1: \documentclass[aps,prd,twocolumn,preprintnumbers,
2: superscriptaddress,showpacs,floatfix]{revtex4}
3:
4: \usepackage{bm}
5: \usepackage{epsfig}
6: \usepackage{graphics}
7: \usepackage{amsmath,amssymb}
8: \usepackage{array}
9:
10: \newcommand{\pdf}{\mathcal{P}}
11: \newcommand{\data}{{\bf D}}
12: \newcommand{\mdl}{\mathcal{M}}
13: \newcommand{\lsim}{\,\raise 0.4ex\hbox{$<$}
14: \kern -0.8em\lower 0.62ex\hbox{$\sim$}\,}
15: \newcommand{\params}{\boldsymbol{\theta}}
16: \newcommand{\mean}{\boldsymbol{\mu}}
17: \newcommand{\like}{L}
18: \newcommand{\lnlike}{\mathcal{L}}
19: \newcommand{\ML}{^*}
20: \newcommand{\dr}{\textrm{d}}
21: \newcommand{\LSP}{\text{LSP}}
22: \newcommand{\be}{\begin{equation}}
23: \newcommand{\ee}{\end{equation}}
24: \newcommand{\Om}{\Omega}
25: \newcommand{\om}{\omega}
26: \newcommand{\Cl}{C_\ell}
27: \newcommand{\ad}{\text{ad}}
28: \newcommand{\is}{\text{is}}
29: \newcommand{\cc}{\text{cc}}
30: \newcommand{\Gev}{\text{GeV}}
31: \newcommand{\Mpc}{\text{Mpc}}
32: \newcommand{\Hbl}{\mathcal{H}}
33: \newcommand{\uf}{\;[10^{-5}]}
34: \newcommand{\ut}{\;[10^{-10}]}
35: \newcommand{\mc}[1]{\mathcal{#1}}
36: \newcommand{\Rrad}{\mc{R}_\text{rad}}
37: \newcommand{\Srad}{\mc{S}_\text{rad}}
38: \newcommand{\EX}[1]{\langle #1 \rangle}
39: \newcommand{\kp}{\sharp}
40:
41: \begin{document}
42:
43: \preprint{UT-STPD-3/04}
44:
45: \title{
46: Constraints on a mixed inflaton and curvaton
47: scenario \\ for the generation of the
48: curvature perturbation}
49:
50: \author{George Lazarides}
51: \email{lazaride@eng.auth.gr}
52: \affiliation{Physics Division, School of
53: Technology, Aristotle University of
54: Thessaloniki, Thessaloniki 54124, Greece}
55: \author{Roberto Ruiz de Austri}
56: \email{r.ruizdeaustri@sheffield.ac.uk}
57: \affiliation{University of Sheffield, Department
58: of Physics and Astronomy, Hicks Building,
59: Hounsfield Road, Sheffield S3 7RH, England}
60: \author{Roberto Trotta}
61: \email{roberto.trotta@physics.unige.ch}
62: \affiliation{D\'epartement de Physique
63: Th\'eorique, Universit\'e de Gen\`eve, 24
64: quai Ernest Ansermet, 1211 Gen\`eve 4,
65: Switzerland}
66:
67: \date{\today}
68:
69: \begin{abstract}
70: We consider a simple supersymmetric grand
71: unified model which naturally solves the
72: strong CP and $\mu$ problems via a Peccei-Quinn
73: symmetry and leads to the standard realization
74: of hybrid inflation. We show that the
75: Peccei-Quinn field of this model can act as a
76: curvaton. In contrast to the standard curvaton
77: hypothesis, both the inflaton and the curvaton
78: contribute to the total curvature perturbation.
79: The model predicts the existence of an
80: isocurvature perturbation too which has mixed
81: correlation with the adiabatic one. The cold
82: dark matter of the universe is mostly
83: constituted by axions, which are produced at
84: the QCD phase transition, plus a small amount
85: of lightest sparticles. The predictions of the
86: model are confronted with the first-year
87: Wilkinson microwave anisotropy probe and other
88: cosmic microwave background radiation data. We
89: analyze in detail two representative choices of
90: parameters for our model and derive bounds on
91: the curvaton contribution to the adiabatic
92: perturbation. We find that, for the choice
93: which provides the best fitting of the data,
94: the curvaton contribution to the amplitude of
95: the adiabatic perturbation must be smaller than
96: about $67\%$ and the amplitude of the partial
97: curvature perturbation from the curvaton
98: smaller than $43.2\times 10^{-5}$ (both at
99: $95\%$ confidence level). The best-fit power
100: spectra are dominated by the adiabatic part of
101: the inflaton contribution. We use Bayesian
102: model comparison to show that this choice of
103: parameters is disfavored with respect to the
104: pure inflaton scale-invariant case with odds of
105: about 50 to 1. For the second choice of
106: parameters examined, the adiabatic mode is
107: dominated by the curvaton, but this choice is
108: strongly disfavored relative to the pure
109: inflaton scale-invariant case (with odds
110: of about $10^7$ to 1). We conclude that in the
111: present framework the perturbations must be
112: dominated by the adiabatic component from the
113: inflaton.
114:
115: \end{abstract}
116:
117: \pacs{12.10.-g,12.60.Jv,98.80.Cq}
118:
119: \maketitle
120:
121: \section{Introduction}
122: \label{intro}
123:
124: \par
125: Inflation, which was originally proposed
126: \cite{guth} as a solution to the outstanding
127: problems of standard big bang cosmology and the
128: problem of unwanted relics, is in good agreement
129: with the recent measurements \cite{wmap1,wmap2}
130: on the angular power spectrum of the cosmic
131: microwave background radiation (CMBR). Moreover,
132: inflation is now established as the most likely
133: origin of the primordial density perturbation
134: from which structure formation in the universe
135: proceeded \cite{llbook}. According to the
136: usual assumption \cite{llbook,lectures}, this
137: perturbation is generated solely by the slowly
138: rolling inflaton field of the usual one-field
139: inflationary models and, thus, is expected to
140: be purely adiabatic. However, although adiabatic
141: perturbations are perfectly consistent with the
142: present data, the presence of a significant
143: isocurvature density perturbation cannot be
144: excluded \cite{wmap2,iso1}. In one-field
145: inflation, the perturbations are almost
146: Gaussian, in agreement with the current upper
147: bounds on non-Gaussianity from the CMBR data,
148: which though cannot exclude the presence of
149: appreciable non-Gaussianity (for a review, see
150: e.g. Ref.~\cite{nongauss}).
151:
152: \par
153: Lately, the alternative possibility
154: \cite{curv1,curv2} that the adiabatic density
155: perturbations originate from the inflationary
156: perturbations of some light ``curvaton'' field
157: different from the inflaton has attracted much
158: attention. The curvaton density perturbations
159: can lead \cite{curv2,curv3}, after curvaton
160: decay, to isocurvature perturbations in the
161: densities of the various components of the
162: cosmic fluid. In the simplest case, the
163: residual isocurvature perturbations are either
164: fully correlated or fully anti-correlated with
165: the adiabatic density perturbation. In general
166: models, however, the correlation can be mixed
167: (see e.g. Ref.~\cite{dllr1}). In the curvaton
168: scenario, significant non-Gaussianity may also
169: appear. The main reason for advocating the
170: curvaton hypothesis is that it makes
171: \cite{dimo} the task of constructing viable
172: models of inflation much easier, since it
173: liberates us from the very restrictive
174: requirement that the inflaton is responsible
175: for the curvature perturbations.
176:
177: \par
178: In a recent paper \cite{dllr1}, a simple
179: extension \cite{rsym} of the minimal
180: supersymmetric standard model (MSSM) which
181: naturally and simultaneously solves the strong
182: CP and $\mu$ problems via a Peccei-Quinn (PQ)
183: \cite{pq} and a continuous R symmetry was
184: considered within the general framework of the
185: standard supersymmetric (SUSY) version
186: \cite{lyth,dss} of hybrid inflation
187: \cite{hybrid}. It was shown that, in this model,
188: the PQ field, which breaks spontaneously the
189: PQ symmetry, can successfully act as a curvaton
190: generating the total curvature perturbation
191: in the universe in accordance with the cosmic
192: background explorer (COBE) measurements
193: \cite{cobe}. The (intermediate) PQ scale
194: is generated by invoking supergravity (SUGRA)
195: and the PQ field corresponds to a flat
196: direction in field space lifted by
197: non-renormalizable interactions. Moreover,
198: the $\mu$ parameter of MSSM is generated
199: \cite{kn} from the PQ scale.
200:
201: \par
202: We feel that the standard curvaton hypothesis
203: \cite{curv1,curv2}, which insists that the
204: total curvature perturbation originates
205: solely from the curvaton, can also be quite
206: restrictive and not so natural. Indeed, in
207: accordance to this hypothesis, one needs to
208: impose \cite{dimo} an upper bound on the
209: inflationary scale in order to ensure that the
210: perturbation from the inflaton is negligible.
211: This bound can be quite strong especially if
212: the slow-roll parameter $\epsilon$ (see e.g.
213: Ref.~\cite{lectures}) happens to be very small,
214: which holds in many cases. In generic models,
215: one would expect that all the scalar fields
216: which are essentially massless during inflation
217: contribute to the total curvature perturbation.
218: So, in the presence a PQ field, which can be
219: kept \cite{dllr1} light during the relevant
220: part of inflation, it is natural to assume that
221: the adiabatic density perturbation is partly
222: due to this field and partly to the inflaton.
223:
224: \par
225: There is yet another reason for abandoning the
226: strict curvaton hypothesis. The recent
227: measurements on the CMBR by the Wilkinson
228: microwave anisotropy probe (WMAP) satellite
229: \cite{wmap1} have considerably strengthened
230: \cite{wmap2,iso1,gm} the bound on the
231: isocurvature perturbation which was obtained
232: \cite{iso2,iso3} by using the pre-WMAP CMBR
233: data. As a consequence, the viability
234: of many curvaton models is in doubt. However,
235: allowing a significant part of the total
236: curvature perturbation in the universe to
237: originate from the inflaton, we can hopefully
238: relax the tension between these models and
239: the present WMAP data without losing the main
240: advantage of the curvaton hypothesis, which
241: is that it facilitate the construction of
242: viable inflationary models (see
243: Ref.~\cite{mixed} for recent investigations of
244: this possibility).
245:
246: \par
247: The PQ curvaton model of Ref.~\cite{dllr1}
248: predicts an isocurvature perturbation of
249: mixed correlation with the curvature
250: perturbation. The extended set of
251: pre-WMAP CMBR and other data which was used
252: \cite{dllr1} to restrict the isocurvature
253: perturbation in the model of
254: Ref.~\cite{dllr1} led to the exclusion of a
255: major part of the available parameter space.
256: Including the more restrictive recent WMAP
257: measurements, the allowed parameter space
258: will certainly be further reduced. It is,
259: moreover, quite possible that the model is
260: even totally excluded by these new data. So,
261: the departure from the strict curvaton
262: hypothesis may prove to be vital for this
263: particular curvaton scheme.
264:
265: \par
266: In this paper, we will extend the PQ curvaton
267: model of Ref.~\cite{dllr1} by embedding it
268: into the concrete SUSY grand unified theory
269: (GUT) model studied in Ref.~\cite{hier}, which
270: is based on the left-right (LR) symmetric gauge
271: group $G_{LR}={\rm SU}(3)_c\times {\rm SU}(2)_L
272: \times{\rm SU}(2)_R\times {\rm U}(1)_{B-L}$.
273: This model leads \cite{hier} naturally to
274: standard SUSY hybrid inflation \cite{lyth,dss}.
275: After the end of inflation, the inflaton
276: performs damped oscillations about the SUSY
277: vacuum and eventually decays into right handed
278: neutrino superfields reheating the universe.
279: The subsequent decay of these superfields to
280: a lepton and an electroweak Higgs superfield
281: generates \cite{leptoinf} a primordial lepton
282: asymmetry \cite{lepto} which is then partly
283: converted into baryon asymmetry by
284: non-perturbative electroweak sphaleron effects.
285: The observed baryon asymmetry of the universe
286: (BAU) can then be easily reproduced \cite{hier}
287: in accord with the data on neutrino masses and
288: mixing. At reheating, gravitinos are also
289: produced thermally. They decay in the late
290: universe leading to lightest sparticles (LSPs),
291: which contribute to the cold dark matter (CDM)
292: in the universe. For simplicity, we assume
293: that this is the only source of relic LSPs
294: neglecting their direct thermal production.
295: Due to the presence of the PQ symmetry, our
296: model contains axions which come into play at
297: the QCD phase transition and can also
298: contribute to CDM.
299:
300: \par
301: The PQ field of our model can acquire a
302: super-horizon spectrum of perturbations from
303: inflation provided that it is effectively
304: massless during the relevant part of inflation.
305: It can thus act as curvaton contributing to
306: the total curvature perturbation together with
307: the inflaton. We study the evolution of the PQ
308: field during and after inflation by including
309: corrections \cite{lyth,crisis,drt95} to the PQ
310: potential which originate from the SUSY
311: breaking in the early universe caused by the
312: presence of a finite energy density. We assume
313: that these corrections are (somewhat)
314: suppressed, which is \cite{noscale} indeed the
315: case if specific K\"{a}hler potentials are
316: used.
317:
318: \par
319: The requirement that the PQ field is
320: essentially massless during inflation yields
321: \cite{dllr1}, for given values of the other
322: parameters, an upper bound on the possible
323: values of this field at the end of inflation.
324: Moreover, it implies that, as inflation
325: terminates, the PQ field emerges \cite{dllr1}
326: with negligible velocity. There is also a
327: lower bound on the value of the PQ field at
328: the end of inflation below which the classical
329: equation of motion during inflation for the
330: mean value of this field in a region of fixed
331: size somewhat bigger than the size of the de
332: Sitter horizon ceases \cite{dllr1} to be valid.
333: This is due to the fact that the quantum
334: perturbations of the PQ field from inflation
335: overshadow its classical kinetic energy
336: density. We will exclude this quantum regime
337: since the calculation of the spectral index of
338: the partial curvature perturbation from the
339: curvaton field in this regime is not so clear.
340:
341: \par
342: The values of the PQ field at the end of
343: inflation which are not excluded by the above
344: considerations can be classified according
345: to whether they lead to the PQ vacua or the
346: trivial (false) vacuum. We find that,
347: generically, there exist two separate bands
348: of such values leading to the correct (PQ)
349: vacua. One of them lies at relatively low
350: values of the curvaton field at the end of
351: inflation, while the other lies at values
352: which are considerably higher. In all other
353: cases, the system ends up in the wrong
354: (trivial) vacuum and thus the corresponding
355: values of the PQ field at the end of
356: inflation must be excluded. Our numerical
357: findings show that the (approximate) COBE
358: constraint on the CMBR can be satisfied
359: only within the upper allowed band. This
360: constraint receives contributions not only
361: from the curvature perturbation but also
362: from the isocurvature one and the cross
363: correlation of the two. Note, however, that
364: this constraint is quite approximate and
365: can be considered only as an indicative
366: criterion.
367:
368: \par
369: The amplitude and spectral index of the
370: partial curvature perturbation from the
371: inflaton are calculated by employing the
372: analysis of Ref.~\cite{lectures} slightly
373: modified to allow for the possibility that
374: the slow-roll conditions are violated and,
375: thus, hybrid inflation ends before
376: reaching the instability point on the
377: inflationary trajectory. The partial
378: curvature perturbation from the curvaton
379: is treated in a more accurate way than in
380: Ref.~\cite{dllr1}. In particular, the
381: evolution during inflation of the
382: perturbation acquired by the PQ field from
383: inflation when our present horizon scale
384: crossed outside the inflationary horizon
385: is considered. It is described by the
386: classical equation of motion for this field
387: in the slow-roll approximation. Solving
388: this equation, we can find the perturbation
389: in the value of the curvaton field at the
390: end of inflation. This same calculation
391: yields the spectral index for the curvaton
392: too. For any given value of the
393: PQ field at the termination of inflation, we
394: take the perturbed value too and follow the
395: subsequent evolution of both these fields
396: until the time of the curvaton decay. This
397: yields the amplitude of the partial curvature
398: perturbation from the curvaton. The total
399: curvature perturbation is then given by the
400: appropriate weighted sum of these two
401: uncorrelated perturbations.
402:
403: \par
404: As mentioned already, the baryons and the
405: LSPs in our model originate from reheating.
406: They thus inherit the partial curvature
407: perturbation of the oscillating and decaying
408: inflaton, which is different from the total
409: curvature perturbation due to the presence of
410: the curvaton. As a consequence, the baryons
411: and LSPs acquire an isocurvature perturbation
412: of mixed correlation with the total curvature
413: perturbation. The CDM in our model contains
414: also axions carrying an isocurvature
415: perturbation which is uncorrelated with the
416: perturbations from the inflaton and the
417: curvaton. The amplitude and spectral index of
418: this isocurvature perturbation is evaluated by
419: following the analysis of Ref.~\cite{dllr1}.
420: We see that, in our model, the correlation of
421: the total adiabatic and isocurvature
422: perturbations is mixed.
423:
424: \par
425: For given values of all the other independent
426: parameters, we take a grid of values of
427: the curvaton field at the end of inflation and
428: the amplitude of partial curvature perturbation
429: from the inflaton which cover the upper or the
430: lower allowed band. We calculate the total CMBR
431: angular power spectrum for each point in this
432: grid. The predictions from each band are
433: confronted with the CMBR temperature (TT) and
434: temperature-polarization (TE) cross correlation
435: angular power spectra from the first-year WMAP
436: observations \cite{wmap1} as well as the CMBR
437: data on smaller scales from the arcminute
438: cosmology bolometer array receiver (ACBAR)
439: \cite{acbar1,acbar2} and the cosmic background
440: imager (CBI) \cite{cbi} experiments. We then
441: study the implications of the resulting
442: restrictions on the various parameters of the
443: model. We also employ Bayesian model testing to
444: compare our model with the standard pure
445: inflaton single-field inflationary model with
446: scale-invariant perturbations. We are
447: particularly interested to see whether the
448: data favor the presence of a non-vanishing
449: curvaton contribution to the adiabatic
450: perturbation.
451:
452: \par
453: The paper is organized as follows. In
454: Sec.~\ref{sec:model}, we outline the salient
455: features of our LR SUSY GUT model which solves
456: the strong CP and $\mu$ problems via a PQ
457: symmetry and leads to the standard version of
458: SUSY hybrid inflation. The evolution of the PQ
459: field during and after inflation as well as its
460: final decay into light particles are sketched
461: in Sec.~\ref{sec:early}.
462: Section~\ref{sec:curvature} is devoted to the
463: evaluation of the total curvature perturbation
464: which, in our case, is partly due to the
465: inflaton and partly to the curvaton. In
466: Sec.~\ref{sec:isocurvature}, we estimate the
467: isocurvature perturbations in the relic
468: density of the baryons, the LSPs and the
469: axions. The total CMBR angular power spectrum
470: predicted by our model is discussed in
471: Sec.~\ref{sec:spectrum}. Our numerical
472: calculation and results are presented and
473: discussed in Sec.~\ref{sec:numerics}, and our
474: conclusions are summarized in
475: Sec.~\ref{sec:conclusion}. Finally, in the
476: Appendix, we review some useful concepts and
477: results from Bayesian statistics.
478:
479:
480: \section{The left-right SUSY GUT model}
481: \label{sec:model}
482:
483: \par
484: We will adopt here the SUSY GUT model of
485: Ref.~\cite{hier} (see also Ref.~\cite{trieste})
486: which is based on the LR symmetric gauge group
487: $G_{LR}$. The ${\rm SU}(2)_{L}$ doublet left
488: handed quark and lepton superfields are denoted
489: by $q_i$ and $l_i$ respectively, whereas the
490: ${\rm SU}(2)_{R}$ doublet antiquark and
491: antilepton superfields by $q_i^c$ and $l^c_i$
492: respectively ($i$=1,2,3 is the family index).
493: The electroweak Higgs superfield $h$ belongs to
494: a bidoublet $(2,2)_0$ representation of
495: ${\rm SU}(2)_L\times {\rm SU}(2)_R\times
496: {\rm U}(1)_{B-L}$.
497:
498: \par
499: The breaking of $G_{LR}$ to the standard model
500: (SM) gauge group $G_{\rm SM}$, at a superheavy
501: scale $M\sim 10^{16}~{\rm{GeV}}$, is achieved
502: through the superpotential
503: \begin{equation}
504: \delta W_1=\kappa S(l_H^c\bar l_H^{c}-M^2),
505: \label{W1}
506: \end{equation}
507: where $l_H^c$, $\bar l_H^{c}$ is a conjugate pair of
508: ${\rm SU}(2)_R$ doublet left handed Higgs superfields
509: with $B-L$ charges equal to $1,-1$ respectively, and
510: $S$ is a gauge singlet left handed superfield. The
511: dimensionless coupling constant $\kappa$ and the
512: mass parameter $M$ can be made real and positive by
513: suitable rephasing of the fields. The SUSY minima of
514: the scalar potential lie on the D-flat direction
515: $l_H^c=\bar l_H^{c*}$ at $\langle S\rangle=0$,
516: $|\langle l_H^c\rangle|=|\langle\bar l_H^{c}\rangle|
517: =M$.
518:
519: \par
520: The model also contains two extra gauge singlet left
521: handed superfields $N$ and $\bar{N}$ for solving
522: \cite{rsym} the $\mu$ problem of MSSM via a
523: PQ symmetry \cite{pq}, which also solves the strong
524: CP problem. They have the following superpotential
525: couplings:
526: \begin{equation}
527: \delta W_2=\frac{\lambda N^2\bar{N}^2}{2m_P}+
528: \frac{\beta N^2h^2}{2m_P},
529: \label{W2}
530: \end{equation}
531: where $\lambda$ and $\beta$ are dimensionless
532: parameters, which can be made real and positive
533: by an appropriate redefinition of the phases of
534: the superfields and $m_{\rm P}\simeq 2.44\times
535: 10^{18}~{\rm GeV}$ is the reduced Planck mass.
536:
537: \par
538: In addition to $G_{LR}$, the model possesses three
539: global ${\rm U}(1)$ symmetries, namely an anomalous
540: PQ symmetry ${\rm U}(1)_{\rm PQ}$, a non-anomalous
541: R symmetry ${\rm U}(1)_R$, and the baryon number
542: symmetry ${\rm U}(1)_B$. The PQ and R charges
543: of the various superfields are
544: \begin{eqnarray}
545: PQ:~q^c,l^c,S,l_H^c,\bar l_H^c(0),~h,\bar N(1),~q,l,
546: N(-1); \nonumber \\
547: R:~h,l_H^c,\bar l_H^c,\bar{N}(0),~q,q^c,l,l^c,N(1/2),
548: ~S(1).
549: \label{charges}
550: \end{eqnarray}
551: Note that global continuous symmetries such as our
552: PQ and R symmetry can effectively arise
553: \cite{continuous} from the rich
554: discrete symmetry groups encountered in many
555: compactified string theories (see e.g.
556: Ref.~\cite{discrete}).
557:
558: \par
559: It is well known that the superpotential in
560: Eq.~(\ref{W1}) leads \cite{lyth,dss} naturally
561: to the standard SUSY realization of hybrid
562: inflation \cite{hybrid}. In particular, the
563: scalar potential which is derived from it
564: possesses a built-in classically flat valley of
565: minima at $l_H^c=\bar l_H^c=0$ and for $|S|$
566: greater than a critical (instability) value
567: $S_c=M$. This valley can serve as inflationary
568: path. Indeed, the constant tree-level potential
569: energy density $\kappa^{2}M^{4}$ on this path
570: can cause exponential expansion of the
571: universe. Moreover, since this constant energy
572: density breaks SUSY, there are important
573: radiative corrections \cite{dss} which provide
574: a logarithmic slope along the inflationary
575: trajectory necessary for driving the system
576: towards the vacua.
577:
578: \par
579: It should be noted that the SUSY GUT model
580: considered here does not predict the existence
581: of topological defects such as magnetic
582: monopoles or cosmic strings. In the opposite
583: case, these defects would have been copiously
584: produced \cite{smooth} at the end of hybrid
585: inflation. The overproduction of magnetic
586: monopoles, in particular, would have led to a
587: cosmological catastrophe and a modification
588: \cite{jean,smooth} of the standard realization
589: of SUSY hybrid inflation would be needed to
590: avoid this problem. This happens in higher
591: gauge groups such as the Pati-Salam group,
592: which predicts the existence of monopoles
593: carrying two units of Dirac magnetic charge
594: \cite{magg}. Cosmic strings, on the other hand,
595: which are generically present in many GUT
596: models \cite{kibble,generic}, would contribute
597: to the cosmological perturbations leading
598: \cite{mairi} to extra restrictions on the
599: parameters of the model. The reason that our
600: model does not predict cosmic strings is that
601: the $G_{LR}$ breaking is achieved by a
602: conjugate pair of ${\rm SU}(2)_R$ doublets
603: with $B-L=1,-1$ which also break the $Z_2$
604: subgroup of ${\rm U}(1)_{B-L}$. This $Z_2$,
605: which does not belong to $G_{\rm SM}$, would
606: have been left unbroken if, alternatively, we
607: had used a pair of ${\rm SU}(2)_R$ triplets
608: with $B-L=2,-2$ for this breaking. This would
609: have led to the presence of $Z_2$ cosmic
610: strings (compare with the $Z_2$ cosmic strings
611: encountered in the ${\rm SO}(10)$ GUT model of
612: Ref.~\cite{kibble}).
613:
614: \par
615: The part of the tree-level scalar potential which
616: is relevant for the PQ symmetry breaking is derived
617: from the superpotential coupling
618: $\lambda N^2\bar{N}^2/2m_{\rm P}$ in Eq.~(\ref{W2})
619: and, after soft SUSY breaking mediated by minimal
620: supergravity (SUGRA), is given by \cite{rsym}
621: \begin{equation}
622: V_{\rm PQ}=\frac{1}{2}m_{3/2}^2\phi^2\left(
623: 1-\frac{|A|\lambda\phi^2}{8m_{3/2}m_P}+
624: \frac{\lambda^2\phi^4}{16m_{3/2}^2m_P^2}\right),
625: \label{PQ-pot}
626: \end{equation}
627: where $m_{3/2}\sim 1~{\rm TeV}$ is the gravitino
628: mass and $A$ is the dimensionless coefficient of
629: the soft SUSY breaking term corresponding to the
630: superpotential term $\lambda N^2\bar{N}^2/
631: 2m_{\rm P}$. Here, the phases $\alpha$,
632: $\varphi$ and $\bar{\varphi}$ of $A$, $N$ and
633: $\bar{N}$ are taken to satisfy the relation
634: $\alpha+2\varphi+2\bar{\varphi}=\pi$ and $\vert
635: N\vert$, $\vert\bar{N}\vert$ are assumed equal,
636: which minimizes the potential. Moreover,
637: rotating $N$ on the real axis by an appropriate
638: R transformation, we defined the canonically
639: normalized real scalar PQ field $\phi=2N$. For
640: $|A|>4$, this potential has a local minimum at
641: $\phi=0$ and absolute minima at
642: \begin{equation}
643: \langle\phi\rangle^2\equiv f_a^2=\frac{2}{3\lambda}
644: \left(|A|+\sqrt{|A|^2-12}\right)m_{3/2}m_{\rm P}
645: \label{eq:fa}
646: \end{equation}
647: with $f_a~(>0)$ being the axion decay constant, i.e.
648: the symmetry breaking scale of ${\rm U}(1)_{\rm PQ}$.
649: Substituting this vacuum expectation value (VEV)
650: into the superpotential coupling $\beta N^2h^2/
651: 2m_{\rm P}$ in Eq.~(\ref{W2}), we obtain a $\mu$
652: term with
653: \begin{equation}
654: \mu=\frac{\beta f_a^2}{4m_{\rm P}}\sim m_{3/2},
655: \label{mu}
656: \end{equation}
657: as desired \cite{kn}. Note that the potential
658: $V_{\rm PQ}$ in Eq.~(\ref{PQ-pot}) should be shifted
659: \cite{dllr1} by adding to it the constant
660: \begin{eqnarray}
661: V_0&=&\frac{1}{108\lambda}\left(|A|+\sqrt{|A|^2-12}
662: \right)\nonumber \\
663: & &\times\left[|A|\left(|A|+
664: \sqrt{|A|^2-12}\right)-24\right]m_{3/2}^3 m_{\rm P},
665: \label{eq:v0}
666: \end{eqnarray}
667: so that it vanishes at its absolute minima.
668:
669: \section{The PQ field in the early universe}
670: \label{sec:early}
671:
672: \par
673: In the early universe, the PQ potential can acquire
674: sizable corrections from the SUSY breaking caused
675: by the presence of a finite energy density
676: \cite{lyth,crisis,drt95}. In particular, during
677: inflation and the subsequent inflaton oscillations,
678: SUSY breaking is transmitted to the PQ system via
679: its coupling to the inflaton given by the SUGRA
680: scalar potential. The resulting corrections, whose
681: scale is set by the Hubble parameter $H$, dominate
682: over the contributions from hidden sector SUSY
683: breaking as long as $H\gg m_{3/2}$. For simplicity,
684: we will ignore the $A$ term type corrections
685: \cite{drt95}. To leading order, we then just obtain
686: a correction $\delta m_{\phi}^2$ to the
687: ${\rm mass}^2$ of the curvaton. For a general
688: K\"{a}hler potential, $\delta m_{\phi}^2\sim H^2$
689: with either sign possible. However, for specific
690: (no-scale like) K\"{a}hler potentials, it might be
691: (partially) cancelled \cite{noscale}. Assuming that
692: $\delta m_{\phi}^2>0$, we write
693: \begin{equation}
694: \delta m_{\phi}^2=\gamma^2 H^2,
695: \label{eq:effec-mass}
696: \end{equation}
697: where $\gamma$ can have different values during
698: inflation and inflaton oscillations. Actually,
699: we must assume that, during inflation, $\gamma
700: \ll 1$ so that the PQ field qualifies as an
701: effectively massless field which acquires
702: perturbations from inflation and thus can act
703: as curvaton. Fortunately, the cancellation of
704: $\delta m_{\phi}^2$ during inflation can, in
705: principle, be ``naturally'' arranged to be
706: exact (see fourth paper in
707: Ref.~\cite{noscale}). So, for simplicity, we
708: could take $\gamma=0$ during inflation. On the
709: other hand, large values of $\gamma$ after the
710: end of inflation would generically lead
711: \cite{dllr2} to a drastic reduction of the
712: density fraction of the PQ field, which thus
713: again would become unable to play the role of
714: curvaton. In view of the fact that, in contrast
715: to the case of inflation, it is not so easy to
716: achieve exact cancellation of
717: $\delta m_{\phi}^2$ during inflaton
718: oscillations, we will only assume that, after
719: the termination of inflation, $\gamma$ is
720: somewhat suppressed, say $\sim 0.1$ (or
721: smaller).
722:
723: \par
724: After reheating, the universe is radiation
725: dominated and, thus, $H\simeq 1/2t\leq 1/
726: 2t_{\rm reh}=\Gamma_{\rm infl}/2$, where $t$ is
727: the cosmic time and
728: $t_{\rm reh}=\Gamma_{\rm infl}^{-1}$ the time at
729: reheating with $\Gamma_{\rm infl}$ being the
730: inflaton decay width. It is easily seen that, in
731: this case, $H\ll m_{3/2}$ as a consequence of the
732: gravitino constraint ($T_{\rm reh}\lesssim 10^9~
733: {\rm GeV}$) \cite{ekn} on the reheat temperature
734: $T_{\rm reh}$, which is given by \cite{lectures}
735: \begin{equation}
736: T_{\rm reh}=\left(\frac{45}{2\pi^2 g_*}
737: \right)^{\frac{1}{4}}
738: (\Gamma_{\rm infl}m_{\rm P})^{\frac{1}{2}},
739: \label{eq:reheat}
740: \end{equation}
741: where $g_*$ is the effective number of degrees
742: of freedom ($g_*=228.75$ for the MSSM spectrum).
743: Thus, the SUSY breaking effects from the finite
744: energy density in the universe are subdominant
745: compared to the hidden sector SUSY breaking
746: effects, whose scale is set by $m_{3/2}$.
747:
748: \par
749: The PQ potential can acquire temperature
750: corrections too. During the era of inflaton
751: oscillations, they originate from the new
752: radiation \cite{reheat} which emerges from the
753: decaying inflaton. It has been shown \cite{dllr1},
754: however, that these corrections are overshadowed
755: by the SUGRA ones which, in the latest stages of
756: this era, are, in turn, overshadowed by the terms
757: from hidden sector SUSY breaking. After reheating,
758: the temperature corrections are less important
759: than the ones from the hidden sector as argued in
760: Refs.~\cite{jean} and \cite{talks}. So, the
761: temperature corrections to the PQ potential can
762: be ignored throughout.
763:
764: \par
765: We see that, in the early universe, the effective
766: scalar potential for the PQ field can be taken
767: to be
768: \begin{equation}
769: V_{\rm PQ}^{\rm eff}=V_{\rm PQ}+\frac{1}{2}
770: \gamma^2 H^2\phi^2+V_0.
771: \label{full-pot}
772: \end{equation}
773: The full effective scalar potential $V$ which is
774: relevant for our analysis here is obtained by
775: adding to $V_{\rm PQ}^{\rm eff}$ the potential
776: for standard SUSY hybrid inflation (see e.g.
777: Ref.~\cite{lectures}).
778:
779: \par
780: The evolution of the field $\phi$ is generally
781: governed by the classical equation of motion
782: \begin{equation}
783: \ddot{\phi}+3H\dot{\phi}+V^{\prime}=0,
784: \label{field-eqn}
785: \end{equation}
786: where overdots and primes denote derivation
787: with respect to the cosmic time $t$ and the
788: PQ field $\phi$ respectively. In particular,
789: Eq.~(\ref{field-eqn}) describes
790: \cite{communication} the evolution during
791: inflation of the mean value of $\phi$ in a
792: comoving region larger than the inflationary
793: horizon. Actually, this equation starts to be
794: valid a short time after this region crosses
795: outside the de Sitter horizon. The mean value
796: of $\phi$, however, in a region of fixed size
797: somewhat bigger than the (almost) constant
798: size of the de Sitter horizon satisfies
799: \cite{notari} this equation during inflation
800: provided that it exceeds a certain value
801: $\phi_Q$ given by
802: \begin{equation}
803: V^{\prime}\sim\frac{3H_{\rm infl}^3}{2\pi},
804: \label{phiQ}
805: \end{equation}
806: where $H_{\rm infl}$ is the almost constant
807: Hubble parameter during inflation.
808: Therefore, if we require that the value
809: $\phi_f$ (considered positive without loss of
810: generality) which is taken at the end of
811: inflation by the mean $\phi$ in a region of
812: fixed size somewhat bigger than
813: $H_{\rm infl}^{-1}$ exceeds $\phi_Q$, we can be
814: sure \cite{dllr1} that the classical evolution
815: equation holds for this mean field until the
816: end of inflation. For values of the mean $\phi$
817: in a region of fixed size somewhat bigger than
818: $H_{\rm infl}^{-1}$ which are smaller than
819: about $\phi_Q$, the random walk executed
820: \cite{rwalk} by this mean field due to the
821: quantum perturbations from inflation cannot be
822: neglected and may overshadow \cite{notari} its
823: classical motion. Thus, in this case, the
824: classical equation of motion during inflation
825: for this mean $\phi$ ceases to be valid. For
826: reasons to be discussed later, we will not
827: consider in our analysis values of $\phi_f$
828: which lie in the quantum regime, i.e. which
829: are smaller than $\phi_Q$ (see
830: Secs.~\ref{sec:curvature} and
831: \ref{sec:evolution}). It should be pointed
832: out in passing that the requirement of
833: complete randomization of the mean $\phi$ in
834: a region of fixed size somewhat bigger than
835: $H_{\rm infl}^{-1}$ as a consequence of its
836: quantum perturbations from inflation implies
837: \cite{prep} an even more stringent bound on
838: this mean $\phi$ given by
839: $V\lesssim H^4_{\rm infl}$.
840:
841: \par
842: Moreover, as explained in the next section, we
843: will have to study only values of $\phi_f$ for
844: which $\phi$ is a slowly rolling field
845: during the relevant part of inflation (i.e.
846: during at least the last $50-60$ e-foldings).
847: It has been shown \cite {dllr1} that, in this
848: case, the PQ field $\phi$ emerges at the end of
849: inflation with negligible velocity (i.e.
850: derivative with respect to cosmic time). Its
851: subsequent evolution during the matter
852: dominated era of damped inflaton oscillations
853: is given by Eqs.~(\ref{full-pot}) and
854: (\ref{field-eqn}) with $H=2/3t$. One finds
855: \cite{dllr1} that, depending on the value of
856: $\phi_f$, the PQ system eventually enters into
857: a phase of damped oscillations about either the
858: trivial (local) minimum of $V_{\rm PQ}$ at
859: $\phi=0$ or one of its PQ (absolute) minima at
860: $\phi=\pm f_a$. Of course, values of $\phi_f$
861: leading to the trivial minimum must be excluded.
862:
863: \par
864: The damped oscillations of the PQ field continue
865: even after reheating, where $H$ becomes equal to
866: $1/2t$. Finally, this field decays via the second
867: coupling in the superpotential of Eq.~(\ref{W2})
868: into a pair of Higgsinos provided that their mass
869: $\mu$ does not exceed half of the mass of the PQ
870: field (see Ref.~\cite{dllr1})
871: \begin{equation}
872: m_{\phi}=\frac{m_{3/2}}{\sqrt{3}}\left(|A|^2-
873: 12\right)^{\frac{1}{4}}\left(|A|+\left(|A|^2-12
874: \right)^{\frac{1}{2}}\right)^{\frac{1}{2}},
875: \label{mphi}
876: \end{equation}
877: which is independent of the parameter $\lambda$.
878: The decay time of the PQ field is
879: $t_{\phi}=\Gamma_{\phi}^{-1}$, where
880: $\Gamma_{\phi}$ is its decay width, which has
881: been found \cite{dllr1} to be given by
882: \begin{equation}
883: \Gamma_{\phi}=\frac{\beta^2f_a^2}
884: {8\pi m_{\rm P}^2}m_{\phi}.
885: \label{Gphi}
886: \end{equation}
887: Note that the coherently oscillating PQ field
888: could evaporate \cite{evap} as a result of
889: scattering with particles in the thermal bath
890: before it decays into Higgsinos. However, one
891: can show \cite{dllr1} that, in the model under
892: discussion here, this does not happen.
893:
894:
895: \section{The curvature perturbation}
896: \label{sec:curvature}
897:
898: \par
899: We will consider here only values of $\phi_f$ for
900: which the PQ field $\phi$ is effectively massless,
901: i.e. $V^{\prime\prime}\leq H^2$, during (at
902: least) the last $50-60$ inflationary e-foldings
903: so that it receives a super-horizon spectrum of
904: perturbations from inflation
905: and can act as curvaton. This requirement also
906: guarantees that $\phi$ is slowly rolling during
907: the relevant part of inflation. The perturbation
908: $\delta\phi$ then evolves at subsequent times and,
909: when $\phi$ settles into damped quadratic
910: oscillations about the PQ vacua, yields a stable
911: perturbation in the energy density of this field.
912: After the PQ field decays, this perturbation is
913: transferred to radiation, thereby contributing to
914: the total curvature perturbation. On the other
915: hand, the radiation, which originates from the
916: inflaton decay, could carry a curvature
917: perturbation prior to the curvaton decay too. It
918: actually inherits the curvature perturbation of
919: the inflaton. Contrary to the standard curvaton
920: hypothesis \cite{curv1}, we make here the more
921: natural assumption that this perturbation is
922: non-zero and, thus, also contributes to the total
923: curvature perturbation.
924:
925: \par
926: The scalar part of the metric perturbation for
927: a flat universe can be written (using the
928: notation of Ref.~\cite{thesis}), in all
929: generality, as follows (for reviews of the
930: gauge invariant theory of cosmological
931: perturbations, see e.g. Ref.~\cite{gipt}):
932: \begin{eqnarray}
933: \delta g_{\mu \nu}d x^\mu d x^\nu &=&
934: -2Adt^2+2aB_{,i}dtdx^i\nonumber \\
935: & &+a^2\left(2C\delta_{ij}
936: +E_{,ij}\right)dx^idx^j,
937: \label{metric}
938: \end{eqnarray}
939: where $\mu,\nu=0,1,2,3$ and $i,j=1,2,3$. Here,
940: $x^0=t$ is the physical time, $x^i$ ($i=1,2,3$)
941: are the comoving spatial coordinates,
942: $Y_{,i}\equiv\partial Y/\partial x^i$
943: ($i=1,2,3$) and $\delta_{ij}$ denotes the
944: Kronecker delta. The dimensionless parameter
945: $a$ is the scale factor of the universe, which
946: is normalized to unity at the present cosmic
947: time. We define the following gauge invariant
948: quantities:
949: \begin{eqnarray}
950: \zeta&\equiv &C-H\frac{\delta\rho}{\dot{\rho}},
951: \label{zetagen}
952: \\
953: \mathcal{R}_{\rm rad}&\equiv &-C-H(B-a^2v),
954: \\
955: \Phi &\equiv &-C-H\left(B-a^2\dot{E}\right).
956: \end{eqnarray}
957: Here, $\rho=-T_0^{~0}$ is the total energy
958: density in the universe with $T_\mu^{~\nu}$
959: being the energy momentum tensor, $\delta\rho
960: =-\delta T_0^{~0}$ is the total density
961: perturbation, and $v$ describes the spatial
962: perturbation in the 4-velocity $u^\mu$ of an
963: observer comoving with the total fluid, i.e.
964: $v^{,i}\equiv -\delta u^i/u^0$. The variable
965: $\zeta$ represents the curvature perturbation
966: on hypersurfaces of uniform density,
967: $\mathcal{R}_{\rm rad}$ is the curvature
968: perturbation in the (total matter) comoving
969: gauge (up to the sign), while $\Phi$ is the
970: Bardeen potential, which is the curvature
971: perturbation (up to the sign) in the
972: longitudinal gauge. These three quantities are
973: related by
974: \begin{equation}
975: \zeta=-\mathcal{R}_{\rm rad}-\frac{2\rho}
976: {9(\rho+p)}\left(\frac{k}{Ha}\right)^2\Phi,
977: \end{equation}
978: where we have used the time-time component of
979: the Einstein equation and $p$ is the total
980: pressure of the universe (i.e. $T_{ij}= p
981: \delta_{ij}$). On super-horizon scales, $k\ll
982: Ha$, the second term on the right hand side of
983: this equation is negligible, and we thus have
984: $\zeta=-\mathcal{R}_{\rm rad}$.
985:
986: \par
987: Using Eq.~(\ref{zetagen}), the total curvature
988: perturbation \cite{bst} in the flat slicing
989: gauge (defined by setting $C=E=0$ in
990: Eq.~(\ref{metric})) is given by
991: \begin{equation}
992: \zeta=\frac{\delta\rho}{3(\rho+p)}.
993: \end{equation}
994: After the curvaton decay, it becomes
995: \cite{curv3}
996: \begin{equation}
997: \zeta=(1-f)\zeta_i+f\zeta_c,
998: \label{zeta}
999: \end{equation}
1000: where $\zeta_i=\delta\rho_r/4\rho_r$ and
1001: $\zeta_c=\delta\rho_\phi/3\rho_\phi$ are the
1002: partial curvature perturbations on spatial
1003: hypersurfaces of constant curvature from the
1004: inflaton and the curvaton respectively at the
1005: curvaton decay with $\rho_r$ and $\rho_\phi$
1006: being the radiation and $\phi$ energy densities
1007: respectively, and $\delta\rho_r$ and $\delta
1008: \rho_\phi$ the corresponding perturbations.
1009: Also,
1010: \begin{equation}
1011: f=\frac{3\rho_{\phi}}{3\rho_{\phi}+4\rho_r}
1012: \label{eq:f}
1013: \end{equation}
1014: evaluated at the time of the curvaton decay. Here,
1015: we assume that
1016: the amplitude of the oscillating $\phi$ has been
1017: sufficiently reduced so that the potential can be
1018: approximately considered as quadratic. Actually,
1019: as shown in Ref.~\cite{dllr2}, this must
1020: necessarily happen before the curvaton decays.
1021: The oscillating curvaton field then behaves like
1022: pressureless matter and $\zeta_c=2\delta\phi_0/
1023: 3\phi_0$, where $\phi_0$ is the amplitude of the
1024: oscillations and $\delta\phi_0$ the perturbation
1025: in this amplitude originating from the
1026: perturbation $\delta\phi_f$ in the value $\phi_f$
1027: of $\phi$ at the end of inflation.
1028:
1029: \par
1030: The comoving curvature perturbation
1031: $\mathcal{R}_{\rm rad}$, for super-horizon
1032: scales, is given by
1033: \begin{equation}
1034: \mathcal{R}_{\rm rad}=(1-f)A_i\left(\frac{k}{H_0}
1035: \right)^{\nu_i}\hat{a}_i+
1036: fA_c\left(\frac{k}{H_0}\right)^{\nu_c}\hat{a}_c,
1037: \label{R}
1038: \end{equation}
1039: where $k$ is the comoving (present physical)
1040: wave number, $H_0$
1041: is the present value of the Hubble parameter
1042: and $\hat{a}_i$, $\hat{a}_c$ are independent
1043: normalized Gaussian random variables. Also,
1044: $A_i$ and $A_c$ are, respectively, the
1045: amplitudes of $-\zeta_i$ and $-\zeta_c$ at the
1046: present horizon scale (i.e. at $k=H_0$), and
1047: $\nu_i$ and $\nu_c$ are the spectral tilts of
1048: the inflaton and curvaton respectively, which
1049: are related to the corresponding spectral
1050: indices $n_i$ and $n_c$ by $n_i=1+2\nu_i$ and
1051: $n_c=1+2\nu_c$. We do not consider running of
1052: the spectral indices, since this is negligible
1053: in our model.
1054:
1055: \par
1056: The amplitude $A_i$ of the partial curvature
1057: perturbation from the inflaton is found \cite{hier}
1058: to be
1059: \begin{equation}
1060: A_i=\left(\frac{2N_{Q}}{3}\right)^{\frac{1}{2}}
1061: \left(\frac{M}{m_{\rm P}}\right)^2x_Q^{-1}y_Q^{-1}
1062: \Lambda(x_Q^2)^{-1}
1063: \label{ai}
1064: \end{equation}
1065: with
1066: \begin{equation}
1067: \Lambda(z)=(z+1)\ln(1+z^{-1})+(z-1)\ln(1-z^{-1})~,
1068: \label{lambda}
1069: \end{equation}
1070: \begin{equation}
1071: y_Q^2=\int_{x_f^2}^{x_Q^2}\frac{dz}{z}
1072: \Lambda(z)^{-1},~y_Q\geq 0~.
1073: \label{yq}
1074: \end{equation}
1075: Here, $N_{Q}$ is the number of e-foldings suffered
1076: by our present horizon scale during inflation,
1077: $x_Q=S_Q/M$ with $S_Q$ being the value
1078: of $\vert S\vert$ when our present horizon scale
1079: crosses outside the inflationary horizon, and
1080: $x_f=S_f/M$ with $S_f$ being the value
1081: of $\vert S\vert$ at the end of inflation.
1082:
1083: \par
1084: In our model, the slow-roll parameters for the
1085: inflaton as functions of $\vert S\vert$ are given
1086: by \cite{lectures}
1087: \begin{equation}
1088: \epsilon_i=\left(\frac{\kappa^2m_{\rm P}}{8\pi^2M}
1089: \right)^2z\Lambda(z)^2,
1090: \label{epsiloni}
1091: \end{equation}
1092: \begin{eqnarray}
1093: \eta_i&=&2\left(\frac{\kappa m_{\rm P}}{4\pi M}
1094: \right)^2\Big[(3z+1)\ln(1+z^{-1})
1095: \nonumber\\
1096: & &+(3z-1)\ln(1-z^{-1})\Big],
1097: \label{etai}
1098: \end{eqnarray}
1099: where $z=x^2$ with $x=\vert S\vert/M$. In the
1100: presence of the curvaton, however, one can show
1101: that the slow-roll conditions (for $\gamma
1102: \rightarrow 0$) take the form $\epsilon,
1103: \vert\eta_i\vert\leq 1$, where
1104: \begin{equation}
1105: \epsilon\equiv -\frac{\dot H}{H^2}=\epsilon_i+
1106: \epsilon_c
1107: \label{epsilon}
1108: \end{equation}
1109: with
1110: \begin{equation}
1111: \epsilon_c=\frac{1}{2}
1112: m_{\rm P}^2\left(\frac{V^{\prime}}{V}\right)^2,
1113: \label{epsilonc}
1114: \end{equation}
1115: and $x_f$ is given by the value of $x$ for
1116: which these conditions are saturated. Actually,
1117: as it turns out, $x_f$ corresponds to
1118: $\eta_i=-1$. For $\kappa\ll 1$, the slow-roll
1119: conditions are violated only extremely close to
1120: the critical point at $x=1$
1121: ($\vert S\vert=S_c$). So, inflation continues
1122: practically until this point is reaches and,
1123: following Ref.~\cite{hier}, we can put $x_f=1$
1124: in Eq.~(\ref{yq}). However, for larger values
1125: of the parameter $\kappa$, inflation can
1126: terminate well before reaching the instability
1127: point.
1128:
1129: \par
1130: Finally, $\kappa$, $N_Q$ are given by
1131: \cite{lectures}
1132: \begin{equation}
1133: \kappa=\frac{2\pi}{\sqrt{N_Q}}
1134: ~y_Q~\frac{M}{m_{\rm P}},
1135: \label{kappa}
1136: \end{equation}
1137: \begin{equation}
1138: N_Q\simeq 55.9+\frac{2}{3}\ln\frac{
1139: \kappa^{\frac{1}{2}}M}{10^{15}~{\rm GeV}}
1140: +\frac{1}{3}\ln\frac{T_{\rm reh}}
1141: {10^9~{\rm GeV}}
1142: \label{NQ}
1143: \end{equation}
1144: (for $H_0\simeq 72~\rm{km}~\rm{sec}^{-1}~
1145: \rm{Mpc}^{-1}$). The spectral index for the
1146: inflaton is $n_i=1-6\epsilon+2\eta_i$ (for
1147: $\gamma\rightarrow 0$), where $\epsilon$ and
1148: $\eta_i$ are evaluated at the time when
1149: our present horizon scale crosses outside
1150: the inflationary horizon. Note that $\epsilon$
1151: enters in this formula, not $\epsilon_i$ as in
1152: the case of pure inflaton. However,
1153: $\epsilon_c$ is normally much smaller than
1154: $\epsilon_i$ which, in turn, is negligible
1155: compared to $\vert\eta_i\vert$. As a
1156: consequence, $n_i\simeq 1+2\eta_i$.
1157:
1158: \par
1159: We now calculate the amplitude $A_c$ of the
1160: partial curvature perturbation from the
1161: curvaton $\phi$. This originates from the
1162: perturbation $\delta\phi_*=(H_*/2\pi)\hat{a}_c$
1163: acquired by $\phi$ from inflation when our
1164: present horizon scale crosses outside the de
1165: Sitter horizon
1166: ($H_*$ is the inflationary Hubble parameter at
1167: that moment). In order to find the evolution of
1168: this perturbation during the subsequent part of
1169: inflation and estimate its value $\delta\phi_f$
1170: at the end of inflation, we must consider the
1171: equation of motion for $\phi$ during inflation
1172: (see Eq.~(\ref{field-eqn})). In the slow-roll
1173: approximation, which is assumed to hold for the
1174: curvaton too, this equation reads
1175: \begin{equation}
1176: 3H\dot\phi+V^{\prime}=0.
1177: \label{slow}
1178: \end{equation}
1179: Taking a small perturbation
1180: $\delta\phi$ of $\phi$, Eq.~(\ref{slow}) gives
1181: \begin{equation}
1182: 3H\delta\dot\phi+3H^{\prime}\delta\phi\dot\phi+
1183: V^{\prime\prime}\delta\phi=0.
1184: \label{dphi1}
1185: \end{equation}
1186: Substituting $\dot\phi$ from Eq.~(\ref{slow}) and
1187: using the Friedmann equation $3H^2m_{\rm P}^2=V$,
1188: Eq.~(\ref{dphi1}) becomes
1189: \begin{equation}
1190: \delta\dot\phi+H(-\epsilon_c+\eta_c)\delta\phi=0,
1191: \label{dphi2}
1192: \end{equation}
1193: where
1194: \begin{equation}
1195: \eta_c=m_{\rm P}^2\frac{V^{\prime\prime}}{V}.
1196: \label{etac}
1197: \end{equation}
1198: Integration of Eq.~(\ref{dphi2}) from the cosmic
1199: time $t_*$ when our present horizon scale crossed
1200: outside the inflationary horizon until the end of
1201: inflation (at time $t_f$) yields
1202: \begin{equation}
1203: \delta\phi_f=\frac{H_*}{2\pi}\,\hat{a}_c
1204: \exp{\int_0^{N_Q}(\epsilon_c-\eta_c)dN},
1205: \label{dphif}
1206: \end{equation}
1207: where we used the relation $dN=-Hdt$ for the
1208: number of e-foldings $N(k)=N_Q+\ln(H_0/k)$
1209: suffered by the scale which corresponds to the
1210: comoving wave number $k$ during hybrid inflation.
1211:
1212: \par
1213: For each value $\phi_f~(>0)$ of the curvaton
1214: field at the end of inflation, we construct the
1215: perturbed field $\phi_f+\delta\phi_f$. We then
1216: follow the evolution of $\phi_f$ and $\phi_f+
1217: \delta\phi_f$ until the time $t_\phi$ of the
1218: curvaton decay and evaluate the value of
1219: $\delta\rho_\phi/\rho_\phi$ at this time. The
1220: amplitude $A_c$ of the partial curvature
1221: perturbation from the curvaton is given by
1222: \begin{equation}
1223: A_c\,\hat{a}_c=\frac{1}{3}\frac{\delta\rho_\phi}
1224: {\rho_\phi}.
1225: \label{Ac}
1226: \end{equation}
1227: We have found numerically that the perturbation
1228: $\delta\phi_0$ in the amplitude of the oscillating
1229: curvaton at $t_\phi$ is proportional to
1230: $\delta\phi_f$. So $\zeta_c$ has the same spectral
1231: tilt as $\delta\phi_f$, which can be found from
1232: Eq.~(\ref{dphif}):
1233: \begin{equation}
1234: \nu_c\equiv\frac{d\ln A_c}{d\ln k}=-\epsilon-
1235: \epsilon_c+\eta_c,
1236: \label{nuc}
1237: \end{equation}
1238: where we used the relation $d\ln k=Hdt$, and
1239: $\epsilon$, $\epsilon_c$ and $\eta_c$ are
1240: evaluated at $t_*$. (Note that a similar
1241: formula has been derived in the first paper
1242: in Ref.~\cite{curv2}, but without the
1243: $-\epsilon_c$ term in the right hand side.)
1244: The spectral index for the
1245: curvaton is then $n_c=1-2\epsilon-2\epsilon_c+
1246: 2\eta_c$ which, in most cases, reduces to $n_c
1247: \simeq 1+2\eta_c$ since we typically have
1248: $\epsilon_i,\epsilon_c\ll\vert\eta_c\vert$.
1249:
1250: \par
1251: It should be pointed out that, in deriving
1252: Eq.~(\ref{nuc}), we assume that, during
1253: inflation, the mean value of $\phi$ in any
1254: region of fixed size (somewhat bigger than)
1255: $H_{\rm infl}^{-1}$ is initially the same and
1256: follows the classical equation of motion
1257: (Eq.~(\ref{field-eqn})). This mean field sets
1258: the (initial) value of the mean $\phi$ in any
1259: comoving region at the time this region exits
1260: the de Sitter horizon. Thus, if it satisfies
1261: the above requirements, we can be sure that,
1262: at any given time during inflation, the
1263: resulting mean $\phi$ in each comoving region
1264: that already exited the horizon is practically
1265: independent from the size of this region. So
1266: the mean $\phi$ in a fixed region of size
1267: $H_{\rm infl}^{-1}$ or in any comoving volume
1268: (after horizon exit) is described by the same
1269: (single-valued) function of the cosmic time
1270: and evolves classically (in fact, it rolls
1271: down slowly in our case). The derivation with
1272: respect to time of the logarithm of the
1273: amplitude of $\delta\phi_f$, which is given by
1274: Eq.~(\ref{dphif}), is then straightforward
1275: leading to Eq.~(\ref{nuc}). On the contrary,
1276: if the mean field in a fixed region of size
1277: somewhat bigger than $H_{\rm infl}^{-1}$
1278: executes a random walk with step of amplitude
1279: $H_{\rm infl}/2\pi$ per Hubble time, the
1280: calculation of the spectral tilt becomes less
1281: clear. So, we decided to exclude the quantum
1282: regime (i.e. the region $\phi_f<\phi_Q$) from
1283: our analysis (see Sec.~\ref{sec:evolution}). It
1284: is true, however, that it is by no means
1285: necessary to avoid the random walk behavior at
1286: all times during inflation. Indeed, only
1287: cosmological scales corresponding to a few
1288: e-foldings after the exit of our present
1289: horizon scale from the de Sitter horizon are
1290: relevant. For simplicity though, we exclude all
1291: the quantum regime so that no random motion of
1292: the mean $\phi$ in a region of size
1293: $H_{\rm infl}^{-1}$ is encountered during
1294: inflation. This, as we will see, has no
1295: influence on our results.
1296:
1297: \section{The isocurvature perturbation}
1298: \label{sec:isocurvature}
1299:
1300: \par
1301: After the termination of inflation, the
1302: inflaton performs damped oscillations about
1303: the SUSY vacuum and eventually decays into
1304: light particles reheating the universe. At
1305: reheating, gravitinos are
1306: thermally produced besides other particles.
1307: They only decay, though, well after the big
1308: bang nucleosynthesis (BBN) since they have
1309: very weak couplings. Each decaying gravitino
1310: yields one sparticle subsequently turning
1311: into the LSP, which is stable. These LSPs
1312: survive until the present time contributing
1313: to the relic abundance of CDM in the universe.
1314: For simplicity, we assume that the thermally
1315: produced LSPs can be neglected, which holds
1316: in many cases. So, all the relic LSPs come
1317: solely from the decaying gravitinos. Baryons
1318: can be produced via a primordial leptogenesis
1319: \cite{lepto} which can occur \cite{leptoinf}
1320: at reheating.
1321:
1322: \par
1323: We see that both the LSPs and the baryons
1324: originate from reheating. Their partial
1325: curvature perturbations, $\zeta_{\rm LSP}$
1326: and $\zeta_B$ respectively, should thus
1327: coincide with the partial curvature
1328: perturbation of the radiation which emerges
1329: from the inflaton decay, i.e.
1330: \begin{equation}
1331: \zeta_{\rm LSP}=\zeta_B=\zeta_i.
1332: \label{zetaLSP}
1333: \end{equation}
1334: The isocurvature perturbation of the LSPs and
1335: the baryons is then given by
1336: \begin{equation}
1337: \mathcal{S}_{{\rm LSP}+B}\equiv 3(
1338: \zeta_{{\rm LSP}+B}-\zeta)=3f(\zeta_i-\zeta_c),
1339: \label{SLSP1}
1340: \end{equation}
1341: where $\zeta_{{\rm LSP}+B}=\zeta_{\rm LSP}=
1342: \zeta_B$ is the partial curvature perturbation
1343: of the LSPs and the baryons and we used
1344: Eq.~(\ref{zeta}). Here, we assume
1345: that the curvature perturbation in radiation
1346: ($\zeta_\gamma$) practically coincides with the
1347: total curvature perturbation. This corresponds to
1348: a negligible neutrino isocurvature perturbation,
1349: which is \cite{curv3} the case provided that, as
1350: in our model, leptogenesis takes place well
1351: before the curvaton decays or dominates the
1352: energy density. Applying the definitions which
1353: follow Eq.~(\ref{R}), Eq.~(\ref{SLSP1}) takes the
1354: form
1355: \begin{equation}
1356: \mathcal{S}_{{\rm LSP}+B}=-3fA_i\left(
1357: \frac{k}{H_0}\right)^{\nu_i}\hat{a}_i+3fA_c\left(
1358: \frac{k}{H_0}\right)^{\nu_c}\hat{a}_c.
1359: \label{SLSP2}
1360: \end{equation}
1361:
1362: \par
1363: Our model contains axions which can also
1364: contribute to the CDM of the universe. They are
1365: produced at the QCD phase transition, which
1366: occurs well after the curvaton decay. They carry
1367: an isocurvature perturbation, which is completely
1368: uncorrelated with the curvature perturbation and
1369: can be written as
1370: \begin{equation}
1371: \mathcal{S}_a=A_a\left(\frac{k}{H_0}
1372: \right)^{\nu_a}\hat{a}_a,
1373: \label{Sa}
1374: \end{equation}
1375: where $A_a$ is its amplitude at the present
1376: horizon scale, $\nu_a$ is the corresponding
1377: spectral tilt (yielding the spectral index
1378: $n_a=1+2\nu_a$) and $\hat{a}_a$ is a normalized
1379: Gaussian random variable which is independent
1380: from $\hat{a}_i$ and $\hat{a}_c$.
1381:
1382: \par
1383: The amplitude $A_a$ is given by \cite{dllr1}
1384: \begin{equation}
1385: A_a=\frac{H_*}{\pi\vert\theta\vert\phi_*},
1386: \label{Aa}
1387: \end{equation}
1388: where $\theta$ is the so-called initial
1389: misalignment angle, i.e. the phase of the complex
1390: PQ field during (and at the end of) inflation and
1391: $\phi_*$ is the value of $\phi$ at $t_*$. In our
1392: case, the angle $\theta$ lies \cite{dllr1} in the
1393: interval $[-\pi/6,\pi/6]$ with all values in it
1394: being equally probable. It is determined by
1395: considering the total CDM abundance
1396: $\Omega_{\rm CDM}h^2$ which, in our model, is
1397: the sum of the relic abundance \cite{km}
1398: \begin{equation}
1399: \Omega_{\rm LSP}h^2\simeq 0.0074\left(\frac
1400: {m_{\rm LSP}}{200~{\rm GeV}}\right)\left(
1401: \frac{T_{\rm reh}}{10^{9}~{\rm GeV}}\right)
1402: \label{OmegaLSP}
1403: \end{equation}
1404: of the LSPs coming from the gravitino decays
1405: and the relic axion abundance \cite{axion}
1406: \begin{equation}
1407: \Omega_ah^2\simeq\theta^2\left(\frac{f_a}
1408: {10^{12}~{\rm GeV}}\right)^{1.175}.
1409: \label{Omegaa}
1410: \end{equation}
1411: Here $\Omega_j=\rho_j/\rho_c$ with $\rho_j$
1412: being the present energy density of the $j$th
1413: species and $\rho_c$ the present critical
1414: energy density of the universe. Furthermore,
1415: $m_{\rm LSP}$ is the LSP mass and
1416: $H_0=100h~\rm{km}~\rm{sec}^{-1}~\rm{Mpc}^{-1}$
1417: the present Hubble parameter. In the numerical
1418: calculation of $M$ and the amplitudes $A_c$
1419: and $A_a$ for given $A_i$ and $\phi_f$ (see
1420: Secs.~\ref{sec:evolution} and \ref{sec:Cell}),
1421: we will fix $h\simeq 0.72$, which is its
1422: best-fit value from the Hubble space telescope
1423: (HST) \cite{hst}. (In the full Monte Carlo (MC)
1424: CMBR analysis, however, we do allow $h$ to vary
1425: as detailed in Sec.~\ref{sec:setup}.) This
1426: approximation has very little influence on the
1427: accuracy of the results (see the remarks at the
1428: end of Sec.~\ref{sec:Cell}). Note that, in
1429: deriving the estimate for $\Omega_a h^2$ in
1430: Eq.~(\ref{Omegaa}), we applied the simplifying
1431: assumptions of Ref.~\cite{dllr1}.
1432:
1433: \par
1434: The spectral tilt $\nu_a$ is evaluated by
1435: observing that the potential of the axion field
1436: remains flat until the QCD transition is reached.
1437: So, there is no evolution of $\theta$ and its
1438: perturbation $\delta\theta=(H_*/2\pi\phi_*)
1439: \hat{a}_a$ after crossing outside the
1440: inflationary horizon and
1441: until the onset of axion oscillations. The
1442: axion isocurvature perturbation, which is
1443: \cite{dllr1} equal to $2\delta\theta/\theta$,
1444: depends on the scale only through $H_*/\phi_*$.
1445: We find
1446: \begin{equation}
1447: \nu_a=-\epsilon+\frac{m_{\rm P}}{\phi_*}
1448: \frac{V^\prime}{V}m_{\rm P}=
1449: -\epsilon+\frac{m_{\rm P}}{\phi_*}
1450: \sqrt{2\epsilon_c},
1451: \label{nua}
1452: \end{equation}
1453: where Eqs.~(\ref{epsilonc}) and (\ref{slow})
1454: were used, and the slow-roll parameters
1455: $\epsilon$ and $\epsilon_c$ are evaluated at
1456: $t_*$. In view of the fact that $\epsilon$ is
1457: normally negligible, Eq.~(\ref{nua}) reduces
1458: to $\nu_a\simeq m_{\rm P}\sqrt{2\epsilon_c}/
1459: \phi_*$ and $n_a\simeq 1+2m_{\rm P}
1460: \sqrt{2\epsilon_c}/\phi_*$.
1461:
1462: \par
1463: Combining Eqs.~(\ref{SLSP2}) and (\ref{Sa}), we
1464: find that the total isocurvature perturbation is
1465: \cite{dllr1}
1466: \begin{equation}
1467: \mathcal{S}_{\rm rad}=\frac{\Omega_{{\rm LSP}+B}}
1468: {\Omega_m}\mathcal{S}_{{\rm LSP}+B}+\frac{\Omega_a}
1469: {\Omega_m}\mathcal{S}_a,
1470: \label{total}
1471: \end{equation}
1472: where we used the definitions $\Omega_{{\rm LSP}
1473: +B}\equiv\Omega_{\rm LSP}+\Omega_B$ and $\Omega_m
1474: \equiv\rho_m/\rho_c=\Omega_{{\rm LSP}+B}+\Omega_a$
1475: with $\rho_m$ being the total matter density at
1476: present.
1477:
1478: \section{The CMBR power spectrum}
1479: \label{sec:spectrum}
1480:
1481: \par
1482: In order to calculate the expected total CMBR
1483: angular power spectrum $C_\ell$ in our model, we
1484: must first specify the various contributions to
1485: the amplitude of the total adiabatic and
1486: isocurvature perturbation as well as the
1487: cross correlation between these two perturbations.
1488: In the following, all the amplitudes are referred
1489: to the pivot scale $k_{\rm P}$, for which we use
1490: the customary value $k_{\rm P}=0.05~{\rm Mpc^{-1}}$
1491: \cite{pivot}. We thus define the amplitudes
1492: of the partial curvature perturbations from the
1493: inflaton ($A_{{\rm P},i}$) and the curvaton
1494: ($A_{{\rm P},c}$), and the amplitude of the
1495: isocurvature perturbation in the axions
1496: ($A_{{\rm P},a}$) at $k=k_{\rm P}$:
1497: \begin{eqnarray*}
1498: A_{{\rm P},i}=A_i\left(\frac{k_{\rm P}}{H_0}
1499: \right)^{\nu_i},~
1500: A_{{\rm P},c}=A_c\left(\frac{k_{\rm P}}{H_0}
1501: \right)^{\nu_c},
1502: \end{eqnarray*}
1503: \begin{equation}
1504: A_{{\rm P},a}=A_a\left(\frac{k_{\rm P}}{H_0}
1505: \right)^{\nu_a},
1506: \label{AP}
1507: \end{equation}
1508: where the amplitudes $A_i$, $A_c$ and $A_a$ at
1509: $k=H_0$ are given in Eqs.~(\ref{ai}), (\ref{Ac})
1510: and (\ref{Aa}) respectively.
1511:
1512: \par
1513: From Eq.~(\ref{R}), we find that the amplitude
1514: squared of the adiabatic perturbation at the
1515: pivot scale is given by
1516: \begin{equation}
1517: R^2=\langle\mathcal{R}_{\rm rad}
1518: \mathcal{R}_{\rm rad}\rangle=R_i^2+R_c^2,
1519: \label{R2}
1520: \end{equation}
1521: where $\mathcal{R}_{\rm rad}$ is evaluated at
1522: $k=k_{\rm P}$, and the inflaton ($R_i^2$) and
1523: curvaton ($R_c^2$) contributions to this amplitude
1524: squared are
1525: \begin{equation}
1526: R_i^2=(1-f)^2A_{{\rm P},i}^2~~{\rm and}~~R_c^2=
1527: f^2A_{{\rm P},c}^2.
1528: \label{Ric}
1529: \end{equation}
1530: The curvaton fractional contribution
1531: to the amplitude of the adiabatic perturbation is
1532: defined as follows:
1533: \begin{equation}
1534: F^{\rm ad}_c=\frac{R_c}{R}.
1535: \label{Fc}
1536: \end{equation}
1537:
1538: \par
1539: The amplitude squared of the isocurvature
1540: perturbation at $k_{\rm P}$ is found from
1541: Eq.~(\ref{total}) to be
1542: \begin{equation}
1543: S^2=\langle\mathcal{S}_{\rm rad}
1544: \mathcal{S}_{\rm rad}\rangle=S_i^2+S_c^2+S_a^2,
1545: \label{S2}
1546: \end{equation}
1547: where
1548: \begin{eqnarray*}
1549: S_i^2=9f^2\left(\frac{\Omega_{{\rm LSP}+B}}
1550: {\Omega_m}\right)^2A_{{\rm P},i}^2,
1551: \end{eqnarray*}
1552: \begin{equation}
1553: S_c^2=9f^2\left(\frac{\Omega_{{\rm LSP}+B}}
1554: {\Omega_m}\right)^2A_{{\rm P},c}^2~~{\rm and}~~
1555: S_a^2=\left(\frac{\Omega_a}{\Omega_m}\right)^2
1556: A_{{\rm P},a}^2
1557: \label{Sica}
1558: \end{equation}
1559: are, respectively, the inflaton, curvaton and
1560: axion contributions to this amplitude squared.
1561:
1562: \par
1563: The cross correlation between the adiabatic and
1564: isocurvature perturbation at the pivot scale
1565: $k_{\rm P}$ is
1566: \begin{equation}
1567: C=\langle\mathcal{R}_{\rm rad}
1568: \mathcal{S}_{\rm rad}\rangle=C_i+C_c,
1569: \label{C}
1570: \end{equation}
1571: where
1572: \begin{equation}
1573: C_i=-R_iS_i~~{\rm and}~~C_c=R_cS_c
1574: \label{Cic}
1575: \end{equation}
1576: are the contributions from the inflaton and the
1577: curvaton respectively. Observe that the axions
1578: do not contribute to the cross correlation (and
1579: the amplitude of the adiabatic perturbation).
1580: Also, note that the isocurvature perturbation
1581: from the inflaton is fully anti-correlated with
1582: the corresponding adiabatic perturbation,
1583: whereas the isocurvature perturbation from the
1584: curvaton is fully correlated with the adiabatic
1585: perturbation from it. So the overall correlation
1586: is mixed. It is thus useful to define \cite{iso3}
1587: the dimensionless cross correlation $\cos\Delta$
1588: as a measure of the correlation between adiabatic
1589: and isocurvature perturbations and the
1590: entropy-to-adiabatic ratio $B$ at $k_{\rm P}$:
1591: \begin{equation}
1592: \cos\Delta=\frac{C}{RS}~~{\rm and}~~B=\frac{S}{R}.
1593: \label{cosDB}
1594: \end{equation}
1595:
1596: \par
1597: The total CMBR angular power spectrum is given,
1598: in the notation of Ref.~\cite{thesis}, by the
1599: superposition
1600: \begin{equation}
1601: C_\ell=C_\ell^{\rm ad}+C_\ell^{\rm is}+
1602: C_\ell^{\rm cc},
1603: \label{Cell}
1604: \end{equation}
1605: where
1606: \begin{equation}
1607: C_\ell^{\rm ad}=R_i^2C_\ell^{{\rm ad},n_i}+
1608: R_c^2C_\ell^{{\rm ad},n_c},
1609: \label{Cellad}
1610: \end{equation}
1611: \begin{equation}
1612: C_\ell^{\rm is}=S_i^2C_\ell^{{\rm is},n_i}+
1613: S_c^2C_\ell^{{\rm is},n_c}+
1614: S_a^2C_\ell^{{\rm is},n_a},
1615: \label{Cellis}
1616: \end{equation}
1617: \begin{equation}
1618: C_\ell^{\rm cc}=C_iC_\ell^{{\rm cc},n_i}+
1619: C_cC_\ell^{{\rm cc},n_c}.
1620: \label{Cellcc}
1621: \end{equation}
1622: The above relations hold for the TT,
1623: E-polarization (EE) and TE cross correlation
1624: angular power spectra.
1625:
1626: \par
1627: The TT power spectrum on large angular scales
1628: (i.e. for $\ell\lesssim 20$), can be approximated
1629: by (see e.g. Ref.~\cite{thesis})
1630: \begin{eqnarray}
1631: C^{\rm TT}_\ell&=&\frac{2\pi^2}{25}
1632: \big[(R_i^2+4S_i^2-4C_i)f(n_i,\ell)
1633: \nonumber\\
1634: & &+(R_c^2+4S_c^2-4C_c)f(n_c,\ell)
1635: \nonumber\\
1636: & &+4S_a^2f(n_a,\ell)\big],
1637: \label{Cobe}
1638: \end{eqnarray}
1639: where
1640: \begin{equation}
1641: f(n,\ell)=(\eta_0k_{\rm P})^{1-n}
1642: \frac{\Gamma(3-n)
1643: \Gamma(\ell-\frac{1}{2}+\frac{n}{2})}{2^{3-n}
1644: \Gamma^2(2-\frac{n}{2})\Gamma(\ell+\frac{5}{2}
1645: -\frac{n}{2})}
1646: \label{fnell}
1647: \end{equation}
1648: with $\eta_0=2H_0^{-1}\simeq 8.33\times 10^3~
1649: {\rm Mpc}$ being the value of conformal time in
1650: the present (matter dominated) universe. The
1651: derivation of Eq.~(\ref{Cobe}) assumes that the
1652: universe is completely matter dominated at the
1653: moment of recombination, and further neglects
1654: the late integrated Sachs-Wolfe (ISW) effect
1655: from the cosmological constant. Therefore this
1656: expression is accurate to about $10-20\%$ and
1657: gives an order of magnitude estimate only.
1658: For scale-invariant spectra, $f(n=1,\ell)=1/
1659: \pi\ell(\ell+1)$ and the Sachs-Wolfe (SW)
1660: plateau is flat. Notice from Eq.~(\ref{Cobe})
1661: that a positively correlated perturbation (as
1662: the perturbation from the curvaton) displays
1663: less power on large angular scales, because
1664: the cross correlation term subtracts power and
1665: can partially cancel the isocurvature
1666: contribution. On the contrary, a negatively
1667: correlated perturbation (such as the one from
1668: the inflaton) presents larger power on the
1669: COBE scales and, therefore, can be more easily
1670: constrained. The COBE measurements give
1671: \cite{cobe10} the following central value
1672: for $\ell=10$:
1673: \begin{equation}
1674: \ell(\ell+1)C^{\rm TT}_\ell/2\pi\big
1675: \vert_{\ell=10}\approx 1.05\times 10^{-10}.
1676: \label{ell10}
1677: \end{equation}
1678: We will apply this COBE constraint with
1679: $C^{\rm TT}_\ell$ evaluated from
1680: Eq.~(\ref{Cobe}) only as an indicative
1681: (approximate) criterion to get a first rough
1682: feeling on the possible compatibility of our
1683: model with the data.
1684:
1685: \section{Numerical calculations and results}
1686: \label{sec:numerics}
1687:
1688: \subsection{The cosmological evolution of
1689: $\phi$}
1690: \label{sec:evolution}
1691:
1692: \par
1693: We are now ready to proceed to the numerical
1694: study of the evolution of the PQ field during
1695: and after inflation. To this end, we fix the
1696: parameter $\kappa$ in the superpotential
1697: $\delta W_1$ in Eq.~(\ref{W1}) for standard
1698: hybrid inflation. Then, for any given value of
1699: the amplitude $A_i$ of the partial curvature
1700: perturbation from the inflaton, we solve
1701: Eqs.~(\ref{ai}), (\ref{kappa}) and (\ref{NQ}),
1702: where $x_f$ is the solution of $\eta_i=-1$
1703: with $\eta_i$ given by Eq.~(\ref{etai}) and
1704: $T_{\rm reh}$, which enters Eq.~(\ref{NQ}), is
1705: taken equal to $10^9~{\rm GeV}$ by saturating
1706: the gravitino bound \cite{ekn}. We thus
1707: determine the mass parameter $M$, which is the
1708: magnitude of the VEV breaking the $G_{LR}$
1709: symmetry. Subsequently, we find the (almost
1710: constant) inflationary Hubble parameter
1711: $H_{\rm infl}=\kappa M^2/\sqrt{3}m_{\rm P}$.
1712: The parameters $M$ and $H_{\rm infl}$ as
1713: functions of the amplitude $A_i$ are shown in
1714: Figs.~\ref{fig:M} and \ref{fig:Hinf}
1715: respectively for $\kappa=3\times 10^{-3}$
1716: (solid line) or $3\times 10^{-2}$ (dashed
1717: line). Note that much smaller values of
1718: $\kappa$ would be considered unnatural. On the
1719: other hand, much bigger $\kappa$'s would push
1720: inflation to higher values of the inflaton
1721: field, where SUGRA corrections become important
1722: and may ruin \cite{lyth} inflation.
1723:
1724: \begin{figure}[tb]
1725: \centering
1726: \includegraphics[width=0.83\linewidth]{M.eps}
1727: \caption{The mass parameter $M$ versus $A_i$
1728: for $\kappa=3\times 10^{-3}$ (solid line) or
1729: $3\times 10^{-2}$ (dashed line).}
1730: \label{fig:M}
1731: \end{figure}
1732:
1733: \par
1734: For any given $A_i$, we chose a value $\phi_f$
1735: of $\phi$ at the end of inflation, which takes
1736: place at cosmic time $t_f=2/3H_{\rm infl}$. We
1737: then solve the classical equation of motion
1738: for $\phi$ during inflation in the slow-roll
1739: approximation going backwards in time
1740: ($t\leq t_f$) by taking $m_{3/2}=300~
1741: {\rm GeV}$ (which is \cite{wang} approximately
1742: its minimal value in the
1743: constrained MSSM with Yukawa unification
1744: \cite{ananth}), $\vert A\vert=5$ and a fixed
1745: value for $\lambda$ ($\sim 10^{-4}$) in the PQ
1746: potential. Note that much bigger values of
1747: $\vert A\vert$ (recall that $\vert A\vert>4)$
1748: or much smaller values of $\lambda$ would not
1749: only be unnatural, but also would lead to an
1750: enhancement of the axion decay constant $f_a$
1751: (see Eq.~(\ref{eq:fa})) and thus require
1752: unnaturally small misalignment angles $\theta$
1753: to achieve the observed total CDM abundance
1754: (see Eq.~(\ref{Omegaa})). On the other hand,
1755: much bigger values of $\lambda$ would reduce
1756: $f_a$ leading to an unacceptably small CDM
1757: abundance. Finally, the parameter $\gamma$
1758: during inflation is put equal to zero for
1759: simplicity (see Sec.~\ref{sec:early}). We find
1760: that, as we move backwards in time, the value
1761: of $\phi$ increases and becomes infinite at a
1762: certain moment. Also, the number of e-foldings
1763: elapsed from this moment of time until the end
1764: of inflation is finite providing an upper bound
1765: $N_{\rm max}$ on the number of e-foldings which
1766: is compatible with our boundary condition
1767: $\phi=\phi_f$ at $t_f$.
1768:
1769: \par
1770: In order to get a (rough) understanding of this
1771: behavior, we approximate the derivative of the
1772: potential $V$ with respect to $\phi$ by
1773: $V^{\prime}\simeq 3\lambda^2\phi^5/16
1774: m_{\rm P}^2$, which holds for $\phi\rightarrow
1775: \infty$. The slow-roll equation for $\phi$
1776: during inflation can then be solved analytically
1777: and yields
1778: \begin{equation}
1779: \phi\simeq\frac{\phi_f}
1780: {\left(1+\frac{\lambda^2\phi_f^4}
1781: {4m_{\rm P}^2H_{\rm infl}}\Delta t
1782: \right)^{\frac{1}{4}}},
1783: \label{analytic}
1784: \end{equation}
1785: where $\Delta t=t-t_f\leq 0$. It is obvious
1786: from this equation that, as $\Delta t
1787: \rightarrow\Delta t_{\rm min}\equiv
1788: -4m_{\rm P}^2H_{\rm infl}/\lambda^2\phi_f^4$,
1789: $\phi\rightarrow\infty$. This implies that
1790: the maximal number of e-foldings which is
1791: allowed for a given value of $\phi_f$ is
1792: $N_{\rm max}\simeq -H_{\rm infl}\Delta
1793: t_{\rm min}=4m_{\rm P}^2H_{\rm infl}^2/
1794: \lambda^2\phi_f^4$. Needless to say that no
1795: approximation for $V^{\prime}$ is used when
1796: we actually calculate $N_{\rm max}$.
1797:
1798: \begin{figure}[tb]
1799: \centering
1800: \includegraphics[width=0.83\linewidth]
1801: {Hinf.eps}
1802: \caption{The inflationary Hubble parameter
1803: $H_{\rm infl}$ versus $A_i$ for
1804: $\kappa=3\times 10^{-3}$ (solid line) or
1805: $3\times 10^{-2}$ (dashed line).}
1806: \label{fig:Hinf}
1807: \end{figure}
1808:
1809: \par
1810: It is clear that we must first impose the
1811: requirement that $N_{\rm max}\geq N_Q$. The
1812: time $t_*$ at which our present horizon scale
1813: crosses outside the inflationary horizon can
1814: then be determined from $N_Q=H_{\rm infl}
1815: (t_f-t_*)$. Furthermore, we demand that, at
1816: $t_*$, $V^{\prime\prime}\leq H_{\rm infl}^2$,
1817: which ensures that this inequality holds for
1818: all times between $t_*$ and $t_f$. This
1819: condition, thus, guarantees that the PQ field
1820: is effectively massless during the relevant
1821: part of inflation and can act as curvaton. It
1822: also provides us with an {\it a posteriori}
1823: justification of the validity of the slow-roll
1824: approximation used and ensures that the
1825: velocity of $\phi$ at the end of inflation is
1826: negligible. This masslessness requirement
1827: yields an upper bound on $\phi_f$ for every
1828: given $A_i$ and fixed values of $\kappa$ and
1829: $\lambda$. The excluded region in the
1830: $A_i-\phi_f$ plane for fixed $\kappa$ and
1831: $\lambda$ is represented as a red/dark shaded
1832: area in the upper part of this plane. In
1833: Figs.~\ref{fig:model-4} and \ref{fig:model-5},
1834: we show this upper red/dark shaded area for
1835: $\kappa=3\times 10^{-3}$, $\lambda=10^{-4}$
1836: (model A) and $\kappa=3\times 10^{-2}$,
1837: $\lambda=10^{-4}$ (model B) respectively. The
1838: lower red/dark shaded area in the $A_i-\phi_f$
1839: plane corresponds to the quantum regime and is
1840: also excluded for the reasons explained at the
1841: end of Sec.~\ref{sec:curvature} (see
1842: Figs.~\ref{fig:model-4} and \ref{fig:model-5}).
1843: This also ensures that, at any given
1844: instant of time during inflation, $\phi$ has
1845: practically the same mean value in all comoving
1846: volumes (which crossed outside the horizon).
1847: So, during inflation, $\phi$ can be simply
1848: considered as a function of time only.
1849:
1850: \par
1851: We start from any given value of $\phi_f$ at
1852: $t_f$ which is not excluded by the above
1853: considerations and assume vanishing time
1854: derivative of $\phi$ at $t_f$. As explained
1855: above, this is a good approximation provided
1856: that $\phi_f$ does not belong to the upper
1857: red/dark shaded area in the $A_i-\phi_f$
1858: plane (see Figs.~\ref{fig:model-4} and
1859: \ref{fig:model-5}). Moreover, we find
1860: numerically that the subsequent evolution of
1861: $\phi$ remains practically unchanged if we
1862: take a small non-vanishing value of $\dot\phi$
1863: at $t_f$. Under these initial conditions at
1864: $t_f$, we follow the evolution of the field at
1865: subsequent times ($t\geq t_f$) by solving the
1866: classical equation of motion in
1867: Eq.~(\ref{field-eqn}) with $H=2/(3t)$ for
1868: $t_f\leq t\leq t_{\rm reh}$ and
1869: $H=1/(2(t-t_{\rm reh}/4))$ for $t_{\rm reh}
1870: \leq t\leq t_\phi$. In the latter expression
1871: for $H$, we subtracted $t_{\rm reh}/4$ from
1872: $t$ in order to achieve continuity of $H$ at
1873: $t=t_{\rm reh}$. The time of reheating
1874: $t_{\rm reh}$ is calculated by using
1875: Eq.~(\ref{eq:reheat}) with $g_*=228.75$ and
1876: the curvaton decay time $t_\phi$ is found by
1877: employing Eq.~(\ref{Gphi}) with $\beta$ from
1878: Eq.~(\ref{mu}), where $\mu$ is taken equal to
1879: $300~{\rm GeV}$, which is smaller than half
1880: the curvaton mass (see Eq.~(\ref{mphi})) so
1881: that the decay of $\phi$ to Higgsinos is not
1882: blocked kinematically. The parameter $\gamma$
1883: in the effective PQ potential of
1884: Eq.~(\ref{full-pot}) is taken equal to $0.1$
1885: after the end of inflation (see
1886: Sec.~\ref{sec:early}). We find that, for fixed
1887: $\kappa$ and $\lambda$, there exist two bands
1888: in the $A_i-\phi_f$ plane which lead to the
1889: desired PQ vacua at $t_\phi$. They are
1890: depicted as an upper and a lower green/lightly
1891: shaded band (see Figs.~\ref{fig:model-4} and
1892: \ref{fig:model-5}). The white (not shaded)
1893: areas in the $A_i-\phi_f$ plane lead to the
1894: false (trivial) minimum at $\phi=0$ and thus
1895: must be excluded. Note that, for all relevant
1896: $\kappa$'s and $\lambda$'s, the quantum
1897: regime overlaps (at most) with the lower
1898: green/lightly shaded band only. However, as
1899: we will see later, this band is anyway
1900: excluded by the data in all cases, the reason
1901: being that it predicts an unacceptably large
1902: isocurvature perturbation. So, the {\it a
1903: priori} exclusion of the quantum regime does
1904: not affect our results in any essential way.
1905:
1906: \begin{figure}[tb]
1907: \centering
1908: \includegraphics[width=0.83\linewidth]
1909: {model-4.eps}
1910: \caption{The two green/lightly shaded bands
1911: in the $A_i-\phi_f$ plane which lead to the
1912: PQ vacua at $t_{\phi}$ for $\kappa=3\times
1913: 10^{-3}$, $\lambda=10^{-4}$ (model A). The
1914: white (not shaded) areas lead to the trivial
1915: vacuum and are thus excluded. The upper
1916: red/dark shaded area is excluded by the
1917: requirement that, at $t_*$,
1918: $V^{\prime\prime}\leq H_{\rm infl}^2$, while
1919: the lower one corresponds to the quantum
1920: regime. The blue/solid line shows the values
1921: of $A_i$, $\phi_f$ which approximately
1922: reproduce the correct value of the CMBR large
1923: scale temperature anisotropy, as measured by
1924: COBE (see Sec.~\ref{sec:toy} for details).}
1925: \label{fig:model-4}
1926: \end{figure}
1927:
1928: \par
1929: Besides models A and B (we summarize the
1930: corresponding parameter values in
1931: Table~\ref{tab:models}), which will be used as
1932: our main examples in this presentation, we have
1933: also studied three extra pairs of values of
1934: $\kappa$ and $\lambda$, namely $\kappa=10^{-3}$
1935: and $\lambda=10^{-4}$, $\kappa=10^{-2}$ and
1936: $\lambda=10^{-4}$, and $\kappa=3\times 10^{-2}$
1937: and $\lambda=3\times 10^{-4}$ (see below). In
1938: the first two cases, the behavior was found to
1939: be quite close to the behavior of model A as
1940: depicted in Fig.~\ref{fig:model-4}, while the
1941: latter case behaves similarly to the model B
1942: (see Fig.~\ref{fig:model-5}).
1943:
1944: \begin{figure}[tb]
1945: \centering
1946: \includegraphics[width=0.83\linewidth]
1947: {model-5.eps}
1948: \caption{The two green/lightly shaded bands
1949: in the $A_i-\phi_f$ plane which lead to the
1950: PQ vacua at $t_{\phi}$ for $\kappa=3\times
1951: 10^{-2}$, $\lambda=10^{-4}$ (model B). The
1952: notation is the same as in
1953: Fig.~\ref{fig:model-4}.}
1954: \label{fig:model-5}
1955: \end{figure}
1956:
1957: \begin{table}
1958: \setlength{\extrarowheight}{5pt} \centering
1959: \begin{tabular}{|>{$}l<{$}c|c c|}
1960: \hline
1961: & Eq. & Model A & ~~~~~~~~~~~~~~~~Model B \\
1962: \hline \hline
1963: \kappa & (\ref{W1}) & $3\times10^{-3}$ &
1964: ~~~~~~~~~~~~~~~~$3\times10^{-2}$ \\
1965: \lambda & (\ref{W2}) &
1966: \multicolumn{2}{c|}{$10^{-4}$}\\
1967: m_{3/2} & (\ref{PQ-pot}) &
1968: \multicolumn{2}{c|}{$300$ GeV}\\
1969: \vert A \vert & (\ref{PQ-pot}) &
1970: \multicolumn{2}{c|}{5}\\
1971: \mu & (\ref{mu}) &
1972: \multicolumn{2}{c|}{$300$ GeV}\\
1973: \gamma & (\ref{eq:effec-mass}) &
1974: \multicolumn{2}{c|}
1975: {0 (0.l) during (after) inflation}\\
1976: T_{\text{reh}} & (\ref{eq:reheat}) &
1977: \multicolumn{2}{c|}{$10^9$ GeV}\\
1978: m_{\text{LSP}} & (\ref{OmegaLSP}) &
1979: \multicolumn{2}{c|}{$200$ GeV}\\
1980: \hline
1981: \end{tabular}
1982: \caption{Summary of the fixed model parameters
1983: for the two representative cases presented in
1984: the text, called model A and B. The equation
1985: where each parameter first appears is also
1986: indicated.
1987: \label{tab:models}}
1988: \end{table}
1989:
1990: \subsection{The calculation of $C_\ell$}
1991: \label{sec:Cell}
1992:
1993: \par
1994: For any fixed pair of values for $\kappa$ and
1995: $\lambda$, we take a grid of values of $A_i$
1996: and $\phi_f$ which span the corresponding upper
1997: or lower green/lightly shaded band. For each
1998: point on this grid, we consider $\phi_f$ and
1999: its perturbed value, which is found by adding
2000: to $\phi_f$ the amplitude of $\delta\phi_f$
2001: from Eq.~(\ref{dphif}). We then follow the
2002: subsequent evolution of both these fields until
2003: the time $t_\phi$ of the curvaton decay, where
2004: we evaluate the amplitude of $\delta\rho_\phi/
2005: \rho_\phi$. The amplitude $A_c$ of the partial
2006: curvature perturbation from the curvaton is
2007: then given by Eq.~(\ref{Ac}).
2008:
2009: \par
2010: We thus obtain a mapping of the values of
2011: $\phi_f$ which are allowed for any given value
2012: of $A_i$ onto the corresponding values of the
2013: amplitude $A_c$ of the partial curvature
2014: perturbation from the curvaton, which is the
2015: relevant variable for the calculation of the
2016: CMBR power spectrum. For the MC
2017: analysis (see Sec.~\ref{sec:MC}), we
2018: therefore use $A_i$ and $A_c$ as base
2019: parameters describing the initial conditions
2020: and limit our grid to values of $A_i$ smaller
2021: than $6\times 10^{-5}$ (since larger values
2022: would overpredict the temperature power of the
2023: SW plateau). The spectral indices $n_i$ and
2024: $n_c$ for each point of this grid are found by
2025: using Eqs.~(\ref{epsiloni}) and (\ref{etai})
2026: applied at $x=x_Q$, and Eqs.~(\ref{epsilonc}),
2027: (\ref{etac}) and (\ref{nuc}) applied at $t_*$.
2028: The fraction $f$ in Eq.~(\ref{eq:f}) is also
2029: evaluated at $t_\phi$.
2030:
2031: \par
2032: The amplitude $A_a$ of the isocurvature
2033: perturbation in the axions is calculated from
2034: Eq.~(\ref{Aa}) with the initial misalignment
2035: angle $\theta$ evaluated, for any given value
2036: of the total CDM abundance $\Omega_{\rm CDM}h^2
2037: =\Omega_{\rm LSP}h^2+\Omega_ah^2$, by using
2038: Eqs.~(\ref{OmegaLSP}) and (\ref{Omegaa}) with
2039: $m_{\rm LSP}=200~{\rm GeV}$. This value of
2040: $m_{\rm LSP}$ corresponds \cite{wang}, in the
2041: constrained MSSM with Yukawa unification
2042: \cite{ananth}, roughly to the value of
2043: $m_{3/2}$ ($=300~{\rm GeV}$) chosen here. Note
2044: that, for our choice of parameters, the LSP
2045: relic abundance is fixed ($\simeq 0.0074$). The
2046: spectral index for axions in each point of the
2047: grid is found by using Eq.~(\ref{nua}).
2048:
2049: \par
2050: In summary, for any fixed pair of values for
2051: $\kappa$ and $\lambda$, we take a grid in the
2052: variables $A_i$ and $\phi_f$ covering the
2053: upper or lower green/lightly shaded band and
2054: numerically map $\phi_f$ into the corresponding
2055: value of the amplitude $A_c$ (which, of course,
2056: depends on $A_i$ too). For any value of
2057: $\Omega_Bh^2$ and $\Omega_ah^2$, we then
2058: calculate the amplitudes squared of the
2059: adiabatic and isocurvature perturbations using
2060: Eqs.~(\ref{R2}), (\ref{Ric}) and
2061: Eqs.~(\ref{S2}), (\ref{Sica}) respectively,
2062: the cross correlation amplitude from
2063: Eqs.~(\ref{C}), (\ref{Cic}) and the total CMBR
2064: temperature and polarization power spectra via
2065: Eqs.~(\ref{Cell})$-$(\ref{Cellcc}). The
2066: curvaton fractional contribution to the
2067: adiabatic amplitude $F^{\rm ad}_c$, the
2068: dimensionless cross correlation $\cos\Delta$
2069: and the entropy-to-adiabatic ratio $B$
2070: are found from Eqs.~(\ref{Fc}) and
2071: (\ref{cosDB}). For the MC analysis (see
2072: Sec.~\ref{sec:MC}), we need to be able to
2073: sample the $A_i-A_c$ space in any point as the
2074: chains evolve. Therefore, we perform a
2075: 2-dimensional (2D) interpolation between the
2076: points on the grid using a bicubic spline. A
2077: little care is needed regarding the present
2078: value of the Hubble parameter, which enters the
2079: computation of the mass scale $M$ and thus
2080: $H_{\rm infl}$ for given amplitude $A_i$ via
2081: its impact on $N_Q$ (note that the first term
2082: in the right hand side of Eq.~(\ref{NQ})
2083: depends on $H_0$). As a consequence, the values
2084: of the amplitudes $A_c$ and $A_a$ will also
2085: depend on the value of $H_0$. However, we have
2086: found numerically that changing the value of
2087: $H_0$ around $H_0=72~\rm{km}~\rm{sec}^{-1}~
2088: \rm{Mpc}^{-1}$ within the HST $1\sigma$ margin
2089: (i.e. by $\pm 8~\rm{km}~\rm{sec}^{-1}~
2090: \rm{Mpc}^{-1}$) has an impact less than about
2091: $3\%$ on the computed quantities. Therefore, in
2092: the computation of $M$ and of the amplitudes
2093: $A_c$ and $A_a$, we fix the present Hubble
2094: parameter to the HST central value $H_0=72~
2095: \rm{km}~\rm{sec}^{-1}~\rm{Mpc}^{-1}$
2096: \cite{hst}. Clearly, we do allow $H_0$ to vary
2097: in the MC analysis (see below). In particular,
2098: the dependence on $H_0$ of the amplitudes
2099: $A_{{\rm P},i}$, $A_{{\rm P},c}$ and
2100: $A_{{\rm P},a}$ at the pivot scale $k_{\rm P}$
2101: (see Eq.~(\ref{AP})) is taken into account.
2102:
2103: \subsection{A toy model for the CMBR}
2104: \label{sec:toy}
2105:
2106: \par
2107: Before deploying the full MC machinery to
2108: derive quantitative constraints on the
2109: various parameters of our model, it is
2110: instructive to consider a toy model for the
2111: pre-WMAP CMBR data which allows us to
2112: understand the salient qualitative features
2113: of the parameter space of our initial
2114: conditions. This approach has the great
2115: advantage of offering a much more transparent
2116: understanding of the parameter constraints
2117: beyond the black box of the numerical MC study.
2118:
2119: \par
2120: A first rough idea about the viability of our
2121: model can be obtained by using the approximate
2122: expressions in Eqs.~(\ref{Cobe}) and
2123: (\ref{fnell}) for the temperature SW plateau
2124: and requiring that the COBE constraint in
2125: Eq.~(\ref{ell10}) is fulfilled. To this end,
2126: we take $\Omega_Bh^2=0.0224$ and
2127: $\Omega_{\rm CDM}h^2=0.1126$, which are the
2128: best-fit values from the WMAP measurements
2129: \cite{wmap1}. (Note that, for this choice of
2130: parameters, the axions constitute about $93\%$
2131: of the CDM in the universe.) We find that the
2132: COBE constraint cannot be satisfied in the
2133: lower green/lightly shaded band in the
2134: $A_i-\phi_f$ plane for any pair of values for
2135: $\kappa$ and $\lambda$. The reason is
2136: that the relatively low values of $\phi_f$ in
2137: this band imply small values for $\phi_*$ too.
2138: The amplitude $A_a$ then turns out to be quite
2139: large as can easily be seen from
2140: Eq.~(\ref{Aa}). This fact combined with the
2141: sizable relic abundance of the axions leads to
2142: large values of $S_a$ (see Eq.~(\ref{Sica})),
2143: which yield an unacceptably large contribution
2144: to the right hand side of Eq.~(\ref{Cobe}). In
2145: the upper green/lightly shaded band in the
2146: $A_i-\phi_f$ plane, on the contrary, the COBE
2147: constraint is generally easily satisfied. The
2148: resulting solution is depicted by a
2149: blue/solid line (see Figs.~\ref{fig:model-4}
2150: and \ref{fig:model-5}).
2151:
2152: \par
2153: We are further interested in deriving ball-park
2154: estimates for the values of $A_i$, $A_c$. To
2155: this goal, we again approximate the SW
2156: temperature plateau by the expression in
2157: Eq.~(\ref{Cobe}). As mentioned, this formula
2158: neglects the effect of the cosmological
2159: constant $\Lambda$ (late ISW term), but our
2160: present purpose is to build an extremely
2161: simplified toy model, not to include all
2162: contributions which are fully taken into
2163: account in the MC analysis below.
2164: Therefore, we take one single datum, namely the
2165: COBE constraint in Eq.~(\ref{ell10}), which
2166: describes the height of the SW plateau, along
2167: with its variance $\ell(\ell+1)\Delta
2168: C^{\rm TT}_{\ell}/2\pi\big\vert_{\ell=10}
2169: \approx 0.2\times 10^{-10}$ \cite{cobe10}.
2170: Furthermore, we drop altogether the dependence
2171: of Eq.~(\ref{Cobe}) on the spectral indices by
2172: taking all perturbations to be scale invariant,
2173: since the value of the indices predicted by our
2174: model is anyway very close to unity (see
2175: Sec.~\ref{sec:parcos}). We need a second datum
2176: to be able to constrain two free parameters
2177: ($A_i$ and $A_c$) and this is given by the
2178: height of the first adiabatic peak as measured
2179: by BOOMERANG \cite{deBernardis}. In the
2180: analysis of Ref.~\cite{Durrer}, this yields
2181: $\ell(\ell+1)C^{\rm TT}_{\ell}/2\pi\big
2182: \vert_{\ell= 212}\approx5.4\times 10^{-10}$
2183: with error $\ell(\ell+1)\Delta
2184: C^{\rm TT}_{\ell}/2\pi\big\vert_{\ell=212}
2185: \approx 0.54\times 10^{-10}$. We model the
2186: theoretical CMBR spectrum at the level of the
2187: first adiabatic peak by retaining the adiabatic
2188: contribution only, since the isocurvature
2189: temperature spectrum drops very fast beyond the
2190: SW plateau. Thus the prediction of our toy PQ
2191: model for the height of the first peak is given
2192: by
2193: \begin{equation}
2194: C_{212}^{\rm{TT,PQ}}\approx\frac{\Delta H}{25}
2195: \left(R_i^2+R_c^2\right),
2196: \end{equation}
2197: where the constant factor $\Delta H\approx 6.5$
2198: approximates the ratio of the first peak height
2199: to the SW plateau for the adiabatic temperature
2200: spectrum of the standard $\Lambda$CDM model.
2201: Once again, this expression is very crude and
2202: does not account for changes in $h$,
2203: $\Omega_\Lambda$ ($\equiv\rho_\Lambda/\rho_c$
2204: with $\rho_\Lambda$ being the dark energy
2205: density), $\Omega_m$, $\Omega_B$ or $\tau_r$
2206: (optical depth to the reionization epoch), all
2207: of which affect the relative height of the peak
2208: to the plateau, but it is sufficient for our
2209: present goal. In summary, the likelihood
2210: function $\like (A_i,A_c)$ of our toy model is
2211: given by
2212: \begin{eqnarray}
2213: -2\ln\like (A_i,A_c)&\approx&\left(
2214: \frac{C_{10}^{\rm{TT,PQ}}
2215: (A_i,A_c)-C^{\rm{TT}}_{10}}
2216: {\Delta C^{\rm{TT}}_{10}}\right)^2
2217: \\
2218: & &+\left(\frac{C_{212}^{\rm{TT,PQ}}(A_i,A_c)-
2219: C^{\rm{TT}}_{212}}{\Delta C^{\rm{TT}}_{212}}
2220: \right)^2,
2221: \nonumber
2222: \end{eqnarray}
2223: where $C_{10}^{\rm{TT,PQ}}$ can be found from
2224: Eq.~(\ref{Cobe}). The posterior (see
2225: Appendix for definitions) is then
2226: \begin{equation}
2227: \pdf(A_i, A_c)\propto\like(A_i, A_c)\pi(A_i)
2228: \pi(A_c).
2229: \end{equation}
2230: We adopt non-informative flat priors on the
2231: amplitudes $A_i, A_c$, so that $\pi(A_i)=
2232: \pi(A_c)=\text{constant}$. An alternative
2233: choice is Jeffreys' prior, which is of the
2234: form $\pi(A_i)=1/A_i$, $\pi(A_c)=1/A_c$
2235: corresponding to a flat prior on $\log A_i$,
2236: $\log A_c$. This prior implies that we do not
2237: have any idea on the scale of $A_i$, $A_c$
2238: before seeing the data and thus represents a
2239: quite extreme choice of prior.
2240:
2241: \begin{figure}[tb]
2242: \centering
2243: \includegraphics[width=\linewidth]{PQToy.ps}
2244: \caption{Posterior marginalized pdf (normalized
2245: at peak value) for the amplitudes $A_i$, $A_c$
2246: from our toy model for two choices of priors:
2247: flat priors (top panels), and Jeffreys' priors
2248: (bottom panels). The black line is for model A
2249: (upper green/lightly shaded band in
2250: Fig.~\ref{fig:model-4}) and the cyan/light gray
2251: line for model B (upper green/lightly shaded
2252: band in Fig.~\ref{fig:model-5}). The
2253: constraints on $A_i$, $A_c$ in model B and on
2254: $A_i$ in model A as well as the upper limit on
2255: $A_c$ in model A are robust with respect to the
2256: choice of priors. Compare the top panels with
2257: Fig.~\ref{fig:1dbase}, which shows the results
2258: of the full MC analysis (with flat
2259: non-informative priors).}
2260: \label{fig:pqtoy}
2261: \end{figure}
2262:
2263: \begin{figure}[tb]
2264: \centering
2265: \includegraphics[width=\linewidth]
2266: {PQCompareBase.ps}
2267: \caption{1-dimensional (1D) marginalized
2268: posterior distribution for the two cases
2269: considered here: model A (black solid line) and
2270: model B (cyan/light gray solid line) with only
2271: the upper green/lightly shaded bands of
2272: Figs.~\ref{fig:model-4} and \ref{fig:model-5}
2273: included. The red/medium gray line is for the
2274: standard pure inflaton single-field
2275: inflationary case with Harrison-Zel'dovich (HZ)
2276: spectrum ($n_i=1.00$ fixed), plotted here for
2277: comparison. We plot non-smoothed histograms
2278: corresponding to top-hat binnings to show the
2279: resolution of our curves. Model A is quite
2280: close to the pure inflaton case, except that the
2281: curvaton contribution helps in reducing the
2282: reionization redshift $z_r$. Model B displays a
2283: preference for non-zero curvaton contribution,
2284: a still lower $z_r$ and larger relic abundance
2285: of baryons ($\omega_B=\Omega_Bh^2$) and CDM
2286: ($\omega_{\rm CDM}=\Omega_{\rm CDM}h^2$). Its
2287: quality of fit is however poorer (see
2288: Sec.~\ref{sec:modelcomp} for details). We also
2289: display as dotted smoothed curves the values of
2290: the mean posterior. All the curves are
2291: normalized at their peak value.}
2292: \label{fig:1dbase}
2293: \end{figure}
2294:
2295: \par
2296: In Fig.~\ref{fig:pqtoy}, we present the
2297: posterior marginalized probability distribution
2298: functions (pdf's) for $A_i$, $A_c$ from our toy
2299: model for the two choices of priors above. We
2300: observe that there is a qualitative difference
2301: between the results for the upper green/lightly
2302: shaded band in Fig.~\ref{fig:model-4} and
2303: \ref{fig:model-5} of model A and B. For model
2304: B, both the inflaton and the curvaton amplitude
2305: are well determined even in our over-simplified
2306: toy model. From the plot, we read off that $A_i
2307: \approx 2.3\times 10^{-5}$, $A_c\approx 70
2308: \times 10^{-5}$. As a consequence, the choice
2309: of prior hardly matters, as we would expect in
2310: a situation where the posterior is dominated by
2311: the likelihood. As for model A, we obtain that
2312: $A_i\approx 4.9\times 10^{-5}$, but we can only
2313: place upper limits on the value of $A_c$.
2314: Since, in this case, $A_c$ is not a
2315: well-determined parameter, the details of its
2316: posterior pdf do depend on the prior.
2317: Nevertheless, it is apparent that the upper
2318: limit is robust, and we can deduce that $A_i
2319: \lsim 50\times 10^{-5}$. These numbers have to
2320: be taken only as ball-park estimates, and this
2321: is why we do not bother to attach errors to
2322: them. However, the comparison of
2323: Fig.~\ref{fig:pqtoy} with
2324: Fig.~\ref{fig:1dbase}, which shows the results
2325: of the quantitative MC analysis including WMAP
2326: and other recent CMBR data and all the other
2327: cosmological parameters, displays an
2328: astonishing agreement between the results of
2329: the above toy model and of the full MC
2330: analysis. The rather precise match of the two
2331: results is actually the fortuitous outcome of
2332: the cancellation between two opposite effects.
2333: On the one hand, the MC analysis compared to
2334: the toy model employs much more precise data,
2335: which reduce the error on the constraints, but,
2336: on the other hand, it also integrates over the
2337: cosmological parameter space (totally ignored
2338: in the toy model), which enlarges the error on
2339: the marginalized quantities.
2340:
2341: \par
2342: From the study of the above toy model, we can
2343: thus draw two conclusions. First, the height of
2344: the first adiabatic peak to the large scale SW
2345: plateau is the key quantity to constraining --
2346: with at least order of magnitude accuracy --
2347: the curvaton to inflaton contribution in the PQ
2348: model. Clearly, quantitatively reliable bounds
2349: need the inclusion of the detailed effect of
2350: the cosmological parameters on the shape of the
2351: power spectra (see next section). Second, we
2352: have seen that the constraints on the inflaton
2353: and curvaton amplitudes are robust with respect
2354: to two different choices of (non-informative)
2355: priors. For this reason, we will, from now on,
2356: adopt a flat prior on $A_i$, $A_c$, which will
2357: be used both for the extraction of the
2358: constraints on the parameters (see
2359: Sec.~\ref{sec:parcos}) and for the purpose of
2360: Bayesian model comparison (see
2361: Sec.~\ref{sec:modelcomp}).
2362:
2363: \subsection{The Monte Carlo CMBR analysis}
2364: \label{sec:MC}
2365:
2366: \subsubsection{The setup}
2367: \label{sec:setup}
2368:
2369: \par
2370: We now proceed to describe the setup of the
2371: full numerical analysis confronting the
2372: predictions of our PQ model with the CMBR
2373: data. We constrain the relevant parameters of
2374: our model by constructing Markov chains using
2375: a modified version of the publicly available
2376: Markov chain MC package \textsc{cosmomc}
2377: \cite{cosmomc} as described in
2378: Ref.~\cite{lewis}. As discussed in
2379: Sec.~\ref{sec:spectrum}, the total CMBR angular
2380: power spectrum is given by a (anti-)correlated
2381: superposition of adiabatic and isocurvature
2382: CMBR modes (see Eq.~(\ref{Cell})). The
2383: adiabatic and isocurvature CMBR transfer
2384: functions are computed in two successive calls
2385: similarly to the technique employed in
2386: Ref.~\cite{Valiviita}. For fixed values of
2387: $\kappa$ and $\lambda$, the initial conditions
2388: are completely specified by $A_i$ and $\phi_f$,
2389: or equivalently by the values of $A_i$ and
2390: $A_c$ as explained in Sec.~\ref{sec:Cell}. The
2391: MC sampling takes as free parameters the
2392: amplitudes $A_i$ and $A_c$, the physical baryon
2393: and axion densities in the present universe
2394: $\omega_B\equiv\Omega_Bh^2$ and
2395: $\omega_a\equiv\Omega_ah^2$ in units of
2396: $1.88\times 10^{-29}~{\rm g/cm}^3$, the
2397: present value of the dimensionless Hubble
2398: parameter $h=H_0/100\text{ km}\text{ sec}^{-1}
2399: \text{ Mpc}^{-1}$, and the redshift $z_r$ at
2400: which the reionization fraction is a half
2401: (assuming sudden reionization). All other
2402: derived quantities are computed from the above
2403: parameter set, as detailed in
2404: Sec.~\ref{sec:Cell}. In particular, the derived
2405: parameters $R_i^2$, $R_c^2$, $F^{\rm ad}_c$,
2406: $S_i^2$, $S_c^2$, $S_a^2$, $C_i$, $C_c$ are
2407: obtained from Eqs.~(\ref{Ric}), (\ref{Fc}),
2408: (\ref{Sica}) and (\ref{Cic}).
2409:
2410: \par
2411: For our choice of $m_{\text{LSP}}$ and
2412: $T_{\rm reh}$, we have $\Omega_{\text{LSP}}h^2
2413: \approx 0.0074=\text{constant}$, and the total
2414: CDM abundance is $\omega_{\rm CDM}\equiv
2415: \Omega_{\rm CDM}h^2=\Omega_{\text{LSP}}h^2+
2416: \Omega_a h^2$. Our analysis considers flat
2417: cosmologies only, thus the cosmological
2418: constant energy density $\Omega_\Lambda$ (in
2419: units of the critical energy density) is a
2420: derived parameter, i.e. $\Omega_\Lambda=1-
2421: (\omega_{\rm CDM}+\omega_B)/h^2$. We assume
2422: three massless neutrino families and no
2423: massive neutrinos (for constraints on these
2424: quantities, see e.g. Ref.~\cite{constraints}).
2425: We neglect the contribution of gravitational
2426: waves to the spectrum, since the tensor to
2427: scalar amplitude ratio at the SW
2428: plateau is proportional to $\epsilon_i$,
2429: which is completely negligible in our case.
2430: In summary, for a fixed choice of $\kappa$
2431: and $\lambda$, our parameter space is six
2432: dimensional:
2433: \begin{equation}
2434: \params=\{\omega_B,\omega_a,h,z_r,A_i,A_c\}\;.
2435: \end{equation}
2436:
2437: \par
2438: We compare the predicted CMBR temperature and
2439: polarization power spectra to the WMAP
2440: first-year data \cite{wmap1} (temperature and
2441: polarization) with the routine for computing
2442: the likelihood supplied by the WMAP team
2443: \cite{Verde}. At a smaller angular scale, we
2444: add the CBI \cite{cbi} and the decorrelated
2445: ACBAR \cite{acbar1,acbar2} band powers as well.
2446: We then run $N=20$ Markov chains starting from
2447: randomly chosen points in the parameter space.
2448: We take particular care in ensuring that the
2449: starting points are spread over a wide range in
2450: the $A_i-A_c$ plane. We then check that all
2451: chains have converged to the same region of
2452: parameter space. This indicates that this
2453: region is indeed a global minimum (at least for
2454: the range explored by the chains). This is the
2455: main reason for using a relatively large number
2456: of chains, since the danger that the chains are
2457: stuck in a local minimum is great when
2458: exploring mixed isocurvature initial conditions
2459: (see e.g. Ref.~\cite{iso2}). A preliminary run
2460: is needed to estimate the covariance matrix,
2461: which is then diagonalized and used to perform
2462: a final run until each chain contains
2463: $M=30,000$ samples. The mixing of the chains is
2464: checked using the Gelman and Rubin criterion
2465: \cite{gelman}, for which we require that the
2466: ratio of the variance of the mean to the mean
2467: of the variance among the chains is $R<0.1$ for
2468: all parameters. The parameter inference is
2469: performed on the merged chains, which contain
2470: around $5\times 10^5$ samples after the burn-in
2471: has been discarded.
2472:
2473: \par
2474: As motivated above, we use flat top-hat priors
2475: on the base parameters
2476: \begin{equation}
2477: \omega_B, \omega_a, z_r, A_i, A_c .
2478: \end{equation}
2479: The limits of the top-hat prior do not matter
2480: for parameter estimation purposes, as long as
2481: we check that the posterior density is
2482: negligible near these limits. However, the
2483: prior range of the accessible parameter space
2484: plays an important role in computing the
2485: Bayes factor for model testing (see
2486: Sec.~\ref{sec:modelcomp}). We limit the maximum
2487: range of the present dimensionless Hubble
2488: parameter $h$ by imposing a top-hat prior $0.40
2489: <h<1.00$ and we use the result of the HST
2490: measurement \cite{hst}
2491: \begin{equation}
2492: \like^{\text{HST}}\propto\exp\left(\frac{h-h_0}
2493: {2\sigma}\right)^2,
2494: \end{equation}
2495: where $h_0=0.72$ and $\sigma=0.08$, by
2496: multiplying the likelihood function for the
2497: CMBR data by the above Gaussian likelihood.
2498:
2499: \par
2500: Parameter constraints will be rather
2501: insensitive to the details of the prior
2502: distribution whenever the posterior is
2503: dominated by the likelihood. As we have seen
2504: from our toy model, the broad lines of the
2505: constraints for the PQ model are indeed
2506: robust with respect to the choice of
2507: non-informative priors. We will see below that
2508: the priors do play an important role in
2509: Bayesian model comparison.
2510:
2511: \begin{table}
2512: \setlength{\extrarowheight}{5pt} \centering
2513: \begin{tabular}{|>{$}l<{$}|>{$}c<{$}>{$}c<{$}|>
2514: {$}c<{$}>{$}c<{$}|}
2515: \hline
2516: \text{param} & \multicolumn{2}{c|}
2517: {\text{model A (upper band)}} &
2518: \multicolumn{2}{c|}
2519: {\text{model B (upper band)}}
2520: \\
2521: \hline
2522: & \text{best fit} & \text{1D }$68\%$
2523: \text{ c.i.} & \text{best fit} &
2524: \text{1D }$68\%$ \text{ c.i.}
2525: \\
2526: \hline\hline
2527: \om_B & 0.024 & 0.024 \pm 0.002 & 0.026 & 0.026
2528: \pm 0.002
2529: \\
2530: \om_a & 0.109 & 0.110 \pm 0.034 & 0.117 & 0.121
2531: \pm 0.040
2532: \\
2533: H_0 & 72.4 & 72.2^{+11.5}_{-9.8} & 64.2 &
2534: 63.4^{+11.0}_{-9.2}
2535: \\
2536: z_r & 13.3 & 13.3^{+8.1}_{-9.3} & 10.4 &
2537: 10.5^{+7.7}_{-6.5}
2538: \\
2539: A_i\uf & 5.1 & 5.1 \pm 0.5 & 3.1 & 3.1\pm 0.4
2540: \\
2541: A_c\uf & 20.3 & <37.9\;(43.2) & 62.0 & 62.8\pm
2542: 10.4
2543: \\
2544: \hline
2545: -\ln L_* & 721.2 & & 732.6 &
2546: \\
2547: \hline
2548: \end{tabular}
2549: \caption{Best-fit parameter values and sample
2550: means with $1\sigma$ confidence intervals
2551: (c.i.) for the 1D marginalized distribution
2552: for the upper band of models A and B. We
2553: indicate upper limits only when the parameter
2554: is not constrained from below, in which case
2555: the first number corresponds to the $68\%$ c.i.
2556: (1 tail) and the number in parenthesis to the
2557: $95\%$ c.i. of the parameter. The present
2558: Hubble parameter $H_0$ is given in $\rm{km}~
2559: \rm{sec}^{-1}~\rm{Mpc}^{-1}$. Finally, $L_*$
2560: is the best-fit likelihood (see Appendix). For
2561: comparison, the Harrison-Zel'dovich (HZ) pure
2562: inflaton model has $-\ln L_*=721.7$.
2563: \label{tab:bestfitbase}}
2564: \end{table}
2565:
2566: \subsubsection{Parameter constraints}
2567: \label{sec:parcos}
2568:
2569: In the Appendix, we summarize
2570: some concepts and results from Bayesian
2571: statistics which will be useful in the
2572: following analysis (see e.g.
2573: Refs.~\cite{MKbook} and \cite{bbook} for reviews
2574: and details). We first consider the
2575: upper green/lightly shaded band of models A and
2576: B. In Tables~\ref{tab:bestfitbase} and
2577: \ref{tab:bestfitderiv}, we present the best-fit
2578: values and parameter constraints obtained from
2579: the MC chains for our base and derived
2580: parameters respectively. The 1-dimensional (1D)
2581: marginalized posterior distributions for the
2582: base and most of the derived parameters are
2583: plotted, respectively, in
2584: Figs.~\ref{fig:1dbase} and \ref{fig:1dderiv},
2585: while the 2D contours of the posterior for
2586: $A_i$, $A_c$ and the adiabatic amplitudes
2587: squared $R_i^2$, $R_c^2$ are displayed in
2588: Fig.~\ref{fig:2dbase}.
2589:
2590: \begin{table}
2591: \setlength{\extrarowheight}{5pt} \centering
2592: \begin{tabular}{|>{$}l<{$}|>{$}c<{$}>{$}c<{$}|>
2593: {$}c<{$}>{$}c<{$}|}
2594: \hline
2595: \text{param} & \multicolumn{2}{c|}
2596: {\text{model A (upper band)}} &
2597: \multicolumn{2}{c|}{\text{model B (upper band)}}
2598: \\
2599: \hline
2600: & \text{best fit} & \text{1D }$68\%$
2601: \text{ c.i.} & \text{best fit} & \text{1D }
2602: $68\%$\text{ c.i.}
2603: \\
2604: \hline \hline
2605: R_i^2\ut & 20.4 & 21.0^{+6.2}_{-4.2} & 7.8 &
2606: 7.7 \pm 2.2
2607: \\
2608: R_c^2\ut & 2.5 & < 10.9\;(14.0)& 14.8 &
2609: 15.3^{+6.3}_{-4.9}
2610: \\
2611: S_i^2\ut & 0.06 & 0.05\pm0.04 & 0.015 &
2612: 0.015 \pm 0.006
2613: \\
2614: S_c^2\ut & 1.2 & < 5.0\;(7.1) & 6.6 &
2615: 6.6^{+3.2}_{-2.1}
2616: \\
2617: S_a^2\ut & 0.4 & 0.4 \pm 0.1 & 6.4 & 6.3
2618: \pm 1.6
2619: \\
2620: C_i\ut & -1.1 & -1.0^{+0.6}_{-0.5}& -0.3
2621: & -0.3 \pm 0.1
2622: \\
2623: C_c\ut & 1.8 & < 6.9\;(9.5) & 9.9 &
2624: 10.0^{+3.6}_{-2.3}
2625: \\
2626: \cos\Delta &0.08 & 0.0^{+0.5}_{-0.2} & 0.55 &
2627: 0.56 \pm 0.10
2628: \\
2629: B & 0.08 & <0.45\;(0.52) & 0.76 & 0.75 \pm 14
2630: \\
2631: \hline
2632: F^{\ad}_c & 0.34 & < 0.60\;(0.67) & 0.81 & 0.81
2633: \pm 0.07
2634: \\
2635: f & 0.08 & 0.07^{+0.02}_{-0.05} & 0.06 & 0.062
2636: \pm 0.004
2637: \\
2638: n_i & 0.988 &- & 0.982 & -
2639: \\
2640: n_c & 1.011 & - & 1.003 & -
2641: \\
2642: n_a & 1.002 & - & 1.000 & -
2643: \\
2644: \hline
2645: \end{tabular}
2646: \caption{As in Table~\ref{tab:bestfitbase}, but
2647: for the derived parameters. We do not give c.i.
2648: for the spectral indices since the variation
2649: in their value is less than $10^{-3}$.
2650: \label{tab:bestfitderiv}}
2651: \end{table}
2652:
2653: \begin{figure}[tb]
2654: \centering
2655: \includegraphics[width=\linewidth]
2656: {PQCompareDeriv.ps}
2657: \caption{As in Fig.~\ref{fig:1dbase}, but for
2658: (most of) the derived parameters (the HZ pure
2659: inflaton model is not included here). The
2660: adiabatic amplitude in model B is dominated by
2661: the curvaton ($F_c^{\rm ad}\approx 0.8$).}
2662: \label{fig:1dderiv}
2663: \end{figure}
2664:
2665: \par
2666: In the upper (green/lightly shaded) band of
2667: model A, the adiabatic amplitude from the
2668: inflaton dominates with only a modest
2669: contribution ($\sim 10\%$ in amplitude squared)
2670: to the adiabatic amplitude from the curvaton
2671: and an even smaller curvaton isocurvature
2672: amplitude and cross correlator. Note that the
2673: axion isocurvature amplitude is negligible in
2674: this case. Due to the small value of the
2675: parameter $f$, the inflaton part of the
2676: isocurvature amplitude squared (correlator) is
2677: suppressed relative to the corresponding part
2678: of the adiabatic amplitude squared by more than
2679: two orders (one order) of magnitude. As a
2680: consequence, the power
2681: spectra are dominated by the adiabatic part of
2682: the inflaton contribution. Moreover, on large
2683: scales, we find that the sum of the adiabatic,
2684: isocurvature and correlation parts of the total
2685: power spectrum coming from the curvaton almost
2686: cancels out, since the contribution from the
2687: curvaton correlator is negative. We can easily
2688: verify the behavior just described with the
2689: help of Fig.~\ref{fig:BFpqMOD4}, where we plot
2690: the best-fit power spectra for model A (upper
2691: band) divided into the inflaton, curvaton and
2692: axion contributions to their adiabatic,
2693: isocurvature and correlator parts. The quality
2694: of the fit, as expressed by the maximum
2695: likelihood value $-\ln L_*=721.2$, is slightly
2696: better than for
2697: the pure inflaton Harrison-Zel'dovich (HZ)
2698: case, which has $-\ln L_*=721.7$ (see, however,
2699: our remarks below regarding model comparison).
2700: This is not surprising since the curvaton
2701: contribution turns out to play a modest role in
2702: this model. In particular, the CMBR data put an
2703: upper bound on the allowed value of the
2704: curvaton amplitude, which is $A_c< 43.2\times
2705: 10^{-5}$ at $95\%$ confidence level (c.l.). The
2706: total temperature power on large scales is
2707: slightly larger than the pure adiabatic part --
2708: the net effect is to increase the height of the
2709: SW plateau compared to the height of the first
2710: adiabatic peak. This mimics the impact of a
2711: larger optical depth (and thus of a larger
2712: $z_r$), and explains why model A shows a
2713: preference for a later reionization than in the
2714: pure inflaton case. The TE spectrum is
2715: dominated by the inflaton adiabatic part, but
2716: on large scales the curvaton and axion
2717: contributions give a net power increase. This
2718: effect helps in better fitting the
2719: ``reionization bump'' (i.e. the power increase
2720: for $\ell\lsim 10$ due to reionization) at low
2721: multipoles reducing the need for a rather
2722: early reionization.
2723:
2724: \begin{figure}[tb]
2725: \centering
2726: \includegraphics[width=\linewidth]
2727: {CompareBase_2D.ps}
2728: \caption{Contours containing 68\% and 95\% of
2729: the probability for the joint posterior pdf for
2730: the upper band of models A and B (same
2731: notation as in Fig.~\ref{fig:1dbase}). The
2732: base primordial parameters $A_i$, $A_c$ are
2733: displayed in the
2734: left panel, while the right panel shows the
2735: derived adiabatic amplitudes squared from the
2736: inflaton ($R_i^2$) and the curvaton ($R_c^2$).
2737: Model B prefers a non-vanishing curvaton
2738: contribution to the adiabatic amplitude, but
2739: its quality of fit is poorer (see text for
2740: details).}
2741: \label{fig:2dbase}
2742: \end{figure}
2743:
2744: \begin{figure}[tb]
2745: \centering
2746: \includegraphics[width=\linewidth]
2747: {BFpqMOD4.peccei.ps}
2748: \caption{Best-fit TT (upper panel) and TE
2749: (lower panel) power spectra for the upper band
2750: of model A (parameters as in
2751: Table~\ref{tab:bestfitbase}). The line codes
2752: are: red/dark gray for the inflaton,
2753: green/medium gray for the curvaton, and
2754: yellow/light gray for the axion contributions;
2755: dotted for the adiabatic, long--dashed for the
2756: isocurvature, and short--dashed for the
2757: correlator parts. The total power corresponds
2758: to the blue/solid line. In the upper panel, the
2759: correlator parts are given in absolute value.
2760: Note though that the inflaton correlator
2761: contribution has to be added (negative
2762: correlation), while the curvaton one has to be
2763: subtracted (positive correlation) to obtain the
2764: total power.}
2765: \label{fig:BFpqMOD4}
2766: \end{figure}
2767:
2768: \begin{figure}[tb]
2769: \centering
2770: \includegraphics[width=\linewidth]
2771: {BFpqMOD5.peccei.ps}
2772: \caption{Best-fit power spectra for the upper
2773: band of model B. The base parameters are
2774: according to Table~\ref{tab:bestfitbase}, and
2775: the notation is as in Fig.~\ref{fig:BFpqMOD4}.}
2776: \label{fig:BFpqMOD5}
2777: \end{figure}
2778:
2779: \begin{figure}[tb]
2780: \centering
2781: \includegraphics[width=\linewidth]{chi_TTEE.ps}
2782: \caption{Cumulative difference between model B
2783: and model A (upper bands) of the quantity
2784: $\chi^2/2\equiv -\ln L_*$ for the WMAP TT
2785: (green/dotted line in the top panel) and TE
2786: data (blue/dashed line in the bottom panel) in
2787: units given on the left vertical axes. Their
2788: sum is given in the top panel by the thick
2789: black solid line. In the top panel, we
2790: superimpose the two
2791: corresponding best-fit TT spectra for model A
2792: (thin black line) and model B (thin cyan/light
2793: gray line) with units on the right vertical
2794: axis. In the bottom panel, we superimpose the
2795: two TE spectra (same notation). To compute the
2796: $\chi^2$ difference as a function of the
2797: multipole $\ell$, we use only the diagonal
2798: elements of the data covariance matrix (for the
2799: MC analysis, we, of course, included also the
2800: off-diagonal elements, which however contribute
2801: only a few percent). We also plot the binned
2802: WMAP TT (top panel) and TE (bottom panel)
2803: errorbars, as a guide to the eye to appreciate
2804: the discriminative power of the WMAP data,
2805: especially around the first acoustic peak in
2806: the temperature spectrum. Model B is a better
2807: fit to the TT SW plateau, since its power
2808: there is smaller, but its TE spectrum in this
2809: region fits the WMAP data worse. Model B cannot
2810: reproduce the overall shape of the first
2811: acoustic peak in temperature with enough accuracy,
2812: and its goodness-of-fit is correspondingly
2813: worse.}
2814: \label{fig:chidiff}
2815: \end{figure}
2816:
2817: \begin{figure}[tb]
2818: \centering
2819: \includegraphics[width=\linewidth]
2820: {PQCompareCorr.ps}
2821: \caption{As in Fig.~\ref{fig:1dbase}, but for
2822: the dimensionless correlator $\cos\Delta$ and
2823: the entropy-to-adiabatic ratio $B$ evaluated
2824: at the pivot scale $k_P$ (see
2825: Eq.~(\ref{cosDB})). (The HZ pure inflaton model
2826: is not included.) The sharp drop for
2827: $B\lsim 0.11$ encountered in model A is a
2828: numerical feature due to the lower boundaries
2829: of our MC run. The curve must accordingly be
2830: interpreted as indicating an upper limit only,
2831: which is given in
2832: Table~\ref{tab:bestfitderiv}.}
2833: \label{fig:1dcorr}
2834: \end{figure}
2835:
2836: \begin{figure}[tb]
2837: \centering
2838: \includegraphics[width=0.5\linewidth]
2839: {CompareCorr_2D.ps}
2840: \caption{Contours containing 68\% and 95\% of
2841: the probability for the joint posterior pdf
2842: for the upper band of models A and B for the
2843: entropy-to-adiabatic ratio $B$ and
2844: the dimensionless correlator $\cos\Delta$ (same
2845: notation as in Fig.~\ref{fig:2dbase}).}
2846: \label{fig:2dcomparecorr}
2847: \end{figure}
2848:
2849: \par
2850: The upper (green/lightly shaded) band of model
2851: B exhibits a preference for a non-vanishing
2852: curvaton amplitude with very high significance
2853: ($A_c=62.8\pm 10.4$) as one can also see from
2854: Fig.~\ref{fig:2dbase}. In this case, the
2855: best-fit spectrum is a superposition of a
2856: dominant curvaton adiabatic part and an
2857: inflaton adiabatic contribution which is around
2858: $50\%$ of the former in amplitude squared (see
2859: Table~\ref{tab:bestfitderiv} and
2860: Fig.~\ref{fig:BFpqMOD5}). This time, the
2861: isocurvature curvaton part is sizable on large
2862: scales, where however it is again cancelled by
2863: the correlator. The large value of $\kappa$
2864: yields a larger $H_*$ (see Fig.~\ref{fig:Hinf})
2865: and thus pushes up the axion isocurvature
2866: contribution (see Eq.~(\ref{Aa})),
2867: which again adds power to the SW plateau. As in
2868: model A (upper band), the isocurvature and
2869: correlator inflaton parts are negligibly small.
2870: For the best-fit parameters, the total power in
2871: temperature in the COBE region is in better
2872: agreement with the data, being slightly smaller
2873: than for model A (upper band), thereby fitting
2874: better the low-multipole region of the
2875: spectrum. However, the overall quality of the
2876: fit is worse ($-\ln L_*=732.6$) because the
2877: model does not reproduce with enough precision
2878: the shape of the first acoustic peak. We
2879: illustrate this in Fig.~\ref{fig:chidiff},
2880: where we compare the two best-fit spectra for
2881: models A and B (upper bands). The WMAP data
2882: around the first temperature peak are of such a
2883: good quality that they can discriminate between
2884: the two models thanks to the slightly different
2885: shape of the peak. The reason why the
2886: curvaton/inflaton mixture does not fit
2887: accurately enough the first temperature peak in
2888: model B is twofold. First, the isocurvature and
2889: correlator contributions are still sizable in
2890: the region of the first peak rise ($\ell\sim
2891: 100$), and this increases slightly the total
2892: power in this part of the spectrum. Second, the
2893: curvaton and inflaton have two slightly
2894: different spectral indices (see
2895: Table~\ref{tab:bestfitderiv}), and the resulting
2896: tilts of the spectra are therefore mismatched.
2897: As for the TE spectrum, the isocurvature axion
2898: part plays an important role in reproducing
2899: the reionization bump. Furthermore, model B
2900: (upper band) shows a preference for a rather
2901: high baryon abundance ($\omega_B\approx
2902: 0.026$), which is in strong tension with the
2903: value indicated by BBN together with
2904: observations of the light elements abundance,
2905: which typically yields $\omega_B\approx 0.020
2906: \pm 0.002$ \cite{Burles}.
2907:
2908: \begin{table}
2909: \setlength{\extrarowheight}{5pt} \centering
2910: \begin{tabular}{|>{$}l<{$}|>{$}c<{$}|>{$}c
2911: <{$}|}
2912: \hline
2913: \text{param} & {\text{model A (lower band)}} &
2914: {\text{model B (lower band)}}
2915: \\
2916: \hline\hline
2917: \om_B & 0.029 & 0.080 \\
2918: \om_a & 0.171 & 0.010\star \\
2919: H_0 & 50.1 & 100.0\star \\
2920: z_r & 19.4 & 30.0\star \\
2921: A_i\uf & 0.10\star & 0.41 \\
2922: A_c\uf & 75.0 & 300.0\star \\
2923: \hline\hline
2924: R_i^2\ut & 0.001 & 0.14\\
2925: R_c^2\ut & 33.25 & 134.1\\
2926: S_i^2\ut & \sim 10^{-4} & 0.002\\
2927: S_c^2\ut & 9.05 & 971.4\\
2928: S_a^2\ut & 14.37& 5512.1\\
2929: C_i\ut & 0.003& 0.02\\
2930: C_c\ut & 17.35& 360.9\\
2931: \cos\Delta & 0.62 & 0.39\\
2932: B & 0.84 & 6.95\\
2933: \hline
2934: %
2935: F^{\ad}_c & 1.00 & 1.00\\
2936: f & 0.077& 0.39\\
2937: n_i & 0.986& 0.982\\
2938: n_c & 1.000&1.000\\
2939: n_a & 1.000 & 1.000\\
2940: \hline
2941: -\ln L_* & 793.4 & 3014.6\\
2942: \hline
2943: \end{tabular}
2944: \caption{Best-fit parameter values for the
2945: lower band of models A and B. An asterisk
2946: indicates that the corresponding parameter has
2947: reached the limit of our top-hat prior in the
2948: MC run, $H_0$ is again in
2949: $\rm{km}~\rm{sec}^{-1}~\rm{Mpc}^{-1}$ and $L_*$
2950: is the best-fit likelihood. Note that, due to
2951: the large value of $\Omega_\Lambda
2952: =0.91$ and the small value of $\Omega_m$ in the
2953: best fit of model B (lower band),
2954: the approximation used in deriving
2955: Eq.~(\ref{Cobe}) is no longer valid.
2956: \label{tab:bestfitbasefirstband}}
2957: \end{table}
2958:
2959: \begin{figure}[tb]
2960: \centering
2961: \includegraphics[width=\linewidth]
2962: {BFpqMOD4FST.peccei.ps}
2963: \caption{Best-fit power spectra for the lower
2964: band of model A. The base parameters are
2965: according to
2966: Table~\ref{tab:bestfitbasefirstband}, and
2967: the notation is as in Fig.~\ref{fig:BFpqMOD4}.}
2968: \label{fig:BFpqMOD4FST}
2969: \end{figure}
2970:
2971: \begin{figure}[tb]
2972: \centering
2973: \includegraphics[width=\linewidth]
2974: {BFpqMOD5FST.peccei.ps}
2975: \caption{Best-fit power spectra for the lower
2976: band of model B. The base parameters are
2977: according to
2978: Table~\ref{tab:bestfitbasefirstband}, and
2979: the notation is as in Fig.~\ref{fig:BFpqMOD4}.}
2980: \label{fig:BFpqMOD5FST}
2981: \end{figure}
2982:
2983: \par
2984: Fig.~\ref{fig:1dcorr} shows the 1D marginalized
2985: constraints on the dimensionless correlator
2986: $\cos\Delta$ and the entropy-to-adiabatic ratio
2987: $B$, while Fig.~\ref{fig:2dcomparecorr} gives
2988: the 2D joint constraints for these two
2989: parameters. In model A (upper band), the
2990: dominance of the inflaton and the low
2991: amplitudes of the correlator modes result in a
2992: value of the effective correlation $\cos\Delta$
2993: roughly centered around zero. As discussed
2994: above, the entropy contributions play a modest
2995: role in model A (upper band), and
2996: correspondingly we obtain the upper limit
2997: $B\lsim 0.5$ (at $95\%$ c.l., 1 tail).
2998: In model B (upper band), the large amplitude of
2999: the curvaton correlator determines a positive
3000: total correlation $\cos\Delta\approx 0.5$.
3001: Together, the curvaton and axion isocurvature
3002: amplitudes constitute a significant fraction of
3003: the total amplitude, so that the
3004: entropy-to-adiabatic ratio is much larger than
3005: zero ($B\approx 0.75$). In summary, model B
3006: (upper band) shows a complex superposition of
3007: modes with a subtly balanced contribution of
3008: adiabatic and isocurvature components.
3009:
3010: \par
3011: We have also performed a MC analysis for the
3012: lower green/lightly shaded band in
3013: Fig.~\ref{fig:model-4} of model A. As expected
3014: from the arguments given above, the quality of
3015: the best fit is poor ($-\ln L_*=793.4$) as one
3016: can see from
3017: Table~\ref{tab:bestfitbasefirstband} and
3018: Fig.~\ref{fig:BFpqMOD4FST}, because
3019: the small value of $\phi_*$ for this band gives
3020: rise to a large axion contribution according
3021: to Eqs.~(\ref{Aa}), (\ref{AP}) and
3022: (\ref{Sica}). Furthermore, the curvaton
3023: amplitude $A_c$ turns out to be much larger
3024: than the amplitude $A_i$ from the inflaton.
3025: This can be understood by observing that, in
3026: contrast to $A_i$, the amplitude $A_c$ from the
3027: curvaton can be large and still give a small
3028: positive or even a negative contribution to the
3029: total power in the SW plateau because the
3030: curvaton correlation is positive and thus
3031: subtracts power from large scales. This is
3032: important since, in this case, the axion
3033: contribution to the plateau is very large
3034: leaving little room for other contributions,
3035: while, at smaller scales, the axion
3036: isocurvature spectrum becomes negligible and
3037: the total temperature spectrum must necessarily
3038: be dominated by either the curvaton or the
3039: inflaton adiabatic contribution. So one of the
3040: two amplitudes $A_i$ or $A_c$ has to be large
3041: enough and, as we saw, this can happen only for
3042: $A_c$. Therefore, at smaller scales, the
3043: spectrum is dominated by the curvaton adiabatic
3044: part. The best-fit inflaton amplitude $A_i$, on
3045: the other hand, corresponds to the lower limit
3046: of our parameter space as one can see from
3047: Table~\ref{tab:bestfitbasefirstband}. The
3048: reason is that the inflaton contribution to the
3049: power spectra is always positive with an
3050: unsuppressed (by factors of $f$ or
3051: $(\Omega_{\rm LSP}+\Omega_B)/\Omega_m$)
3052: adiabatic part. Thus the inflaton amplitude
3053: $A_i$ must be very small in order not to
3054: overpredict the large-scale power. Furthermore,
3055: the TE spectrum has a very pronounced
3056: reionization bump, which results from a rather
3057: early reionization epoch and the large axion
3058: isocurvature contribution.
3059:
3060: \par
3061: The lower band of model B is totally incapable
3062: of producing a spectrum in reasonable agreement
3063: with the data (see
3064: Table~\ref{tab:bestfitbasefirstband}). For
3065: values of $A_i$, $A_c$ corresponding to this
3066: band, the axion contribution is huge and gives
3067: temperature fluctuations at large angular
3068: scales which are a few orders of magnitude
3069: larger than what is observed (see
3070: Fig.~\ref{fig:BFpqMOD5FST}).
3071:
3072: \par
3073: Regarding the issue of non-Gaussianity, we note
3074: that our parameter $f$, defined in
3075: Eq.~(\ref{eq:f}), is $\gg 10^{-2}$. Therefore,
3076: non-Gaussianity from the curvaton partial
3077: curvature perturbation is well within the
3078: current bounds from WMAP \cite{wmapgauss} (see
3079: Ref.~\cite{curv3}). Actually, the values of $f$
3080: in our model are high enough to ensure that
3081: this statement remains true even in the limit
3082: of pure curvaton. Moreover, the
3083: non-Gaussianity of the isocurvature
3084: perturbation in the axions is also negligible.
3085: This is because the perturbation
3086: $\delta\theta=H_*/2\pi\phi_*$ acquired during
3087: inflation by the initial misalignment angle
3088: $\theta$ (see e.g. Ref.~\cite{dllr1}) is always
3089: much smaller than $\theta$. As a consequence,
3090: terms which are second order in this
3091: perturbation can be safely neglected (see
3092: Ref.~\cite{axionnongauss}). Finally, the
3093: non-Gaussian component from the inflaton is
3094: also negligibly small.
3095:
3096: \par
3097: We have performed two other MC runs with the
3098: same value for $\lambda=10^{-4}$ as in model A
3099: and $\kappa$ slightly larger or smaller, i.e.
3100: $\kappa=10^{-2}$ or $\kappa=10^{-3}$,
3101: recovering a behavior which is qualitatively
3102: similar to the behavior of model A. We have
3103: thus chosen to present our results for this
3104: particular value of $\kappa$ ($=3\times
3105: 10^{-3}$) as representative of the behavior of
3106: this class of models. We have also tried a
3107: slightly larger value of $\lambda$
3108: ($=3\times 10^{-4}$) for the same value of
3109: $\kappa=3\times 10^{-2}$ as in model B, and
3110: found a behavior qualitatively similar to
3111: model B.
3112:
3113: \subsubsection{Bayesian model comparison}
3114: \label{sec:modelcomp}
3115:
3116: \par
3117: So far we have been concerned with the
3118: problem of deriving parameter constraints from
3119: the data, given an underlying model for the
3120: generation of the primordial fluctuations.
3121: Models A and B (upper bands) actually both
3122: belong to a continuous class of models
3123: (belonging to our PQ model) which is
3124: parameterized by $\kappa$ and $\lambda$.
3125: However, since in this work we have chosen to
3126: fix the values of $\kappa$ and $\lambda$, we
3127: may as well consider models A and B (upper
3128: bands) as two discrete disconnected models.
3129: The question is then to compare these two
3130: models with the standard pure inflaton HZ model
3131: and decide which of these three models is most
3132: favored by data. It should be stressed that
3133: model comparison (or model testing) is a
3134: different issue than parameter extraction, and
3135: indeed it represents a further step in Bayesian
3136: inference. In fact, it can very well be that
3137: model testing arrives at a different conclusion
3138: than parameter estimation. Indeed, it can be
3139: that the estimated value of a parameter under
3140: a model $\mdl_1$ is far from the null value
3141: predicted by model $\mdl_2$, but $\mdl_1$ is
3142: disfavored against $\mdl_2$ by Bayesian model
3143: testing, a fact sometimes called ``Bartlett's
3144: paradox'' \cite{Lindley}. This is exactly the
3145: case for $A_c$ in model B (upper band) compared
3146: to the pure inflaton model with flat spectrum
3147: of perturbations, as we will show below.
3148:
3149: \par
3150: Sampling statistics (i.e. the frequentist
3151: approach to parameter estimation) uses the
3152: notion of the goodness-of-fit test to
3153: assess the viability of a model without the
3154: need of specifying an alternative hypothesis.
3155: This usually reduces to the $\chi^2$ statistics
3156: for the observed data, presuming the model
3157: under consideration, $\mdl$, is true.
3158: $\mdl$ is then rejected if the value of the
3159: goodness-of-fit falls above a certain
3160: threshold in the tail of the distribution. If
3161: we take this criterion at face value and use
3162: the $\chi^2$ statistics for the WMAP data, the
3163: best-fit pure inflaton $\Lambda$CDM model with
3164: $n_i\neq 1$ having $\chi^2=1431$ for $\nu=1342$
3165: degrees of freedom (see last paper in
3166: Ref.~\cite{wmap1}) would be rejected with a
3167: type I error (i.e. probability of falsely
3168: rejecting a true model) of about $5\%$. Notice
3169: that this does not mean, as sometimes stated,
3170: that ``the probability of the model is $5\%$''.
3171: We will see below that Bayesian model
3172: comparison is more informative and can give
3173: useful guidance for model building.
3174:
3175: \par
3176: It is clear that models with a very poor best
3177: fit can be discarded simply by inspection. An
3178: extreme example is certainly the lower band of
3179: model B presented above. The goodness-of-fit
3180: for the lower band of model A is also
3181: very poor, even though this is not readily
3182: distinguishable from Fig.~\ref{fig:BFpqMOD4FST}
3183: due to the logarithmic scale. In fact, the
3184: best-fit point has $-\ln L_*=793.4$, compared
3185: to $-\ln L_*=721.2$ for the upper band of this
3186: model. Again, we can dismiss the lower band
3187: without further analysis. However, we need a
3188: more powerful tool if we are to decide, on the
3189: basis of the available data, between the
3190: viability of our PQ model as opposed to the
3191: pure inflaton HZ model. Bayesian inference
3192: offers a natural tool for model comparison in
3193: the form of the evidence in favor of the model
3194: (see Appendix for details and precise
3195: definitions).
3196:
3197: \begin{figure}[tb]
3198: \centering
3199: \includegraphics[width=\linewidth]
3200: {LaplaceApprox.ps}
3201: \caption{Illustration of the Laplace
3202: approximation to the (non-normalized) posterior
3203: used to compute the model evidence. For our
3204: base parameters, we plot the logarithm of the
3205: 1D marginalized posterior from the MC samples
3206: (crosses corresponding to the histograms in
3207: Fig.~\ref{fig:1dbase}) along with the
3208: corresponding Laplace approximation (solid
3209: smooth lines) of Eq.~(\ref{eq:laplace}) for the
3210: upper bands of model A (black) and model B
3211: (cyan/light gray). Clearly, the Laplace
3212: approximation is very good for model B, and of
3213: acceptable quality for model A. All the curves
3214: are normalized to zero at their peak value.
3215: In particular, $A_c$ is a rather non-Gaussian
3216: direction for model A (upper band), since the
3217: posterior from the MC samples only gives an
3218: upper limit on this parameter.}
3219: \label{fig:laplace}
3220: \end{figure}
3221:
3222: \par
3223: We thus compute the Bayes factor for models A
3224: and B (upper bands), comparing each of them to
3225: the scale-invariant (i.e. $n_i = 1.00$) pure
3226: inflaton model, which we call in the following
3227: the HZ inflation model. We approximate the
3228: integrals involved in the evaluation of the
3229: Bayes factor by using the Laplace approximation
3230: (see Eq.~(\ref{eq:laplace})). We must first
3231: check that a multi-dimensional Gaussian is a
3232: reasonable approximation to the actual shape of
3233: our (non-normalized) posterior pdf. This is
3234: shown in Fig.~\ref{fig:laplace}, where we plot
3235: the logarithm of the 1D marginalized
3236: non-normalized posterior from the MC samples
3237: along with the corresponding
3238: Laplace approximation. For model B (upper band),
3239: the Gaussian approximation is quite accurate
3240: along all directions, while for model
3241: A (upper band) we notice that $A_c$ is a rather
3242: non-Gaussian direction, which is not surprising
3243: since this model only gives upper bounds on
3244: the amplitude $A_c$. However, inspection of the
3245: figure does seem to support the view that it is
3246: reasonably accurate to use the Laplace
3247: approximation to compute the model evidence.
3248: Another qualitative criterion is the similarity
3249: of the marginalized 1D posterior and the mean
3250: posterior. If the posterior is a
3251: multi-dimensional Gaussian, then the two curves
3252: coincide (see e.g. Ref.~\cite{lewis}).
3253: From Fig.~\ref{fig:1dbase}, we can indeed
3254: confirm that the two curves are very similar,
3255: again strengthening the conclusion that the
3256: Laplace approximation is legitimate in our
3257: case, more so for model B.
3258:
3259: \par
3260: We first compare the upper band of model B
3261: ($\mdl_1$ in the notation used in the
3262: Appendix) against the HZ inflationary case
3263: ($\mdl_2$), and compute the
3264: Bayes factor $B_{12}$ in favor of model B
3265: (upper band). The term $\mathcal{L}_{12}$ (see
3266: Eq.~(\ref{eq:L12})) is just the difference of
3267: the best-fit log-likelihoods yielding
3268: $\mathcal{L}_{12}=-10.9$, and clearly disfavors
3269: model B (upper band), whose fit is worse. The
3270: term $\mathcal{C}_{12}$ (see Eq.~(\ref{eq:C12}))
3271: describes the ratio of the occupied volumes in
3272: parameter space by the posterior pdf's of the
3273: two models times a factor taking into account
3274: the different dimensionality of the two
3275: parameter spaces and is found to be
3276: $\mathcal{C}_{12}=3.0$. Further considerations
3277: are needed to evaluate $\mathcal{F}_{12}$ (see
3278: Eq.~(\ref{eq:F12})), the term which reflects
3279: the prior available volume of parameter space
3280: under each model. In the present case, we do
3281: not need to specify the top-hat range of the
3282: priors on the parameters which are common to
3283: both model B (upper band) and the HZ inflation
3284: model. This is because, whatever prior belief
3285: we hold about the possible range for these
3286: parameters, it will be the same for both
3287: models. The parameters common to both models
3288: are $\omega_B$, $\omega_{\rm CDM}$, $h$, $z_r$
3289: and $A_i$. In the PQ model, the prior range for
3290: the CDM abundance actually applies to the axion
3291: component of CDM only, but the difference,
3292: which is caused by the presence in the universe
3293: of the LSPs, is very small and insignificant
3294: for our considerations here. Also, we could
3295: actually assign to the inflaton amplitude $A_i$
3296: two different prior ranges for the PQ and the
3297: HZ inflation models if we had any reason to
3298: believe that the ranges can be significantly
3299: different in the two cases. Here we take the
3300: view that $A_i$ is essentially a free parameter
3301: in both models, and thus whatever prior range
3302: we assign to it will cancel out from
3303: $\mathcal{F}_{12}$. As a consequence, the only
3304: prior range which does not cancel out from
3305: $\mathcal{F}_{12}$ is the one for the extra
3306: parameter of our PQ model, i.e. $A_c$. So we
3307: have
3308: \begin{equation}
3309: \mathcal{F}_{12}=\ln\frac{1}{\Delta A_c},
3310: \end{equation}
3311: where $\Delta A_c$ is the top-hat prior range
3312: of $A_c$ for the PQ model which is allowed in
3313: model B (upper band). The bounds of the prior
3314: on $A_c$ correspond to the limits on $A_c$ in
3315: the upper band of model B. The lower limit is
3316: of no importance, since we can simply set it
3317: equal to zero. The upper limit is achieved on
3318: the boundary of the upper band, which gives
3319: $A_c^{\rm{max}}\approx 0.1$. From these
3320: considerations and writing $A_c$ in units of
3321: $10^{-5}$ (which are the same units used in
3322: the covariance matrix), we have $\Delta A_c
3323: \approx 10^4$, and thus $\mathcal{F}_{12}=
3324: -9.2$. Putting everything together, we find
3325: $\ln B_{12}=-17.1$ and thus the Bayes factor
3326: disfavors model B (upper band) against the HZ
3327: inflation model with odds of about $10^7$
3328: against $1$. Notice that the
3329: reason for such high odds comes partly from the
3330: worse quality of fit of model B (upper band),
3331: and partly from an ``Occam's razor'' penalty of
3332: the PQ model, because it introduces a new scale
3333: in the problem (the curvaton amplitude $A_c$)
3334: which has a very wide prior range. We comment
3335: further on this aspect below.
3336:
3337: \begin{table}
3338: \setlength{\extrarowheight}{5pt}\centering
3339: \begin{tabular}{|>{$}l<{$}|>{$}c<{$}|>{$}c<{$}|
3340: }
3341: \hline
3342: & {\text{model A}} &
3343: {\text{model B}}
3344: \\
3345: & {\text{(upper band)}} &
3346: {\text{(upper band)}}
3347: \\
3348: & {\text{versus HZ}} & {\text{versus HZ}}
3349: \\
3350: \hline \hline
3351: \mathcal{L}_{12}
3352: & 0.5 & -10.9
3353: \\
3354: \mathcal{C}_{12}
3355: & 4.8 & 3.0
3356: \\
3357: \mathcal{F}_{12}
3358: & -9.2& -9.2
3359: \\
3360: \hline
3361: \ln B_{12}
3362: & -3.9& -17.1
3363: \\
3364: \text{posterior odds}
3365: & 1:50& 1:10^7
3366: \\
3367: \hline
3368: \end{tabular}
3369: \caption{Results of the Bayes factors analysis
3370: for comparing the PQ models A and B (upper
3371: bands) to the HZ inflation model. When
3372: considering CMBR data only, the odds are
3373: in favor of the HZ model against the PQ models.
3374: In particular, model B (upper band) is very
3375: strongly disfavored.
3376: \label{tab:evidence}}
3377: \end{table}
3378:
3379: \par
3380: We now compute the Bayes factor for the upper
3381: band of model A ($\mdl_1$) against the HZ
3382: inflation model ($\mdl_2$). As mentioned above,
3383: in this case we expect the Laplace
3384: approximation to be less accurate, because
3385: $A_c$ is now a non-Gaussian direction. The term
3386: $\mathcal{L}_{12}$ is now slightly in favor of
3387: model A (upper band), since its fit is
3388: marginally better than the one of the HZ
3389: inflation model, giving $\mathcal{L}_{12}=0.5$.
3390: From the covariance matrices of the two models,
3391: we obtain $\mathcal{C}_{12}=4.8$, while the
3392: same reasoning as above gives again a very
3393: small value for $\mathcal{F}_{12}$ as a
3394: consequence of the large allowed prior range of
3395: $A_c$, i.e. $\mathcal{F}_{12}=-9.2$. This last
3396: term again heavily penalizes the PQ model,
3397: resulting in $\ln B_{12}=-3.9$, or odds of
3398: $50:1$ against the PQ model A (upper band). The
3399: Laplace approximation is however likely to
3400: underestimate the actual volume occupied by the
3401: likelihood function in parameter space, due to
3402: the fact that $A_c$ is a rather non-Gaussian
3403: direction for model A. As a consequence,
3404: the above odds can be interpreted as an upper
3405: limit for the evidence against model A. We
3406: summarize these results on the PQ models A and
3407: B (upper bands) in Table~\ref{tab:evidence}.
3408:
3409: \par
3410: One must be very careful when interpreting the
3411: above results for the evidence. In fact, the HZ
3412: inflationary model, which we used for the
3413: comparison, is a natural benchmark model as far
3414: as the CMBR data are concerned. However, one
3415: must keep in mind that the PQ model which we
3416: describe in this work addresses many
3417: fundamental issues which lie outside the scope
3418: of the phenomenological HZ inflation model,
3419: such as the strong CP and $\mu$ problems of
3420: MSSM, the generation of the observed BAU and
3421: the nature of the CDM in the universe. Our
3422: approach was to consider fundamental (SUSY GUT)
3423: models of particle physics within the
3424: cosmological framework by merging together
3425: requirements and constraints both from the
3426: particle physics and the cosmology side. In
3427: general, it is clear that any viable particle
3428: physics model has many free parameters about
3429: which little or no experimental information is
3430: available at the moment, such as the parameters
3431: of the electroweak Higgs sector, the sparticle
3432: mass spectrum or the boundary conditions from
3433: supergravity which also determine the radiative
3434: electroweak symmetry breaking. Nevertheless, we
3435: do have some indication about their order of
3436: magnitude by applying criteria of simplicity,
3437: naturalness and elegance to our fundamental
3438: theories. At this stage, the exquisite
3439: cosmological data nowadays at our disposal can
3440: provide significant constraints on the
3441: parameter space of the model. Our evidence
3442: calculation, in fact, takes into account only a
3443: part of our knowledge (i.e. the CMBR data)
3444: neglecting all the other issues which are
3445: addressed at a fundamental level by our model.
3446: It seems fair to say that, on the whole (i.e.
3447: considering both cosmology and particle
3448: physics), the PQ model presented in this work
3449: aims at a broader explanatory power than a
3450: phenomenological inflation model. We thus
3451: conclude from our analysis that the present
3452: CMBR data can strongly disfavor certain regions
3453: of parameter space, as it is the case for model
3454: B. In this case, we found robust evidence that
3455: a mixture of curvaton and inflaton
3456: contributions to the cosmological perturbations
3457: is in disagreement with the present CMBR data.
3458: On the other hand, the odds against model A
3459: should be regarded while keeping in mind the
3460: above considerations. In conclusion, it seems
3461: to us that, at the present stage, we cannot
3462: reject the possibility of a subdominant curvaton
3463: contribution to predominantly inflaton-dominated
3464: adiabatic perturbations.
3465:
3466: \par
3467: On a more phenomenological level, our treatment
3468: of the evidence highlights a generic feature of
3469: any model which introduces a second (or
3470: several) scale-free parameter(s) to describe an
3471: extra non-adiabatic component, namely that
3472: Occam's razor of Bayesian model comparison will
3473: always penalize such a model with respect to
3474: the pure inflaton single-field HZ inflationary
3475: model by assigning to it a very small
3476: $\mathcal{F}_{12}$ term. This reflects the fact
3477: that the extra amplitude parameter
3478: (describing an isocurvature contribution) can
3479: {\it a priori} assume any value at all, and
3480: therefore there is no hard-wired justification
3481: why its value should be smaller than $10^{-5}$,
3482: or indeed close to but different from zero. So,
3483: from a Bayesian point of view, it is certainly
3484: unjustified to increase the model's complexity
3485: to achieve only a minuscule gain (if any at
3486: all) on the quality of fit. Traditionally, most
3487: attention has been devoted to the maximum
3488: likelihood value as the criterion to judge
3489: whether a certain new parameter is useful or
3490: not. But, from the point of view of model
3491: building and Bayesian model testing,
3492: restricting the prior volume of parameter space
3493: could end up to be as useful, for an
3494: inflationary model which would predict the
3495: value of $A_i$ to be in the ball-park of
3496: $10^{-5}$ would have very favorable posterior
3497: odds against a model in which $A_i$ is
3498: essentially a free parameter.
3499:
3500: \par
3501: Finally, the example of model B (upper band)
3502: above strikingly illustrates that Bayesian
3503: ``credible intervals'' cannot be interpreted as
3504: ``significance levels'', for, just by looking
3505: at the constraints on the amplitude $A_c$ for
3506: the upper band of model B, one could have
3507: (erroneously) deduced a $6\sigma$ detection of
3508: a non-zero curvaton amplitude. Clearly, from
3509: the poor quality of the best fit
3510: ($-\ln L_*=732.6$), it would have been
3511: immediately obvious, without the need of
3512: computing the Bayes factor, that this
3513: particular model could not be favored by data.
3514: However, the conceptual point remains: the
3515: question whether a certain parameter value
3516: differs from the null reference value cannot be
3517: answered by looking at the confidence intervals
3518: (c.i.), but must be tackled by proper model
3519: comparison procedures. While this point is
3520: certainly clear to data analysts working in the
3521: field, it seems appropriate to stress it once
3522: more in view of a widespread misinterpretation
3523: of this concept.
3524:
3525: \section{Conclusions}
3526: \label{sec:conclusion}
3527:
3528: \par
3529: We considered a simple and concrete SUSY GUT
3530: model which automatically and naturally solves
3531: the strong CP and $\mu$ problems of MSSM via a
3532: PQ and a continuous R symmetry. This model
3533: also leads to the standard SUSY realization of
3534: hybrid inflation. The PQ field of this model,
3535: which corresponds to a flat direction in field
3536: space lifted by non-renormalizable interactions,
3537: can act as curvaton contributing together with
3538: the inflaton to the total curvature
3539: perturbation in the universe. In contrast to
3540: the standard curvaton hypothesis, we did not
3541: suppress the contribution from the inflaton.
3542:
3543: \par
3544: The CDM in the universe predicted by this model
3545: consists predominantly of axions which are
3546: produced at the QCD phase transition. It also
3547: contains LSPs originating from the gravitinos
3548: which were thermally produced at reheating and
3549: decayed well after BBN. For simplicity, we
3550: assumed that
3551: there are no thermally produced LSPs in the
3552: model. The baryons, which are generated via a
3553: primordial leptogenesis occurring at reheating,
3554: as well as the LSPs inherit the partial
3555: curvature perturbation of the inflaton. This is
3556: different from the total curvature perturbation,
3557: which receives a contribution from the curvaton
3558: too. Therefore, the baryons and LSPs carry also
3559: an isocurvature perturbation, whose correlation
3560: with the total curvature perturbation is mixed.
3561: The axions, as usual, contribute only to the
3562: isocurvature perturbation. The resulting total
3563: isocurvature perturbation has mixed
3564: correlation with the adiabatic one.
3565:
3566: \par
3567: Most of the parameters of the model but two
3568: were chosen by using criteria of
3569: naturalness and simplicity. We considered two
3570: representative cases for the two remaining
3571: parameters, $\kappa$ and $\lambda$, and
3572: compared the predictions with the first-year
3573: WMAP observations and other CMBR measurements.
3574: We found that, in one of the two cases (model
3575: B), the curvaton and axions contributions to
3576: the CMBR power spectra are important, and that
3577: this leads to a significant disagreement with
3578: the data. In fact, Bayesian model comparison
3579: disfavors this case as compared to the
3580: scale-invariant pure inflaton model with odds
3581: of $10^7$ against 1. The other possibility
3582: which we studied (model A) predicts a
3583: predominantly adiabatic power spectrum from the
3584: inflaton, where the curvaton and axions
3585: contributions are subdominant. We derived upper
3586: bounds for the amplitude of the partial
3587: curvature perturbation from the curvaton in
3588: this case ($A_c\lsim 43.2\times 10^{-5}$ at
3589: 95\% c.l.), and noticed that the interplay of
3590: the non-inflaton contributions results in a
3591: later reionization redshift
3592: ($z_r=13.3^{+8.1}_{-9.3}$ at 68\% c.l.). Even
3593: though the best-fit likelihood for this case
3594: ($-\ln L_* = 721.2$) is better than in the pure
3595: inflaton HZ case, evaluation of the evidence
3596: under the Laplace approximation gives odds of
3597: about 50 to 1 against model A compared to the
3598: pure inflaton HZ case. These odds must be
3599: interpreted with caution, since they do not
3600: take into account the fact that the PQ model
3601: aims at a more fundamental explanatory power
3602: and that it addresses many particle physics
3603: issues which are outside the scope of
3604: inflationary models.
3605:
3606: \par
3607: In summary, we have shown that certain regions
3608: of the parameter space ($\kappa\approx 3\times
3609: 10^{-2}$ and $10^{-4}\lsim\lambda\lsim 3\times
3610: 10^{-4}$) can be excluded on the grounds of
3611: present-day CMBR observations. Quantitative
3612: bounds on the allowed values of $\kappa$ and
3613: $\lambda$ could be derived by treating them as
3614: free parameters in the MC analysis, an
3615: exploration left for future work. Our approach
3616: -- embedding particle physics model building in
3617: the cosmological framework -- pursued the issue
3618: of merging together fundamental theories and
3619: cosmological constraints in a realistic and
3620: viable model for the generation of the
3621: cosmological perturbations. It is encouraging
3622: that modern cosmological observations are now
3623: informative enough as to give useful and robust
3624: guidance along this path.
3625:
3626: \section*{ACKNOWLEDGMENTS}
3627: We thank K. Dimopoulos, R. Durrer, S. Leach, D.
3628: Lyth, Ch. Ringeval, M. Sakellariadou and D.
3629: Wands for useful discussions. We are
3630: particularly grateful to K. Dimopoulos and D.
3631: Lyth for communicating to us their results on
3632: the randomization regime for the curvaton field
3633: prior to publication. This work was supported
3634: by the European Union under RTN contracts
3635: HPRN-CT-2000-00148 and HPRN-CT-2000-00152 and
3636: was performed on the Myri cluster owned and
3637: operated by the University of Geneva. R.T. was
3638: partially supported by the European Network
3639: CMBNet and the Swiss National Science
3640: Foundation.
3641:
3642: \appendix
3643:
3644: \renewcommand{\theequation}
3645: {\thesection\arabic{equation}}
3646: \setcounter{equation}{0}
3647: \renewcommand{\thefigure}
3648: {\thesection\arabic{figure}}
3649: \setcounter{figure}{0}
3650:
3651: \newcommand{\gsim}{\,\raise 0.4ex\hbox{$>$}
3652: \kern -0.7em\lower 0.62ex\hbox{$\sim$}\,}
3653:
3654: \section{Bayesian inference: a primer}
3655:
3656: \subsection{Bayesian parameter estimation}
3657: \label{appendix:params}
3658:
3659: \par
3660: Bayesian inference is based on Bayes' theorem,
3661: which is nothing more than rewriting the
3662: definition of conditional probability:
3663: \begin{equation}
3664: \pdf(A\vert B)=\frac{\pdf(B\vert A)\pdf(A)}
3665: {\pdf(B)}
3666: \quad
3667: \text{(Bayes' theorem).}
3668: \end{equation}
3669: In order to clarify the meaning of this
3670: relation, let us write $\params$ (a vector of
3671: $d$ parameters under a model $\mdl$) for
3672: $A$ and $\data$ (the data at hand) for $B$,
3673: obtaining
3674: \begin{eqnarray}
3675: \pdf(\params\vert\data,\mdl)&=&
3676: \frac{\like(\data\vert\params,\mdl)
3677: \pi(\params,\mdl)}{\int_{\Omega}
3678: \like(\data\vert\params,\mdl)
3679: \pi(\params,\mdl)\dr\params}
3680: \nonumber\\
3681: &=&\frac{\like(\data\vert\params,\mdl)
3682: \pi(\params,\mdl)}{\pdf(\data\vert\mdl)},
3683: \label{eq:Bayes_Theorem}
3684: \end{eqnarray}
3685: where $\Omega$ designates the parameter space
3686: (of dimensionality $d$) under model $\mdl$.
3687: This equation relates the {\it posterior
3688: probability} $\pdf(\params\vert\data,\mdl)$ for
3689: the parameters $\params$ of the model $\mdl$
3690: given the data $\data$ to the {\it likelihood
3691: function} $\like(\data\vert\params,\mdl)$ if
3692: the {\it prior probability distribution
3693: function} $\pi(\params,\mdl)$ for the
3694: parameters under the model is known. The latter
3695: is called ``prior'' for short. The quantity in
3696: the denominator is independent of $\params$ and
3697: is called the {\it evidence} of the data for
3698: the model $\mdl$ \cite{MKbook}. The evidence
3699: is the central quantity for model comparison,
3700: as we explain below, but, in the context of
3701: parameter estimation within a model, it is just
3702: an overall multiplicative constant which does
3703: not matter. In short,
3704: \begin{equation}
3705: \text{posterior}=
3706: \frac{\text{likelihood}\times\text{prior}}
3707: {\text{evidence}} .
3708: \end{equation}
3709:
3710: \par
3711: The prior distribution contains all the
3712: knowledge about the parameters
3713: before observing the data, i.e. our physical
3714: understanding of the model, our insight into
3715: the experimental setup and its performance,
3716: and in short all our prior scientific
3717: experience. This information is then updated
3718: via Bayes' theorem to the posterior
3719: distribution by multiplying the prior with the
3720: likelihood function which contains the
3721: information coming from the data. The posterior
3722: probability is the base for inference about
3723: $\params$. The most probable value for the
3724: parameters is the one for which the posterior
3725: probability is largest.
3726:
3727: \par
3728: Bayes' postulate states that, in the absence of
3729: other arguments, the prior probability should
3730: be assumed to be equal for all values of the
3731: parameters over a certain range
3732: ($\params_{\text{min}}\leq\params\leq
3733: \params_{\text{max}}$). This is called a ``flat
3734: prior'', i.e.
3735: \begin{equation}
3736: \pi(\params,\mdl)=\left[ H(\params-
3737: \params_{\text{min}})H(\params_{\text{max}}-
3738: \params)\right]\prod_{j=1}^{d}\frac{1}
3739: {\Delta\theta_j},
3740: \label{eq:flat_prior}
3741: \end{equation}
3742: where $H$ is the Heaviside step function and
3743: $\Delta\theta_j\equiv\theta_{\text{max},j}-
3744: \theta_{\text{min},j}>0$, $\forall \;j$.
3745: Clearly, a flat prior on $\params$ does not
3746: correspond to a flat prior on some other set
3747: $\boldsymbol{\alpha}(\params)$ obtained via a
3748: non-linear transformation, since the two prior
3749: distributions are related via
3750: \begin{equation}
3751: \pi(\params,\mdl)=
3752: \pi(\boldsymbol{\alpha},\mdl)
3753: \frac{\dr\boldsymbol{\alpha}(\params)}
3754: {\dr\params}.
3755: \label{eq:prior}
3756: \end{equation}
3757: A recurrent criticism is that the final
3758: inference depends on the prior which one
3759: chooses to use. However, in a situation in
3760: which the data exhibit a clear preference for a
3761: certain value for a parameter, the posterior is
3762: effectively dominated by the likelihood, and
3763: the choice of prior will not matter much. This
3764: is currently the case, as far as the CMBR is
3765: concerned, for high ``signal to noise''
3766: parameters such as the baryon density.
3767: Constraints on other parameters such as the
3768: curvaton amplitude $A_c$ in model A (upper
3769: band) of our PQ scenario are likely to
3770: depend slightly on the details of the chosen
3771: prior distribution. In other words, constraints
3772: on parameters which are not clearly determined
3773: will suffer from a certain degree of
3774: subjectivity, depending on what prior
3775: $\pi(\params,\mdl)$ we choose on the right
3776: hand side of Eq.~(\ref{eq:Bayes_Theorem}). This
3777: fact should be interpreted as a warning,
3778: telling us that the data are not powerful
3779: enough to clearly single out the parameter
3780: under consideration.
3781:
3782: \subsection{Bayes factors}
3783: \label{appendix:factors}
3784:
3785: \par
3786: Let us consider two competing models $\mdl_1$
3787: and $\mdl_2$ and ask what is the posterior
3788: probability of each model given the data
3789: $\data$. By Bayes' theorem we have
3790: \begin{equation}
3791: \pdf(\mdl_i\vert\data)\propto
3792: \pdf(\data\vert\mdl_i)\pi(\mdl_i)~~(i=1,2),
3793: \end{equation}
3794: where $\pdf(\data\vert\mdl_i)$ is the
3795: evidence of the data under model $\mdl_i$ and
3796: $\pi(\mdl_i)$ is the prior probability of the
3797: $i$th model before we see the data. The ratio
3798: of the posterior odds for the two competing
3799: models is called {\it Bayes factor}
3800: \cite{Kass}:
3801: \begin{equation}
3802: B_{12}\equiv\frac{\pdf(\mdl_1\vert\data)}
3803: {\pdf(\mdl_2\vert\data)}.
3804: \label{eq:b12}
3805: \end{equation}
3806: Usually, we do not hold any prior beliefs about
3807: the two models and therefore
3808: $\pi(\mdl_1)=\pi(\mdl_2)=1/2$, so that the
3809: Bayes factor reduces to the ratio of the
3810: evidences. The Bayes factor can be interpreted
3811: as an automatic Occam's razor, which disfavors
3812: complex models involving many parameters (see
3813: Ref.~\cite{MKbook} for details) as we discuss
3814: below and demonstrate in the text.
3815:
3816: \par
3817: The evidence in favor of $\mdl$ can be
3818: evaluated by performing the integral
3819: \begin{eqnarray}
3820: \pdf(\data\vert\mdl)&=&\int_{\Omega}
3821: \like(\data\vert\params,\mdl)
3822: \pi(\params,\mdl)\dr\params
3823: \nonumber \\
3824: &=&\int_{\Omega}
3825: \bar{\pdf}(\params\vert\data,\mdl)\dr\params,
3826: \end{eqnarray}
3827: where $\bar{\pdf}(\params\vert\data,\mdl)$
3828: designates the non-normalized posterior
3829: probability (i.e. the numerator in the right
3830: hand side of Eq.~(\ref{eq:Bayes_Theorem})).
3831: Computing the above integral from
3832: the MC samples is difficult, since there will
3833: be very few or no samples in the tails of the
3834: likelihood. There are however a number of
3835: approximate methods which can be applied
3836: \cite{DiCiccio}. Most of them rely on the fact
3837: that, for a large number of data points, the
3838: likelihood function will tend to be a
3839: multi-dimensional Gaussian distribution. One
3840: simple approximation is then to expand the
3841: logarithm of the non-normalized posterior to
3842: second order around its peak, which (for flat
3843: prior) occurs at the best-fit parameter choice
3844: $\params_*$, where the likelihood is maximized.
3845: We obtain
3846: \begin{equation}
3847: \ln\frac{\bar{\pdf}(\params\vert\data,\mdl)}
3848: {\bar{\pdf}(\params_*\vert\data,\mdl)}\approx
3849: -\frac{1}{2}(\params-\params_*)^T
3850: {\bf C}^{-1}(\params-\params_*),
3851: \label{eq:laplace}
3852: \end{equation}
3853: where ${\bf C}$ is the covariance matrix
3854: of the model $\mdl$ evaluated at the best-fit
3855: point, which can be estimated from the MC
3856: samples. This is called Laplace approximation
3857: and can be expected to give sensible results if
3858: the non-normalized posterior is reasonably well
3859: described by the multi-dimensional Gaussian
3860: Eq.~(\ref{eq:laplace}). It is then
3861: straightforward to evaluate the evidence by
3862: using the approximate form in
3863: Eq.~(\ref{eq:laplace}) for the non-normalized
3864: posterior, obtaining
3865: \begin{eqnarray}
3866: \pdf(\data\vert\mdl)&=&\int_{\Omega}
3867: \bar{\pdf}(\params\vert\data,\mdl)
3868: \dr\params
3869: \nonumber\\
3870: &\approx&(2\pi)^{\frac{d}{2}}
3871: \bar{\pdf}(\params_*\vert\data,\mdl)
3872: \sqrt{\det{\bf C}}.
3873: \end{eqnarray}
3874: For flat separable priors of the
3875: form in Eq.~(\ref{eq:flat_prior}) we can simply
3876: write, abbreviating ${\bf \Delta}\params
3877: \equiv\prod_{j=1}^{d}\Delta\theta_j$,
3878: \begin{equation}
3879: \bar{\pdf}(\params_*\vert\data,\mdl)=
3880: \like(\data\vert\params_*,\mdl)\frac{1}
3881: {{\bf \Delta}\params},
3882: \label{eq:laplace_2}
3883: \end{equation}
3884: an expression which is still approximately
3885: correct even if we used non-flat priors, and
3886: interpret $\Delta\theta_j$ as the
3887: typical width of the prior pdf (say the
3888: standard deviation along the direction of the
3889: $j$th parameter for a Gaussian distributed
3890: prior). For a Gaussian prior, it is easy to
3891: derive an exact expression analogous to
3892: Eq.~(\ref{eq:laplace_2}), but for simplicity
3893: we will stick to the above form.
3894:
3895: \par
3896: For the logarithm of the Bayes factor in the
3897: Laplace approximation, we finally obtain the
3898: following handy expression:
3899: \begin{equation}
3900: \ln B_{12}\approx
3901: \mathcal{L}_{12}+\mathcal{C}_{12}+
3902: \mathcal{F}_{12},
3903: \label{eq:B12}
3904: \end{equation}
3905: where we have defined
3906: \begin{eqnarray}
3907: \mathcal{L}_{12}&\equiv&\ln
3908: \frac{\like(\data\vert\params^{(1)}_*,\mdl_1)}
3909: {\like(\data\vert\params^{(2)}_*,\mdl_2)},
3910: \label{eq:L12}
3911: \\
3912: \mathcal{C}_{12}&\equiv&
3913: \frac{1}{2}\left(\ln \left[(2\pi)^{d^{(1)}-
3914: d^{(2)}}\right]+\ln\frac{\det{\bf C}^{(1)}}
3915: {\det{\bf C}^{(2)}}\right),
3916: \label{eq:C12}
3917: \\
3918: \mathcal{F}_{12}&\equiv&
3919: \ln\frac{{\bf \Delta}\params^{(2)}}
3920: {{\bf \Delta}\params^{(1)}}
3921: \label{eq:F12},
3922: \end{eqnarray}
3923: where quantities referring to the model
3924: $\mdl_i$ ($i=1,2$) are indicated by a
3925: superscript $(i)$. The term
3926: $\mathcal{L}_{12}$ is the logarithm of the
3927: ratio of the best-fit likelihoods. From a
3928: frequentist point of view, this quantity is
3929: approximately $\chi^2$ distributed, and thus
3930: it is common practice to apply to it a
3931: goodness-of-fit test to assess whether the
3932: extra parameters have produced a
3933: ``significant'' increase of the quality of
3934: fit. If the model $\mdl_1$ contains more
3935: parameters than the model $\mdl_2$, then
3936: $\mdl_1$ should show an improved fit to the
3937: data, i.e. we should have $\mathcal{L}_{12}>0$,
3938: unless the extra parameters are useless, in
3939: which case $\mathcal{L}_{12}=0$. In any case,
3940: a goodness-of-fit test alone does not say
3941: anything about the structure of the parameter
3942: space for the model under consideration, since
3943: it is limited to the maximum likelihood point.
3944: But Bayesian evidence does contain two further
3945: pieces of information, $\mathcal{C}_{12}$ and
3946: $\mathcal{F}_{12}$, which taken together are
3947: sometimes referred to as ``Occam's factor''.
3948: Here we prefer to consider these terms
3949: separately to help distinguishing their
3950: different origin. The term $\mathcal{C}_{12}$
3951: describes the structure of the posterior shape
3952: in the Gaussian approximation. Since the
3953: determinant is the product of the eigenvalues
3954: of the covariance matrix, which represent the
3955: standard deviations squared along the
3956: corresponding eigenvectors in the parameter
3957: space of the model, it follows that if $\mdl_1$
3958: has a narrower posterior than $\mdl_2$, then
3959: $\mathcal{C}_{12}<0$, thereby disfavoring
3960: $\mdl_1$. This apparent contradiction (how can
3961: a model with smaller errors display a smaller
3962: evidence?) is resolved when we take into
3963: account the term $\mathcal{F}_{12}$, which
3964: describes the prior available parameter space
3965: under each model. The sum of the terms
3966: $\mathcal{C}_{12}$ and $\mathcal{F}_{12}$ thus
3967: disfavors the model with the largest volume of
3968: ``wasted'' parameter space when the data
3969: arrive. A more complex model $\mdl_1$ -- having
3970: a large number of parameters and thus a large
3971: volume of prior accessible parameter space --
3972: will naturally fit the data better due to its
3973: flexibility, i.e. we will have
3974: $\mathcal{L}_{12}>0$, but it will be penalized
3975: for introducing extra dimensions in parameter
3976: space, i.e. the sum $\mathcal{C}_{12}+
3977: \mathcal{F}_{12}$ will be negative. In summary,
3978: the Bayes factor tends to select the model
3979: which exhibits an optimal trade-off between
3980: simplicity and quality of fit.
3981:
3982: \par
3983: It should be clear that the choice of priors
3984: plays an important role in Bayesian model
3985: comparison (testing) via its impact on the term
3986: $\mathcal{F}_{12}$. In particular, prior pdf's
3987: used in evaluating the Bayes factor must be
3988: proper, i.e. normalizable, so that we can
3989: impose the normalization condition
3990: \begin{equation}
3991: \int_{\Omega}\pi(\params,\mdl)\dr\params=1.
3992: \end{equation}
3993: Although generally a strong dependence on the
3994: choice of priors is regarded as suspicious, in
3995: this case, we should consider the role of
3996: $\mathcal{F}_{12}$ as a way to implement
3997: {\it a priori} model features into the Bayes
3998: factor, as we show in the text of the paper.
3999:
4000: \def\ijmp#1#2#3{{Int. Jour. Mod. Phys.}
4001: {\bf #1},~#3~(#2)}
4002: \def\plb#1#2#3{{Phys. Lett. B }{\bf #1},~#3~(#2)}
4003: \def\zpc#1#2#3{{Z. Phys. C }{\bf #1},~#3~(#2)}
4004: \def\prl#1#2#3{{Phys. Rev. Lett.}
4005: {\bf #1},~#3~(#2)}
4006: \def\rmp#1#2#3{{Rev. Mod. Phys.}
4007: {\bf #1},~#3~(#2)}
4008: \def\prep#1#2#3{{Phys. Rep. }{\bf #1},~#3~(#2)}
4009: \def\prd#1#2#3{{Phys. Rev. D }{\bf #1},~#3~(#2)}
4010: \def\npb#1#2#3{{Nucl. Phys. }{\bf B#1},~#3~(#2)}
4011: \def\npps#1#2#3{{Nucl. Phys. B (Proc. Sup.)}
4012: {\bf #1},~#3~(#2)}
4013: \def\mpl#1#2#3{{Mod. Phys. Lett.}
4014: {\bf #1},~#3~(#2)}
4015: \def\arnps#1#2#3{{Annu. Rev. Nucl. Part. Sci.}
4016: {\bf #1},~#3~(#2)}
4017: \def\sjnp#1#2#3{{Sov. J. Nucl. Phys.}
4018: {\bf #1},~#3~(#2)}
4019: \def\jetp#1#2#3{{JETP Lett. }{\bf #1},~#3~(#2)}
4020: \def\app#1#2#3{{Acta Phys. Polon.}
4021: {\bf #1},~#3~(#2)}
4022: \def\rnc#1#2#3{{Riv. Nuovo Cim.}
4023: {\bf #1},~#3~(#2)}
4024: \def\ap#1#2#3{{Ann. Phys. }{\bf #1},~#3~(#2)}
4025: \def\ptp#1#2#3{{Prog. Theor. Phys.}
4026: {\bf #1},~#3~(#2)}
4027: \def\apjl#1#2#3{{Astrophys. J. Lett.}
4028: {\bf #1},~#3~(#2)}
4029: \def\n#1#2#3{{Nature }{\bf #1},~#3~(#2)}
4030: \def\apj#1#2#3{{Astrophys. J.}
4031: {\bf #1},~#3~(#2)}
4032: \def\anj#1#2#3{{Astron. J. }{\bf #1},~#3~(#2)}
4033: \def\apjs#1#2#3{{Astrophys. J. Suppl.}
4034: {\bf #1},~#3~(#2)}
4035: \def\mnras#1#2#3{{Mon. Not. Roy. Astron. Soc.}
4036: {\bf #1},~#3~(#2)}
4037: \def\grg#1#2#3{{Gen. Rel. Grav.}
4038: {\bf #1},~#3~(#2)}
4039: \def\s#1#2#3{{Science }{\bf #1},~#3~(#2)}
4040: \def\baas#1#2#3{{Bull. Am. Astron. Soc.}
4041: {\bf #1},~#3~(#2)}
4042: \def\ibid#1#2#3{{\it ibid. }{\bf #1},~#3~(#2)}
4043: \def\cpc#1#2#3{{Comput. Phys. Commun.}
4044: {\bf #1},~#3~(#2)}
4045: \def\astp#1#2#3{{Astropart. Phys.}
4046: {\bf #1},~#3~(#2)}
4047: \def\epjc#1#2#3{{Eur. Phys. J. C}
4048: {\bf #1},~#3~(#2)}
4049: \def\nima#1#2#3{{Nucl. Instrum. Meth. A}
4050: {\bf #1},~#3~(#2)}
4051: \def\jhep#1#2#3{{J. High Energy Phys.}
4052: {\bf #1},~#3~(#2)}
4053: \def\lnp#1#2#3{{Lect. Notes Phys.}
4054: {\bf #1},~#3~(#2)}
4055: \def\appb#1#2#3{{Acta Phys. Polon. B}
4056: {\bf #1},~#3~(#2)}
4057: \def\fcp#1#2#3{{Fund. Cos. Phys.}
4058: {\bf #1},~#3~(#2)}
4059: \def\ptps#1#2#3{{Prog. Theor. Phys. Suppl.}
4060: {\bf #1},~#3~(#2)}
4061: \def\ss#1#2#3{{Statist. Sci.}
4062: {\bf #1},~#3~(#2)}
4063: \def\jasa#1#2#3{{J. Amer. Statist. Assoc.}
4064: {\bf #1},~#3~(#2)}
4065: \def\jcap#1#2#3{{J. Cosmol. Astropart. Phys.}
4066: {\bf #1},~#3~(#2)}
4067: \def\bm#1#2#3{{Biometrika}
4068: {\bf #1},~#3~(#2)}
4069: \def\ams#1#2#3{{Ann. Math. Stat.}
4070: {\bf #1},~#3~(#2)}
4071:
4072:
4073: \begin{thebibliography}{}
4074:
4075: \bibitem{guth} A. Guth, \prd{23}{1981}{347}.
4076: %%CITATION = PHRVA,D23,347;%%
4077:
4078: \bibitem{wmap1}
4079: C.L. Bennett {\it et al.}, \apjs{148}{2003}{1};
4080: %%CITATION = ASTRO-PH 0302207;%%
4081: G. Hinshaw {\it et al.}, \ibid{148}{2003}{135};
4082: %%CITATION = ASTRO-PH 0302217;%%
4083: A. Kogut {\it et al.}, \ibid{148}{2003}{161};
4084: %%CITATION = ASTRO-PH 0302213;%%
4085: D.N. Spergel {\it et al.}, \ibid{148}{2003}{175}.
4086: %%CITATION = ASTRO-PH 0302209;%%
4087:
4088: \bibitem{wmap2}
4089: H.V. Peiris {\it et al.}, \apjs{148}{2003}{213}.
4090: %%CITATION = ASTRO-PH 0302225;%%
4091:
4092: \bibitem{llbook}
4093: A.R. Liddle and D.H. Lyth, {\it Cosmological
4094: inflation and large-scale structure}
4095: (Cambridge University Press, Cambridge, U.K.,
4096: 2000).
4097:
4098: \bibitem{lectures} G. Lazarides,
4099: \lnp{592}{2002}{351} [hep- ph/0111328];
4100: %%CITATION = HEP-PH 0111328;%%
4101: hep-ph/0204294.
4102: %%CITATION = HEP-PH 0204294;%%
4103:
4104: \bibitem{iso1}
4105: C. Gordon and A. Lewis, \prd{67}{2003}{123513};
4106: %%CITATION = ASTRO-PH 0212248;%%
4107: P. Crotty, J. Garc{\'\i}a-Bellido, J. Lesgourgues
4108: and A. Riazuelo, \prl{91}{2003}{171301};
4109: %%CITATION = ASTRO-PH 0306286;%%
4110: M. Bucher, J. Dunkley, P.G. Ferreira, K. Moodley
4111: and C. Skordis, \ibid{93}{2004}{081301};
4112: %%CITATION = ASTRO-PH 0401417;%%
4113: K. Moodley, M. Bucher, J. Dunkley, P.G. Ferreira
4114: and C. Skordis, astro-ph/0407304.
4115: %%CITATION = ASTRO-PH 0407304;%%
4116:
4117: \bibitem{nongauss}
4118: N. Bartolo, E. Komatsu, S. Matarrese and
4119: A. Riotto, astro-ph/0406398.
4120: %%CITATION = ASTRO-PH 0406398;%%
4121:
4122: \bibitem{curv1}
4123: S. Mollerach, \prd{42}{1990}{313};
4124: %%CITATION = PHRVA,D42,313;%%
4125: A.D. Linde and V. Mukhanov,
4126: \ibid{56}{1997}{535}.
4127: %%CITATION = ASTRO-PH 9610219;%%
4128:
4129: \bibitem{curv2}
4130: D.H. Lyth and D. Wands, \plb{524}{2002}{5};
4131: %%CITATION = HEP-PH 0110002;%%
4132: T. Moroi and T. Takahashi, \ibid{522}{2001}{215};
4133: {\bf 539}, 303(E) (2002).
4134: %%CITATION = HEP-PH 0110096;%%
4135:
4136: \bibitem{curv3}
4137: D.H. Lyth, C. Ungarelli and D. Wands,
4138: \prd{67}{2003}{023503}.
4139: %%CITATION = ASTRO-PH 0208055;%%
4140:
4141: \bibitem{dllr1}
4142: K. Dimopoulos, G. Lazarides, D. Lyth and R.
4143: Ruiz de Austri, \jhep{05}{2003}{057}.
4144: %%CITATION = HEP-PH 0303154;%%
4145:
4146: \bibitem{dimo}
4147: K. Dimopoulos and D.H. Lyth, Phys. Rev. D
4148: {\bf 69}, 123509 (2004).
4149: %\prd{69}{2004}{123509}.
4150: %%CITATION = HEP-PH 0209180;%%
4151:
4152: \bibitem{rsym} G. Lazarides and Q. Shafi,
4153: Phys. Rev. D {\bf 58}, 071702 (1998).
4154: %\prd{58}{1998}{071702}.
4155: %%CITATION = HEP-PH 9803397;%%
4156:
4157: \bibitem{pq} R. Peccei and H. Quinn,
4158: Phys. Rev. Lett. {\bf 38}, 1440 (1977);
4159: %\prl{38}{1977}{1440};
4160: %%CITATION = PRLTA,38,1440;%%
4161: S. Weinberg, \ibid{40}{1978}{223};
4162: %%CITATION = PRLTA,40,223;%%
4163: F. Wilczek, \ibid{40}{1978}{279}.
4164: %%CITATION = PRLTA,40,279;%%
4165:
4166: \bibitem{lyth} E.J. Copeland, A.R. Liddle,
4167: D.H. Lyth, E.D. Stewart and D. Wands,
4168: \prd{49}{1994}{6410}.
4169: %%CITATION = ASTRO-PH 9401011;%%
4170:
4171: \bibitem{dss} G. Dvali, Q. Shafi and
4172: R. Schaefer, \prl{73}{1994}{1886}.
4173: %%CITATION = HEP-PH 9406319;%%
4174:
4175: \bibitem{hybrid} A.D. Linde, \plb{259}{1991}{38};
4176: %%CITATION = PHLTA,B259,38;%%
4177: \prd{49}{1994}{748}.
4178: %%CITATION = ASTRO-PH 9307002;%%
4179:
4180: \bibitem{cobe}
4181: C.L. Bennett {\it et al.}, \apj{464}{1996}{L1}.
4182: %%CITATION = ASTRO-PH 9601067;%%
4183:
4184: \bibitem{kn} J.E. Kim and H.P. Nilles,
4185: \plb{138}{1984}{150}.
4186: %%CITATION = PHLTA,B138,150;%%
4187:
4188: \bibitem{gm}
4189: C. Gordon and K.A. Malik, Phys. Rev. D {\bf 69},
4190: 063508 (2004);
4191: %%CITATION = ASTRO-PH 0311102;%%
4192: M. Beltr\'{a}n, J. Garc{\'\i}a-Bellido, J.
4193: Lesgourgues and A. Riazuelo, astro-ph/0409326.
4194: %%CITATION = ASTRO-PH 0409326;%%
4195:
4196: \bibitem{iso2}
4197: E. Pierpaoli, J. Garc{\'\i}a-Bellido and
4198: S. Borgani, \jhep{10}{1999}{015};
4199: %%CITATION = HEP-PH 9909420;%%
4200: R. Trotta, A. Riazuelo and R. Durrer,
4201: \prl{87}{2001}{231301}; \prd{67}{2003}{063520}.
4202: %%CITATION = ASTRO-PH 0104017;%%
4203: %%CITATION = ASTRO-PH 0211600;%%
4204:
4205: \bibitem{iso3}
4206: L. Amendola, C. Gordon, D. Wands and M. Sasaki,
4207: \prl{88}{2002}{211302}.
4208: %%CITATION = ASTRO-PH 0107089;%%
4209:
4210: \bibitem{mixed}
4211: D. Langlois and F. Vernizzi, astro-ph/0403258;
4212: %%CITATION = ASTRO-PH 0403258;%%
4213: F. Ferrer, S. R\"{a}s\"{a}nen and J.
4214: V\"{a}liviita, astro-ph/0407300.
4215: %%CITATION = ASTRO-PH 0407300;%%
4216:
4217: \bibitem{hier}
4218: G. Lazarides and N.D. Vlachos,
4219: %\plb{459}{1999}{482}.
4220: Phys. Lett. B {\bf 459}, 482 (1999).
4221: %%CITATION = HEP-PH 9903511;%%
4222:
4223: \bibitem{leptoinf}
4224: G. Lazarides and Q. Shafi, \plb{258}{1991}{305};
4225: %%CITATION = PHLTA,B258,305;%%
4226: G. Lazarides, R.K. Schaefer and Q. Shafi,
4227: \prd{56}{1997}{1324}.
4228: %%CITATION = HEP-PH 9608256;%%
4229:
4230: \bibitem{lepto}
4231: M. Fukugita and T. Yanagida, Phys. Lett. B
4232: {\bf 174}, 45 (1986).
4233: %\plb{174}{1986}{45}.
4234: %%CITATION = PHLTA,B174,45;%%
4235:
4236: \bibitem{crisis}
4237: M. Dine, W. Fischler and D. Nemeschansky,
4238: \plb{136}{1984}{169};
4239: %%CITATION = PHLTA,B136,169;%%
4240: G.D. Coughlan, R. Holman, P. Ramond and
4241: G.G. Ross, \ibid{140}{1984}{44}.
4242: %%CITATION = PHLTA,B140,44;%%
4243:
4244: \bibitem{drt95}
4245: M. Dine, L. Randall and S. Thomas,
4246: \prl{75}{1995}{398};
4247: %%CITATION = HEP-PH 9503303;%%
4248: \npb{458}{1996}{291}.
4249: %%CITATION = HEP-PH 9507453;%%
4250:
4251: \bibitem{noscale}
4252: E.D. Stewart, \prd{51}{1995}{6847};
4253: %%CITATION = HEP-PH 9405389;%%
4254: M.K. Gaillard, H. Murayama and K.A. Olive,
4255: \plb{355}{1995}{71};
4256: %%CITATION = HEP-PH 9504307;%%
4257: M.K. Gaillard, D.H. Lyth and H. Murayama,
4258: \prd{58}{1998}{123505};
4259: %%CITATION = HEP-TH 9806157;%%
4260: C. Panagiotakopoulos, \plb{459}{1999}{473};
4261: %%CITATION = HEP-PH 9904284;%%
4262: R. Jeannerot, S. Khalil and G. Lazarides,
4263: \jhep{07}{2002}{069}.
4264: %%CITATION = HEP-PH 0207244;%%
4265:
4266: \bibitem{acbar1}
4267: J.H. Goldstein {\it et al.},
4268: \apj{599}{2003}{773};
4269: %%CITATION = ASTRO-PH 0212517;%%
4270: C.-l. Kuo {\it et al.}, \ibid{600}{2004}{32}.
4271: %%CITATION = ASTRO-PH 0212289;%%
4272:
4273: \bibitem{acbar2}
4274: \textsf{http://cosmologist.info/ACBAR}.
4275:
4276: \bibitem{cbi}
4277: T.J. Pearson {\it et al.}, \apj{591}{2003}{556}.
4278: %%CITATION = ASTRO-PH 0205388;%%
4279:
4280: \bibitem{trieste}
4281: G. Lazarides, PRHEP-trieste99/008
4282: [hep-ph/9905450].
4283: %%CITATION = HEP-PH 9905450;%%
4284:
4285: \bibitem{continuous}
4286: G. Lazarides, C. Panagiotakopoulos and
4287: Q. Shafi, \prl{56}{1986}{432}.
4288: %%CITATION = PRLTA,56,432;%%
4289:
4290: \bibitem{discrete}
4291: N. Ganoulis, G. Lazarides and Q. Shafi,
4292: \npb{323}{1989}{374};
4293: %%CITATION = NUPHA,B323,374;%%
4294: G. Lazarides and Q. Shafi,
4295: \ibid{B329}{1990}{182}.
4296: %%CITATION = NUPHA,B329,182;%%
4297:
4298: \bibitem{smooth}
4299: G. Lazarides and C. Panagiotakopoulos,
4300: \prd{52}{1995}{559}.
4301: %%CITATION = HEP-PH 9506325;%%
4302:
4303: \bibitem{jean}
4304: R. Jeannerot, S. Khalil, G. Lazarides and Q. Shafi,
4305: \jhep{10}{2000}{012}.
4306: %%CITATION = HEP-PH 0002151;%%
4307:
4308: \bibitem{magg}
4309: G. Lazarides, M. Magg and Q. Shafi,
4310: \plb{97}{1980}{87}.
4311: %%CITATION = PHLTA,B97,87;%%
4312:
4313: \bibitem{kibble}
4314: T.W.B. Kibble, G. Lazarides and Q. Shafi,
4315: \plb{113}{1982}{237}.
4316: %%CITATION = PHLTA,B113,237;%%
4317:
4318: \bibitem{generic}
4319: R. Jeannerot, J. Rocher and M. Sakellariadou,
4320: \prd{68}{2003}{103514}.
4321: %%CITATION = HEP-PH 0308134;%%
4322:
4323: \bibitem{mairi}
4324: J. Rocher and M. Sakellariadou, hep-ph/0405133.
4325: %%CITATION = HEP-PH 0405133;%%
4326:
4327: \bibitem{dllr2}
4328: K. Dimopoulos, G. Lazarides, D. Lyth and R.
4329: Ruiz de Austri, \prd{68}{2003}{123515}.
4330: %%CITATION = HEP-PH 0308015;%%
4331:
4332: \bibitem{ekn}
4333: J.R. Ellis, J.E. Kim and D.V. Nanopoulos,
4334: \plb{145}{1984}{181};
4335: %%CITATION = PHLTA,B145,181;%%
4336: J.R. Ellis, D.V. Nanopoulos and S. Sarkar,
4337: \npb{259}{1985}{175};
4338: %%CITATION = NUPHA,B259,175;%%
4339: J.R. Ellis, G.B. Gelmini, J.L. Lopez,
4340: D.V. Nanopoulos and S. Sarkar,
4341: \ibid{B373}{1992}{399}.
4342: %%CITATION = NUPHA,B373,399;%%
4343:
4344: \bibitem{reheat}
4345: R.J. Scherrer and M.S. Turner, Phys. Rev. D
4346: {\bf 31}, 681 (1985).
4347: %%CITATION = PHRVA,D31,681;%%
4348:
4349: \bibitem{talks}
4350: G. Lazarides, in {\it Recent developments
4351: in particle physics and cosmology}, edited by
4352: G.C. Branco, Q. Shafi and J.I. Silva-Marcos
4353: (Kluwer Academic Publishers, Dordrecht, 2001),
4354: p. 399 [hep-ph/0011130];
4355: %%CITATION = HEP-PH 0011130;%%
4356: R. Jeannerot, S. Khalil and G. Lazarides, in
4357: {\it The proceedings of Cairo international
4358: conference on high energy physics}, edited by
4359: S. Khalil, Q. Shafi and H. Tallat (Rinton
4360: Press Inc., Princeton, 2001), p. 254
4361: [hep-ph/0106035].
4362: %%CITATION = HEP-PH 0106035;%%
4363:
4364: \bibitem{communication}
4365: K. Dimopoulos and D.H. Lyth, private
4366: communication.
4367:
4368: \bibitem{notari}
4369: K. Dimopoulos, D.H. Lyth, A. Notari and A. Riotto,
4370: \jhep{07}{2003}{053}.
4371: %%CITATION = HEP-PH 0304050;%%
4372:
4373: \bibitem{rwalk}
4374: D.H. Lyth and E.D. Stewart, Phys. Rev. D
4375: {\bf 46}, 532 (1992).
4376: %\prd{46}{1992}{532}.
4377: %%CITATION = PHRVA,D46,532;%%
4378:
4379: \bibitem{prep}
4380: K. Dimopoulos and D.H. Lyth, in preparation.
4381:
4382: \bibitem{evap}
4383: R. Allahverdi, B.A. Campbell and J.R. Ellis,
4384: \npb{579}{2000}{355};
4385: %%CITATION = HEP-PH 0001122;%%
4386: A. Anisimov and M. Dine,
4387: {\it ibid. }{\bf B619}, 729 (2001).
4388: %%CITATION = HEP-PH 0008058;%%
4389:
4390: \bibitem{thesis}
4391: R. Trotta, {\it Cosmic microwave anisotropies:
4392: beyond standard parameters} (Ph.D. dissertation
4393: at the University of Geneva, 2004), in
4394: \textsf{http://theory.physics.unige.ch/$\tilde{\phantom{a}}$trotta}.
4395:
4396: \bibitem{gipt}
4397: J.M. Bardeen, \prd{22}{1980}{1882};
4398: %%CITATION = PHRVA,D22,1882;%%
4399: H. Kodama and M. Sasaki, \ptps{78}{1984}{1};
4400: %%CITATION = PTPSA,78,1;%%
4401: V.F. Mukhanov, H.A. Feldman and R.H. Brandenberger,
4402: \prep{215}{1992}{203};
4403: %%CITATION = PRPLC,215,203;%%
4404: R. Durrer, \fcp{15}{1994}{209} [astro-ph/9311041].
4405: %%CITATION = ASTRO-PH 9311041;%%
4406:
4407: \bibitem{bst}
4408: J.M. Bardeen, P.J. Steinhardt and M.S. Turner,
4409: \prd{28}{1983}{679};
4410: %%CITATION = PHRVA,D28,679;%%
4411: J. Martin and D.J. Schwarz, \ibid{57}{1998}{3302};
4412: %%CITATION = GR-QC 9704049;%%
4413: D. Wands, K.A. Malik, D.H. Lyth and A.R. Liddle,
4414: \ibid{62}{2000}{043527}.
4415: %%CITATION = ASTRO-PH 0003278;%%
4416:
4417: \bibitem{km}
4418: M. Kawasaki and T. Moroi, Prog. Theor. Phys.
4419: {\bf 93}, 879 (1995).
4420: %\ptp{93}{1995}{879}.
4421: %%CITATION = HEP-PH 9403364;%%
4422:
4423: \bibitem{axion}
4424: M.S. Turner, \prd{33}{1986}{889}.
4425: %%CITATION = PHRVA,D33,889;%%
4426:
4427: \bibitem{hst}
4428: W.L. Freedman {\it et al.}, \apj{553}{2001}{47}.
4429: %%CITATION = ASTRO-PH 0012376;%%
4430:
4431: \bibitem{pivot}
4432: E.J. Copeland, I.J. Grivell and A.R. Liddle,
4433: astro-ph/ 9712028.
4434: %%CITATION = ASTRO-PH 9712028;%%
4435:
4436: \bibitem{cobe10}
4437: M. Tegmark and A.J.S. Hamilton, astro-ph/9702019.
4438: %%CITATION = ASTRO-PH 9702019;%%
4439:
4440: \bibitem{wang}
4441: B. Ananthanarayan, Q. Shafi and X.M. Wang,
4442: \prd{50}{1994}{5980};
4443: %%CITATION = HEP-PH 9311225;%%
4444: M.E. Gomez, G. Lazarides and C. Pallis,
4445: \ibid{61}{2000}{123512}.
4446: %%CITATION = HEP-PH 9907261;%%
4447:
4448: \bibitem{ananth}
4449: B. Ananthanarayan, G. Lazarides and Q. Shafi,
4450: \prd{44}{1991}{1613};
4451: %%CITATION = PHRVA,D44,1613;%%
4452: \plb{300}{1993}{245}.
4453: %%CITATION = PHLTA,B300,245;%%
4454:
4455: \bibitem{deBernardis}
4456: P. de Bernardis {\it et al.},
4457: \apj{564}{2002}{559}.
4458: %%CITATION = ASTRO-PH 0105296;%%
4459:
4460: \bibitem{Durrer}
4461: R. Durrer, B. Novosyadlyj and S. Apunevych,
4462: \apj{583}{2003}{33}.
4463: %%CITATION = ASTRO-PH 0111594;%%
4464:
4465: \bibitem{cosmomc}
4466: \textsf{http://cosmologist.info/cosmomc/}.
4467:
4468: \bibitem{lewis}
4469: A. Lewis and S. Bridle, \prd{66}{2002}{103511}.
4470: %%CITATION = ASTRO-PH 0205436;%%
4471:
4472: \bibitem{Valiviita}
4473: J. V\"{a}liviita and V. Muhonen,
4474: Phys. Rev. Lett. {\bf 91}, 131302 (2003).
4475: %\prl{91}{2003}{131302}.
4476: %%CITATION = ASTRO-PH 0304175;%%
4477:
4478: \bibitem{constraints}
4479: R. Bowen, S.H. Hansen, A. Melchiorri, J. Silk
4480: and R. Trotta, \mnras{334}{2002}{760};
4481: %%CITATION = ASTRO-PH 0110636;%%
4482: P. Crotty, J. Lesgourgues and S. Pastor,
4483: \prd{67}{2003}{123005};
4484: %%CITATION = ASTRO-PH 0302337;%%
4485: E. Pierpaoli, \mnras{342}{2003}{L63};
4486: %%CITATION = ASTRO-PH 0302465;%%
4487: S. Hannestad, \jcap{05}{2003}{004}.
4488: %%CITATION = ASTRO-PH 0303076;%%
4489:
4490: \bibitem{Verde}
4491: L. Verde {\it et al.}, \apjs{148}{2003}{195}.
4492: %%CITATION = ASTRO-PH 0302218;%%
4493:
4494: \bibitem{gelman}
4495: A. Gelman and D. Rubin, \ss{7}{1992}{457}.
4496: %Statist. Sci.{\bf 7}, 457 (1992).
4497:
4498: \bibitem{MKbook}
4499: D.J. MacKay, {\it Information theory,
4500: inference, and learning algorithms}
4501: (Cambridge University Press, Cambridge, U.K.,
4502: 2003).
4503:
4504: \bibitem{bbook}
4505: G.E. Box and G.C. Tiao, {\it Bayesian inference
4506: in statistical analysis}, Addison-Wesley series
4507: in behavioral science: quantitative methods
4508: (Addison-Wesley, Reading, Massachusetts,
4509: U.S.A., 1973).
4510:
4511: \bibitem{Burles}
4512: S. Burles, K.M. Nollett and M.S. Turner,
4513: \prd{63}{2001}{063512}.
4514: %%CITATION = ASTRO-PH 0008495;%%
4515:
4516: \bibitem{wmapgauss}
4517: E. Komatsu {\it et al.},
4518: \apjs{148}{2003}{119}.
4519: %%CITATION = ASTRO-PH 0302223;%%
4520:
4521: \bibitem{axionnongauss}
4522: D.H. Lyth, \prd{45}{1992}{3394}.
4523: %%CITATION = PHRVA,D45,3394;%%
4524:
4525: \bibitem{Lindley}
4526: D.V. Lindley, \bm{44}{1957}{187};
4527: M.S. Bartlett, \ibid{44}{1957}{533}.
4528:
4529: \bibitem{Kass}
4530: R.E. Kass and A.E. Raftery,
4531: \jasa{90}{1995}{773}.
4532:
4533: \bibitem{DiCiccio}
4534: T.J. DiCiccio, R.E. Kass, A.E. Raftery and
4535: L. Wasserman, \jasa{92}{1997}{903}.
4536:
4537: \end{thebibliography}
4538:
4539: \end{document}
4540: