev.\
4632:   D {\bf 52} (1995) 3993;\\
4633:   A. Ilakovac, Phys.\ Rev.\ D {\bf 54} (1996) 5653;\\
4634:   M. Frank and H.  Hamidian, Phys.\ Rev.\ D {\bf 54} (1996) 6790;\\
4635:   P.  Kalyniak and I. Mello, Phys.\ Rev.\ D {\bf 55} (1997) 1453;\\
4636:   Z. Gagyi-Palffy, A.  Pilaftsis and K.  Schilcher, Nucl.\ Phys.\ B {\bf
4637:   513} (1998) 517;\\
4638:   S.  Fajfer and A.  Ilakovac, Phys.\ Rev.\ D {\bf 57} (1998) 4219;\\
4639:   M.  Raidal and A.  Santamaria, Phys.\ Lett.\ B {\bf 421} (1998) 250;\\
4640:   M.~Czakon, M.~Zralek and J.~Gluza, Nucl.\ Phys.\ B {\bf 573} (2000) 57;\\
4641:   A. Ioannisian and A. Pilaftsis, Phys.\ Rev.\ D {\bf 62}
4642:   (2000) 066001;\\
4643:   J.~I.~Illana and T.~Riemann, Phys.\ Rev.\ D {\bf 63} (2001) 053004;\\
4644:   G.~Cvetic, C.~Dib, C.~S.~Kim and J.~D.~Kim, Phys.\ Rev.\ D {\bf 66} (2002)
4645:   034008; hep-ph/0504126;\\
4646:   A.~Masiero, S.~K.~Vempati and O.~Vives, New J.\ Phys.\  {\bf 6}
4647:   (2004) 202.
4648: 
4649: \bibitem{LFVN} For recent studies, see,\\
4650: T.~Fujihara, S.~K.~Kang, C.~S.~Kim, D.~Kimura and T.~Morozumi,
4651: %``Low-scale seesaw model and lepton flavor violating rare B decays,''
4652:   hep-ph/0512010;\\
4653: F.~Deppisch, T.~S.~Kosmas and J.~W.~F.~Valle,
4654:   %``Enhanced mu- e- conversion in nuclei in the inverse seesaw model,''
4655:   arXiv:hep-ph/0512360.
4656: 
4657: \bibitem{LLfit} C.~P.~Burgess, S.~Godfrey, H.~Konig, D.~London and
4658: I.~Maksymyk, Phys.\ Rev.\ D {\bf 49} (1994) 6115;\\
4659: E.~Nardi, E.~Roulet and D.~Tommasini, Phys.\ Lett.\ B~{\bf 327} (1994) 319;\\
4660: D.~Tommasini, G.~Barenboim, J.~Bernabeu and C.~Jarlskog, Nucl.\ Phys.\
4661: B {\bf 444} (1995) 451;\\
4662: S.~Bergmann and A.~Kagan, Nucl.\ Phys.\ B~{\bf 538} (1999) 368.
4663: 
4664: \bibitem{GGP} S.~Gopalakrishna, A.~de Gouvea and W.~Porod,
4665:   %``Right-handed sneutrinos as nonthermal dark matter,''
4666:   hep-ph/0602027.
4667: 
4668: \bibitem{mcdonald}  J.~McDonald, Phys.\ Rev.\ D {\bf 50} (1994) 3637.
4669: 
4670: \bibitem{pospelov} C.P.~Burgess, M.~Pospelov and T.~ter Veldhuis,
4671:   Nucl.\ Phys.\ B {\bf 619} (2001) 709.
4672: 
4673: \bibitem{Sabine} For recent studies within the context of the
4674: CP-violating MSSM, see,\\
4675: M.~E.~Gomez, T.~Ibrahim, P.~Nath and S.~Skadhauge,
4676:   %``WMAP dark matter constraints and Yukawa unification in SUGRA models  with
4677:   %CP phases,''
4678:   Phys.\ Rev.\ D {\bf 72} (2005) 095008;\\
4679: G.~Belanger, F.~Boudjema, S.~Kraml, A.~Pukhov and A.~Semenov,
4680: %``Relic density of neutralino dark matter in the MSSM with CP violation,''
4681:   arXiv:hep-ph/0604150.
4682: 
4683: \bibitem{EWBAU} For recent analyses, see,\\
4684: M.~Carena, M.~Quiros, M.~Seco and C.~E.~M.~Wagner,
4685:   %``Improved results in supersymmetric electroweak baryogenesis,''
4686:   Nucl.\ Phys.\ B {\bf 650} (2003) 24;\\
4687: T.~Konstandin, T.~Prokopec and M.~G.~Schmidt,
4688: %``Kinetic description of fermion flavor mixing and CP-violating sources  for
4689: %baryogenesis,''
4690:   Nucl.\ Phys.\ B {\bf 716} (2005) 373;\\ M.~Carena, A.~Megevand,
4691: M.~Quiros and C.~E.~M.~Wagner, Nucl.\ Phys.\ B {\bf 716} (2005)~319;\\
4692: T.~Konstandin, T.~Prokopec, M.~G.~Schmidt and M.~Seco,
4693: %``MSSM electroweak baryogenesis and flavour mixing in transport equations,''
4694: Nucl.\ Phys.\ B {\bf 738} (2006)~1.
4695: 
4696: 
4697: \bibitem{CKP}
4698: D.~Chang, W.~Y.~Keung and A.~Pilaftsis,
4699: %``New two-loop contribution to electric dipole moment in supersymmetric
4700: %theories,''
4701: Phys.\ Rev.\ Lett.\  {\bf 82} (1999) 900;\\
4702: A.~Pilaftsis,
4703: %``Higgs-mediated electric dipole moments in the MSSM: An application to
4704: %baryogenesis and Higgs searches,''
4705: Nucl.\ Phys.\ B {\bf 644} (2002) 263;\\
4706:  D.~Chang, W.~F.~Chang and W.~Y.~Keung,
4707: %``New constraint from electric dipole moments on chargino baryogenesis in
4708: %MSSM,''
4709:   Phys.\ Rev.\ D {\bf 66} (2002) 116008.
4710: 
4711: \bibitem{CPX} M.~Carena, J.~R.~Ellis, A.~Pilaftsis and C.~E.~M.~Wagner,
4712:   %``CP-violating MSSM Higgs bosons in the light of LEP 2,''
4713:   Phys.\ Lett.\ B {\bf 495} (2000) 155.
4714: 
4715: \bibitem{DP} D.~K.~Ghosh, R.~M.~Godbole and D.~P.~Roy,
4716: %``Probing the CP-violating light neutral Higgs in the charged Higgs decay at
4717: %the LHC,''
4718:   Phys.\ Lett.\ B {\bf 628} (2005) 131;\\
4719: D.~K.~Ghosh and S.~Moretti,
4720:   %``Probing the light neutral Higgs boson scenario of the CP-violating MSSM
4721:   %Higgs sector at the LHC,''
4722:   Eur.\ Phys.\ J.\ C {\bf 42} (2005) 341.
4723: 
4724: 
4725: \end{thebibliography}
4726: 
4727: \end{document}
4728: