FIVE-YEAR WILKINSON MICROWAVE ANISOTROPY PROBE* OBSERVATIONS: COSMOLOGICAL INTERPRETATION

, , , , , , , , , , , , , , , , , , and

Published 2009 February 11 © 2009. The American Astronomical Society. All rights reserved.
, , Citation E. Komatsu et al 2009 ApJS 180 330 DOI 10.1088/0067-0049/180/2/330

0067-0049/180/2/330

ABSTRACT

The Wilkinson Microwave Anisotropy Probe (WMAP) 5-year data provide stringent limits on deviations from the minimal, six-parameter Λ cold dark matter model. We report these limits and use them to constrain the physics of cosmic inflation via Gaussianity, adiabaticity, the power spectrum of primordial fluctuations, gravitational waves, and spatial curvature. We also constrain models of dark energy via its equation of state, parity-violating interaction, and neutrino properties, such as mass and the number of species. We detect no convincing deviations from the minimal model. The six parameters and the corresponding 68% uncertainties, derived from the WMAP data combined with the distance measurements from the Type Ia supernovae (SN) and the Baryon Acoustic Oscillations (BAO) in the distribution of galaxies, are: Ωbh2 = 0.02267+0.00058−0.00059, Ωch2 = 0.1131 ± 0.0034, ΩΛ = 0.726 ± 0.015, ns = 0.960 ± 0.013, τ = 0.084 ± 0.016, and $\Delta _{\cal R}^2 = (2.445\pm 0.096)\times 10^{-9}$ at k = 0.002 Mpc-1. From these, we derive σ8 = 0.812 ± 0.026, H0 = 70.5 ± 1.3 km s-1 Mpc−1, Ωb = 0.0456 ± 0.0015, Ωc = 0.228 ± 0.013, Ωmh2 = 0.1358+0.0037−0.0036, zreion = 10.9 ± 1.4, and t0 = 13.72 ± 0.12 Gyr. With the WMAP data combined with BAO and SN, we find the limit on the tensor-to-scalar ratio of r < 0.22(95%CL), and that ns > 1 is disfavored even when gravitational waves are included, which constrains the models of inflation that can produce significant gravitational waves, such as chaotic or power-law inflation models, or a blue spectrum, such as hybrid inflation models. We obtain tight, simultaneous limits on the (constant) equation of state of dark energy and the spatial curvature of the universe: −0.14 < 1 + w < 0.12(95%CL) and −0.0179 < Ωk < 0.0081(95%CL). We provide a set of "WMAP distance priors," to test a variety of dark energy models with spatial curvature. We test a time-dependent w with a present value constrained as −0.33 < 1 + w0 < 0.21 (95% CL). Temperature and dark matter fluctuations are found to obey the adiabatic relation to within 8.9% and 2.1% for the axion-type and curvaton-type dark matter, respectively. The power spectra of TB and EB correlations constrain a parity-violating interaction, which rotates the polarization angle and converts E to B. The polarization angle could not be rotated more than −5fdg9 < Δα < 2fdg4 (95% CL) between the decoupling and the present epoch. We find the limit on the total mass of massive neutrinos of ∑mν < 0.67 eV(95%CL), which is free from the uncertainty in the normalization of the large-scale structure data. The number of relativistic degrees of freedom (dof), expressed in units of the effective number of neutrino species, is constrained as Neff = 4.4 ± 1.5 (68%), consistent with the standard value of 3.04. Finally, quantitative limits on physically-motivated primordial non-Gaussianity parameters are −9 < flocalNL < 111 (95% CL) and −151 < fequilNL < 253 (95% CL) for the local and equilateral models, respectively.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Measurements of microwave background fluctuations by the Cosmic Background Explorer (COBE; Smoot et al. 1992; Bennett et al. 1994, 1996), the Wilkinson Microwave Anisotropy Probe (WMAP; Bennett et al. 2003a, 2003b), and ground and balloon-borne experiments (Miller et al. 1999, 2002; de Bernardis et al. 2000; Hanany et al. 2000; Netterfield et al. 2002; Ruhl et al. 2003; Mason et al. 2003; Sievers et al. 2003, 2007; Pearson et al. 2003; Readhead et al. 2004; Dickinson et al. 2004; Kuo et al. 2004, 2007; Reichardt et al. 2008; Jones et al. 2006; Montroy et al. 2006; Piacentini et al. 2006) have addressed many of the questions that were the focus of cosmology for the past 50 years: How old is the universe? How fast is it expanding? What is the size and shape of the universe? What is the composition of the universe? What seeded the formation of galaxies and large-scale structure?

By accurately measuring the statistical properties of the microwave background fluctuations, WMAP has helped establish a standard cosmology: a flat Λ cold dark matter (CDM) model composed of atoms, dark matter, and dark energy, with nearly scale-invariant adiabatic Gaussian fluctuations. With our most recent measurements, WMAP has measured the basic parameters of this cosmology to high precision: with the WMAP 5-year data alone, we find the density of dark matter (21.4%), the density of atoms (4.4%), the expansion rate of the universe, the amplitude of density fluctuations, and their scale dependence, as well as the optical depth due to reionization (Dunkley et al. 2009; also see Table 1 for summary).

Table 1. Summary of the Cosmological Parameters of ΛCDM Model and the Corresponding 68% Intervals

Class Parameter WMAP 5 Year MLa WMAP+BAO+SN ML WMAP 5 Year Meanb WMAP+BAO+SN Mean
Primary 100Ωbh2 2.268 2.262 2.273 ± 0.062 2.267+0.058−0.059
  Ωch2 0.1081 0.1138 0.1099 ± 0.0062 0.1131 ± 0.0034
  ΩΛ 0.751 0.723 0.742 ± 0.030 0.726 ± 0.015
  ns 0.961 0.962 0.963+0.014−0.015 0.960 ± 0.013
  τ 0.089 0.088 0.087 ± 0.017 0.084 ± 0.016
  $\Delta ^2_{\cal R}(k_0^{\rm c})$ 2.41 × 10−9 2.46 × 10−9 (2.41 ± 0.11) × 10−9 (2.445 ± 0.096) × 10−9
Derived σ8 0.787 0.817 0.796 ± 0.036 0.812 ± 0.026
  H0 72.4 km s-1 Mpc−1 70.2 km s-1 Mpc−1 71.9+2.6−2.7 km s−1 Mpc−1 70.5 ± 1.3 km s Mpc
  Ωb 0.0432 0.0459 0.0441 ± 0.0030 0.0456 ± 0.0015
  Ωc 0.206 0.231 0.214 ± 0.027 0.228 ± 0.013
  Ωmh2 0.1308 0.1364 0.1326 ± 0.0063 0.1358+0.0037−0.0036
  zdreion 11.2 11.3 11.0 ± 1.4 10.9 ± 1.4
  te0 13.69 Gyr 13.72 Gyr 13.69 ± 0.13 Gyr 13.72 ± 0.12 Gyr

Notes. aDunkley et al. (2009). "ML" refers to the Maximum Likelihood parameters. bDunkley et al. (2009). "Mean" refers to the mean of the posterior distribution of each parameter. ck0 = 0.002 Mpc−1. $\Delta ^2_{\cal R}(k)=k^3P_{\cal R}(k)/(2\pi ^2)$ (Equation (15)). d"Redshift of reionization," if the universe was reionized instantaneously from the neutral state to the fully ionized state at zreion. eThe present-day age of the universe.

Download table as:  ASCIITypeset image

Cosmologists are now focused on a new set of questions: What is the nature of the dark energy? What is the dark matter? Did inflation seed the primordial fluctuations? If so, what is the class of the inflationary model? How did the first stars form? Microwave background observations from WMAP, Planck, and from the upcoming generation of cosmic microwave background (CMB) ground and balloon-borne experiments will play an important role in addressing these questions.

This paper will discuss how the WMAP results, particularly when combined with other astronomical observations (mainly the distance measurements), are now providing new insights into these questions through constraints on gravitational waves and nonadiabatic (entropic) fluctuations, measurements of primordial non-Gaussianity, accurate determination of the primordial spectral index and the geometry of the universe, and limits on parity-violating interactions (see Table 2 for summary).

Table 2. Summary of the 95% Confidence Limits on Deviations from the Simple (Flat, Gaussian, Adiabatic, Power-Law) ΛCDM Model

Section Name Type WMAP 5 Year WMAP+BAO+SN
Section 3.2 Gravitational wavea No running index r < 0.43b r < 0.22
Section 3.1.3 Running index No grav. wave −0.090 < dns/dln k < 0.019c −0.068 < dns/dln k < 0.012
Section 3.4 Curvatured   −0.063 < Ωk < 0.017e −0.0179 < Ωk < 0.0081f
  Curvature radiusg Positive curv. Rcurv > 12 h−1 Gpc Rcurv > 22 h−1 Gpc
    Negative curv. Rcurv > 23 h−1 Gpc Rcurv > 33 h−1 Gpc
Section 3.5 Gaussianity Local −9 < flocalNL < 111h N/A
    Equilateral −151 < fequilNL < 253i N/A
Section 3.6 Adiabaticity Axion α0 < 0.16j α0 < 0.072k
    Curvaton α−1 < 0.011l α−1 < 0.0041m
Section 4 Parity violation Chern–Simonsn −5fdg9 < Δα < 2fdg4 N/A
Section 5 Dark energy Constant wo −1.37 < 1 + w < 0.32p −0.14 < 1 + w < 0.12
    Evolving w(z)q N/A −0.33 < 1 + w0 < 0.21r
Section 6.1 Neutrino masss   mν < 1.3 eVt mν < 0.67 eVu
Section 6.2 Neutrino species   Neff > 2.3v Neff = 4.4 ± 1.5w (68%)

Notes. aIn the form of the tensor-to-scalar ratio, r, at k = 0.002 Mpc−1. bDunkley et al. (2009). cDunkley et al. (2009). d(Constant) dark energy equation of state allowed to vary (w ≠ −1). eWith the HST prior, H0 = 72 ± 8 km s-1 Mpc−1. For w = −1, −0.052 < Ωk < 0.013(95%CL). fFor w = −1, −0.0178 < Ωk < 0.0066(95%CL). g$R_{\rm curv}=(c/H_0)/\sqrt{|\Omega _k|}=3/\sqrt{|\Omega _k|}\,h^{-1}$ Gpc. hCleaned V + W map with lmax = 500 and the KQ75 mask, after the point-source correction. iCleaned V + W map with lmax = 700 and the KQ75 mask, after the point-source correction. jDunkley et al. (2009). kIn terms of the adiabaticity deviation parameter, $\delta _{adi}^{(c,\gamma)}=\sqrt{\alpha }/3$ (Equation (39)), the axion-like dark matter and photons are found to obey the adiabatic relation (Equation (36)) to 8.9%. lDunkley et al. (2009). mIn terms of the adiabaticity deviation parameter, $\delta _{adi}^{(c,\gamma)}=\sqrt{\alpha }/3$ (Equation (39)), the curvaton-like dark matter and photons are found to obey the adiabatic relation (Equation (36)) to 2.1%. nFor an interaction of the form given by $[\phi (t)/M]F_{\alpha \beta }\tilde{F}^{\alpha \beta }$, the polarization rotation angle is $\Delta \alpha =M^{-1}\int \frac{dt}{a} \dot{\phi }$. oFor spatially curved universes (Ωk ≠ 0). pWith the HST prior, H0 = 72 ± 8 km s-1 Mpc−1. qFor a flat universe (Ωk = 0). rw0w(z = 0). smν = 94(Ωνh2) eV. tDunkley et al. (2009). uFor w = −1. For w ≠ −1, ∑mν < 0.80 eV(95%CL). vDunkley et al. (2009). wWith the HST prior, H0 = 72 ± 8 km s-1 Mpc−1. The 95% limit is 1.8 < Neff < 7.6.

Download table as:  ASCIITypeset image

This paper is one of seven papers on the analysis of the WMAP 5-year data: Hinshaw et al. (2009) reported on the data processing, map-making, and systematic error limits; Hill et al. (2009) on the physical optics modeling of beams and the 5-year window functions (beam transfer functions); Gold et al. (2009) on the modeling, understanding, and subtraction of the temperature and polarized foreground emission; Wright et al. (2009) on the catalogue of point sources detected in the 5-year temperature data; Nolta et al. (2009) on the measurements of the temperature and polarization power spectra; and Dunkley et al. (2009) on the parameter estimation methodology, the cosmological parameters inferred from the WMAP data alone, and comparison between different cosmological data sets.

This paper is organized as follows. In Section 2, we briefly summarize new aspects of our analysis of the WMAP 5-year temperature and polarization data. In Section 3, we constrain the spatial curvature of the observable universe, Gaussianity/adiabaticity/scale-invariance of the primordial fluctuations, and the amplitude of primordial gravitational waves. We discuss their implications for the physics of the early, primordial universe. In Section 4, we demonstrate that the power spectra of TB and EB correlations,15 which are usually ignored in the cosmological analysis, can be used to constrain a certain parity-violating interaction that couples to photons. In Section 5, we explore the nature of dark energy, and in Section 6, we study the properties of neutrinos in cosmology. We conclude in Section 7.

2. SUMMARY OF 5-YEAR ANALYSIS

2.1. WMAP 5-year Data: Temperature and Polarization

With 5 years of observations of Jupiter and an extensive physical optics modeling of beams (Hill et al. 2009), our understanding of the beam transfer function, bl, has improved significantly: the fractional beam errors, Δbl/bl, have been nearly halved in most differencing assemblies (DAs). In some cases, for example W4, the errors have been reduced by as much as a factor of 4.

Many of the small-scale CMB experiments have been calibrated to the WMAP 3-year data at high multipoles. Since the new beam model raises the 5-year power spectrum almost uniformly by ∼2.5% relative to the 3-year power spectrum over l ≳ 200 (Hill et al. 2009), these small-scale CMB experiments have been undercalibrated by the same amount, that is, ∼2.5% in power and 1.2% in temperature. For example, the latest Arcminute Cosmology Bolometer Array Receiver (ACBAR) data (Reichardt et al. 2008) report on the calibration error of 2.23% in temperature (4.5% in power), which is twice as large as the magnitude of miscalibration; thus, we expect the effect of miscalibration to be subdominant in the error budget. Note that the change in the beam is fully consistent with the 1σ error of the previous WMAP beam reported in Page et al. (2003). Since the ACBAR calibration error includes the previous WMAP beam error, the change in the beam should have a minimal impact on the current ACBAR calibration.

While we use only V and W bands for the cosmological analysis of the temperature data, the treatment of bl in Q band affects our determination of the point-source contamination to the angular power spectrum, Aps.16 The 5-year estimate of the point-source correction, Aps = 0.011 ± 0.001 μK2 sr (Nolta et al. 2009), is slightly lower than the 3-year estimate, Aps = 0.014 ±  0.003 μK2 sr (Hinshaw et al. 2007), partly because more sources have been detected and masked by the 5-year source mask (390 sources have been detected in the 5-year temperature data, whereas 323 sources were detected in the 3-year data (Wright et al. 2009)).

Note that the uncertainty in Aps has been reduced by a factor of 3. The uncertainty in the previous estimate was inflated to include the lower value found by Huffenberger et al. (2006) (0.011) and higher value from our original estimate (0.017). Much of the discrepancy between these estimates is due to the multipole range over which Aps is fit. With the improved beam model from the 5-year analysis, the dependence on the multipole range has disappeared, and thus we no longer need to inflate our uncertainty. See Nolta et al. (2009) for more details.

The method for cleaning foreground emission in both temperature and polarization data is the same as we used for the 3-year data, that is, the template-based cleaning method described in Section 5.3 of Hinshaw et al. (2007) for temperature and Section 4.3 of Page et al. (2007) for polarization. Gold et al. (2009) described the results from the template cleaning of the 5-year data with the new coefficients. In addition, Gold et al. (2009) and Dunkley et al. (2009) explored alternative modelings of the foreground emission. All of these methods gave consistent results. Gold et al. (2009) also described definitions of the new masks, KQ75 and KQ85, that replace the previous masks, Kp0 and Kp2, that are recommended for the analysis of Gaussianity tests and the power spectrum, respectively.

The method for measuring the TT and TE spectra at higher multipoles, that is, l ⩾ 33 for TT and l ⩾ 24 for TE, is also the same as we used for the 3-year data (Hinshaw et al. 2007). As for the estimation of the cosmological parameters from these spectra, we now include the weak gravitational lensing effect of CMB due to the intervening matter fluctuations (see Lewis & Challinor 2006, for a review), which was not included in the 3-year analysis. We continue to marginalize over a potential contribution from the Sunyaev–Zel'dovich effect (SZE), using exactly the same template SZE power spectrum that we used for the 3-year analysis: CSZEl from Komatsu & Seljak (2002) with Ωm = 0.26, Ωb = 0.044, h = 0.72, ns = 0.97, and σ8 = 0.80 (see also Section 2.1 of Spergel et al. 2007). We continue to use the V- and W-band data for estimating the high-l temperature power spectrum, and the Q- and V-band data for the high-l polarization power spectra.

We have improved our treatment of the temperature and polarization power spectra at lower multipoles, as described below.

Low-l temperature. We use the Gibbs sampling technique and the Blackwell–Rao (BR) estimator to evaluate the likelihood of the temperature power spectrum at l ⩽ 32 (Jewell et al. 2004; Wandelt 2003; Wandelt et al. 2004; O'Dwyer et al. 2004; Eriksen et al. 2004a, 2007c, 2007b; Chu et al. 2005; Larson et al. 2007). For the 3-year analysis, we used the resolution 4 Internal Linear Combination (ILC) temperature map (Nside = 16) with a Gaussian smoothing of 9fdg183 (FWHM). Since the ILC map has an intrinsic Gaussian smoothing of 1°, we have added an extra smoothing of 9fdg1285. We then directly evaluated the likelihood in the pixel space for a given Cl. For the 5-year analysis, we use a higher-resolution map, the resolution 5 ILC map (Nside = 32) with a smaller Gaussian smoothing of 5° (FWHM). The potential foreground leakage due to smoothing is, therefore, reduced. The BR estimator has an advantage of being much faster to compute, which is why we have adopted the Gibbs sampling and the BR estimator for the 5-year data release. We have confirmed that both the resolution 4 pixel-based likelihood and the resolution 5 Gibbs-based likelihood yield consistent results (see Dunkley et al. 2009 for details). Both options are made publicly available in the released likelihood code.

Low-l polarization. While we continue to use the direct evaluation of the likelihood of polarization power spectra in pixel space from coadded resolution 3 (Nside = 8) polarization maps (Stokes Q and U maps), we now add the Ka band data to the coadded maps; we used only Q- and V-band data for the 3-year analysis. We believe that we understand the polarized foreground emission (dominated by synchrotron, traced well by the K-band data) in the Ka band data well enough to justify the inclusion of the Ka band (Gold et al. 2009). This, with 2 years of more integration, has led to a significant reduction of the noise power spectra (averaged over l = 2 − 7) in the polarization EE and BB power spectra by a factor of as much as 2.3 compared to the 3-year analysis. As a result, the EE power spectrum averaged over l = 2 − 7 exceeds the noise by a factor of 10, that is, our measurement of the EE power spectrum averaged over l = 2 − 7 is now limited by cosmic variance and the possibility of residual foreground emission and/or systematic errors,17 rather than by noise. In addition, we have added a capability of computing the likelihood of TB and EB power spectra to the released likelihood code. This allows us to test models in which nonzero TB and EB correlations can be generated. We discuss this further in Section 4.

We continue to use the Markov Chain Monte Carlo (MCMC) technique to explore the posterior distribution of cosmological parameters given the measured temperature and polarization power spectra. For details on the implementation and convergence criteria, see Dunkley et al. (2009).

2.2. Comments on Systematic Errors in the Cosmological Parameters Derived from the WMAP 5-year Data

Hinshaw et al. (2009) gave extensive descriptions of our limits on the systematic errors in the WMAP 5-year temperature and polarization data. With the improved treatment of beams and gain calibrations, we are confident that the instrumental systematic errors in the cosmological results derived from the temperature data are negligible compared to the statistical errors. As for the polarization data, we find that the W-band polarization data still contain the systematic errors that we do not fully understand, and thus we do not use the W-band polarization for the cosmological analysis. We do not find any evidence for the unaccounted instrumental systematic errors in the other bands that we use for cosmology.

The most dominant systematic errors in the cosmological analysis are the foreground emission. Since CMB dominates over the foreground emission in the temperature data in V and W bands outside the galaxy mask, and we also reduce the subdominant foreground contamination at high Galactic latitudes further by using the K- and Ka-band data (for synchrotron emission), the external Hα map (for free–free emission), and the external dust map, the systematic errors from the foreground emission are unimportant for the temperature data, even at the lowest multipoles where the foreground is most important (Gold et al. 2009).

We, however, find that the uncertainty in our modeling of the polarized foreground is not negligible compared to the statistical errors. For the 5-year polarization analysis, we have used two independent foreground-cleaning algorithms: one based upon the template fitting (as developed for the 3-year analysis; see Page et al. 2007) and the other based upon the Gibbs sampling (Dunkley et al. 2008). The optical depth, τ, is the parameter that is most affected by the uncertainty in the polarized foreground model. The template fitting method gives τ = 0.087 ± 0.017. The Gibbs sampling method gives a range of values from τ = 0.085 ± 0.025 to τ = 0.103 ± 0.018, depending upon various assumptions made about the properties of the polarized synchrotron and dust emission. Therefore, the systematic error in τ is comparable to the statistical error.

This has an implication for the determination of the primordial tilt, ns, as there is a weak correlation between ns and τ (see Figure 1): for τ = 0.087, we find ns = 0.963, while for τ = 0.105, we find ns = 0.98. Since the statistical error of ns is 0.015, the systematic error in ns (from the polarized foreground) is comparable to the statistical one. The other parameters that are correlated with ns, that is, the baryon density (Figure 1), the tensor-to-scalar ratio (Figure 2), and the amplitude of nonadiabatic fluctuations (Sections 3.6.3 and 3.6.4), would be similarly affected. For the parameters that are not correlated with ns or τ, the systematic errors are insignificant.

Figure 1.

Figure 1. Constraint on the primordial tilt, ns (Section 3.1.2). No running index or gravitational waves are included in the analysis. (Left) One-dimensional marginalized constraint on ns from the WMAP-only analysis. (Middle) Two-dimensional joint marginalized constraint (68% and 95% CL), showing a strong correlation between ns and Ωbh2. (Right) A mild correlation with τ. None of these correlations are reduced significantly by including BAO or SN data, as these datasets are not sensitive to Ωbh2 or τ; however, the situation changes when the gravitational wave contribution is included (see Figure 2).

Standard image High-resolution image
Figure 2.

Figure 2. Constraint on the tensor-to-scalar ratio, r, at k = 0.002 Mpc−1 (Section 3.2.4). No running index is assumed. See Figure 4 for r with the running index. In all panels, we show the WMAP-only results in blue and WMAP+BAO+SN in red. (Left) One-dimensional marginalized distribution of r, showing the WMAP-only limit, r < 0.43(95%CL), and WMAP+BAO+SN, r < 0.22(95%CL). (Middle) Joint two-dimensional marginalized distribution (68% and 95% CL), showing a strong correlation between ns and r. (Right) Correlation between ns and Ωmh2. The BAO and SN data help to break this correlation which, in turn, reduces the correlation between r and ns, resulting in a factor of 2.2 better limit on r.

Standard image High-resolution image

2.3. External Data Sets: Hubble Constant, Luminosity, and Angular Diameter Distances

Aside from the CMB data (including the small-scale CMB measurements), the main external astrophysical results that we shall use in this paper for the joint cosmological analysis are the following distance-scale indicators.

  • 1.  
    A Gaussian prior on the present-day Hubble's constant from the Hubble Key Project final results, H0 = 72 ± 8  km s-1 Mpc−1 (Freedman et al. 2001). While the uncertainty is larger than the WMAP's determination of H0 for the minimal ΛCDM model (see Table 1), this information improves upon limits on the other models, such as models with nonzero spatial curvature.
  • 2.  
    The luminosity distances out to Type Ia supernovae (SNe) with their absolute magnitudes marginalized over uniform priors. We use the "union" SN samples compiled by Kowalski et al. (2008). The union compilation contains 57 nearby (0.015 < z ≲ 0.15) Type Ia SNe and 250 high-z Type Ia SNe, after selection cuts. The high-z samples contain the Type Ia SNe from the Hubble Space Telescope (HST; Knop et al. 2003; Riess et al. 2004, 2007), the SuperNova Legacy Survey (SNLS; Astier et al. 2006), the Equation of State: SupErNovae trace Cosmic Expansion (ESSENCE) survey (Wood-Vasey et al. 2007), as well as those used in the original papers of the discovery of the acceleration of the universe (Riess et al. 1998; Perlmutter et al. 1999), and the samples from Barris et al. (2004) and Tonry et al. (2003). The nearby Type Ia SNe are taken from Hamuy et al. (1996), Riess et al. (1999), Jha et al. (2006), Krisciunas et al. (2001, 2004a, 2004b). Kowalski et al. (2008) have processed all of these samples using the same light curve fitter called {\sf SALT} (Guy et al. 2005), which allowed them to combine all the data in a self-consistent fashion. The union compilation is the largest to date. The previous compilation by Davis et al. (2007) used a smaller number of Type Ia SNe, and did not use the same light curve fitter for all the samples. We examine the difference in the derived ΛCDM cosmological parameters between the union compilation and Davis et al.'s compilation in Appendix D. While we ignore the systematic errors when we fit the Type Ia SN data, we examine the effects of systematic errors on the ΛCDM parameters and the dark energy parameters in Appendix D.
  • 3.  
    Gaussian priors on the distance ratios, rs/DV(z), at z = 0.2 and 0.35 measured from the Baryon Acoustic Oscillations (BAO) in the distribution of galaxies (Percival et al. 2007). The CMB observations measure the acoustic oscillations in the photon-baryon plasma, which can be used to measure the angular diameter distance to the photon decoupling epoch. The same oscillations are imprinted on the distribution of matter, traced by the distribution of galaxies, which can be used to measure the angular distances to the galaxies that one observes from galaxy surveys. While both CMB and galaxies measure the same oscillations, in this paper, we shall use the term, BAO, to refer only to the oscillations in the distribution of matter, for definiteness.

Here, we describe how we use the BAO data in more detail. The BAO can be used to measure not only the angular diameter distance, DA(z), through the clustering perpendicular to the line of sight, but also the expansion rate of the universe, H(z), through the clustering along the line of sight. This is a powerful probe of dark energy; however, the accuracy of the current data does not allow us to extract DA(z) and H(z) separately, as one can barely measure BAO in the spherically-averaged correlation function (Okumura et al. 2008).

The spherical average gives us the following effective distance measure (Eisenstein et al. 2005):

Equation (1)

where DA(z) is the proper (not comoving) angular diameter distance:

Equation (2)

where fk[x] = sin x, x, and sinh x for Ωk < 0 (k = 1), Ωk = 0 (k = 0), and Ωk > 0 (k = −1), respectively.

There is an additional subtlety. The peak positions of the (spherically averaged) BAO depend actually on the ratio of DV(z) to the sound horizon size at the so-called drag epoch, zd, at which baryons were released from photons. Note that there is no reason why the decoupling epoch of photons, z*, needs to be the same as the drag epoch, zd. They would be equal only when the energy density of photons and that of baryons were equal at the decoupling epoch—more precisely, they would be equal only when R(z) ≡ 3ρb/(4ργ) = (3Ωb/4Ωγ)/(1 + z) ≃ 0.67(Ωbh2/0.023)[1090/(1 + z)] was unity at z = z*. Since we happen to live in the universe in which Ωbh2 ≃ 0.023, this ratio is less than unity, and thus the drag epoch is slightly later than the photon decoupling epoch, zd < z*. As a result, the sound horizon size at the drag epoch happens to be slightly larger than that at the photon decoupling epoch. In Table 3, we give the CMB decoupling epoch, BAO drag epoch, as well as the corresponding sound horizon radii that are determined from the WMAP 5-year data.

Table 3. Sound Horizon Scales Determined by the WMAP 5-year Data

  Quantity Equation 5 Year WMAP
CMB z* (66) 1090.51 ± 0.95
CMB rs(z*) (6) 146.8 ± 1.8 Mpc
Matter zd (3) 1020.5 ± 1.6
Matter rs(zd) (6) 153.3 ± 2.0 Mpc

Notes. CMB: the sound horizon scale at the photon decoupling epoch, z*, imprinted on the CMB power spectrum; matter: the sound horizon scale at the baryon drag epoch, zd, imprinted on the matter (galaxy) power spectrum.

Download table as:  ASCIITypeset image

We use a fitting formula for zd proposed by Eisenstein & Hu (1998):

Equation (3)

where

Equation (4)

Equation (5)

In this paper, we frequently combine the WMAP data with rs(zd)/DV(z) extracted from the Sloan Digital Sky Survey (SDSS) and the Two Degree Field Galaxy Redshift Survey (2dFGRS; Percival et al. 2007), where rs(z) is the comoving sound horizon size given by

Equation (6)

where Ωγ = 2.469 × 10−5h−2 for Tcmb = 2.725 K, and

Equation (7)

The radiation density parameter, Ωr, is the sum of photons and relativistic neutrinos,

Equation (8)

where Neff is the effective number of neutrino species (the standard value is 3.04). For more details on Neff, see Section 6.2. When neutrinos are nonrelativistic at a, one needs to reduce the value of Neff accordingly. Also, the matter density must contain the neutrino contribution when they are nonrelativistic,

Equation (9)

where Ων is related to the sum of neutrino masses as

Equation (10)

For more details on the neutrino mass, see Section 6.1.

All the density parameters refer to the values at the present epoch, and add up to unity:

Equation (11)

Throughout this paper, we shall use ΩΛ to denote the dark energy density parameter at present:

Equation (12)

Here, weff(a) is the effective equation of state of dark energy given by

Equation (13)

and w(a) is the usual dark energy equation of state, that is, the dark energy pressure divided by the dark energy density:

Equation (14)

For vacuum energy (cosmological constant), w does not depend on time, and w = −1.

Percival et al. (2007) have determined rs(zd)/DV(z) out to two redshifts, z = 0.2 and 0.35, as rs(zd)/DV(z = 0.2) = 0.1980 ± 0.0058 and rs(zd)/DV(z = 0.35) = 0.1094 ± 0.0033, with a correlation coefficient of 0.39. We follow the description given in Appendix A of Percival et al. (2007) to implement these constraints in the likelihood code. We have checked that our calculation of rs(zd) using the above formulae (including zd) matches the value that they quote,18111.426 h−1 Mpc, to within 0.2 h−1 Mpc, for their quoted cosmological parameters, Ωm = 0.25, Ωbh2 = 0.0223, and h = 0.72.

We have decided to use these results, as they measure only the distances, and are not sensitive to the growth of structure. This property enables us to identify the information added by the external astrophysical results more clearly. In addition to these, we shall also use the BAO measurement by Eisenstein et al. (2005)19 and the flux power spectrum of Lyα forest from Seljak et al. (2006) in the appropriate context.

For the 3-year data analysis in Spergel et al. (2007), we also used the shape of the galaxy power spectra measured from the SDSS main sample and the Luminous Red Galaxies (Tegmark et al. 2004b, 2006), and 2dFGRS (Cole et al. 2005). We have found some tension between these data sets, which could be indicative of the degree by which our understanding of nonlinearities, such as the nonlinear matter clustering, nonlinear bias, and nonlinear redshift space distortion, is still limited at low redshifts, that is, z ≲ 1. See Dunkley et al. (2009) for more detailed study on this issue. Also see Sánchez & Cole (2008) on the related subject. The galaxy power spectra should provide us with important information on the growth of structure (which helps constrain the dark energy and neutrino masses) as our understanding of nonlinearities improves in the future. In this paper, we do not combine these datasets because of the limited understanding of the consequences of nonlinearities.

3. FLAT, GAUSSIAN, ADIABATIC, POWER-LAW ΛCDM MODEL, AND ITS ALTERNATIVES

The theory of inflation, the idea that the universe underwent a brief period of rapid accelerated expansion (Starobinsky 1979, 1982; Kazanas 1980; Guth 1981; Sato 1981; Linde 1982; Albrecht & Steinhardt 1982), has become an indispensable building block of the standard model of our universe.

Models of the early universe must explain the following observations: the universe is nearly flat and the fluctuations observed by WMAP appear to be nearly Gaussian (Komatsu et al. 2003), scale-invariant, super-horizon, and adiabatic (Spergel & Zaldarriaga 1997; Spergel et al. 2003; Peiris et al. 2003). Inflation models have been able to explain these properties successfully (Mukhanov & Chibisov 1981; Hawking 1982; Starobinsky 1982; Guth & Pi 1982; Bardeen et al. 1983).

Although many models have been ruled out observationally (see Kinney et al. 2006; Alabidi & Lyth 2006a; Martin & Ringeval 2006, for recent surveys), there are more than 100 candidate inflation models available (see Liddle & Lyth 2000; Bassett et al. 2006; Linde 2008, for reviews). Therefore, we now focus on the question, "which model is the correct inflation model?" This must be answered by the observational data.

However, an inflationary expansion may not be the only way to solve cosmological puzzles and create primordial fluctuations. Contraction of the primordial universe followed by a bounce to expansion can, in principle, make a set of predictions that are qualitatively similar to those of inflation models (Khoury et al. 2001, 2002a, 2002b, 2003; Buchbinder et al. 2007, 2008; Koyama & Wands 2007; Koyama et al. 2007; Creminelli & Senatore 2007), although building concrete models and making robust predictions have been challenging (Kallosh et al. 2001a, 2001b, 2008; Linde 2002).

There is also a fascinating possibility that one can learn something about the fundamental physics from cosmological observations. For example, recent progress in implementing de Sitter vacua and inflation in the context of string theory (see McAllister & Silverstein 2008, for a review) makes it possible to connect the cosmological observations to the ingredients of the fundamental physics via their predicted properties of inflation such as the shape of the power spectrum, spatial curvature of the universe, and non-Gaussianity of primordial fluctuations.

3.1. Power Spectrum of Primordial Fluctuations

3.1.1. Motivation and Analysis

The shape of the power spectrum of primordial curvature perturbations, $P_{\cal R}(k)$, is one of the most powerful and practical tool for distinguishing among inflation models. Inflation models with featureless scalar-field potentials usually predict that $P_{\cal R}(k)$ is nearly a power law (Kosowsky & Turner 1995)

Equation (15)

Here, $\Delta ^2_{\cal R}(k)$ is a useful quantity, which gives an approximate contribution of ${\cal R}$ at a given scale per logarithmic interval in k to the total variance of ${\cal R}$, as $\langle {\cal R}^2({\mathbf{x}})\rangle = \int d\ln k \Delta ^2_{\cal R}(k)$. It is clear that the special case with ns = 1 and dns/dln k = 0 results in the "scale-invariant spectrum," in which the contributions of ${\cal R}$ at any scales per logarithmic interval in k to the total variance are equal (hence, the term "scale invariance"). Following the usual terminology, we shall call ns and dns/dln k the tilt of the spectrum and the running index, respectively. We shall take k0 to be 0.002 Mpc−1.

The significance of ns and dns/dln k is that different inflation models motivated by different physics make specific, testable predictions for the values of ns and dns/dln k. For a given shape of the scalar field potential, V(ϕ), of a single-field model, for instance, one finds that ns is given by a combination of the first derivative and second derivative of the potential, 1 − ns = 3M2pl(V'/V)2 − 2M2pl(V''/V) (where M2pl = 1/(8πG) is the reduced Planck mass), and dns/dln k is given by a combination of V'/V, V''/V, and V‴/V (see Liddle & Lyth 2000, for a review).

This means that one can reconstruct the shape of V(ϕ) up to the first three derivatives in this way. As the expansion rate squared is proportional to V(ϕ) via the Friedmann equation, H2 = V/(3M2pl), one can reconstruct the expansion history during inflation by measuring the shape of the primordial power spectrum.

How generic are ns and dns/dln k? They are physically motivated by the fact that most inflation models satisfy the slow-roll conditions, and thus deviations from a pure power-law, scale-invariant spectrum, ns = 1 and dns/dln k = 0, are expected to be small, and the higher-order derivative terms such as V'''' and higher are negligible. However, there is always a danger of missing some important effects, such as sharp features, when one relies too much on a simple parametrization like this. Therefore, a number of people have investigated various, more general ways of reconstructing the shape of $P_{\cal R}(k)$ (Matsumiya et al. 2002, 2003; Mukherjee & Wang 2003b, 2003a, 2003c; Bridle et al. 2003; Kogo et al. 2004, 2005; Hu & Okamoto 2004; Hannestad 2004; Shafieloo & Souradeep 2004, 2008; Sealfon et al. 2005; Tocchini-Valentini et al. 2005, 2006; Spergel et al. 2007; Verde & Peiris 2008) and V(ϕ) (Lidsey et al. 1997; Grivell & Liddle 2000; Kadota et al. 2005; Covi et al. 2006; Lesgourgues & Valkenburg 2007; Powell & Kinney 2007).

These studies have indicated that the parametrized form (Equation (15)) is basically a good fit, and no significant features were detected. Therefore, we do not repeat this type of analysis in this paper, but focus on constraining the parametrized form given by Equation (15).

Finally, let us comment on the choice of priors. We impose uniform priors on ns and dns/dln k, but there are other possibilities for the choice of priors. For example, one may impose uniform priors on the slow-roll parameters, epsilon = (M2pl/2)(V'/V)2, η = M2pl(V''/V) and ξ = M4pl(V'V‴/V2), and on the number of e-foldings, N, rather than on ns and dns/dln k (Peiris & Easther 2006a, 2006b; Easther & Peiris 2006). It has been found that, as long as one imposes a reasonable lower bound on N, N > 30, both approaches yield similar results.

To constrain ns and dns/dln k, we shall use the WMAP 5-year temperature and polarization data, the small-scale CMB data, and/or BAO and SN distance measurements. In Table 4, we summarize our results presented in Sections 3.1.2, 3.1.3, and 3.2.4.

Table 4. Primordial Tilt ns, Running Index dns/dln k, and Tensor-to-Scalar Ratio r

Section Model Parametera 5 Year WMAPb 5 Year WMAP = +CMBc 5 Year WMAP+ACBAR08d 5 Year WMAP+BAO+SN
Section 3.1.2 Power law ns 0.963+0.014−0.015 0.960 ± 0.014 0.964 ± 0.014 0.960 ± 0.013
Section 3.1.3 Running ns 1.031+0.054−0.055e 1.059+0.051−0.049 1.031 ± 0.049 1.017+0.042−0.043f
    dns/dln k −0.037 ± 0.028 −0.053 ± 0.025 −0.035+0.024−0.025 −0.028 ± 0.020g
Section 3.2.4 Tensor ns 0.986 ± 0.022 0.979 ± 0.020 0.985+0.019−0.020 0.970 ± 0.015
    r <0.43(95%CL) <0.36(95%CL) <0.40(95%CL) <0.22(95%CL)
Section 3.2.4 Running ns 1.087+0.072−0.073 1.127+0.075−0.071 1.083+0.063−0.062 1.089+0.070−0.068
  +Tensor r <0.58(95%CL) <0.64(95%CL) <0.54(95%CL) <0.55(95%CL)h
    dns/dln k −0.050 ± 0.034 −0.072+0.031−0.030 −0.048 ± 0.027 −0.053 ± 0.028i

Notes. aDefined at k0 = 0.002 Mpc−1. bDunkley et al. (2009). c"CMB" includes the small-scale CMB measurements from CBI (Mason et al. 2003; Sievers et al. 2003, 2007; Pearson et al. 2003; Readhead et al. 2004), VSA (Dickinson et al. 2004), ACBAR (Kuo et al. 2004, 2007), and BOOMERanG (Ruhl et al. 2003; Montroy et al. 2006; Piacentini et al. 2006). d"ACBAR08" is the complete ACBAR data set presented in Reichardt et al. (2008). We used the ACBAR data in the multipole range of 900 < l < 2000. eAt the pivot point for WMAP only, kpivot = 0.080 Mpc−1, where ns and dns/dln k are uncorrelated, ns(kpivot) = 0.963 ± 0.014. fAt the pivot point for WMAP+BAO+SN, kpivot = 0.106 Mpc−1, where ns and dns/dln k are uncorrelated, ns(kpivot) = 0.961 ± 0.014. gWith the Lyα forest data (Seljak et al. 2006), dns/dln k = −0.012 ± 0.012. hWith the Lyα forest data (Seljak et al. 2006), r < 0.28 (95% CL). iWith the Lyα forest data (Seljak et al. 2006), dns/dln k = −0.017+0.014−0.013.

Download table as:  ASCIITypeset image

3.1.2. Results: Tilt

First, we test the inflation models with dns/dln k = 0 and negligible gravitational waves. The WMAP 5-year temperature and polarization data yield ns = 0.963+0.014−0.015, which is slightly above the 3-year value with a smaller uncertainty, ns(3yr) = 0.958 ± 0.016 (Spergel et al. 2007). We shall provide the reason for this small upward shift in Section 3.1.3.

The scale-invariant, Harrison–Zel'dovich–Peebles spectrum, ns = 1, is at 2.5 standard deviations away from the mean of the likelihood for the WMAP-only analysis. The significance increases to 3.1 standard deviations for WMAP+BAO+SN. Looking at the two-dimensional constraints that include ns, we find that the most dominant correlation that is still left is the correlation between ns and Ωbh2 (see Figure 1). The larger the Ωbh2 is, the smaller the second peak becomes, and the larger the ns is required to compensate it. Also, the larger the Ωbh2 is, the larger the Silk damping (diffusion damping) becomes, and the larger the ns is required to compensate it.

This argument suggests that the constraint on ns should improve as we add more information from the small-scale CMB measurements that probe the Silk damping scales; however, the current data do not improve the constraint very much yet: ns = 0.960 ± 0.014 from WMAP+CMB, where "CMB" includes the small-scale CMB measurements from CBI (Mason et al. 2003; Sievers et al. 2003, 2007; Pearson et al. 2003; Readhead et al. 2004), VSA (Dickinson et al. 2004), ACBAR (Kuo et al. 2004, 2007), and BOOMERanG (Ruhl et al. 2003; Montroy et al. 2006; Piacentini et al. 2006), all of which go well beyond the WMAP angular resolution, so that their small-scale data are statistically independent of the WMAP data.

We find that the small-scale CMB data do not improve the determination of ns because of their relatively large statistical errors. We also find that the calibration and beam errors are important. Let us examine this using the latest ACBAR data (Reichardt et al. 2008). The WMAP+ACBAR yields 0.964 ± 0.014. When the beam error of ACBAR is ignored, we find ns = 0.964 ± 0.013. When the calibration error is ignored, we find ns = 0.962 ± 0.013. Therefore, both the beam and calibration error are important in the error budget of the ACBAR data.

The Big Bang nucleosynthesis (BBN), combined with measurements of the deuterium-to-hydrogen ratio, D/H, from quasar absorption systems, has been extensively used for determining Ωbh2, independent of any other cosmological parameters (see Steigman 2007, for a recent summary). The precision of the latest determination of Ωbh2 from BBN (Pettini et al. 2008) is comparable to that of the WMAP data-only analysis. More precise measurements of D/H will help reduce the correlation between ns and Ωbh2, yielding a better determination of ns.

There is still a bit of correlation left between ns and the electron-scattering optical depth, τ (see Figure 1). While a better measurement of τ from future WMAP observations as well as the Planck satellite should help reduce the uncertainty in ns via a better measurement of τ, the effect of Ωbh2 is much larger.

We find that the other datasets, such as BAO, SN, and the shape of galaxy power spectrum from SDSS or 2dFGRS, do not improve our constraints on ns, as these datasets are not sensitive to Ωbh2 or τ; however, this will change when we include the running index, dns/dln k (Section 3.1.3) and/or the tensor-to-scalar ratio, r (Section 3.2.4).

3.1.3. Results: Running Index

Next, we explore more general models in which a sizable running index may be generated (we still do not include gravitational waves; see Section 3.2 for the analysis that includes gravitational waves). We find no evidence for dns/dln k from WMAP only, −0.090 < dns/dln k < 0.019(95%CL), or WMAP+BAO+SN −0.068 < dns/dln k < 0.012(95%CL). The improvement from WMAP-only to WMAP+BAO+SN is only modest.

We find a slight upward shift from the 3-year result, dns/dln k = −0.055+0.030−0.031 (68% CL; Spergel et al. 2003), to the 5-year result, dns/dln k = −0.037 ± 0.028 (68% CL; WMAP only). This is caused by a combination of three effects:

  • 1.  
    The 3-year number for dns/dln k was derived from an older analysis pipeline for the temperature data, namely the resolution 3 (instead of 4) pixel-based low-l temperature likelihood and a higher point-source amplitude, Aps = 0.017 μK2sr.
  • 2.  
    With 2 years of more integration, we have a better signal-to-noise near the third acoustic peak, whose amplitude is slightly higher than the 3-year determination (Nolta et al. 2009).
  • 3.  
    With the improved beam model (Hill et al. 2009), the temperature power spectrum at l ≳ 200 has been raised nearly uniformly by ∼2.5% (Hill et al. 2009; Nolta et al. 2009).

All of these effects go in the same direction, that is, to increase the power at high multipoles and reduce a negative running index. We find that these factors contribute to the upward shift in dns/dln k at comparable levels.

Note that an upward shift in ns for a power-law model, 0.958 to 0.963 (Section 3.1.2), is not subject to (1) because the 3-year number for ns was derived from an updated analysis pipeline using the resolution 4 low-l code and Aps = 0.014 μK2 sr. We find that (2) and (3) contribute to the shift in ns at comparable levels. An upward shift in σ8 from the 3-year value, 0.761, to the 5-year value, 0.796, can be similarly explained.

We do not find any significant evidence for the running index when the WMAP data and small-scale CMB data (CBI, VSA, ACBAR07, BOOMERanG) are combined, −0.1002 < dns/dln k < −0.0037(95%CL), or the WMAP data and the latest results from the analysis of the complete ACBAR data (Reichardt et al. 2008) are combined, −0.082 < dns/dln k < 0.015(95%CL).

Our best 68% CL constraint from WMAP+BAO+SN shows no evidence for the running index, dns/dln k = −0.028 ± 0.020. In order to improve upon the limit on dns/dln k, one needs to determine ns at small scales, as dns/dln k is simply given by the difference between ns's measured at two different scales, divided by the logarithmic separation between two scales. The Lyα forest provides such information (see Section 7; also Table 4).

3.2. Primordial Gravitational Waves

3.2.1. Motivation

The presence of primordial gravitational waves is a robust prediction of inflation models, as the same mechanism that generated primordial density fluctuations should also generate primordial gravitational waves (Grishchuk 1975; Starobinsky 1979). The amplitude of gravitational waves relative to that of density fluctuations is model-dependent.

The primordial gravitational waves leave their signatures imprinted on the CMB temperature anisotropy (Rubakov et al. 1982; Fabbri & Pollock 1983; Abbott & Wise 1984; Starobinsky 1985; Crittenden et al. 1993), as well as on polarization (Basko & Polnarev 1980; Bond & Efstathiou 1984; Polnarev 1985; Crittenden et al. 1993, 1995; Coulson et al. 1994).20 The spin-2 nature of gravitational waves leads to two types of a polarization pattern on the sky (Zaldarriaga & Seljak 1997; Kamionkowski et al. 1997a): (1) the curl-free mode (E mode), in which the polarization directions are either purely radial or purely tangential to hot/cold spots in temperature, and (2) the divergence-free mode (B mode), in which the pattern formed by polarization directions around hot/cold spots possess nonzero vorticity.

In the usual gravitational instability picture, in the linear regime (before shell crossing of fluid elements), velocity perturbations can be written in terms of solely a gradient of a scalar velocity potential u, $\vec{v}=\vec{\nabla }u$. This means that no vorticity would arise, and, therefore, no B mode polarization can be generated from density or velocity perturbations in the linear regime. However, primordial gravitational waves can generate both E and B mode polarization; thus, the B mode polarization offers a smoking-gun signature for the presence of primordial gravitational waves (Seljak & Zaldarriaga 1997; Kamionkowski et al. 1997a). This gives us a strong motivation to look for signatures of the primordial gravitational waves in CMB.

In Table 4, we summarize our constraints on the amplitude of gravitational waves, expressed in terms of the tensor-to-scalar ratio, r, defined by Equation (20).

3.2.2. Analysis

We quantify the amplitude and spectrum of primordial gravitational waves in the following form:

Equation (16)

where we have ignored a possible scale dependence of nt(k), as the current data cannot constrain it. Here, by Ph(k), we mean

Equation (17)

where $\tilde{h}_{ij}({\bf k})$ is the Fourier transform of the tensor metric perturbations, gij = a2ij + hij), which can be further decomposed into the two polarization states (+ and ×) with the appropriate polarization tensor, e(+,×)ij, as

Equation (18)

with the normalization that e+ije+,ij = e×ije×,ij = 2 and e+ije×,ij = 0. Unless there was a parity-violating interaction term such as $f(\phi) R_{\mu \nu \rho \sigma }\tilde{R}^{\mu \nu \rho \sigma }$, where f(ϕ) is an arbitrary function of a scalar field, Rμνρσ is the Riemann tensor, and $\tilde{R}^{\mu \nu \rho \sigma }$ is a dual tensor (Lue et al. 1999), both polarization states are statistically independent with the same amplitude, meaning

Equation (19)

This implies that parity-violating correlations, such as the TB and EB correlations, must vanish. We shall explore such parity-violating correlations in Section 4 in a slightly different context. For limits on the difference between $\langle |\tilde{h}_+|^2\rangle$ and $\langle |\tilde{h}_\times |^2\rangle$ from, respectively, the TB and EB spectra of the WMAP 3-year data, see Saito et al. (2007).

In any case, this definition suggests that Ph(k) is given by $P_h(k) = 4\langle |\tilde{h}|^2\rangle$. Notice a factor of 4. This is the definition of Ph(k) that we have been consistently using since the first year release (Peiris et al. 2003; Spergel et al. 2003, 2007; Page et al. 2007). We continue to use this definition.

With this definition of Δ2h(k) (Equation (16)), we define the tensor-to-scalar ratio, r, at k = k0, as

Equation (20)

where $\Delta _{\cal R}(k)$ is the curvature perturbation spectrum given by Equation (15). We shall take k0 to be 0.002 Mpc−1. In this paper, we sometimes loosely call this quantity the "amplitude of gravitational waves."

What about the tensor spectral tilt, nt? In single-field inflation models, there exists the so-called consistency relation between r and nt (see Liddle & Lyth 2000, for a review)

Equation (21)

In order to reduce the number of parameters, we continue to impose this relation at k = k0 = 0.002 Mpc−1. For a discussion on how to impose this constraint in a more self-consistent way, see Peiris & Easther (2006a).

To constrain r, we shall use the WMAP 5-year temperature and polarization data, the small-scale CMB data, and/or BAO and SN distance measurements.

3.2.3. How WMAP Constrains the Amplitude of Gravitational Waves

Let us show how the gravitational wave contribution is constrained by the WMAP data (see Figure 3). In this pedagogical analysis, we vary only r and τ, while adjusting the overall amplitude of fluctuations such that the height of the first peak of the temperature power spectrum is always held fixed. All the other cosmological parameters are fixed at Ωk = 0, Ωbh2 = 0.02265, Ωch2 = 0.1143, H0 = 70.1 km s-1 Mpc−1, and ns = 0.960. Note that the limit on r from this analysis should not be taken as our best limit, as this analysis ignores the correlation between r and the other cosmological parameters, especially ns. The limit on r from the full analysis will be given in Section 3.2.4.

  • 1.  
    (The gray contours in the left panel and the upper right of the right panel of Figure 3.) The low-l polarization data (TE/EE/BB) at l ≲ 10 are unable to place meaningful limits on r. A large r can be compensated by a small τ, producing nearly the same EE power spectrum at l ≲ 10. (Recall that the gravitational waves also contribute to EE.) As a result, r that is as large as 10 is still allowed within 68% CL.21
  • 2.  
    (The red contours in the left panel and the lower left of the right panel of Figure 3.) Such a high value of r, however, produces too negative a TE correlation between 30 ≲ l ≲ 150. Therefore, we can improve the limit on r significantly—by nearly an order of magnitude—by simply using the high-l TE data. The 95% upper bound at this point is still as large as r ∼ 2.22
  • 3.  
    (The blue contours in the left panel and the upper left of the right panel of Figure 3.) Finally, the low-l temperature data at l ≲ 30 severely limit the excess low-l power due to gravitational waves, bringing the upper bound down to r ∼ 0.2. Note that this bound is about a half of what we actually obtain from the full MCMC of the WMAP-only analysis, r < 0.43(95%CL). This is because we have fixed ns, and thus ignored the correlation between r and ns shown by Figure 2.
Figure 3.

Figure 3. How the WMAP temperature and polarization data constrain the tensor-to-scalar ratio, r. (Left) The contours show 68% and 95% CL. The gray region is derived from the low-l polarization data (TE/EE/BB at l ⩽ 23) only, the red region from the low-l polarization plus the high-l TE data at l ⩽ 450, and the blue region from the low-l polarization, the high-l TE, and the low-l temperature data at l ⩽ 32. (Right) The gray curves show (r, τ) = (10, 0.050), the red curves (r, τ) = (1.2, 0.075), and the blue curves (r, τ) = (0.20, 0.080), which are combinations of r and τ that give the upper edge of the 68% CL contours shown on the left panel. The vertical lines indicate the maximum multipoles below which the data are used for each color. The data points with 68% CL errors are the WMAP 5-year measurements (Nolta et al. 2009). Note that the BB power spectrum at l ∼ 130 is consistent with zero within 95% CL.

Standard image High-resolution image

3.2.4. Results

Having understood which parts of the temperature and polarization spectra constrain r, we obtain the upper limit on r from the full exploration of the likelihood space using the MCMC. Figure 2 shows the one-dimensional constraint on r and the two-dimensional constraint on r and ns, assuming a negligible running index. With the WMAP 5-year data alone, we find r < 0.43(95%CL). Since the B-mode contributes little here, and most of the information essentially comes from TT and TE, our limit on r is highly degenerate with ns, and thus we can obtain a better limit on r only when we have a better limit on ns.

When we add BAO and SN data, the limit improves significantly to r < 0.22(95%CL). This is because the distance information from BAO and SN reduces the uncertainty in the matter density, and thus it helps to determine ns better because ns is also degenerate with the matter density. This "chain of correlations" helped us improve our limit on r significantly from the previous results. This limit, r < 0.22(95%CL), is the best limit on r to date.23 With the new data, we are able to get more stringent limits than our earlier analyses (Spergel et al. 2007) that combined the WMAP data with measurements of the galaxy power spectrum and found r < 0.30 (95% CL).

A dramatic reduction in the uncertainty in r has an important implication for ns as well. Previously, ns > 1 was within the 95% CL when the gravitational wave contribution was allowed, owing to the correlation between ns and r. Now, we are beginning to disfavor ns > 1 even when r is nonzero: with WMAP+BAO+SN, we find −0.0022 < 1 − ns < 0.0589(95%CL).24

However, these stringent limits on r and ns weaken to −0.246 < 1 − ns < 0.034(95%CL) and r < 0.55(95%CL) when a sizable running index is allowed. The BAO and SN data helped reduce the uncertainty in dns/dln k (Figure 4), but not enough to improve on the other parameters compared to the WMAP-only constraints. The Lyα forest data can improve the limit on dns/dln k even when r is present (see Section 7; also Table 4).

Figure 4.

Figure 4. Constraint on the tensor-to-scalar ratio, r, the tilt, ns, and the running index, dns/dln k, when all of them are allowed to vary (Section 3.2.4). In all panels, we show the WMAP-only results in blue and WMAP+BAO+SN in red. (Left) Joint two-dimensional marginalized distribution of ns and r at k = 0.002 Mpc−1 (68% and 95% CL). (Middle) ns and dns/dln k. (Right) dns/dln k and r. We find no evidence for the running index. While the inclusion of the running index weakens our constraint on ns and r, the data do not support any need for treating the running index as a free parameter: changes in χ2 between the power-law model and the running model are χ2(running) − χ2(power-law) ≃ −1.8 with and without the tensor modes for WMAP5+BAO+SN, and 1.2 for WMAP5.

Standard image High-resolution image

3.3. Implications for Inflation Models

How do the WMAP 5-year limits on ns and r constrain inflationary models?25 In the context of single-field models, one can write down ns and r in terms of the derivatives of potential, V(ϕ), as (Liddle & Lyth 2000)

Equation (22)

Equation (23)

where $M_{pl}=1/\sqrt{8\pi G}$ is the reduced Planck mass, and the derivatives are evaluated at the mean value of the scalar field at the time that a given scale leaves the horizon. These equations may be combined to give a relation between ns and r:

Equation (24)

This equation indicates that it is the curvature of the potential that divides models on the ns-r plane; thus, it makes sense to classify inflation models on the basis of the sign and magnitude of the curvature of the potential (Peiris et al. 2003).26

What is the implication of our bound on r for inflation models? Equation (24) suggests that a large r can be generated when the curvature of the potential is positive, that is, V''>0, at the field value that corresponds to the scales probed by the WMAP data. Therefore, it is a set of positive curvature models that we can constrain from the limit on r. However, negative curvature models are more difficult to constrain from r, as they tend to predict small r (Peiris et al. 2003). We shall not discuss negative curvature models in this paper: many of these models, including those based upon the Coleman–Weinberg potential, fit the WMAP data (see, e.g., Dvali et al. 1994; Shafi & Senoguz 2006).

Here we shall pick three simple, but representative, forms of V(ϕ) that can produce V''>0:27

  • 1.  
    Monomial (chaotic-type) potential, V(ϕ) ∝ ϕα. This form of the potential was proposed by, and is best known for, Linde's chaotic inflation models (Linde 1983). This model also approximates a pseudo Nambu–Goldstone boson potential (natural inflation; Freese et al. 1990; Adams et al. 1993) with the negative sign, V(ϕ) ∝ 1 − cos(ϕ/f), when ϕ/f ≪ 1, or with the positive sign, V(ϕ) ∝ 1 + cos(ϕ/f), when ϕ/f ∼ 1.28 This model can also approximate the Landau–Ginzburg type of spontaneous symmetry breaking potential, V(ϕ) ∝ (ϕ2v2)2, in the appropriate limits.
  • 2.  
    Exponential potential, $V(\phi)\propto \exp [-(\phi /M_{pl})\sqrt{2/p}]$. A unique feature of this potential is that the dynamics of inflation is exactly solvable, and the solution is a power-law expansion, a(t) ∝ tp, rather than an exponential one. For this reason, this type of model is called power-law inflation (Abbott & Wise 1984; Lucchin & Matarrese 1985). They often appear in models of scalar-tensor theories of gravity (Accetta et al. 1985; La & Steinhardt 1989; Futamase & Maeda 1989; Steinhardt & Accetta 1990; Kalara et al. 1990).
  • 3.  
    ϕ2 plus vacuum energy, V(ϕ) = V0 +  m2ϕ2/2. These models are known as Linde's hybrid inflation (Linde 1994). This model is a "hybrid" because the potential combines the chaotic-type (with α = 2) with a Higgs-like potential for the second field (which is not shown here). This model behaves as if it were a single-field model until the second field terminates inflation when ϕ reaches some critical value. When ϕ ≫ (2V0)1/2/m, this model is the same as model 1 with α = 2, although one of Linde's motivation was to avoid having such a large field value that exceeds Mpl.

These potentials29 make the following predictions for r and ns as a function of their parameters:

  • 1.  
    $r=8(1-n_s)\frac{\alpha }{\alpha +2}$
  • 2.  
    r = 8(1 − ns)
  • 3.  
    $r=8(1-n_s)\frac{\tilde{\phi }^2}{2\tilde{\phi }^2-1}$.

Here, for 3, we have defined a dimensionless variable, $\tilde{\phi }\equiv m\phi /(2V_0)^{1/2}$. This model approaches model 1 with α = 2 for $\tilde{\phi }\gg 1$ and yields the scale-invariant spectrum, ns = 1, when $\tilde{\phi }=1/\sqrt{2}$.

We summarize our findings below, and in Figure 5.

Figure 5.

Figure 5. Constraint on three representative inflation models whose potential is positively curved, V''>0 (Section 3.3). The contours show the 68% and 95% CL derived from WMAP+BAO+SN. (Top) The monomial, chaotic-type potential, V(ϕ) ∝ ϕα (Linde 1983), with α = 4 (solid) and α = 2 (dashed) for single-field models, and α = 2 for multiaxion field models with β = 1/2 (Easther & McAllister 2006; dotted). The symbols show the predictions from each of these models with the number of e-folds of inflation equal to 50 and 60. The λϕ4 potential is excluded convincingly, the m2ϕ2 single-field model lies outside of (at the boundary of) the 68% region for N = 50 (60), and the m2ϕ2 multiaxion model with N = 50 lies outside of the 95% region. (Middle) The exponential potential, $V(\phi)\propto \exp [-(\phi /M_{pl})\sqrt{2/p}]$, which leads to a power-law inflation, a(t) ∝ tp (Abbott & Wise 1984; Lucchin & Matarrese 1985). All models but p ∼ 120 are outside of the 68% region. The models with p < 60 are excluded at more than 99% CL, and those with p < 70 are outside of the 95% region. For multifield models these limits can be translated into the number of fields as pnpi, where pi is the p-parameter of each field (Liddle et al. 1998). The data favor n ∼ 120/pi fields. (Bottom) The hybrid-type potential, $V(\phi)=V_0+(1/2)m^2\phi ^2=V_0(1+\tilde{\phi }^2)$, where $\tilde{\phi }\equiv m\phi /(2V_0)^{1/2}$ (Linde 1994). The models with $\tilde{\phi }<2/3$ drive inflation by the vacuum energy term, V0, and are disfavored at more than 95% CL, while those with $\tilde{\phi }>1$ drive inflation by the quadratic term, and are similar to the chaotic type (the left panel with α = 2). The transition regime, $2/3<\tilde{\phi }<1$ are outside of the 68% region, but still within the 95% region.

Standard image High-resolution image

  • 1.  
    Assuming that the monomial potentials are valid to the end of inflation including the reheating of the universe, one can relate ns and r to the number of e-folds of inflation, N ≡ ln(aend/aWMAP), between the expansion factors at the end of inflation, aend, and the epoch when the wavelength of fluctuations that we probe with WMAP leave the horizon during inflation, aWMAP. The relations are (Liddle & Lyth 2000)
    Equation (25)
    We take N = 50 and 60 as a reasonable range (Liddle & Leach 2003). For α = 4, that is, inflation by a massless self-interacting scalar field V(ϕ) = (λ/4)ϕ4, we find that both N = 50 and 60 are far away from the 95% region and they are excluded convincingly at more than 99% CL. For α = 2, that is, inflation by a massive free scalar field V(ϕ) = (1/2)m2ϕ2, the model with N = 50 lies outside of the 68% region, whereas the model with N = 60 is at the boundary of the 68% region. Therefore, both of these models are consistent with the data within the 95% CL. While this limit applies to a single massive free field, Easther & McAllister (2006) showed that a model with many massive axion fields (N-flation model; Dimopoulos et al. 2005) can shift the predicted ns further away from unity,
    Equation (26)
    where "N.f" refers to "N fields" and "s.f." to "single field," and β is a free parameter of the model. Easther & McAllister (2006) argued that β ∼ 1/2 is favored, for which 1 − ns is larger than the single-field prediction by as much as 25%. The prediction for the tensor-to-scalar ratio, r, is the same as the single-field case (Alabidi & Lyth 2006b). Therefore, this model lies outside of the 95% region for N = 50. As usual, however, these monomial potentials can be made a better fit to the data by invoking a nonminimal coupling between the inflaton and gravity, as the nonminimal coupling can reduce r to negligible levels (Komatsu & Futamase 1999; Hwang & Noh 1998; Tsujikawa & Gumjudpai 2004). Piao (2006) has shown that N-flation models with monomial potentials, V(ϕ) ∝ ϕα, generically predict ns that is smaller than the corresponding single-field predictions.
  • 2.  
    For an exponential potential, r and ns are uniquely determined by a single parameter, p, that determines a power-law index of the expansion factor, a(t) ∝ tp, as
    Equation (27)
    We find that p < 60 is excluded at more than 99% CL, 60 < p < 70 is within the 99% region but outside of the 95% region, and p > 70 is within the 95% region. The models with p ∼ 120 lie on the boundary of the 68% region, but other parameters are not within the 68% CL. This model can be thought of as a single-field inflation with p ≫ 1, or multifield inflation with n fields, each having pi ∼ 1 or even pi < 1 (assisted inflation; Liddle et al. 1998). In this context, therefore, one can translate the above limits on p into the limits on the number of fields. The data favor n ∼ 120/pi fields.
  • 3.  
    For this model, we can divide the parameter space into three regions, depending upon the value of $\tilde{\phi }$ that corresponds to the field value when the wavelength of fluctuations that we probe with WMAP leaves the horizon. When $\tilde{\phi }\ll 1$, the potential is dominated by a constant term, which we call "Flat Potential Regime." When $\tilde{\phi }\gg 1$, the potential is indistinguishable from the chaotic-type (model 1) with α = 2. We call this region "Chaotic Inflation-like Regime." When $\tilde{\phi }\sim 1$, the model shows a transitional behavior, and thus we call it "Transition Regime." We find that the flat potential regime with $\tilde{\phi }\lesssim 2/3$ lies outside of the 95% region. The transition regime with $2/3\lesssim \tilde{\phi }\lesssim 1$ is within the 95% region, but outside of the 68% region. Finally, the chaotic-like regime contains the 68% region. Since inflation in this model ends by the second field whose dynamics depends on other parameters, there is no constraint from the number of e-folds.

These examples show that the WMAP 5-year data, combined with the distance information from BAO and SN, begin to disfavor a number of popular inflation models.

3.4. Curvature of the Observable Universe

3.4.1. Motivation

The flatness of the observable universe is one of the predictions of conventional inflation models. How much curvature can we expect from inflation? The common view is that inflation naturally produces the spatial curvature parameter, Ωk, on the order of the magnitude of quantum fluctuations, that is, Ωk ∼ 10−5. However, the current limit on Ωk is of order 10−2; thus, the current data are not capable of reaching the level of Ωk that is predicted by the common view.

Would a detection of Ωk rule out inflation? It is possible that the value of Ωk is just below our current detection limit, even within the context of inflation: inflation may not have lasted for so long, and the curvature radius of our universe may just be large enough for us not to see the evidence for curvature within our measurement accuracy, yet. While this sounds like fine-tuning, it is a possibility.

This is something we can test by constraining Ωk better. There is also a revived (and growing) interest in measurements of Ωk, as Ωk is degenerate with the equation of state of dark energy, w. Therefore, a better determination of Ωk has an important implication for our ability to constrain the nature of dark energy.

3.4.2. Analysis

Measurements of the CMB power spectrum alone do not strongly constrain Ωk. More precisely, any experiments that measure the angular diameter or luminosity distance to a single redshift are not able to constrain Ωk uniquely, as the distance depends not only on Ωk, but also on the expansion history of the universe. For a universe containing matter and vacuum energy, it is essential to combine at least two absolute distance indicators, or the expansion rates, out to different redshifts, in order to constrain the spatial curvature well. Note that CMB is also sensitive to ΩΛ, via the late-time integrated Sachs–Wolfe (ISW) effect, and to Ωm, via the signatures of gravitational lensing in the CMB power spectrum. These properties can be used to reduce the correlation between Ωk and Ωm (Stompor & Efstathiou 1999) or ΩΛ (Ho et al. 2008; Giannantonio et al. 2008).

It has been pointed out by a number of people (e.g., Eisenstein et al. 2005) that a combination of distance measurements from BAO and CMB is a powerful way to constrain Ωk. One needs more distances, if dark energy is not a constant but dynamical.

In this section, we shall make one important assumption that the dark energy component is vacuum energy, that is, a cosmological constant. In Section 5, we shall study the case in which the equation of state, w, and Ωk are varied simultaneously.

3.4.3. Results

Figure 6 shows the limits on ΩΛ and Ωk. While the WMAP data alone cannot constrain Ωk (see the left panel), the WMAP data combined with the HST's constraint on H0 tighten the constraint significantly, to −0.052 < Ωk < 0.013(95%CL). The WMAP data combined with SN yield ∼50% better limits, −0.0316 < Ωk < 0.0078(95%CL), compared to WMAP+HST. Finally, the WMAP+BAO yields the smallest statistical uncertainty, −0.0170 < Ωk < 0.0068(95%CL), which is a factor of 2.6 and 1.7 better than WMAP+HST and WMAP+SN, respectively. This shows how powerful the BAO is in terms of constraining the spatial curvature of the universe; however, this statement needs to be re-evaluated when dynamical dark energy is considered, for example, w ≠ −1. We shall come back to this point in Section 5.

Figure 6.

Figure 6. Joint two-dimensional marginalized constraint on the vacuum energy density, ΩΛ, and the spatial curvature parameter, Ωk (Section 3.4.3). The contours show the 68% and 95% CL. (Left) The WMAP-only constraint (light blue) compared with WMAP+BAO+SN (purple). Note that we have a prior on ΩΛ, ΩΛ > 0. This figure shows how powerful the extra distance information is for constraining Ωk. (Middle) A blow-up of the region within the dashed lines in the left panel, showing WMAP-only (light blue), WMAP+HST (gray), WMAP+SN (dark blue), and WMAP+BAO (red). The BAO provides the most stringent constraint on Ωk. (Right) One-dimensional marginalized constraint on Ωk from WMAP+HST, WMAP+SN, and WMAP+BAO. We find the best limit, −0.0178 < Ωk < 0.0066(95%CL), from WMAP+BAO+SN, which is essentially the same as WMAP+BAO. See Figure 12 for the constraints on Ωk when dark energy is dynamical, that is, w ≠ −1, with time-independent w. Note that neither BAO nor SN alone is able to constrain Ωk: they need the WMAP data for lifting the degeneracy. Also note that BAO+SN is unable to lift the degeneracy either, as BAO needs the sound horizon size measured by the WMAP data.

Standard image High-resolution image

Finally, when WMAP, BAO, and SN are combined, we find −0.0178 < Ωk < 0.0066(95%CL). As one can see from the right panel of Figure 6, the constraint on Ωk is totally dominated by that from WMAP+BAO; thus, the size of the uncertainty does not change very much from WMAP+BAO to WMAP+BAO+SN. Note that the above result indicates that we have reached 1.3% accuracy (95% CL) in determining Ωk, which is rather good. The future BAO surveys at z ∼ 3 are expected to yield an order of magnitude better determination, that is, 0.1% level, of Ωk (Knox 2006).

It is instructive to convert our limit on Ωk to the limits on the curvature radius of the universe. As Ωk is defined as Ωk = −kc2/(H20R2curv), where Rcurv is the present-day curvature radius, one can convert the upper bounds on Ωk into the lower bounds on Rcurv, as $R_{\rm curv}=(c/H_0)/\sqrt{\left|\Omega _k\right|}=3/\sqrt{\left|\Omega _k\right|}\,h^{-1}\,{\rm Gpc}$. For negatively curved universes, we find Rcurv > 37 h−1 Gpc, whereas for positively curved universes, Rcurv > 22 h−1 Gpc. Incidentally these values are greater than the particle horizon at present, 9.7 h−1 Gpc (computed for the same model).

The 68% limits from the 3-year data (Spergel et al. 2007) were Ωk = −0.012 ± 0.010 from WMAP-3 yr+BAO (where BAO is from the SDSS LRG of Eisenstein et al. (2005)), and Ωk = −0.011 ± 0.011 from WMAP-3 yr+SN (where SN is from the SNLS data of Astier et al. 2006). The 68% limit from WMAP-5 yr+BAO+SN (where both BAO and SN have more data than for the 3-year analysis) is Ωk = −0.0050+0.0061−0.0060. A significant improvement in the constraint is due to a combination of the better WMAP, BAO, and SN data.

We conclude that, if dark energy is vacuum energy (cosmological constant) with w = −1, we do not find any deviation from a spatially flat universe.

3.4.4. Implications for the Duration of Inflation

What does this imply for inflation? Since we do not detect any finite curvature radius, inflation had to last for a long enough period in order to make the observable universe sufficiently flat within the observational limits. This argument allows us to find a lower bound on the total number of e-foldings of the expansion factor during inflation, from the beginning to the end, Ntot ≡ ln(aend/abegin) (see also Section 4.1 of Weinberg 2008).

When the curvature parameter, Ωk, is much smaller than unity, it evolves with the expansion factor, a, as Ωka−2, a2, and a during inflation, radiation, and matter era, respectively. Therefore, the observed Ωk is related to Ωk at the beginning of inflation as30

Equation (28)

Equation (29)

where g* is the effective number of relativistic dof contributing to entropy, zeq is the matter–radiation equality redshift, and Tend and Teq are the reheating temperature of the universe at the end of inflation31 and the temperature at the equality epoch, respectively. To within 10% accuracy, we take 1 + zeq = 3200, Teq = 0.75 eV, and g*,eq = 3.9. We find

Equation (30)

Here, it is plausible that g*,end ∼ 100 in the standard model of elementary particles, and ∼200 when the supersymmetric partners are included. The difference between these two cases gives the error of only ΔNtot = −0.2, and thus can be ignored.

The curvature parameter at the beginning of inflation must be below of order unity, as inflation would not begin otherwise. However, it is plausible that Ωbegink was not too much smaller than 1; otherwise, we have to explain why it was so small before inflation, and probably we would have to explain it by inflation before inflation. In that case, Ntot would refer to the sum of the number of e-foldings from two periods of inflation. From this argument, we shall take Ωbegink ∼ 1.

The reheating temperature can be anywhere between 1 MeV and 1016 GeV. It is more likely that it is between 1 TeV and 108 GeV for various reasons, but the allowed region is still large. If we scale the result to a reasonably conservative lower limit on the reheating temperature, Tend ∼ 1 TeV, then we find, from our limit on the curvature of the universe,

Equation (31)

A factor of 10 improvement in the upper limit on |Ωbegink| will raise this limit by ΔNtot = 1.2.

Again, Ntot here refers to the total number of e-foldings of inflation. In Section 3.3, we use N ≡ ln(aend/aWMAP), which is the number of e-foldings between the end of inflation and the epoch when the wavelength of fluctuations that we probe with WMAP leaves the horizon during inflation. Therefore, by definition, N is less than Ntot.

3.5. Primordial Non-Gaussianity

3.5.1. Motivation and Background

In the simplest model of inflation, the distribution of primordial fluctuations is close to a Gaussian with random phases. The level of deviation from a Gaussian distribution and random phases, called non-Gaussianity, predicted by the simplest model of inflation is well below the current limit of measurement. Thus, any detection of non-Gaussianity would be a significant challenge to the currently favored models of the early universe.

The assumption of Gaussianity is motivated by the following view: the probability distribution of quantum fluctuations, P(φ), of free scalar fields in the ground state of the Bunch–Davies vacuum, φ, is a Gaussian distribution; thus, the probability distribution of primordial curvature perturbations (in the comoving gauge), ${\cal R}$, generated from φ (in the flat gauge) as ${\cal R}=-[H(\phi)/\dot{\phi _0}]\varphi$ (Mukhanov & Chibisov 1981; Hawking 1982; Starobinsky 1982; Guth & Pi 1982; Bardeen et al. 1983), would also be a Gaussian distribution. Here, H(ϕ) is the expansion rate during inflation and ϕ0 is the mean field, that is, ϕ = ϕ0 + φ.

This argument suggests that non-Gaussianity can be generated when (a) scalar fields are not free, but have some interactions, (b) there are nonlinear corrections to the relation between ${\cal R}$ and φ, and (c) the initial state is not in the Bunch–Davies vacuum.

For (a), one can think of expanding a general scalar field potential V(ϕ) to the cubic order or higher, $V(\phi)=\bar{V}+V^{\prime }\varphi +(1/2)V^{\prime \prime }\varphi ^2+(1/6)V^{\prime \prime \prime }\varphi ^3+\cdots$. The cubic (or higher-order) interaction terms can yield non-Gaussianity in φ (Falk et al. 1993). When perturbations in gravitational fields are included, there are many more interaction terms that arise from expanding the Ricci scalar to the cubic order, with coefficients containing derivatives of V and ϕ0, such as $\dot{\phi _0}V^{\prime \prime }, \dot{\phi _0}^3/H$, etc. (Maldacena 2003).

For (b), one can think of this relation, ${\cal R}=-[H(\phi)/\dot{\phi _0}]\varphi$, as the leading-order term of a Taylor series expansion of the underlying nonlinear (gauge) transformation law between ${\cal R}$ and φ. Salopek & Bond (1990) showed that, in the single-field models, ${\cal R}=4\pi G\int _{\phi _0}^{\phi _0+\varphi }d\phi \left(\partial \ln H/\partial \phi \right)^{-1}$. Therefore, even if φ is precisely Gaussian, ${\cal R}$ can be non-Gaussian due to nonlinear terms such as φ2 in a Taylor series expansion of this relation. One can write this relation in the following form, up to second order in ${\cal R}$,

Equation (32)

where ${\cal R}_{\rm L}$ is a linear part of the curvature perturbation. We thus find that the second term makes ${\cal R}$ non-Gaussian, even when ${\cal R}_{\rm L}$ is precisely Gaussian. This formula has also been found independently by other researchers, extended to multifield cases, and often referred to as the "δN formalism" (Sasaki & Stewart 1996; Lyth et al. 2005; Lyth & Rodriguez 2005).

The observers like us, however, do not measure the primordial curvature perturbations, ${\cal R}$, directly. A more observationally-relevant quantity is the curvature perturbation during the matter era, Φ. At the linear order, these quantities are related by $\Phi =(3/5){\cal R}_{\rm L}$ (e.g., Kodama & Sasaki 1984), but the actual relation is more complicated at the nonlinear order (see Bartolo et al. 2004, for a review). In any case, this argument has motivated our defining the "local nonlinear coupling parameter," flocalNL, as (Komatsu & Spergel 2001)32

Equation (33)

If we take equation (32), for example, we find flocalNL = −(5/24πG)(∂2ln H/∂ϕ2) (Komatsu 2001). Here, we have followed the terminology proposed by Babich et al. (2004) and called flocalNL the "local" parameter, as both sides of Equation (33) are evaluated at the same location in space. (Hence, the term "local.")

Let us comment on the magnitude of the second term in Equation (33). Since Φ ∼ 10−5, the second term is smaller than the first term by 10−5flocalNL; thus, the second term is only 0.1% of the first term for flocalNL ∼ 102. As we shall see below, the existing limits on flocalNL have already reached this level of Gaussianity, and thus it is clear that we are already talking about a tiny deviation from Gaussian fluctuations. This limit is actually better than the current limit on the spatial curvature, which is only on the order of 1%. Therefore, Gaussianity tests offer a stringent test of early universe models.

In the context of single-field inflation in which the scalar field slowly rolls down the potential, the quantities, H, V, and ϕ, change slowly. Therefore, one generically expects that flocalNL is small, on the order of the so-called slow-roll parameters epsilon and η, which are typically of order 10−2 or smaller. In this sense, the single-field, slow-roll inflation models are expected to result in a tiny amount of non-Gaussianity (Salopek & Bond 1990; Falk et al. 1993; Gangui et al. 1994; Maldacena 2003; Acquaviva et al. 2003). These contributions from the epoch of inflation are much smaller than those from the ubiquitous, second-order cosmological perturbations, that is, the nonlinear corrections to the relation between Φ and ${\cal R}$, which result in flocalNL of order unity (Liguori et al. 2006; Smith & Zaldarriaga 2006). Also see Bartolo et al. (2004) for a review on this subject.

One can use the cosmological observations, such as the CMB data, to constrain flocalNL (Verde et al. 2000; Komatsu & Spergel 2001). While the temperature anisotropy, ΔT/T, is related to Φ via the Sachs–Wolfe formula as ΔT/T = −Φ/3 at the linear order on very large angular scales (Sachs & Wolfe 1967), there are nonlinear corrections (nonlinear Sachs–Wolfe effect, nonlinear Integrated Sachs–Wolfe effect, gravitational lensing, etc.) to this relation, which add terms of order unity to flocalNL by the time we observe it in CMB (Pyne & Carroll 1996; Mollerach & Matarrese 1997). On smaller angular scales, one must include the effects of acoustic oscillations of photon-baryon plasma by solving the Boltzmann equations. The second-order corrections to the Boltzmann equations can also yield flocalNL of order unity (Bartolo et al. 2006, 2007).

Any detection of flocalNL at the level that is currently accessible would have a profound implication for the physics of inflation. How can a large flocalNL be generated? We essentially need to break either (a) single field or (b) slow-roll. For example, a multifield model, known as the curvaton scenario, can result in much larger values of flocalNL (Linde & Mukhanov 1997; Lyth et al. 2003), and so can the models with field-dependent (variable) decay widths for reheating of the universe after inflation (Dvali et al. 2004b, 2004a). A more violent, nonlinear reheating process called "preheating" can give rise to a large flocalNL (Enqvist et al. 2005; Jokinen & Mazumdar 2006; Chambers & Rajantie 2008).

Although breaking of slow-roll usually results in a premature termination of inflation, it is possible to break it temporarily for a brief period, without terminating inflation, by some features (steps, dips, etc.) in the shape of the potential. In such a scenario, a large non-Gaussianity may be generated at a certain limited scale at which the feature exists (Kofman et al. 1991; Wang & Kamionkowski 2000; Komatsu et al. 2003). The structure of non-Gaussianity from features is much more complex and model-dependent than flocalNL (Chen et al. 2007a, 2008).

There is also a possibility that non-Gaussianity can be used to test alternatives to inflation. In a collapsing universe followed by a bounce (e.g., new Ekpyrotic scenario), flocalNL is given by the inverse (as well as inverse-squared) of slow-roll parameters; thus, flocalNL as large as of order 10–102 is a fairly generic prediction of this class of models (Creminelli & Senatore 2007; Koyama et al. 2007; Buchbinder et al. 2008; Lehners & Steinhardt 2008a, 2008b).

Using the angular bispectrum,33 the harmonic transform of the angular three-point correlation function, Komatsu et al. (2002) have obtained the first observational limit on flocalNL from the COBE 4-year data (Bennett et al. 1996), finding −3500 < flocalNL < 2000 (95% CL). The uncertainty was large due to a relatively large beam size of COBE, which allowed us to go only to the maximum multipole of lmax = 20. Since the signal-to-noise ratio of flocalNL is proportional to lmax, it was expected that the WMAP data would yield a factor of ∼50 improvement over the COBE data (Komatsu & Spergel 2001).

The full bispectrum analysis was not feasible with the WMAP data, as the computational cost scales as N5/2pix, where Npix is the number of pixels, which is on the order of millions for the WMAP data. The "KSW" estimator (Komatsu et al. 2005) has solved this problem by inventing a cubic statistic that combines the triangle configurations of the bispectrum optimally so that it is maximally sensitive to flocalNL.34 The computational cost of the KSW estimator scales as N3/2pix. We give a detailed description of the method that we use in this paper in Appendix A.

We have applied this technique to the WMAP 1-year and 3-year data, and found −58 < flocalNL < 134 (l = 265; Komatsu et al. 2003) and −54 < flocalNL < 114 (lmax = 350; Spergel et al. 2007), respectively, at 95% CL. Creminelli et al. performed an independent analysis of the WMAP data and found similar limits: −27 < flocalNL < 121 (lmax = 335; Creminelli et al. 2006) and −36 < flocalNL < 100 (l = 370; Creminelli et al. 2007) for the 1-year and 3-year data, respectively. These constraints are slightly better than the WMAP team's, as their estimator for flocalNL was improved from the original KSW estimator.

While these constraints are obtained from the KSW-like fast bispectrum statistics, many groups have used the WMAP data to measure flocalNL using various other statistics, such as the Minkowski functionals (Komatsu et al. 2003; Spergel et al. 2007; Gott et al. 2007; Hikage et al. 2008), real-space three-point function (Gaztañaga & Wagg 2003; Chen & Szapudi 2005), integrated bispectrum (Cabella et al. 2006), 2-1 cumulant correlator power spectrum (Chen & Szapudi 2006), local curvature (Cabella et al. 2004), and spherical Mexican hat wavelet (Mukherjee & Wang 2004). The suborbital CMB experiments have also yielded constraints on flocalNL: MAXIMA (Santos et al. 2003), VSA (Smith et al. 2004), Archeops (Curto et al. 2007), and BOOMERanG (De Troia et al. 2007).

We stress that it is important to use different statistical tools to measure flocalNL if any signal is found, as different tools are sensitive to different systematics. The analytical predictions for the Minkowski functionals (Hikage et al. 2006) and the angular trispectrum (the harmonic transform of the angular 4-point correlation function; Okamoto & Hu 2002; Kogo & Komatsu 2006) as a function of flocalNL are now available. Studies on the forms of the trispectrum from inflation models have just begun, and some important insights have been obtained (Boubekeur & Lyth 2006; Huang & Shiu 2006; Byrnes et al. 2006; Seery & Lidsey 2007; Seery et al. 2007; Arroja & Koyama 2008). It is now understood that the trispectrum is at least as important as the bispectrum in discriminating inflation models: some models do not produce any bispectra but produce significant trispectra, and other models produce similar amplitudes of the bispectra but produce very different trispectra (Huang & Shiu 2006; Buchbinder et al. 2008).

In this paper, we shall use the estimator that further improves upon Creminelli et al. (2006) by correcting an inadvertent numerical error of a factor of 2 in their derivation (Yadav et al. 2008). Yadav & Wandelt (2008) used this estimator to measure flocalNL from the WMAP 3-year data. We shall also use the Minkowski functionals to find a limit on flocalNL.

In addition to flocalNL, we shall also estimate the "equilateral nonlinear coupling parameter," fequilNL, which characterizes the amplitude of the three-point function (i.e., the bispectrum) of the equilateral configurations, in which the lengths of all the three wave vectors forming a triangle in Fourier space are equal. This parameter is useful and highly complementary to the local one: while flocalNL mainly characterizes the amplitude of the bispectrum of the squeezed configurations, in which two wave vectors are large and nearly equal and the other wave vector is small, and thus it is fairly insensitive to the equilateral configurations, fequilNL is mainly sensitive to the equilateral configurations with little sensitivity to the squeezed configurations. In other words, it is possible that one may detect flocalNL without any detection of fequilNL and vice versa.

These two parameters cover a fairly large class of models. For example, fequilNL can be generated from inflation models in which the scalar field takes on the nonstandard (noncanonical) kinetic form, such as ${\cal L}=P(X,\phi)$, where X = (∂ϕ)2. In this class of models, the effective sound speed of ϕ can be smaller than the speed of light, c2s = [1 + 2X(∂2P/∂X2)/(∂P/∂X)]−1 < 1. While the sign of fequilNL is negative for the DBI inflation, fequilNL ∼ −1/c2s < 0 in the limit of cs ≪ 1, it can be positive or negative for more general models (Seery & Lidsey 2005; Chen et al. 2007b; Cheung et al. 2008; Li et al. 2008). Such models can be realized in the context of String Theory via the noncanonical kinetic action called the DBI form (Alishahiha et al. 2004), and in the context of an IR modification of gravity called the ghost condensation (Arkani-Hamed et al. 2004).

The observational limits on fequilNL have been obtained from the WMAP 1-year and 3-year data as −366 < fequilNL < 238 (l = 405; Creminelli et al. 2006) and −256 < fequilNL < 332 (lmax = 475; Creminelli et al. 2007), respectively.

There are other forms, too. Warm inflation might produce a different form of fNL(Moss & Xiong 2007; Moss & Graham 2007). Also, the presence of particles at the beginning of inflation, that is, a departure of the initial state of quantum fluctuations from the Bunch–Davies vacuum, can result in an enhanced non-Gaussianity in the "flattened" triangle configurations (Chen et al. 2007b; Holman & Tolley 2008). We do not consider these forms of non-Gaussianity in this paper.

In this paper, we do not discuss the non-Gaussian signatures that cannot be characterized by flocalNL, fequilNL, or bsrc (the point-source bispectrum amplitude). There have been many studies on non-Gaussian signatures in the WMAP data in various forms (Chiang et al. 2003, 2007; Naselsky et al. 2007; Park 2004; de Oliveira-Costa et al. 2004; Tegmark et al. 2003; Larson & Wandelt 2004; Eriksen et al. 2004a, 2004b, 2004c, 2007a; Copi et al. 2004, 2006, 2007; Schwarz et al. 2004; Gordon et al. 2005; Bielewicz et al. 2005; Jaffe et al. 2005, 2006; Vielva et al. 2004; Cruz et al. 2005, 2006, 2007b, 2007a; Cayón et al. 2005; Bridges et al. 2008; Wiaux et al. 2008; Räth et al. 2007; Land & Magueijo 2005b, 2005a, 2007; Rakić & Schwarz 2007; Park et al. 2007; Bernui et al. 2007; Hajian & Souradeep 2003, 2006; Hajian et al. 2005; Prunet et al. 2005; Hansen et al. 2004a, 2004b), many of which are related to the large-scale features at l ≲ 20. We expect these features to be present in the WMAP 5-year temperature map, as the structure of CMB anisotropy in the WMAP data on such large angular scales has not changed very much since the 3-year data.

3.5.2. Analysis

The largest concern in measuring primordial non-Gaussianity from the CMB data is the potential contamination from the Galactic diffuse foreground emission. To test how much the results would be affected by this, we measure fNL parameters from the raw temperature maps and from the foreground-reduced maps.

We shall mainly use the KQ75 mask, the new mask that is recommended for tests of Gaussianity (Gold et al. 2009). The important difference between the new mask and the previous Kp0 mask (Bennett et al. 2003c) is that the new mask is defined by the difference between the K-band map and the ILC map, and that between the Q band and ILC. Therefore, the CMB signal was absent when the mask was defined, which removes any concerns regarding a potential bias in the distribution of CMB on the masked sky.35

To carry out tests of Gaussianity, one should use the KQ75 mask, which is slightly more conservative than Kp0, as the KQ75 mask cuts slightly more sky: we retain 71.8% of the sky with KQ75, while 76.5% with Kp0. To see how sensitive we are to the details of the mask, we also tried Kp0 as well as the new mask that is recommended for the power spectrum analysis, KQ85, which retains 81.7% of the sky. The previous mask that corresponds to KQ85 is the Kp2 mask, which retains 84.6% of the sky.

In addition, we use the KQ75p1 mask, which replaces the point-source mask of KQ75 with the one that does not mask the sources identified in the WMAP K-band data. Our point-source selection at K band removes more sources and sky in regions with higher CMB flux. We estimate the amplitude of this bias by using the KQ75p1 mask, which does not use any WMAP data for the point-source identification. The small change in flocalNL shows that this is a small bias.

The unresolved extra-galactic point sources also contribute to the bispectrum (Refregier et al. 2000; Komatsu & Spergel 2001; Argüeso et al. 2003; Serra & Cooray 2008), and they can bias our estimates of primordial non-Gaussianity parameters such as flocalNL and fequilNL. We estimate the bias by measuring flocalNL and fequilNL from Monte Carlo simulations of point sources, and list them as ΔflocalNL and ΔfequilNL in Tables 5 and 7, respectively. As the errors in these estimates of the bias are limited by the number of Monte Carlo realizations (which is 300), one may obtain a better estimate of the bias using more realizations.

Table 5. Clean-Map Estimates and the Corresponding 68% Intervals of the Local form of Primordial Non-Gaussianity, flocalNL, the Point-Source Bispectrum Amplitude, bsrc (in units of 10−5 μK3 sr2), and Monte-Carlo Estimates of Bias Due to Point Sources, ΔflocalNL

Band Mask lmax flocalNL ΔflocalNL bsrc
V+W KQ85 400 50 ± 29 1 ± 2 0.26 ± 1.5
V+W KQ85 500 61 ± 26 2.5 ± 1.5 0.05 ± 0.50
V+W KQ85 600 68 ± 31 3 ± 2 0.53 ± 0.28
V+W KQ85 700 67 ± 31 3.5 ± 2 0.34 ± 0.20
V+W Kp0 500 61 ± 26 2.5 ± 1.5  
V+W KQ75p1a 500 53 ± 28 4 ± 2  
V+W KQ75 400 47 ± 32 3 ± 2 −0.50 ± 1.7
V+W KQ75 500 55 ± 30 4 ± 2 0.15 ± 0.51
V+W KQ75 600 61 ± 36 4 ± 2 0.53 ± 0.30
V+W KQ75 700 58 ± 36 5 ± 2 0.38 ± 0.21

Note. aThis mask replaces the point-source mask in KQ75 with the one that does not mask the sources identified in the WMAP K-band data.

Download table as:  ASCIITypeset image

We give a detailed description of our estimators for flocalNL, fequilNL, and bsrc, the amplitude of the point-source bispectrum, as well as of Monte Carlo simulations in Appendix A.

3.5.3. Results: Bispectrum

In Table 5, we show our measurement of flocalNL from the template-cleaned V+W map (Gold et al. 2009) with four different masks, KQ85, Kp0, KQ75p1, and KQ75, in the increasing order of the size of the mask. For KQ85 and KQ75, we show the results from different maximum multipoles used in the analysis, lmax = 400, 500, 600, and 700. The WMAP 5-year temperature data are limited by cosmic variance to l ∼ 500.

We find that both KQ85 and Kp0 for lmax = 500 show evidence for flocalNL > 0 at more than 95% CL, 9 < flocalNL < 113 (95% CL), before the point-source bias correction, and 6.5 < flocalNL < 110.5 (95% CL) after the correction. For a higher lmax, lmax = 700, we still find evidence for flocalNL > 0, 1.5 < flocalNL < 125.5 (95% CL), after the correction.36

This evidence is, however, reduced when we use larger masks, KQ75p1 and KQ75. For the latter, we find −5 < flocalNL < 115 (95% CL) before the source bias correction, and −9 < flocalNL < 111 (95% CL) after the correction, which we take as our best estimate. This estimate improves upon our previous estimate from the 3-year data, −54 < flocalNL < 114 (95% CL; Spergel et al. 2007, for l = 350), by cutting much of the allowed region for flocalNL < 0. To test whether the evidence for flocalNL > 0 can also be seen with the KQ75 mask, we need more years of WMAP observations.

Let us study the effect of mask further. We find that the central value of flocalNL (without the source correction) changes from 61 for KQ85 to 55 for KQ75 at lmax = 500. Is this change expected? To study this, we have computed flocalNL from each of the Monte Carlo realizations using KQ85 and KQ75. We find the root mean square (rms) scatter of 〈(flocalNLKQ85flocalNLKQ75)21/2MC = 13, 12, 15, and 15, for lmax = 400, 500, 600, and 700, respectively. Therefore, the change in flocalNL measured from the WMAP data is consistent with a statistical fluctuation. For the other masks at lmax = 500, we find 〈(flocalNLKp0flocalNLKQ75)21/2MC = 9.7 and 〈(flocalNLKQ75p1flocalNLKQ75)21/2MC = 4.0.

In Table 6, we summarize the results from various tests. As a null test, we have measured flocalNL from the difference maps such as Q−W and V−W, which are sensitive to non-Gaussianity in noise and the residual foreground emission, respectively. Since the difference maps do not contain the CMB signal, which is a source of a large cosmic variance in the estimation of flocalNL, the errors in the estimated flocalNL are much smaller. Before the foreground cleaning ("Raw" in the second column), we see negative values of flocalNL, which is consistent with the foreground emission having positively skewed temperature distribution and flocalNL > 0 mainly generating negative skewness. We do not find any significant signal of flocalNL at more than 99% CL for raw maps, or at more than 68% CL for cleaned maps, which indicates that the temperature maps are quite clean outside of the KQ75 mask.

Table 6. Null Tests, Frequency Dependence, and Raw-Map Estimates of the Local form of Primordial Non-Gaussianity, flocalNL, for lmax = 500

Band Foreground Mask flocalNL
Q−W Raw KQ75 −0.53 ± 0.22
V−W Raw KQ75 −0.31 ± 0.23
Q−W Clean KQ75 0.10 ± 0.22
V−W Clean KQ75 0.06 ± 0.23
Q Raw KQ75p1a −42 ± 45
V Raw KQ75p1 38 ± 34
W Raw KQ75p1 43 ± 33
Q Raw KQ75 −42 ± 48
V Raw KQ75 41 ± 35
W Raw KQ75 46 ± 35
Q Clean KQ75p1 9 ± 45
V Clean KQ75p1 47 ± 34
W Clean KQ75p1 60 ± 33
Q Clean KQ75 10 ± 48
V Clean KQ75 50 ± 35
W Clean KQ75 62 ± 35
V+W Raw KQ85 9 ± 26
V+W Raw Kp0 48 ± 26
V+W Raw KQ75p1 41 ± 28
V+W Raw KQ75 43 ± 30

Note. aThis mask replaces the point-source mask in KQ75 with the one that does not mask the sources identified in the WMAP K-band data.

Download table as:  ASCIITypeset image

From the results presented in Table 6, we find that the raw-map results yield more scatter in flocalNL estimated from various data combinations than the clean-map results. From these studies, we conclude that the clean-map results are robust against the data combinations, as long as we use only the V- and W-band data.

In Table 7, we show the equilateral bispectrum, fequilNL, from the template-cleaned V+W map with the KQ75 mask. We find that the point-source bias is much more significant for fequilNL: we detect the bias in fequilNL at more than the 5σ level for lmax = 600 and 700. After correcting for the bias, we find −151 < fequilNL < 253 (95% CL; lmax = 700) as our best estimate. Our estimate improves upon the previous one, −256 < fequilNL < 332 (95% CL; Creminelli et al. 2006, for l = 475), by reducing the allowed region from both above and below by a similar amount.

Table 7. Clean-Map Estimates and the Corresponding 68% Intervals of the Equilateral Form of Primordial Non-Gaussianity, fequilNL, and Monte-Carlo Estimates of Bias Due to Point Sources, ΔfequilNL

Band Mask lmax fequilNL ΔfequilNL
V+W KQ75 400 77 ± 146 9 ± 7
V+W KQ75 500 78 ± 125 14 ± 6
V+W KQ75 600 71 ± 108 27 ± 5
V+W KQ75 700 73 ± 101 22 ± 4

Download table as:  ASCIITypeset image

Finally, the bispectrum from very high multipoles, for example, lmax = 900, can be used to estimate the amplitude of residual point-source contamination. One can use this information to check for a consistency between the estimate of the residual point sources from the power spectrum and that from the bispectrum. In Table 8, we list our estimates of bsrc. The raw maps and cleaned maps yield somewhat different values, indicating a possible leakage from the diffuse foreground to an estimate of bsrc. Our best estimate in the Q band is bsrc = 4.3 ± 1.3 μK3 sr2 (68% CL). See Nolta et al. (2009) for the comparison between bsrc, Cps, and the point-source counts.

Table 8. Point-Source Bispectrum Amplitude, bsrc, for lmax = 900

Band Foreground Mask bsrc (10−5μK3 sr2)
Q Raw KQ75p1a 11.1 ± 1.3
V Raw KQ75p1 0.83 ± 0.31
W Raw KQ75p1 0.16 ± 0.24
V+W Raw KQ75p1 0.28 ± 0.16
Q Raw KQ75 6.0 ± 1.3
V Raw KQ75 0.43 ± 0.31
W Raw KQ75 0.12 ± 0.24
V+W Raw KQ75 0.14 ± 0.16
V+W Raw KQ85 0.20 ± 0.15
Q Clean KQ75p1 8.7 ± 1.3
V Clean KQ75p1 0.75 ± 0.31
W Clean KQ75p1 0.16 ± 0.24
V+W Clean KQ75p1 0.28 ± 0.16
Q Clean KQ75 4.3 ± 1.3
V Clean KQ75 0.36 ± 0.31
W Clean KQ75 0.13 ± 0.24
V+W Clean KQ75 0.14 ± 0.16
V+W Clean KQ85 0.13 ± 0.15

Note. aThis mask replaces the point-source mask in KQ75 with the one that does not mask the sources identified in the WMAP K-band data.

Download table as:  ASCIITypeset image

Incidentally, we also list bsrc from the KQ75p1 mask, whose source mask is exactly the same as we used for the first-year analysis. We find bsrc = 8.7 ± 1.3 μK3 sr2 in the Q band, which is in an excellent agreement with the first-year result, bsrc = 9.5 ± 4.4 μK3 sr2 (Komatsu et al. 2003).

3.5.4. Results: Minkowski Functionals

For the analysis of the Minkowski functionals, we follow the method described in Komatsu et al. (2003) and Spergel et al. (2007). In Figure 7, we show all of the three Minkowski functionals (Gott et al. 1990; Mecke et al. 1994; Schmalzing & Buchert 1997; Schmalzing & Gorski 1998; Winitzki & Kosowsky 1998) that one can define on a two-dimensional sphere: the cumulative surface area (bottom), the contour length (middle), and the Euler characteristics (which is also known as the genus; top), as a function the "threshold," ν, which is the number of σ of hot and cold spots, defined by

Equation (34)

where σ0 is the standard deviation of the temperature data (which includes both signal and noise) at a given resolution of the map that one works with. We compare the Minkowski functionals measured from the WMAP data with the mean and dispersion of Gaussian realizations that include CMB signal and noise. We use the KQ75 mask and the V+W-band map.

Figure 7.

Figure 7. Minkowski functionals from the WMAP 5-year data, measured from the template-cleaned V+W map at Nside = 128 (28' pixels) outside of the KQ75 mask. From the top to bottom panels, we show the Euler characteristics (also known as the genus), the contour length, and the cumulative surface area, as a function of the threshold (the number of σ's of hot and cold spots), ν ≡ ΔT0. (Left) The data (symbols) are fully consistent with the mean and dispersion of Gaussian realizations that include CMB and noise. The gray bands show the 68% intervals of Gaussian realizations. (Right) The residuals between the WMAP data and the mean of the Gaussian realizations. Note that the residuals are highly correlated from bin to bin. From this result, we find flocalNL = −57 ± 60 (68% CL). From Nside = 64, we find flocalNL = −68 ± 69 (68% CL).

Standard image High-resolution image

While Figure 7 shows the results at resolution 7 (Nside = 128), we have carried out Gaussianity tests using the Minkowski functionals at six different resolutions from resolution 3 (Nside = 8) to resolution 8 (Nside = 256). We find no evidence for departures from Gaussianity at any resolutions, as summarized in Table 9; in this table, we list the values of χ2 of the Minkowski functionals relative to the Gaussian predictions:

Equation (35)

where FiWMAP and Fisim are the ith Minkowski functionals measured from the WMAP data and Gaussian simulations, respectively, the angular bracket denotes the average over realizations, and $\Sigma _{\nu _1\nu _2}^{ij}$ is the covariance matrix estimated from the simulations. We use 15 different thresholds from ν = −3.5 to ν = +3.5, as indicated by the symbols in Figure 7, and thus the number of dof in the fit is 15 × 3 = 45. We show the values of χ2WMAP and the dof in the second column, and the probability of having χ2 that is larger than the measured value, F(>χ2WMAP), in the third column. The smallest probability is 0.1 (at Nside = 8), and thus we conclude that the Minkowski functionals measured from the WMAP 5-year data are fully consistent with Gaussianity.

Table 9. χ2 Analysis of the Minkowski Functionals for the Template-Cleaned V+W map

Nside χ2WMAP/dof F(>χ2WMAP)
256 51.5/45 0.241
128 40.0/45 0.660
64 54.2/45 0.167
32 46.8/45 0.361
16 44.7/45 0.396
8 61.3/45 0.104

Note. The results from the area, contour length, and Euler characteristics are combined.

Download table as:  ASCIITypeset image

What do these results imply for flocalNL? We find that the absence of non-Gaussianity at Nside = 128 and 64 gives the 68% limits on flocalNL as flocalNL = −57 ± 60 and −68 ± 69, respectively. The 95% limit from Nside = 128 is −178 < flocalNL < 64. The errors are larger than those from the bispectrum analysis given in Section 3.5.3 by a factor of 2, which is partly because we have not used the Minkowski functional at all six resolutions to constrain flocalNL. For a combined analysis of the WMAP 3-year data, see Hikage et al. (2008).

It is intriguing that the Minkowski functionals prefer a negative value of flocalNL, flocalNL ∼ −60, whereas the bispectrum prefers a positive value, flocalNL ∼ 60. In the limit that non-Gaussianity is weak, the Minkowski functionals are sensitive to three "skewness parameters": (1) 〈(ΔT)3〉, (2) 〈(ΔT)2[∂2T)]〉, and (3) 〈[∂(ΔT)]2[∂2T)]〉, all of which can be written in terms of the weighted sum of the bispectrum; thus, the Minkowski functionals are sensitive to some selected configurations of the bispectrum (Hikage et al. 2006). It would be important to study where an apparent "tension" between the Minkowski functionals and the KSW estimator comes from. This example shows how important it is to use different statistical tools to identify the origin of non-Gaussian signals on the sky.

3.6. Adiabaticity of Primordial Fluctuations

3.6.1. Motivation

"Adiabaticity" of primordial fluctuations offers important tests of inflation as well as clues to the origin of matter in the universe. The negative correlation between the temperature and E-mode polarization (TE) at l ∼ 100 is a generic signature of adiabatic superhorizon fluctuations (Spergel & Zaldarriaga 1997; Peiris et al. 2003). The improved measurement of the TE power spectrum and the temperature power spectrum from the WMAP 5-year data, combined with the distance information from BAO and SN, now provide tight limits on deviations of primordial fluctuations from adiabaticity.

Adiabaticity may be loosely defined as the following relation between fluctuations in radiation density and those in matter density:

Equation (36)

This version37 of the condition guarantees that the entropy density (dominated by radiation, sr ∝ ρ3/4r) per matter particle is unperturbed, that is, δ(sr/nm) = 0.

There are two situations in which the adiabatic condition may be satisfied: (1) there is only one degree of freedom in the system, foe example, both radiation and matter were created from decay products of a single scalar field that was solely responsible for generating fluctuations, and (2) matter and radiation were in thermal equilibrium before any nonzero conserving quantum number (such as baryon number minus lepton number, BL) was created (e.g., Weinberg 2004).

Therefore, detection of any nonadiabatic fluctuations, that is, any deviation from the adiabatic condition (Equation (36)), would imply that there were multiple scalar fields during inflation, and either matter (baryon or dark matter) was never in thermal equilibrium with radiation, or a nonzero conserving quantum number associated with matter was created well before the era of thermal equilibrium. In any case, the detection of nonadiabatic fluctuations between matter and radiation has a profound implication for the physics of inflation and, perhaps more importantly, the origin of matter.

For example, axions, a good candidate for dark matter, generate nonadiabatic fluctuations between dark matter and photons, as axion density fluctuations could be produced during inflation independent of curvature perturbations (which were generated from inflaton fields, and responsible for CMB anisotropies that we observe today), and were not in thermal equilibrium with radiation in the early universe (see Kolb & Turner 1990; Sikivie 2008, for reviews). We can, therefore, place stringent limits on the properties of axions by looking at a signature of deviation from the adiabatic relation in the CMB temperature and polarization anisotropies.

In this paper, we focus on the nonadiabatic perturbations between CDM and CMB photons. Nonadiabatic perturbations between baryons and photons are exactly the same as those between CDM and photons, up to an overall constant; thus, we shall not consider them separately in this paper. For neutrinos and photons, we consider only adiabatic perturbations. In other words, we consider only three standard neutrino species (i.e., no sterile neutrinos) and assume that the neutrinos were in thermal equilibrium before the lepton number was generated.

The basic idea behind this study is not new, and adiabaticity has been extensively constrained using the WMAP data since the first-year release, including general (phenomenological) studies without references to specific models (Peiris et al. 2003; Crotty et al. 2003a; Bucher et al. 2004; Moodley et al. 2004; Lazarides et al. 2004; Kurki-Suonio et al. 2005; Beltrán et al. 2005a; Dunkley et al. 2005; Bean et al. 2006; Trotta 2007; Keskitalo et al. 2007), as well as constraints on specific models such as double inflation (Silk & Turner 1987; Polarski & Starobinsky 1992, 1994), axion (Weinberg 1978; Wilczek 1978; Seckel & Turner 1985; Linde 1985, 1991; Turner & Wilczek 1991), and curvaton (Lyth & Wands 2003; Moroi & Takahashi 2001, 2002; Bartolo & Liddle 2002), all of which can be constrained from the limits on nonadiabatic fluctuations (Gordon & Lewis 2003; Gordon & Malik 2004; Beltrán et al. 2004, 2005b, 2007; Lazarides 2005; Parkinson et al. 2005; Kawasaki & Sekiguchi 2008).

We shall use the WMAP 5-year data, combined with the distance information from BAO and SN, to place more stringent limits on two types of nonadiabatic CDM fluctuations: (1) axion-type and (2) curvaton-type. Our study given below is similar to that by Kawasaki & Sekiguchi (2008) for the WMAP 3-year data.

3.6.2. Analysis

We define the nonadiabatic, or entropic, perturbation between the CDM and photons, ${\cal S}_{c,\gamma }$, as

Equation (37)

and report on the limits on the ratio of the power spectrum of ${\cal S}_{c,\gamma }, P_{\cal S}(k)$, to the curvature perturbation, $P_{\cal R}(k)$, at a given pivot wavenumber, k0, given by (e.g., Bean et al. 2006)

Equation (38)

We shall take k0 to be 0.002 Mpc−1.

While α parametrizes the ratio of the entropy power spectrum to the curvature power spectrum, it may be more informative to quantify "how much the adiabatic relation (Equation (36)) can be violated." To quantify this, we introduce the adiabaticity deviation parameter, δadi, given by

Equation (39)

which can be used to say, "the deviation from the adiabatic relation between dark matter and photons must be less than 100δ(c,γ)adi%." The numerator is just the definition of the entropy perturbation, ${\cal S}_{c,\gamma }$, whereas the denominator is given by

Equation (40)

Therefore, we find, up to the first order in ${\cal S}/{\cal R}$,

Equation (41)

for α ≪ 1.

There could be a significant correlation between ${\cal S}_{c,\gamma }$ and ${\cal R}$ (Langlois 1999; Langlois & Riazuelo 2000; Gordon et al. 2001). We take this into account by introducing the cross-correlation coefficient, as

Equation (42)

where $P_{{\cal S},\cal R}(k)$ is the cross-correlation power spectrum.

Here, we have a negative sign on the left-hand side because of the following reason. In our notation that we used for Gaussianity analysis, the sign convention of the curvature perturbation is such that it gives temperature anisotropy on large scales (the Sachs–Wolfe limit) as $\Delta T/T=-(1/5){\cal R}$. However, those who investigate correlations between ${\cal S}$ and ${\cal R}$ usually use an opposite sign convention for the curvature perturbation, such that the temperature anisotropy is given by

Equation (43)

on large angular scales (Langlois 1999), where $\tilde{\cal R}\equiv -{\cal R}$, and define the correlation coefficient by

Equation (44)

Therefore, in order to use the same sign convention for β as most of the previous work, we shall use Equation (42), and call β = +1 "totally correlated" and β = −1 "totally anticorrelated," entropy perturbations.

It is also useful to understand how the correlation or anticorrelation affects the CMB power spectrum at low multipoles. By squaring Equation (43) and taking the average, we obtain

Equation (45)

Therefore, the "correlation," β>0, reduces the temperature power spectrum on low multipoles, whereas the "anticorrelation," β < 0, increases the power. This point will become important when we interpret our results: namely, models with β < 0 will result in a positive correlation between α and ns (Gordon & Lewis 2003). Note that this property is similar to that of the tensor mode: as the tensor mode adds a significant power only to l ≲ 50, the tensor-to-scalar ratio, r, is degenerate with ns (see Figure 2).

Finally, we specify the power spectrum of S as a pure power law,

Equation (46)

in analogy to the curvature power spectrum, $P_{\cal R}(k)\propto k^{n_s-4}$. Note that β does not depend on k for this choice of $P_{{\cal S},{\cal R}}(k)$. With this parametrization, it is straightforward to compute the angular power spectra of the temperature and polarization of CMB.

In this paper, we shall pay attention to two limiting cases that are physically motivated: totally uncorrelated (β = 0) entropy perturbations, axion-type (Seckel & Turner 1985; Linde 1985, 1991; Turner & Wilczek 1991), and totally anticorrelated (β = −1) entropy perturbations, curvaton-type (Linde & Mukhanov 1997; Lyth & Wands 2003; Moroi & Takahashi 2001, 2002; Bartolo & Liddle 2002). Then, we shall use α0 to denote α for β = 0 and α−1 for β = −1.

3.6.3. Results: Implications for Axion

First, let us consider the axion case in which ${\cal S}$ and ${\cal R}$ are totally uncorrelated, that is, β = 0. This case represents the entropy perturbation between photons and axions, with axions accounting for some fraction of dark matter in the universe. For simplicity, we take the axion perturbations to be scale invariant, that is, m = 1. In Appendix B, we show that this choice corresponds to taking one of the slow-roll parameters, epsilon, to be less than 10−2 or adding a tiny amount of gravitational waves, r ≪ 0.1, which justifies our ignoring gravitational waves in the analysis.

The left panel of Figure 8 shows that we do not find any evidence for the axion entropy perturbations. The limits are α0 < 0.16(95%CL) and α0 < 0.072(95%CL) for the WMAP-only analysis and WMAP+BAO+SN, respectively. The latter limit is the most stringent to date, from which we find the adiabaticity deviation parameter of δcadi < 0.089 (Equation (39)); thus, we conclude that the axion dark matter and photons should obey the adiabatic relation (Equation (36)) to 8.9%, at the 95% CL.

Figure 8.

Figure 8. Constraint on the axion entropy perturbation fraction, α0 (Section 3.6.3). In all panels, we show the WMAP-only results in blue and WMAP+BAO+SN in red. (Left) One-dimensional marginalized constraint on α0, showing WMAP-only and WMAP+BAO+SN. (Middle) Joint two-dimensional marginalized constraint (68% and 95% CL), showing the correlation between α0 and ns for WMAP-only and WMAP+BAO+SN. (Right) Correlation between ns and Ωmh2. The BAO and SN data help to reduce this correlation which, in turn, reduces correlation between α0 and ns, resulting in a factor of 2.2 better limit on α0.

Standard image High-resolution image

We find that ns and α0 are strongly degenerate (see the middle panel of Figure 8). It is easy to understand the direction of correlation. As the entropy perturbation with a scale-invariant spectrum adds power to the temperature anisotropy on large angular scales only, the curvature perturbation tries to compensate it by reducing power on large scales with a larger tilt, ns. However, since a larger ns produces too much power on small angular scales, the fitting tries to increase Ωbh2 to suppress the second peak and reduce Ωch2 to suppress the third peak. Overall, Ωmh2 needs to be reduced to compensate an increase in ns, as shown in the right panel of Figure 8.

Adding the distance information from the BAO and SN helps to break the correlation between Ωmh2 and ns by constraining Ωmh2, independent of ns. Therefore, with WMAP+BAO+SN we find an impressive, factor of 2.2 improvement in the constraint on α0.

What does this imply for the axions? It has been shown that the limit on the axion entropy perturbation can be used to place a constraint on the energy scale of inflation which, in turn, leads to a stringent constraint on the tensor-to-scalar ratio, r (Kain 2006; Beltrán et al. 2007; Sikivie 2008; Kawasaki & Sekiguchi 2008).

In Appendix B, we study a particular axion cosmology called the "misalignment angle scenario," in which the Pecci–Quinn symmetry breaking occurred during inflation and was never restored after inflation. In other words, we assume that the Pecci–Quinn symmetry breaking scale set by the axion decay constant, fa, which has been constrained to be greater than 1010 GeV from the SN 1987 A (Yao et al. 2006), is at least greater than the reheating temperature of the universe after inflation. This is a rather reasonable assumption, as the reheating temperature is usually taken to be as low as 108 GeV in order to avoid overproduction of unwanted relics (Pagels & Primack 1982; Coughlan et al. 1983; Ellis et al. 1986). Such a low reheating temperature is natural also because a coupling between inflaton and matter had to be weak; otherwise, it would terminate inflation prematurely.

There is another constraint. The Hubble parameter during inflation needs to be smaller than fa, that is, Hinffa; otherwise, the Pecci–Quinn symmetry would be restored by quantum fluctuations (Lyth & Stewart 1992).

In this scenario, axions acquired quantum fluctuations during inflation, in the same way that inflaton fields would acquire fluctuations. These fluctuations were then converted to mass density fluctuations when axions acquired mass at the QCD phase transition at ∼200 MeV. We observe a signature of the axion mass density fluctuations via CDM-photon entropy perturbations imprinted in the CMB temperature and polarization anisotropies.

We find that the tensor-to-scalar ratio, r, the axion density, Ωa, the CDM density, Ωc, the phase of the Pecci–Quinn field within our observable universe, θa, and α0 are related as (for an alternative expression that has fa left instead of θa, see Equation (B7))

Equation (47)

Equation (48)

where γ ⩽ 1 is a "dilution factor" representing the amount by which the axion density parameter, Ωah2, would have been diluted due to a potential late-time entropy production by, for example, decay of some (unspecified) heavy particles, between 200 MeV and the epoch of nucleosynthesis, 1 MeV. Here, we have used the limit on the CDM density parameter, Ωch2, from the axion entropy perturbation model that we consider here, Ωch2 = 0.1052+0.0068−0.0070, as well as the observational fact that α0 ≪ 1.

With our limit, α0 < 0.072(95%CL), we find a limit on r within this scenario as

Equation (49)

Therefore, in order for the axion dark matter scenario that we have considered here to be compatible with Ωc ∼ Ωa and the limits on the nonadiabaticity and Ωch2, the energy scale of inflation should be low, and hence the gravitational waves are predicted to be negligible, unless the axion density was diluted severely by a late-time entropy production, γ ∼ 0.8 × 10−7 (for θa ∼ 1), the axion phase (or the misalignment angle) within our observable universe was close to zero, θa ∼ 3 × 10−9 (for γ ∼ 1), or both γ and θa were close to zero with lesser degree. All of these possibilities would give r ∼ 0.01, a value that could be barely detectable in the foreseeable future. One can also reverse Equation (49) to obtain

Equation (50)

Therefore, the axion density would be negligible for the detectable r, unless θa, or γ, or both are tuned to be small.

Whether such an extreme production of entropy is highly unlikely, or such a tiny angle is an undesirable fine-tuning, can be debated. In any case, it is clear that the cosmological observations, such as the CDM density, entropy perturbations, and gravitational waves, can be used to place a rather stringent limit on the axion cosmology based upon the misalignment scenario, one of the most popular scenarios for axions to become a dominant dark matter component in the universe.

3.6.4. Results: Implications for Curvaton

Next, let us consider one of the curvaton models in which ${\cal S}$ and $\tilde{\cal R}$ are totally anticorrelated, that is, β = −1 (Lyth & Wands 2003; Moroi & Takahashi 2001, 2002; Bartolo & Liddle 2002). One can also write ${\cal S}$ as ${\cal S}=B{\cal R}=-B\tilde{\cal R}$ where B > 0; thus, B2 = α−1/(1 − α−1).38 We take the spectral index of the curvaton entropy perturbation, m, to be the same as that of the adiabatic perturbation, ns, that is, ns = m.

The left panel of Figure 9 shows that we do not find any evidence for the curvaton entropy perturbations, either. The limits, α−1 < 0.011(95%CL) and α−1 < 0.0041(95%CL) for the WMAP-only analysis and WMAP+BAO+SN, respectively, are more than a factor of 10 better than those for the axion perturbations. The WMAP-only limit is better than the previous limit by a factor of 4 (Bean et al. 2006). From the WMAP+BAO+SN limit, we find the adiabaticity deviation parameter of δ(c,γ)adi < 0.021 (Equation (39)); thus, we conclude that the curvaton dark matter and photons should obey the adiabatic relation (Equation (36)) to 2.1% at the 95% CL.

Figure 9.

Figure 9. Constraint on the curvaton entropy perturbation fraction, α−1 (Section 3.6.4). In all panels, we show the WMAP-only results in blue and WMAP+BAO+SN in red. (Left) One-dimensional marginalized constraint on α−1, showing WMAP-only and WMAP+BAO+SN. (Middle) Joint two-dimensional marginalized constraint (68% and 95% CL), showing the correlation between α−1 and ns for WMAP-only and WMAP+BAO+SN. (Right) Correlation between ns and Ωmh2. The BAO and SN data help to reduce this correlation which, in turn, reduces correlation between α−1 and ns, resulting in a factor of 2.7 better limit on α−1. These properties are similar to those of the axion dark matter presented in Figure 8.

Standard image High-resolution image

Once again, adding the distance information from the BAO and SN helps to reduce the correlation between ns and Ωmh2 (see the right panel of Figure 9) and reduces the correlation between ns and α−1. The directions in which these parameters are degenerate are similar to those for the axion case (see Figure 8), as the entropy perturbation with β = −1 also increases the CMB temperature power spectrum on large angular scales, as we described in Section 3.6.2.

What is the implication for this type of curvaton scenario, in which β = −1? This scenario would arise when CDM was created from the decay products of the curvaton field. One then finds a prediction (Lyth et al. 2003)

Equation (51)

where ρcurvaton and ρtotal are the curvaton density and total density at the curvaton decay, respectively. Note that there would be no entropy perturbation if curvaton dominated the energy density of the universe completely at the decay. The reason is simple: in such a case, all of the curvaton perturbation would become the adiabatic perturbation, so would the CDM perturbation. Our limit, α−1 < 0.0041(95%CL), indicates that ρcurvatontotal is close to unity, which simplifies the relation (Equation (51)) to give

Equation (52)

Note that it is the adiabaticity deviation parameter given by Equation (39) that gives the deviation of ρcurvatontotal from unity. From this result, we find

Equation (53)

As we mentioned in Section 3.5.1, the curvaton scenario is capable of producting the local form of non-Gaussianity, and flocalNL is given by (Lyth & Rodriguez 2005, and references therein39)

Equation (54)

which gives −1.25 ⩽ flocalNL(curvaton) ≲ −1.21, for α−1 < 0.0041(95%CL). While we need to add additional contributions from postinflationary, nonlinear gravitational perturbations of order unity to this value in order to compare with what we measure from CMB, the limit from the curvaton entropy perturbation is consistent with the limit from the measured flocalNL (see Section 3.5.3).

However, should the future data reveal flocalNL ≫ 1, then either this scenario would be ruled out (Beltrán 2008; Li et al. 2008b) or the curvaton dark matter must have been in thermal equilibrium with photons.

For the other possibilities, including possible baryon entropy perturbations, see Gordon & Lewis (2003).

4. PROBING PARITY VIOLATION OF THE UNIVERSE: TB AND EB CORRELATION

4.1. Motivation

Since the temperature and E-mode polarization are parity-even and the B-mode polarization is parity-odd, the TB and EB correlations should vanish in a universe that conserves parity (Kamionkowski et al. 1997a, 1997b; Seljak & Zaldarriaga 1997; Zaldarriaga & Seljak 1997). For this reason, the TB and EB correlations are usually used to check for systematics, and not widely used as a cosmological probe.

However, parity is violated in the weak interactions (Lee & Yang 1956; Wu et al. 1957). Why cannot parity be violated at cosmological scales?

Polarization of photons offers a powerful way of probing the cosmological parity violation or the "cosmological birefringence" (Lue et al. 1999; Carroll 1998). Let us consider a parity-violating interaction term in the Lagrangian such as the Chern–Simons term, ${\cal L}_{\rm CS}=-(1/2)p_\alpha A_\beta \tilde{F}^{\alpha \beta }$, where Fαβ and Aβ are the usual electromagnetic tensor and vector potential, respectively, $\tilde{F}^{\alpha \beta }=(1/2)\epsilon ^{\alpha \beta\,\mu \nu }F_{\mu \nu }$ is a dual tensor, and pα is an arbitrary timelike four-vector.40 Carroll et al. (1990) have shown that the Chern–Simons term makes two polarization states of photons propagate with different group velocities, causing the polarization plane to rotate by an angle Δα.

What would pα be? We may take this to be a derivative of a light scalar field, pα = 2(∂αϕ)/M, where M is some unspecified energy scale. In this case, the rotation angle is given by $\Delta \alpha =\int \frac{dt}{a} \dot{\phi }/M=(\Delta \phi)/M$ (Carroll et al. 1990; Carroll 1998; Liu et al. 2006; Xia et al. 2008). Such a field might have something to do with dark energy, for example. We are, therefore, looking at a potential parity-violating interaction between the visible section (i.e., photons) and dark sector (i.e., dark energy).

Such an unusual rotation of polarization vectors has been constrained by observations of radio galaxies and quasars (Carroll 1998); one of the best datasets available today at a single redshift is 3C9 at z = 2.012, which gives a limit on the rotation angle, Δα = 2° ± 3° (68% CL). There are about ten measurements between z = 0.425 and z = 2.012, whose average is Δα = −0fdg6 ± 1fdg5 (68% CL).

The rotation of the polarization plane converts the E-mode polarization to the B-mode. As a result, B modes can be produced from E modes even if inflation did not produce much B modes. This is similar to the gravitational lensing effect, which also produces B modes from E modes (Zaldarriaga & Seljak 1998), but there is an important difference: the lensing does not violate parity, but this interaction does. As a result, the lensing does not yield nonzero TB or EB, but this interaction yields both TB and EB.

We shall constrain Δα between the reionization epoch, z ∼ 10, and the present epoch, as well as Δα between the decoupling epoch, z ≃ 1090, and the present epoch, using the TB and EB spectra that we measure from the WMAP 5-year data.

4.2. Analysis

Before we proceed, we should remember that the magnitude of polarization rotation angle, Δα, depends on the path length over which photons experienced a parity-violating interaction. As pointed out by Liu et al. (2006), this leads to the polarization angle that depends on l. We can divide this l-dependence in two regimes:

  • 1.  
    l ≲ 20: the polarization signal was generated during reionization (Zaldarriaga 1997). We are sensitive only to the polarization rotation between the reionization epoch and present epoch.
  • 2.  
    l ≳ 20: the polarization signal was generated at the decoupling epoch. We are sensitive to the polarization rotation between the decoupling epoch and present epoch; thus, we have the largest path length in this case.

Below, we shall explore two cases separately. Note that we shall use only the polarization spectra: TE, TB, EE, BB, and EB, and do not use the temperature spectrum, as the temperature spectrum is not affected by the parity-violating interaction.

Moreover, for the analysis at l ⩽ 23, we only vary the polarization angle, Δα, and the optical depth, τ, and fix the other parameters at Ωk = 0, Ωbh2 = 0.02265, Ωch2 = 0.1143, H0 = 70.1 km s-1 Mpc−1, and ns = 0.960. At each value of τ, we readjust the overall normalization of power spectra such that the first peak of the temperature spectrum is held fixed. For the analysis at l ⩾ 24, we fix τ at 0.085 and vary only Δα, as there is no correlation between Δα and τ at high multipoles. We ignore EE, BB, and EB at l ⩾ 24, as they are much noisier than TE and TB and thus do not add much information.

When the polarization plane is rotated by Δα, the intrinsic (primordial) TE, EE, and BB spectra are converted into TE, TB, EE, BB, and EB spectra as (Lue et al. 1999; Feng et al. 2005)

Equation (55)

Equation (56)

Equation (57)

Equation (58)

Equation (59)

where Cl's are the primordial power spectra in the absence of parity violation, while Cobsl's are what we would observe in the presence of parity violation. To simplify the problem and maximize our sensitivity to a potential signal of Δα, we ignore the primordial BB and use only a reduced set:

Equation (60)

Equation (61)

Equation (62)

Equation (63)

Equation (64)

Therefore, TB and EB will be produced via the "leakage" from TE and EE, respectively. Note that E and B are totally correlated in this case: (CEB,obsl)2 = CEE,obslCBB,obsl.

Several groups have constrained Δα from the WMAP 3-year data and from the BOOMERanG data (Feng et al. 2006; Liu et al. 2006; Kostelecky & Mewes 2007; Cabella et al. 2007; Xia et al. 2008). All but Liu et al. (2006) assumed that Δα is constant at all multipoles, which is acceptable when they consider the TB and EB data at l ≳ 20, that is, the BOOMERanG data and high-l WMAP data. However, this requires care when one considers the low-l WMAP data. Moreover, all of the authors used a Gaussian form of the likelihood function for Cl, which is again acceptable at high multipoles, but it is inaccurate at low multipoles.

For the 5-year data release, we have added capabilities of computing the likelihood of TB and EB spectra at low multipoles, 2 ⩽ l ⩽ 23, exactly, and that of the TB spectrum at high multipoles, 24 ⩽ l ⩽ 450, using the MASTER (pseudo-Cl) algorithm. We shall use this code to obtain the limit on Δα from the 5-year WMAP polarization data. For the low-l polarization, we use the Ka-, Q-, and V-band data, whereas for the high-l polarization, we use the Q- and V-band data.

4.3. Results

Figure 10 shows our limit on Δα between (1) the reionization epoch and present epoch from the low-l polarization data (dark blue), (2) between the decoupling epoch and present epoch from the high-l polarization data (light blue), and (3) combined constraints from the low-l and high-l data assuming a constant Δα across the entire multipole range (red). We find no evidence for parity-violating interactions: the 95% CL (68% CL) limits are −22fdg2 < Δα < 7fdg2 (Δα = −7fdg5 ± 7fdg3) for (1), −5fdg5 < Δα < 3fdg1 (Δα = −1fdg2 ± 2fdg2) for (2), and −5fdg9 < Δα < 2fdg4 (Δα = −1fdg7 ± 2fdg1) for (3).

Figure 10.

Figure 10. Constraint on the polarization rotation angle, Δα, due to a parity-violating interaction that rotates the polarization angle of CMB (Section 4.3). We have used the polarization spectra (TE/TB/EE/BB/EB at l ⩽ 23 and TE/TB at l ⩾ 24) and did not use the TT power spectrum. (Left) One-dimensional marginalized constraint on Δα in units of degrees. The dark blue, light blue, and red curves show the limits from the low-l (2 ⩽ l ⩽ 23), high-l (24 ⩽ l ⩽ 450), and combined (2 ⩽ l ⩽ 450) analysis of the polarization data, respectively. (Right) Joint two-dimensional marginalized constraint on τ and Δα (68% and 95% CL). The bigger contours are from the low-l analysis, while the smaller ones are from the combined analysis. The vertical dotted line shows the best-fitting optical depth in the absence of parity violation (τ = 0.086), whereas the horizontal dotted line shows Δα = 0 to guide eyes.

Standard image High-resolution image

The previous 95% CL (68% CL) limits on Δα are largely based upon the high-l TB and EB data from the WMAP 3-year data and/or BOOMERanG: −13fdg7 < Δα < 1fdg9 (Δα = −6fdg0 ± 4fdg0; Feng et al. 2006), −25° < Δα < 2° (Δα = −12° ± 7°; Kostelecky & Mewes 2007) −8fdg5 < Δα < 3fdg5 (Δα = −2fdg5 ± 3fdg0; Cabella et al. 2007), and −13fdg8 < Δα < 1fdg4 (Δα = −6fdg2 ± 3fdg8; Xia et al. 2008). Our limits from the WMAP 5-year data are tighter than the previous ones by a factor of 1.5–2, and already comparable to those from the polarization data of radio galaxies and quasars (see Section 4.1). Note that the radio galaxies and quasars measure the rotation of polarization between up to z = 2 and the present epoch, whereas our limits measure the rotation between the decoupling epoch, z ≃ 1090, and the present epoch.

These results show that the TB and EB polarization data can provide interesting limits on parity-violating interaction terms. The future data will be able to place more stringent limits (Xia et al. 2008). In particular, adding the Ka and W-band data to the high-l polarization should improve our limit significantly.

5. DARK ENERGY

5.1. Motivation

Dark energy is one of the most mysterious observations in physics today. The issue is the following: when the luminosity distances out to Type Ia SNe (Riess et al. 1998; Perlmutter et al. 1999) and the angular diameter distances measured from the BAO (Eisenstein et al. 2005) as well as CMB (Bennett et al. 2003b) are put together in the context of homogeneous and isotropic cosmological models, one cannot fit these distances without having an accelerated expansion of the universe today. A straightforward interpretation of this result is that we need an additional energy component in the universe that has a large negative pressure, which causes the expansion to accelerate.

However, we do not know much about dark energy. A study of review articles written over the past 20 years reveals a growing circle of ignorance (Weinberg 1989; Carroll et al. 1992; Sahni & Starobinsky 2000; Padmanabhan 2003, 2005; Peebles & Ratra 2003; Copeland et al. 2006); physicists first struggled to understand why the cosmological constant or vacuum energy term was so close to zero and then to understand why it was nonzero. Cosmologists then explored the possibility that dark energy was dynamical, for example, in a form of some light scalar field (Ford 1987; Wetterich 1988; Ratra & Peebles 1988; Peebles & Ratra 1988; Fujii & Nishioka 1990; Chiba et al. 1997; Caldwell et al. 1998; Copeland et al. 1998; Ferreira & Joyce 1998; Zlatev et al. 1999). Recently, there has been significant interest in modifications to general relativity, in the context of explaining the acceleration of the universe (Dvali et al. 2000; Deffayet et al. 2002).

Currently, the properties of dark energy are mainly constrained by the distance information. There are other promising ways of finding dark energy independent of distances: the expansion rate of the universe at higher (z ≳ 0.5) redshifts, the ISW effect, and a slow-down of the growth of the large-scale structure in the universe due to dark energy. While these tools are powerful in principle, the current data are not accurate enough to distinguish between the effects of dark energy and spatial curvature of the universe, owing to the degeneracy between them (e.g., Nesseris & Perivolaropoulos 2008; Ho et al. 2008; Giannantonio et al. 2008).

Indeed, the properties of dark energy, such as the present-day density and its evolution, for example the equation of state parameter w, are degenerate with the spatial curvature of the universe, Ωk. In this section, we shall explore both flat and curved universes when we report on our limits on the dark energy properties.

In Sections 5.2 and 5.3, we explore constraints on a time-independent (i.e., constant) equation of state, w, assuming flat (Ωk = 0) and curved (Ωk ≠ 0) geometries, respectively. In Section 5.4, we introduce a set of "WMAP distance priors," and use them to explore a wider range of model space that has a time-dependent equation of state, w = w(z). Throughout Section 5.4, we use the distance information only to constrain the properties of dark energy. We thus assume the standard homogeneous and isotropic Friedmann–Lemaitre–Robertson–Walker–universe, and do not consider modifications of gravity or local inhomogeneity, as the distance information alone cannot discriminate between these models and the accelerated expansion due to dark energy. Finally, in Section 5.5, we introduce a "WMAP normalization prior."

5.2. Constant Equation of State: Flat Universe

What are we doing by assuming a flat universe, when we constrain the dark energy equation of state, w? Most inflation models in which the inflationary periods last for much longer than 60 e-folds predict Ωk ∼ 10−5, which is three orders of magnitude below the current constraint (see Section 3.4). In this subsection, we use a "strong inflation prior," imposing a flatness prior, and explore dark energy models in the context of such inflation models. We shall explore curved universes in Section 5.3.

Figure 11 shows the constraints on w and the present-day dark energy density, ΩΛ. The WMAP data alone cannot constrain this parameter space very well, as certain combinations of w and ΩΛ can produce very similar angular diameter distances out to the decoupling epoch.

Figure 11.

Figure 11. Constraint on the time-independent (constant) dark energy equation of state, w, and the present-day dark energy density, ΩΛ, assuming a flat universe, Ωk = 0 (Section 5.2). Note that we have imposed a prior on w, w > − 2.5. (Left) Joint two-dimensional marginalized distribution of w and Ωk. The contours show the 68% and 95% CL. The WMAP-only constraint (light blue) is compared with WMAP+HST (gray), WMAP+BAO (red), WMAP+SN (dark blue), and WMAP+BAO+SN (purple). This figure shows how powerful a combination of the WMAP data and the current SN data is for constraining w. (Middle) One-dimensional marginalized constraint on w for a flat universe from WMAP+HST (gray), WMAP+BAO (red), and WMAP+SN (dark blue). The WMAP+BAO+SN result (not shown) is essentially the same as WMAP+SN. (Right) One-dimensional marginalized constraints on ΩΛ for a flat universe from WMAP+HST (gray), WMAP+BAO (red), and WMAP+SN (dark blue). The WMAP+BAO+SN result (not shown) is essentially the same as WMAP+SN. See Figure 12 for the constraints on w for nonflat universes. Note that neither BAO nor SN alone is able to constrain w: they need the WMAP data for lifting the degeneracy. Note also that BAO+SN is unable to lift the degeneracy either, as BAO needs the sound horizon size measured by the WMAP data.

Standard image High-resolution image

The HST prior helps a little bit −0.47 < 1 + w < 0.42(95%CL) by constraining ΩΛ: the WMAP data measure Ωmh2 and a flatness prior imposes a constraint, ΩΛ = 1 − (Ωmh2)/h2; thus, an additional constraint on h from the HST Key Project helps determine ΩΛ better.

The current angular diameter distance measurements from the BAO do not quite break the degeneracy between w and ΩΛ, as they constrain the distances at relatively low redshifts, z = 0.2 and 0.35, whereas the transition from matter to dark energy domination, which is sensitive to w, happens at earlier times. Therefore, the future BAO surveys at higher redshifts should be more sensitive to w. The WMAP+BAO yields −0.68 < 1 + w < 0.21(95%CL).41

Finally, the Type Ia SN data break the degeneracy nicely, as their constraint on this parameter space is nearly orthogonal to what is determined by the CMB data: WMAP+SN yields −0.12 < 1 + w < 0.14(95%CL).42

With a flatness prior, the constraint on w from SN is so powerful that WMAP+SN is similar to WMAP+BAO+SN. We conclude that, when the equation of state does not depend on redshifts, dark energy is consistent with vacuum energy, with −0.12 < 1 + w < 0.13(95%CL)43 (from WMAP+BAO+SN), in the context of a flat universe at the level of curvature that is predicted by long-lasting inflation models.

5.3. Constant Equation of State: Curved Universe

In this subsection, we do not assume a flat universe, but do assume a constant equation of state (for a time-dependent equation of state, see Section 5.4.2). As we discussed in Section 3.4, the WMAP data alone are unable to place meaningful constraints on the spatial curvature of the universe; however, two or more distance or expansion rate measurements break the degeneracy between Ωk and Ωm. As Figure 6 shows, the combination of the WMAP measurement of the distance to the decoupling epoch at z ≃ 1090, and the distance measurements out to z = 0.2 and 0.35 from BAO, strongly constrains the curvature, at the level of 1%–2%.

However, when dark energy is dynamical, we need three distance indicators that cover a wide range of redshift. As the current SN data cover a wider range in redshifts, 0.02 ≲ z ⩽ 1.7, than the BAO data, the SN data help to constrain the evolution of dark energy, that is, w.

Figure 12 shows the constraints on w and Ωk from the WMAP 5-year data alone, WMAP+HST, WMAP+BAO, WMAP+SN, as well as WMAP+BAO+SN. The middle panel is particularly illuminating. The WMAP+BAO combination fixes Ωk, nearly independent of w.44 The WMAP+SN combination yields a degeneracy line that is tilted with respect to the WMAP+BAO line. The WMAP+BAO and WMAP+SN lines intersect at Ωk ∼ 0 and w ∼ −1, and the combined constraints are −0.0179 < Ωk < 0.0081(95%CL) and −0.14 < 1 + w < 0.12(95%CL), respectively.45 It is remarkable that the limit on Ωk is as good as that for a vacuum energy model, −0.0178 < Ωk < 0.0066(95%CL). This is because the BAO and SN yield constraints on Ωk and w that are complementary to each other, breaking the degeneracy effectively.

Figure 12.

Figure 12. Joint two-dimensional marginalized constraint on the time-independent (constant) dark energy equation of state, w, and the curvature parameter, Ωk (Section 5.3). Note that we have imposed a prior on w, w > − 2.5. The contours show the 68% and 95% CL. (Left) The WMAP-only constraint (light blue; 95% CL) compared with WMAP+BAO+SN (purple; 68% and 95% CL). This figure shows how powerful the extra distance information from BAO and SN is for constraining Ωk and w simultaneously. (Middle) A blow-up of the left panel, showing WMAP+HST (gray), WMAP+BAO (red), WMAP+SN (dark blue), and WMAP+BAO+SN (purple). This figure shows that we need both BAO and SN to constrain Ωk and w simultaneously: WMAP+BAO fixes Ωk and WMAP+SN fixes w. (Right) The same as the middle panel, but with the BAO prior re-weighted by a weaker BAO prior from the SDSS LRG sample (Eisenstein et al. 2005). The BAO data used in the other panels combine the SDSS main and LRG, as well as the 2dFGRS data (Percival et al. 2007). The constraints from these are similar, and thus our results are not sensitive to the exact form of the BAO data sets. Note that neither BAO nor SN alone is able to constrain w or Ωk: they need the WMAP data for lifting the degeneracy. Also note that BAO+SN is unable to lift the degeneracy either, as BAO needs the sound horizon size measured by the WMAP data.

Standard image High-resolution image

These limits give the lower bounds to the curvature radii of the observable universe as Rcurv > 33 h−1 Gpc and Rcurv > 22 h−1 Gpc for negatively and positively curved universes, respectively.

Is the apparent "tension" between the WMAP+BAO limit and the WMAP+SN limit in Figure 12 the signature of new physics? We have checked this by the BAO distance scale out to z = 0.35 from the SDSS LRG sample, obtained by Eisenstein et al. (2005), instead of the z = 0.2 and z = 0.35 constraints based on the combination of SDSS LRGs with the SDSS main sample and 2dFGRS (Percival et al. 2007). While it is not an independent check, it does provide some measurement of the sensitivity of the constraints to the details of the BAO dataset.

The right panel of Figure 12 shows that the results are not sensitive to the exact form of the BAO datasets.46 The Eisenstein et al. (2005) BAO prior is a bit weaker than Percival et al.'s, and thus the WMAP+BAO contours extend more to w ≳ −1. The important point is that the direction of degeneracy does not change. Therefore, the combined limits from WMAP, SN, and the Eisenstein et al. BAO, −0.15 < 1 + w < 0.13(95%CL), and −0.0241 < Ωk < 0.0094(95%CL) are similar to those with Percival et al.'s BAO, −0.14 < 1 + w < 0.12(95%CL), and −0.0179 < Ωk < 0.0081(95%CL). As expected, a weaker BAO prior resulted in a weaker limit on Ωk.

While the above argument suggests that there is no serious tension between WMAP+BAO and WMAP+SN constraints, would it be possible that the tension, if any, could be caused by the WMAP data? As the BAO data use the sound horizon size measured by the WMAP data, rs(zd), some systematic errors causing the miscalculation of rs(zd) could lead to a misinterpretation of the BAO data. The current measurement errors in rs(zd)/DV(z) from the BAO data are 2.9% at z = 0.2 and 3.0% at z = 0.35. However, WMAP measures rs(zd) with 1.3% accuracy (see Table 3). We are confident that the systematic error in rs(zd), if any, is smaller than the statistical error; thus, it is unlikely that WMAP causes a misinterpretation of the BAO data.

From these studies, we are able to place rather stringent, simultaneous limits on Ωk (to a 1%–2% level, depending upon the sign) and w (to a 14% level). The spatial curvature is consistent with zero and the dark energy is consistent with vacuum energy. How does this conclusion change when we allow w to vary?

5.4. WMAP Distance Priors for Testing Dark Energy Models

5.4.1. Motivation

Dark energy influences the distance scales as well as the growth of structure. The CMB power spectra are sensitive to both, although sensitivity to the growth of structure is fairly limited, as it influences the CMB power spectrum via the ISW effect at low multipoles (l ≲ 10), whose precise measurement is hampered by a large cosmic variance.

However, CMB is sensitive to the distance to the decoupling epoch via the locations of peaks and troughs of the acoustic oscillations, which can be measured precisely. More specifically, CMB measures two distance ratios: (1) the angular diameter distance to the decoupling epoch divided by the sound horizon size at the decoupling epoch, DA(z*)/rs(z*), and (2) the angular diameter distance to the decoupling epoch divided by the Hubble horizon size at the decoupling epoch, DA(z*)H(z*)/c. This consideration motivates our using these two distance ratios to constrain various dark energy models, in the presence of the spatial curvature, on the basis of distance information (Wang & Mukherjee 2007; Wright 2007).

We shall quantify the first distance ratio, DA(z*)/rs(z*), by the "acoustic scale," lA, defined by

Equation (65)

where a factor of (1 + z*) arises because DA(z*) is the proper (physical) angular diameter distance (Equation (2)), whereas rs(z*) is the comoving sound horizon at z* (Equation (6)). Here, we shall use the fitting function of z* proposed by Hu & Sugiyama (1996):

Equation (66)

where

Equation (67)

Equation (68)

Note that one could also use the peak of the probability of last scattering of photons, that is, the peak of the visibility function, to define the decoupling epoch, which we denote as zdec. Both quantities yield similar values. We shall adopt z* here because it is easier to compute, and, therefore, it allows one to implement the WMAP distance priors in a straightforward manner.

The second distance ratio, DA(z*)H(z*)/c, is often called the "shift parameter," R, given by (Bond et al. 1997)

Equation (69)

This quantity is different from DA(z*)H(z*)/c by a factor of $\sqrt{1+z_*}$, and also ignores the contributions from radiation, curvature, or dark energy to H(z*). Nevertheless, we shall use R to follow the convention in the literature.

We give the 5-year WMAP constraints on lA, R, and z* that we recommend as the WMAP distance priors for constraining dark energy models. However, we note an important caveat. As pointed out by Elgarøy & Multamäki (2007) and Corasaniti & Melchiorri (2008), the derivation of the WMAP distance priors requires us to assume the underlying cosmology first, as all of these quantities are derived parameters from fitting the CMB power spectra. Therefore, one must be careful about which model one is testing. Here, we give the WMAP distance priors, assuming the following model.

  • 1.  
    The standard Friedmann–Lemaitre–Robertson–Walker universe with matter, radiation, dark energy, and spatial curvature.
  • 2.  
    Neutrinos with the effective number of neutrinos equal to 3.04, and the minimal mass (mν ∼ 0.05 eV).
  • 3.  
    Nearly power-law primordial power spectrum of curvature perturbations, |dns/dln k| ≪ 0.01.
  • 4.  
    Negligible primordial gravitational waves relative to the curvature perturbations, r ≪ 0.1.
  • 5.  
    Negligible entropy fluctuations relative to the curvature perturbations, α ≪ 0.1.

In Figure 13, we show the constraints on w and Ωk from the WMAP distance priors (combined with BAO and SN). We find a good agreement with the full MCMC results (compare the middle and right panels of Figure 12 with the left and right panels of Figure 13, respectively). The constraints from the WMAP distance priors are slightly weaker than the full MCMC, as the distance priors use only a part of the information contained in the WMAP data.

Figure 13.

Figure 13. Joint two-dimensional marginalized constraint on the time-independent (constant) dark energy equation of state, w, and the curvature parameter, Ωk, derived solely from the WMAP distance priors (lA, R, z*) (see Section 5.4.1), combined with either BAO (red), SN (dark blue), or both (purple). The contours show the Δχ2total = 2.30 (68.3% CL) and Δχ2total = 6.17 (95.4% CL). The left (BAO data from Percival et al. 2007) and right (BAO data from Eisenstein et al. 2005) panels should be compared with the middle and right panels of Figure 12, respectively, which have been derived from the full WMAP data combined with the same BAO and SN data. While the WMAP distance priors capture most of the information in this parameter space, the constraint is somewhat weaker than that from the full analysis.

Standard image High-resolution image

Of course, the agreement between Figures 12 and 13 does not guarantee that these priors yield good results for the other, more complex dark energy models with a time-dependent w; however, the previous studies indicate that, under the assumptions given above, these priors can be used to constrain a wide variety of dark energy models (Wang & Mukherjee 2007; Elgarøy & Multamäki 2007; Corasaniti & Melchiorri 2008). Also see Li et al. (2008a) for the latest comparison between the WMAP distance priors and the full analysis.

Here is the prescription for using the WMAP distance priors.

  • 1.  
    For a given Ωbh2 and Ωmh2, compute z* from Equation (66).
  • 2.  
    For a given H0, Ωmh2, Ωrh2 (which includes Neff = 3.04), ΩΛ, and w(z), compute the expansion rate, H(z), from Equation (7), as well as the comoving sound horizon size at z*, rs(z*), from Equation (6).
  • 3.  
    For a given Ωk and H(z) from the previous step, compute the proper angular diameter distance, DA(z), from Equation (2).
  • 4.  
    Use Equations (65) and (69) to compute lA(z*) and R(z*), respectively.
  • 5.  
    Form a vector containing xi = (lA, R, z*) in this order.
  • 6.  
    Use Table 10 for the data vector, di = (lWMAPA, RWMAP, zWMAP*). We recommend the maximum likelihood (ML) values.
  • 7.  
    Use Table 11 for the inverse covariance matrix, (C−1)ij.
  • 8.  
    Compute the likelihood, L, as χ2WMAP ≡ −2ln L = (xidi)(C−1)ij(xjdj).
  • 9.  
    Add this to the favorite combination of the cosmological datasets. In this paper we add χ2WMAP to the BAO and SN data, that is, χ2total = χ2WMAP + χ2BAO + χ2SN.
  • 10.  
    Marginalize the posterior distribution over Ωbh2, Ωmh2, and H0 with uniform priors. Since the WMAP distance priors combined with the BAO and SN data provide tight constraints on these parameters, the posterior distribution of these parameters is close to a Gaussian distribution. Therefore, the marginalization is equivalent to minimizing χ2total with respect to Ωbh2, Ωmh2, and H0 (also see Cash 1976; Wright 2007). We use a downhill simplex method for minimization (amoeba routine in Numerical Recipes; Press et al. 1992). The marginalization over Ωbh2, Ωmh2, and H0 leaves us with the the marginalized posterior distribution of the dark energy function, w(a), and the curvature parameter, Ωk.

Table 10. WMAP Distance Priors Obtained from the WMAP 5-year Fit to Models with Spatial Curvature and Dark Energy

  5 Year MLa 5 Year Meanb Error, σ
lA(z*) 302.10 302.45 0.86
R(z*) 1.710 1.721 0.019
zc* 1090.04 1091.13 0.93

Notes. The correlation coefficients are: $r_{l_A,R}=0.1109, r_{l_A,z_*}=0.4215$, and $r_{R,z_*}=0.6928$. aMaximum likelihood values (recommended). bMean of the likelihood. cEquation (66).

Download table as:  ASCIITypeset image

Table 11. Inverse Covariance Matrix for the WMAP Distance Priors

  lA(z*) R(z*) z*
lA(z*) 1.800 27.968 −1.103
R(z*)   5667.577 −92.263
z*     2.923

Download table as:  ASCIITypeset image

Note that this prescription eliminates the need for running the MCMC entirely, and thus the computational cost for evaluating the posterior distribution of w(a) and Ωk is not demanding. In Section 5.4.2, we shall apply the WMAP distance priors to constrain the dark energy equation of state that depends on redshifts, w = w(z).

For those who wish to include an additional prior on Ωbh2, we give the inverse covariance matrix for the "extended" WMAP distance priors: (lA(z*), R(z*), z*, 100Ωbh2), as well as the maximum likelihood value of 100Ωbh2, in Table 12. We note, however, that it is sufficient to use the reduced set, (lA(z*), R(z*), z*), as the extended WMAP distance priors give very similar constraints on dark energy (see Wang 2008).

Table 12. Inverse Covariance Matrix for the Extended WMAP Distance Priors

  lA(z*) R(z*) z* 100Ωbh2
lA(z*) 31.001 −5015.642 183.903 2337.977
R(z*)   876807.166 −32046.750 −403818.837
z*     1175.054 14812.579
100Ωbh2       187191.186

Note. The maximum likelihood value of Ωbh2 is 100Ωbh2 = 2.2765.

Download table as:  ASCIITypeset image

5.4.2. Application of the WMAP Distance Priors: Variable Equation of State

In this subsection, we explore a time-dependent equation of state of dark energy, w(z). We use the following parametrized form:

Equation (70)

where $\tilde{w}(a)=\tilde{w}_0+(1-a)\tilde{w}_a$. We give a motivation, derivation, and detailed discussion on this form of w(a) in Appendix C. This form has a number of desirable properties, as follows.

  • 1.  
    w(a) approaches −1 at early times, a < atrans, where atrans = 1/(1 + ztrans) is the "transition epoch" and ztrans is the transition redshift. Therefore, the dark energy density tends to a constant value at a < atrans.
  • 2.  
    The dark energy density remains totally subdominant relative to the matter density at the decoupling epoch.
  • 3.  
    We recover the widely-used linear form, w(a) = w0 + (1 − a)wa (Chevallier & Polarski 2001; Linder 2003), at late times, a > atrans.
  • 4.  
    The early-time behavior is consistent with some of scalar field models classified as the "thawing models" (Caldwell & Linder 2005) in which a scalar field was moving very slowly at early times and then began to move faster at recent times.
  • 5.  
    Since the late-time form of w(a) allows w(a) to go below −1, our form is more general than models based upon a single scalar field.
  • 6.  
    The form is simple enough to give a closed, analytical form of the effective equation of state, weff(a) = (ln a)−1ln a0dln a'w(a') (Equation C6), which determines the evolution of the dark energy density, $\rho _{\rm de}(a)=\rho _{\rm de}(0)a^{-3[1+w_{\rm eff}(a)]}$; thus, it allows one to easily compute the evolution of the expansion rate and cosmological distances.

While this form contains three free parameters, $\tilde{w}_0,\tilde{w}_a$, and ztrans, we shall give constraints on the present-day value of w, w0w(a = 1), and the first derivative of w at present, w' ≡.dw/dz|z = 0, instead of $\tilde{w}_0$ and $\tilde{w}_a$, as they can be compared to the previous results in the literature more directly. We find that the results are not sensitive to the exact values of ztrans.

In Figure 14, we present the constraint on w0 and w' that we have derived from the WMAP distance priors (lA, R, z*), combined with the BAO and SN data. Note that we have assumed a flat universe in this analysis, although it is straightforward to include the spatial curvature. Wang & Mukherjee (2007) and Wright (2007) showed that the two-dimensional distribution extends more towards southeast, that is, w > − 1 and w' < 0, when the spatial curvature is allowed.

Figure 14.

Figure 14. Constraint on models of time-dependent dark energy equation of state, w(z) (Equation (70)), derived from the WMAP distance priors (lA, R, and z*) combined with the BAO and SN distance data (Section 5.4.2). There are three parameters: w0 is the value of w at the present epoch, w0w(z = 0), w' is the first derivative of w with respect to z at z = 0, w' ≡ dw/dz|z = 0, and ztrans is the transition redshift above which w(z) approaches −1. Here, we assume flatness of the universe, Ωk = 0. (Left) Joint two-dimensional marginalized distribution of w0 and w' for ztrans = 10. The constraints are similar for the other values of ztrans. The contours show the Δχ2total = 2.30 (68.3% CL) and Δχ2total = 6.17 (95.4% CL). (Middle) One-dimensional marginalized distribution of w0 for ztrans = 0.5 (dotted), 2 (dashed), and 10 (solid). (Right) One-dimensional marginalized distribution of w' for ztrans = 0.5 (dotted), 2 (dashed), and 10 (solid). The constraints are similar for all ztrans. We do not find evidence for the evolution of dark energy. Note that neither BAO nor SN alone is able to constrain w0 or w': they need the WMAP data for lifting the degeneracy. Also note that BAO+SN is unable to lift the degeneracy either, as BAO needs the sound horizon size measured by the WMAP data.

Standard image High-resolution image

The 95% limit on w0 for ztrans = 10 is −0.33 < 1 + w0 < 0.21.47 Our results are consistent with the previous work using the WMAP 3-year data (see Zhao et al. 2007; Wang & Mukherjee 2007; Wright 2007; Lazkoz et al. 2008, for recent work and references therein). The WMAP 5-year data help tighten the upper limit on w' and the lower limit on w0, whereas the lower limit on w' and the upper limit on w0 come mainly from the Type Ia SN data. As a result, the lower limit on w' and the upper limit on w0 are sensitive to whether we include the systematic errors in the SN data. For this investigation, see Appendix D.

Alternatively, one may take the linear form, w(a) = w0 + (1 − a)wa, literally and extend it to an arbitrarily high redshift. This can result in an undesirable situation in which the dark energy is as important as the radiation density at the epoch of the BBN; however, one can severely constrain such a scenario by using the limit on the expansion rate from BBN (Steigman 2007). We follow Wright (2007) to adopt a Gaussian prior on

Equation (71)

where we have kept Ωm and Ωk for definiteness, but they are entirely negligible compared to the radiation density at the redshift of BBN, zBBN = 109. Figure 15 shows the constraint on w0 and w' for the linear evolution model derived from the WMAP distance priors, the BAO and SN data, and the BBN prior. The 95% limit on w0 is −0.29 < 1 + w0 < 0.21,48 which is similar to what we have obtained above.

Figure 15.

Figure 15. Constraint on the linear evolution model of dark energy equation of state, w(z) = w0 + w'z/(1 + z), derived from the WMAP distance priors (lA, R, and z*) combined with the BAO and SN distance data as well as the BBN prior (Equation (71)). Here, we assume flatness of the universe, Ωk = 0. (Left) Joint two-dimensional marginalized distribution of w0 and w'. The contours show the Δχ2total = 2.30 (68.3% CL) and Δχ2total = 6.17 (95.4% CL). (Middle) One-dimensional marginalized distribution of w0. (Right) One-dimensional marginalized distribution of w'. We do not find evidence for the evolution of dark energy. Note that Linder (2003) defines w' as the derivative of w at z = 1, whereas we define it as the derivative at z = 0. They are related by w'linder = 0.5w'WMAP.

Standard image High-resolution image

5.5. WMAP Normalization Prior

So far, we have been mainly using the distance information to constrain the properties of dark energy; however, this is not the only information that one can use to constrain the properties of dark energy. The amplitude of fluctuations is a powerful tool for distinguishing between dark energy and modifications to gravity (Ishak et al. 2006; Koyama & Maartens 2006; Amarzguioui et al. 2006; Doré et al. 2007; Linder & Cahn 2007; Upadhye 2007; Zhang et al. 2007; Yamamoto et al. 2007; Chiba & Takahashi 2007; Bean et al. 2007b; Hu & Sawicki 2007; Song et al. 2007; Daniel et al. 2008; Jain & Zhang 2008; Bertschinger & Zukin 2008; Amin et al. 2008; Hu 2008), as well as for determining the mass of neutrinos (Hu et al. 1998; Lesgourgues & Pastor 2006).

The microwave background observations measure the amplitude of fluctuations at the decoupling epoch. By combining this measurement with the amplitude measured from various low redshift tracers, one can learn more about the dark energy properties and the mass of neutrinos.

The overall amplitude of CMB anisotropy is set by the amplitude of primordial curvature perturbations, ${\cal R}$. For example, on very large angular scales where the Sachs–Wolfe limit can be used, the temperature anisotropy is given by $\Delta T/T=-{\cal R}/5$ or, in terms of the curvature perturbation during the matter era, Φ, it is given by ΔT/T = −Φ/3. On small angular scales where the acoustic physics must be taken into account, we have the acoustic oscillation whose amplitude is also given by ${\cal R}$.

This motivates our reporting the "WMAP normalization," a measurement of the overall normalization of the curvature perturbations expressed in terms of $\Delta ^2_{\cal R}(k_{\it WMAP})$, where $\Delta ^2_{\cal R}(k)\equiv k^3P_{\cal R}(k)/(2\pi ^2)$ is a contribution to the total variance of ${\cal R}, \langle {\cal R}^2\rangle$, per logarithmic interval of k (also see Equation (15) and descriptions below it).

Here, kWMAP is different from k0 = 0.002 Mpc−1 that we used to define ns, dns/dln k, r, and $\Delta ^2_{\cal R}(k_0)$ reported in Tables 1 and 4. The goal in this subsection is to give the normalization that is as model independent as possible.

At k0 = 0.002 Mpc−1, for example, we find $10^9\Delta _{\cal R}^2(k_0)=2.48, 2.41$, and 2.46 for a flat ΛCDM model, a curved ΛCDM model, and a flat ΛCDM model with massive neutrinos. The scatter between these values comes solely from the fact that k0 = 0.002 Mpc−1 is not a right place to define the normalization. In other words, this is not the pivot scale of the WMAP data.

We find that kWMAP = 0.02 Mpc−1, that is, a factor of 10 larger than k0, gives similar values of $\Delta _{\cal R}^2(k_{\it WMAP})$ for a wide range of models, as summarized in Table 13. From these results, we give the WMAP normalization prior:

Equation (72)

which is valid for models with Ωk ≠ 0, w ≠ −1, or mν > 0. However, we find that these normalizations cannot be used for the models that have the tensor modes, r > 0, or the running index, dns/dln k ≠ 0. We failed to find a universal normalization for these models. Nevertheless, our WMAP normalization, given by Equation (72), is still valid for a wide range of cosmological models.

Table 13. Amplitude of Curvature Perturbations, ${\cal R}$, Measured by WMAP at kWMAP = 0.02 Mpc−1

Model $10^9\times \Delta _{\cal R}^2(k_{\it WMAP})$
Ωk = 0 and w = −1 2.211 ± 0.083
Ωk ≠ 0 and w = −1 2.212 ± 0.084
Ωk = 0 and w ≠ −1 2.208 ± 0.087
Ωk ≠ 0 and w ≠ −1 2.210 ± 0.084
Ωk = 0, w = −1 and mν > 0 2.212 ± 0.083
Ωk = 0, w ≠ −1 and mν > 0 2.218 ± 0.085
WMAP Normalization Prior 2.21 ± 0.09

Download table as:  ASCIITypeset image

How can one use the WMAP normalization? In order to predict the linear matter density power spectrum, Plin(k), one needs to relate the primordial curvature perturbations, ${\cal R}_{\mathbf{k}}$, to the linear matter density fluctuations at arbitrary redshifts, δm,k(z). From Einstein's equations, we find (see, e.g., Appendix C of Takada et al. 2006)

Equation (73)

where D(k,z) and T(k) are the linear growth rate and the matter transfer function normalized such that T(k) → 1 as k → 0 and (1 + z)D(k, z) → 1 as k → 0 during the matter era (e.g., z = 30, where the radiation density is less than 1% of the matter density), respectively. Note that D(k, z) does not depend on k when neutrinos are massless; however, it depends on k when they are massive (e.g., Hu & Eisenstein 1998). The linear matter density power spectrum is given by

Equation (74)

One application of the WMAP normalization is the computation of the present-day normalization of matter fluctuations, which is commonly expressed in terms of σ8, given by

Equation (75)

where R = 8 h−1 Mpc. Both the dark energy properties (or modified gravity) and the mass of neutrinos change the value of D(k, z = 0). The transfer function, T(k), is much less affected, as long as neutrinos are still relativistic at the decoupling epoch, and the dark energy or modified gravity effect is unimportant at the decoupling epoch.

Ignoring the mass of neutrinos and modifications to gravity, one can obtain the growth rate by solving the following differential equation (Wang & Steinhardt 1998; Linder & Jenkins 2003):

Equation (76)

where

Equation (77)

Equation (78)

Equation (79)

Equation (80)

During the matter era, g(a) does not depend on a; thus, the natural choice for the initial conditions are g(ainitial) = 1 and $\left. dg/d\ln a\right|_{a=a_{\rm initial}}=0$, where ainitial must be taken during the matter era, for example, ainitial = 1/31 (i.e., z = 30).

In Figure 16, we show the predicted values of σ8 as a function of w in a flat universe (left panel) and curved universes (right panel; see the middle panel of Figure 17 for σ8 as a function of the mass of neutrinos). Here, we have used the full information in the WMAP data. The normalization information alone is unable to give meaningful predictions for σ8, which depends not only on $\Delta ^2_{\cal R}(k_{\it WMAP})$, but also on the other cosmological parameters via T(k) and D(k, z = 0), especially w and Ωmh2. While the predictions from the WMAP data alone are still weak, adding the extra distance information from the BAO and SN data helps improve the predictions. We find σ8 = 0.807+0.045−0.044 for a flat universe and σ8 = 0.795 ± 0.046 for curved universes.

Figure 16.

Figure 16. Predictions for the present-day amplitude of matter fluctuations, σ8, as a function of the (constant) dark energy equation of state parameter, w, derived from the full WMAP data (blue) and from WMAP+BAO+SN (red). The contours show the 68% and 95% CL. (Left) Flat universe, Ωk = 0. (Right) Curved universe, Ωk ≠ 0.

Standard image High-resolution image
Figure 17.

Figure 17. Constraint on the total mass of neutrinos, ∑mν (Section 6.1.3). In all panels, we show the WMAP-only results in blue and WMAP+BAO+SN in red. (Left) Joint two-dimensional marginalized distribution of H0 and ∑mν (68% and 95% CL). The additional distance information from BAO helps reduce the correlation between H0 and ∑mν. (Middle) The WMAP data, combined with the distances from BAO and SN, predict the present-day amplitude of matter fluctuations, σ8, as a function of ∑mν. An independent determination of σ8 will help determine ∑mν tremendously. (Right) Joint two-dimensional marginalized distribution of w and ∑mν. No significant correlation is observed. Note that we have a prior on w, w > − 2.5, and thus the WMAP-only lower limit on w in this panel cannot be trusted.

Standard image High-resolution image

By combining these results with σ8 measured from various low redshift tracers, one can reduce the remaining correlation between w and σ8 to obtain a better limit on w. The precision of the current data from weak lensing surveys is comparable to these predictions, for example, σ8m/0.25)0.64 = 0.785 ± 0.043 (Canada-France-Hawaii Telescope Legacy Survey (CFHTLS); Fu et al. 2008). The weak lensing surveys will soon become powerful enough to yield smaller uncertainties in σ8 than predicted from WMAP+BAO+SN.

6. NEUTRINOS

In this section, we shall use the WMAP data, combined with the distance information from BAO and SN observations, to place limits on the total mass of massive neutrino species (Section 6.1), as well as on the effective number of neutrino-like species that were still relativistic at the decoupling epoch (Section 6.2)

6.1. Neutrino Mass

6.1.1. Motivation

The existence of nonzero neutrino masses has been firmly established by the experiments detecting atmospheric neutrinos (Hirata et al. 1992; Fukuda et al. 1994, 1998; Allison et al. 1999; Ambrosio et al. 2001), solar neutrinos (Davis et al. 1968; Cleveland et al. 1998; Hampel et al. 1999; Abdurashitov et al. 1999; Fukuda et al. 2001b, 2001a; Ahmad et al. 2002; Ahmed et al. 2004), reactor neutrinos (Eguchi et al. 2003; Araki et al. 2005), and accelerator beam neutrinos (Ahn et al. 2003; Michael et al. 2006). These experiments have placed stringent limits on the squared mass differences between the neutrino mass eigenstates, Δm221 ≃ 8 × 10−5 eV2, from the solar and reactor experiments, and Δm232 ≃ 3 × 10−3 eV2, from the atmospheric and accelerator beam experiments.

One needs different experiments to measure the absolute masses. The next-generation tritium β-decay experiment, KATRIN,49 is expected to reach the electron neutrino mass of as small as ∼0.2 eV. Cosmology has also been providing useful limits on the mass of neutrinos (see Dolgov 2002; Elgarøy & Lahav 2005; Tegmark 2005; Lesgourgues & Pastor 2006; Fukugita 2006; Hannestad 2006b, for reviews). Since the determination of the neutrino mass is of fundamental importance in physics, there is enough motivation to pursue cosmological constraints on the neutrino mass.

How well can CMB constrain the mass of neutrinos? We do not expect massive neutrinos to affect the CMB power spectra very much (except through the gravitational lensing effect), if they were still relativistic at the decoupling epoch, z ≃ 1090. This means that, for massive neutrinos to affect the CMB power spectra, at least one of the neutrino masses must be greater than the mean energy of relativistic neutrinos per particle at z ≃ 1090 when the photon temperature of the universe is Tγ ≃ 3000 K ≃ 0.26 eV. Since the mean energy of relativistic neutrinos is given by 〈E〉 = (7π4Tν)/(180ζ(3)) ≃ 3.15Tν = 3.15(4/11)1/3Tγ, we need at least one neutrino species whose mass satisfies mν > 3.15(4/11)1/3Tγ ≃ 0.58 eV; thus, it would not be possible to constrain the neutrino mass using the CMB data alone, if the mass of the heaviest neutrino species is below this value.

If the neutrino mass eigenstates are degenerate with the effective number of species equal to 3.04, this argument suggests that ∑mν ∼ 1.8 eV would be the limit to which the CMB data are sensitive. Ichikawa et al. (2005) argued that ∑mν ∼ 1.5 eV would be the limit for the CMB data alone, which is fairly close to the value given above.

In order to go beyond ∼1.5 eV, therefore, one needs to combine the CMB data with the other cosmological probes. We shall combine the WMAP data with the distance information from BAO and SN to place a limit on the neutrino mass. We shall not use the galaxy power spectrum in this paper, and, therefore, our limit on the neutrino mass is free from the uncertainty in the galaxy bias. We discuss this in more detail in Section 6.1.2.

6.1.2. Analysis

We assume that, for definiteness, the neutrino mass eigenstates are degenerate, which means that all of the three neutrino species have equal masses.50 We measure the neutrino mass density parameter, Ωνh2, and convert it to the total mass, ∑mν, via

Equation (81)

where Nν is the number of massive neutrino species. We take it to be 3.04. Note that in this case, the mass density parameter is the sum of baryons, CDM, and neutrinos: Ωm = Ωb + Ωc + Ων.

Since the release of the 1 year (Spergel et al. 2003) and 3 year (Spergel et al. 2007) results on the cosmological analysis of the WMAP data, there have been a number of studies with regard to the limits on the mass of neutrinos (Hannestad 2003; Elgarøy & Lahav 2003; Allen et al. 2003; Tegmark et al. 2004a; Barger et al. 2004; Hannestad & Raffelt 2004; Crotty et al. 2004; Seljak et al. 2005a, 2005b; Ichikawa et al. 2005; Hannestad 2005; Lattanzi et al. 2005; Hannestad & Raffelt 2006; Goobar et al. 2006; Feng et al. 2006; Lesgourgues et al. 2007). These analyses reached different limits depending upon (1) the choice of datasets and (2) the parameters in the cosmological model.

The strongest limits quoted on neutrino masses come from combining CMB measurements with measurements of the amplitude of density fluctuations in the recent universe. Clustering of galaxies and Lyα forest observations have been used to obtain some of the strongest limits on neutrino masses (Seljak et al. 2005b, 2006; Viel et al. 2006). As the neutrino mass increases, the amplitude of mass fluctuations on small scales decreases (Bond et al. 1980; Bond & Szalay 1983; Ma 1996; also see the middle panel of Figure 17), which can be used to weigh neutrinos in the universe (Hu et al. 1998; Lesgourgues & Pastor 2006).

These analyses are sensitive to correctly calculating the relationship between the level of observed fluctuations in galaxies (or gas) and the mass fluctuation with the strongest limits coming from the smallest scales in the analyses. With the new WMAP data, these limits are potentially even stronger. There are, however, several potential concerns with this limit: there is already "tension" between the high level of fluctuations seen in the Lyman alpha forest and the amplitude of mass fluctuations inferred from WMAP (Lesgourgues et al. 2007), the relationship between gas temperature and density appears to be more complicated than assumed in the previous Lyα forest analysis (Kim et al. 2007; Bolton et al. 2008), and additional astrophysics could potentially be the source of some of the small-scale fluctuations seen in the Lyα forest data. Given the power of the Lyα forest data, it is important to address these issues; however, they are beyond the scope of this paper.

In this paper, we take the more conservative approach and use only the WMAP data and the distance measures to place limits on the neutrino masses. Our approach is more conservative than Goobar et al. (2006), who have found a limit of 0.62 eV on the sum of the neutrino mass from the WMAP 3-year data, the SDSS–LRG BAO measurement, the SNLS SN data, and the shape of the galaxy power spectra from the SDSS main sample and 2dFGRS. While we use the WMAP 5-year data, the SDSS+2dFGRS BAO measurements, and the union SN data, we do not use the shape of the galaxy power spectra. See Section 2.3 in this paper or Dunkley et al. (2009) for more detail on this choice.

In summary, we do not use the amplitude or shape of the matter power spectrum, but exclusively rely on the CMB data and the distance measurements. As a result, our limits are weaker than the strongest limits in the literature.

Next, let us discuss (2), the choice of parameters. A few correlations between the neutrino mass and other cosmological parameters have been identified. Hannestad (2005) has found that the limit on the neutrino mass degrades significantly when the dark energy equation of state, w, is allowed to vary (also see Figure 18 of Spergel et al. 2007). This correlation would arise only when the amplitude of the galaxy or Lyα forest power spectrum was included, as the dark energy equation of state influences the growth rate of the structure formation. Since we do not include them, our limit on the neutrino mass is not degenerate with w. We shall come back to this point later in the Section 6.1.3. Incidentally, our limit is not degenerate with the running index, dns/dln k, or the tensor-to-scalar ratio, r.

6.1.3. Results

Figure 17 summarizes our limits on the sum of neutrino masses, ∑mν.

With the WMAP data alone we find ∑mν < 1.3 eV(95%CL) for the ΛCDM model in which w = −1, and ∑mν < 1.5 eV(95%CL) for w ≠ −1. We assume a flat universe in both cases. These constraints are very similar, which means that w and ∑mν are not degenerate. We show this more explicitly on the right panel of Figure 17.

When the BAO and SN data are added, our limits improve significantly, by a factor 2, to ∑mν < 0.67 eV(95%CL) for w = −1, and ∑mν < 0.80 eV(95%CL) for w ≠ −1. Again, we do not observe much correlation between w and ∑mν. While the distances out to either BAO or SN cannot reduce correlation between Ωm (or H0) and w, a combination of the two can reduce this correlation effectively, leaving little correlation left on the right panel of Figure 17.

What information do BAO and SN add to improve the limit on ∑mν? It's the Hubble constant, H0, as shown in the left panel of Figure 17. This effect has been explained by Ichikawa et al. (2005) as follows.

The massive neutrinos modify the CMB power spectrum by their changing the matter-to-radiation ratio at the decoupling epoch. If the sum of degenerate neutrino masses is below 1.8 eV, the neutrinos were still relativistic at the decoupling epoch. However, they are definitely nonrelativistic at the present epoch, as the neutrino oscillation experiments have shown that at least one neutrino species is heavier than 0.05 eV. This means that the Ωm that we measure must be the sum of Ωb, Ωc, and Ων; however, at the decoupling epoch, neutrinos were still relativistic, and thus the matter density at the decoupling epoch was actually smaller than a naive extrapolation from the present value.

As the matter-to-radiation ratio was smaller than one would naively expect, it would accelerate the decay of gravitational potential around the decoupling epoch. This leads to an enhancement in the so-called early integrated Sachs–Wolfe (ISW) effect. The larger ∑mν is, the larger early ISW becomes, as long as the neutrinos were still relativistic at the decoupling epoch, that is, ∑mν ≲ 1.8 eV.

The large ISW causes the first peak position to shift to lower multipoles by adding power at l ∼ 200; however, this shift can be absorbed by a reduction in the value of H0.51 This is why ∑mν and H0 are anticorrelated (see Ichikawa et al. 2005, for a further discussion on this subject).

It is the BAO distance that provides a better limit on H0, as BAO is an absolute distance indicator. The SN is totally insensitive to H0, as their absolute magnitudes have been marginalized over (SN is a relative distance indicator); however, the SN data do help reduce the correlation between w and H0 when w is allowed to vary. As a result, we have equally tight limits on ∑mν regardless of w.

Our limit, ∑mν < 0.67 eV(95%CL) (for w = −1), is weaker than the best limit quoted in the literature, as we have not used the information on the amplitude of fluctuations traced by the large-scale structure. The middle panel of Figure 17 shows how the WMAP data combined with BAO and SN predict the present-day amplitude of matter fluctuations, σ8, as a function of ∑mν. From this, it is clear that an accurate, independent measurement of σ8 will reduce the correlation between σ8 and ∑mν, and provide a significant improvement in the limit on ∑mν.

Improving upon our understanding of nonlinear astrophysical effects, such as those raised by Bolton et al. (2008) for the Lyα forest data and Sánchez & Cole (2008) for the SDSS and 2dFGRS data, is a promising way to improve upon the numerical value of the limit, as well as the robustness of the limit, on the mass of neutrinos.

6.2. Effective Number of Neutrino Species

6.2.1. Motivation

While the absolute mass of neutrinos is unknown, the number of neutrino species is well known: it is 3. The high precision measurement of the decay width of Z into neutrinos (the total decay width minus the decay width to quarks and charged leptons), carried out by LEP using the production of Z in e+e collisions, has yielded Nν = 2.984 ± 0.008 (Yao et al. 2006). However, are there any other particles that we do not know yet, and that are relativistic at the photon decoupling epoch?

Such extra relativistic particle species can change the expansion rate of the universe during the radiation era. As a result, they change the predictions from the BBN for the abundance of light elements such as helium and deuterium (Steigman et al. 1977). One can use this property to place a tight bound on the relativistic dof, expressed in terms of the "effective number of neutrino species," Neff (see Equation (84) for the precise definition). As the BBN occurred at the energy of ∼0.1 MeV, which is later than the neutrino decoupling epoch immediately followed by e+e annihilation, the value of Neff for three neutrino species is slightly larger than 3. With other subtle corrections included, the current standard value is Nstandardeff = 3.04 (Dicus et al. 1982; Gnedin & Gnedin 1998; Dolgov et al. 1999; Mangano et al. 2002). The 2σ interval for Neff from the observed helium abundance, YP = 0.240 ± 0.006, is 1.61 < Neff < 3.30 (see Steigman 2007, for a recent summary).

Many people have been trying to find evidence for the extra relativistic dof in the universe, using the cosmological probes such as CMB and the large-scale structure (Pierpaoli 2003; Hannestad 2003; Crotty et al. 2003b, 2004; Barger et al. 2003; Trotta & Melchiorri 2005; Lattanzi et al. 2005; Cirelli & Strumia 2006; Hannestad 2006a; Ichikawa et al. 2007; Mangano et al. 2007; Hamann et al. 2007; de Bernardis et al. 2008). There is a strong motivation to seek the answer for the following question, "can we detect the cosmic neutrino background, and confirm that the signal is consistent with the expected number of neutrino species that we know?" Although we cannot detect the cosmic neutrino background directly yet, there is a possibility that we can detect it indirectly by looking for the signatures of neutrinos in the CMB power spectrum.

In this section, we shall revisit this classical problem by using the WMAP 5-year data as well as the distance information from BAO and SN, and Hubble's constant measured by HST.

6.2.2. Analysis

It is common to write the energy density of neutrinos (including antineutrinos), when they were still relativistic, in terms of the effective number of neutrino species, Neff, as

Equation (82)

where Tν is the temperature of neutrinos. How do we measure Neff from CMB?

The way that we use CMB to determine Neff is relatively simple. The relativistic particles that stream freely influence CMB in two ways: (1) their energy density changing the matter-radiation equality epoch and (2) their anisotropic stress acting as an additional source for the gravitational potential via Einstein's equations. Incidentally, the relativistic particles that do not stream freely, but interact with matter frequently, do not have a significant anisotropic stress because they isotropize themselves via interactions with matter; thus, anisotropic stress of photons before the decoupling epoch was very small. Neutrinos, on the other hand, decoupled from matter much earlier (∼2 MeV), and thus their anisotropic stress was significant at the decoupling epoch.

The amount of the early ISW effect changes as the equality redshift changes. The earlier the equality epoch is, the more the ISW effect CMB photons receive. This effect can be measured via the height of the third acoustic peak relative to the first peak. Therefore, the equality redshift, zeq, is one of the fundamental observables that one can extract from the CMB power spectrum.

One usually uses zeq to determine Ωmh2 from the CMB power spectrum, without noticing that it is actually zeq that they are measuring. However, the conversion from zeq to Ωmh2 is automatic only when one knows the radiation content of the universe exactly—in other words, when one knows Neff exactly:

Equation (83)

where Ωγh2 = 2.469 × 10−5 is the present-day photon energy density parameter for Tcmb = 2.725 K. Here, we have used the standard relation between the photon temperature and neutrino temperature, Tν = (4/11)1/3Tγ, derived from the entropy conservation across the electron–positron annihilation (see, e.g., Weinberg 1972; Kolb & Turner 1990).

However, if we do not know Neff precisely, it is not possible to use zeq to measure Ωmh2. In fact, we lose our ability to measure Ωmh2 from CMB almost completely if we do not know Neff. Likewise, if we do not know Ωmh2 precisely, it is not possible to use zeq to measure Neff. As a result, Neff and Ωmh2 are linearly correlated (degenerate), with the width of the degeneracy line given by the uncertainty in our determination of zeq.

The distance information from BAO and SN provides us with an independent constraint on Ωmh2, which helps to reduce the degeneracy between zeq and Ωmh2.

The anisotropic stress of neutrinos also leaves distinct signatures in the CMB power spectrum, which is not degenerate with Ωmh2 (Hu et al. 1995; Bashinsky & Seljak 2004). Trotta & Melchiorri (2005; also see Melchiorri & Serra 2006) have reported on evidence for the neutrino anisotropic stress at slightly more than 95% CL. They have parametrized the anisotropic stress by the viscosity parameter, c2vis (Hu 1998), and found c2vis > 0.12 (95% CL). However, they had to combine the WMAP 1-year data with the SDSS data to see the evidence for nonzero c2vis.

In Dunkley et al. (2009), we reported on the lower limit to Neff solely from the WMAP 5-year data. In this paper, we shall combine the WMAP data with the distance information from BAO and SN as well as Hubble's constant from HST to find the best-fitting value of Neff.

6.2.3. Results

Figure 18 shows our constraint on Neff. The contours in the left panel lie on the expected linear correlation between Ωmh2 and Neff given by

Equation (84)

which follows from Equation (83). Here, Ωmh2 = 0.1308 and zeq = 3138 are the maximum likelihood values from the simplest ΛCDM model. The width of the degeneracy line is given by the accuracy of our determination of zeq, which is given by zeq = 3141+154−157 (WMAP-only) for this model. Note that the mean value of zeq for the simplest ΛCDM model with Neff = 3.04 is zeq = 3176+151−150, which is close. This confirms that zeq is one of the fundamental observables, and Neff is merely a secondary parameter that can be derived from zeq. The middle panel of Figure 18 shows this clearly: zeq is determined independently of Neff. For each value of Neff along a constant zeq line, there is a corresponding Ωmh2 that gives the same value of zeq along the line.

Figure 18.

Figure 18. Constraint on the effective number of neutrino species, Neff (Section 6.2.3). Note that we have imposed a prior on Neff, 0 < Neff < 10. In all panels, we show the WMAP-only results in blue and WMAP+BAO+SN in red. (Left) Joint two-dimensional marginalized distribution (68% and 95% CL), showing a strong degeneracy between Ωmh2 and Neff. This degeneracy line is given by the equality redshift, 1 + zeq = Ωmr = (4.050 × 104mh2/(1 + 0.2271Neff). The thick solid lines show the 68% and 95% limits calculated from the WMAP-only limit on zeq: zeq = 3141+154−157 (68% CL). The 95% CL contours do not follow the lines below Neff ∼ 1.5 but close there, which shows a strong evidence for the cosmic neutrino background from its effects on the CMB power spectrum via the neutrino anisotropic stress. The BAO and SN provide an independent constraint on Ωmh2, which helps reduce the degeneracy between Neff and Ωmh2. (Middle) When we transform the horizontal axis of the left panel to zeq, we observe no degeneracy. The vertical solid lines show the one-dimensional marginalized 68% and 95% distributions calculated from the WMAP-only limit on zeq: zeq = 3141+154−157 (68% CL). Therefore, the left panel is simply a rotation of this panel using a relation among zeq, Ωmh2, and Neff. (Right) One-dimensional marginalized distribution of Neff from WMAP-only and WMAP+BAO+SN+HST. Note that a gradual decline of the likelihood toward Neff ≳ 6 for the WMAP-only constraint should not be trusted, as it is affected by the hard prior, Neff < 10. The WMAP+BAO+SN+HST constraint is robust. This figure shows that the lower limit on Neff comes solely from the WMAP data. The 68% interval from WMAP+BAO+SN+HST, Neff = 4.4 ± 1.5, is consistent with the standard value, 3.04, which is shown by the vertical line.

Standard image High-resolution image

However, the contours do not extend all the way down to Neff = 0, although Equation (84) predicts that Neff should go to zero when Ωmh2 is sufficiently small. This indicates that we see the effect of the neutrino anisotropic stress at a high significance. While we need to repeat the analysis of Trotta & Melchiorri (2005) in order to prove that our finding of Neff > 0 comes from the neutrino anisotropic stress, we believe that there is a strong evidence that we see nonzero Neff via the effect of neutrino anisotropic stress, rather than via zeq.

While the WMAP data alone can give a lower limit on Neff (Dunkley et al. 2009), they cannot give an upper limit owing to the strong degeneracy with Ωmh2. Therefore, we use the BAO, SN, and HST data to break the degeneracy. We find Neff = 4.4 ± 1.5 (68%) from WMAP+BAO+SN+HST, which is fully consistent with the standard value, 3.04 (see the right panel of Figure 18).

7. CONCLUSION

With 5-years of integration, the WMAP temperature and polarization data have improved significantly. An improved determination of the third acoustic peak has enabled us to reduce the uncertainty in the amplitude of matter fluctuation, parametrized by σ8, by a factor of 1.4 from the WMAP 3-year result. The E-mode polarization is now detected at five standard deviations (see 3.0 standard deviations for the 3-year data; Page et al. 2007), which rules out an instantaneous reionization at zreion = 6 at the 3.5σ level. Overall, the WMAP 5-year data continue to support the simplest, six-parameter ΛCDM model (Dunkley et al. 2009).

In this paper, we have explored our ability to limit deviations from the simplest picture, namely non-Gaussianity, nonadiabatic fluctuations, nonzero gravitational waves, nonpower-law spectrum, nonzero curvature, dynamical dark energy, parity-violating interactions, nonzero neutrino mass, and nonstandard number of neutrino species. Detection of any of these items will immediately lead us to the new era in cosmology and a better understanding of the physics of our universe.

From these studies, we conclude that we have not detected any convincing deviations from the simplest six-parameter ΛCDM model at the level greater than 99% CL. By combining WMAP data with the distance information from BAO and SN, we have improved the accuracy of the derived cosmological parameters. As the distance information provides strong constraints on the matter density (both BAO and SN) and Hubble's constant (BAO), the uncertainties in Ωmh2 and H0 have been reduced by factors of 1.7 and 2, respectively, from the WMAP-only limits. The better determination of H0 reduces the uncertainty in ΩΛ (as well as Ωb and Ωc) by a factor of 2, and the better determination of Ωmh2 reduces the uncertainty in σ8 by a factor of 1.4. These results are presented visually in Figure 19. Also see Table 1 for the summary of the cosmological parameters of the ΛCDM model. The addition of BAO and SN does not improve the determinations of Ωbh2 or τ as expected. Since ns is mainly degenerate with Ωbh2 and τ, with the former being more degenerate, the addition of BAO and SN does not improve our determination of ns, when we consider the simplest six-parameter ΛCDM model.

Figure 19.

Figure 19. Four representative cosmological parameters that have improved significantly by adding the BAO and SN data. (Also see Table 1.) The contours show the 68% and 95% CL. The WMAP-only constraint is shown in blue, while WMAP+BAO+SN in red. (Left) The distance information from BAO and SN provides a better determination of Ωmh2, which results in 40% better determination of σ8. (Right) The BAO data, being the absolute distance indicator, provides a better determination of H0, which results in a factor of 2 better determination of ΩΛ, Ωb, and Ωc.

Standard image High-resolution image

To find the limits on various deviations from the simplest model, we have explored the parameter space by combining the WMAP 5-year data with the distance measurements from the BAO and Type Ia SN observations. Here, we summarize significant findings from our analysis (also see Table 2).

  • 1.  
    Gravitational waves and primordial power spectrum. Improved simultaneous constraint on the amplitude of primordial gravitational waves and the shape of the primordial power spectrum (from WMAP+BAO+SN). In terms of the tensor-to-scalar ratio, r, we have found r < 0.22(95%CL), which is the tightest bound to date. A blue primordial spectrum index, ns > 1, now begins to be disfavored even in the presence of gravitational waves. We find no evidence for the running index, dns/dln k. The parameter space allowed for inflation models has shrunk significantly since the 3-year data release (Section 3.3), most notably for models that produce significant gravitational waves, such as chaotic or power-law inflation models, and models that produce ns > 1, such as hybrid inflation models.
  • 2.  
    Dark energy and curvature. Improved simultaneous constraint on the dark energy equation of state, w, and the spatial curvature of the universe, Ωk (from WMAP+BAO+SN). We find −0.0179 < Ωk < 0.0081(95%CL) and −0.14 < 1 + w < 0.12(95%CL). The curvature radius of the universe should be greater than Rcurv > 22 and 33 h−1 Gpc for positive and negative curvatures, respectively. The combination of WMAP, BAO, and SN is particularly powerful for constraining w and Ωk simultaneously.
  • 3.  
    Time-dependent equation of state. Using the WMAP distance priors (lA, R, z*) combined with the BAO and SN distance data, we have constrained time-dependent w. The present-day value of w, w0, is constrained as −0.33 < 1 + w0 < 0.21 (95% CL), for a flat universe.
  • 4.  
    Non-Gaussianity. Improved constraints on primordial non-Gaussianity parameters, −9 < flocalNL < 111 and −151 < fequilNL < 253 (95% CL), from the WMAP temperature data. The Gaussianity tests show that the primordial fluctuations are Gaussian to the 0.1% level, which provides the strongest evidence for the quantum origin of the primordial fluctuations.
  • 5.  
    Nonadiabaticity. Improved constraints on nonadiabatic fluctuations. The photon and matter fluctuations are found to obey the adiabatic relation to 8.9% and 2.1% for axion- and curvaton-type nonadiabatic fluctuations, respectively.
  • 6.  
    Parity violation. WMAP's limits on the TB and EB correlations indicate that parity-violating interactions that couple to photons could not have rotated the polarization angle by more than −5fdg9 < Δα < 2fdg4 between the decoupling epoch and present epoch.
  • 7.  
    Neutrino mass. With the WMAP data combined with the distance information from BAO and SN, we find a limit on the neutrino mass, ∑mν < 0.67 eV(95%CL), which is better than the WMAP-only limit by a factor of 2, owing to an additional constraint on H0 provided by BAO. The limit does not get worse very much even when w is allowed to vary, as the SN data reduce correlation between H0 and w effectively. Since we rely only on the CMB data and distance information, our limit is not sensitive to our understanding of nonlinear astrophysical effects in the large-scale structure data.
  • 8.  
    Number of neutrino species. With the WMAP data alone, we find evidence for nonzero Neff (Dunkley et al. 2009), which is likely coming from our measurement of the effect of neutrino anisotropic stress on the CMB power spectrum. With the BAO, SN, and HST added, we break the degeneracy between Ωmh2 and Neff, and find Neff = 4.4 ± 1.5 (68% CL) and 1.8 < Neff < 7.6(95%CL), which are consistent with the standard value, Neff = 3.04; thus, we do not find any evidence for the extra relativistic species in the universe.

The limits that we have obtained from our analysis in this paper are already quite stringent; however, we emphasize that they should still be taken as a prototype of what we can achieve in the future, including more integration of the WMAP observations.

A smaller noise level in the temperature data will reduce the uncertainty in non-Gaussianity parameters. An improved determination of the TE spectrum increases our sensitivity to nonadiabatic fluctuations and to the primordial gravitational waves. The E-mode polarization will be more dominated by the signal, to the point where we begin to constrain the detailed history of the reionization of the universe beyond a simple parametrization. Our limit on the B-mode polarization continues to improve, which will provide us with a better understanding of the polarized foreground. The improved TB and EB correlations will provide better limits on the cosmological birefringence.

While we have chosen to combine the WMAP data mainly with the distance indicators, one should be able to put even more stringent limits on important parameters such as r, ns, dns/dln k, w(z), mν, and Neff, by including the other datasets that are sensitive to the amplitude of density fluctuations, such as the amplitude of the galaxy power spectrum, Lyα forest, weak lensing, and cluster abundance. With the Lyα forest data from Seljak et al. (2006), for example, the limit on the running index improves from −0.068 < dns/dln k < 0.012(95%CL) to −0.034 < dns/dln k < 0.011 (95% CL) for r = 0, and −0.11484 < dns/dln k < −0.00079(95%CL) to −0.0411 < dns/dln k < 0.0067 (95% CL) for r ≠ 0. A better understanding of systematic errors in these methods will be crucial in improving our understanding of the physics of our universe. Hints for new physics may well be hidden in the deviations from the simplest six-parameter ΛCDM model.

The WMAP mission is made possible by the support of the Science Mission Directorate Office at NASA Headquarters. This research was additionally supported by NASA grants NNG05GE76G, NNX07AL75G S01, LTSA03-000-0090, ATPNNG04GK55G, and ADP03-0000-092. E.K. acknowledges support from an Alfred P. Sloan Research Fellowship. We thank M. Greason and N. Odegard for their help on the analysis of the WMAP data, and B. Griswold for the artwork. We thank U. Seljak, A. Slosar, and P. McDonald for providing the Lyα forest likelihood code, W. Percival for useful information on the implementation of the BAO data, E. Hivon for providing a new HEALPix routine to remove the monopole and dipole from the weighted temperature maps, and K. Smith for his help on implementing the bispectrum optimization algorithm. We thank A. Conley for valuable comments on our original treatment of the Type Ia SN data, and M. Kowalski, J. Guy, and A. Slosar for making the union SN data available to us. Computations for the analysis of non-Gaussianity in Section 3.5 were carried out by the Terascale Infrastructure for Groundbreaking Research in Engineering and Science (TIGRESS) at the Princeton Institute for Computational Science and Engineering (PICSciE). This research has made use of NASA's Astrophysics Data System Bibliographic Services. We acknowledge use of the HEALPix (Gorski et al. 2005), CAMB (Lewis et al. 2000), and CMBFAST (Seljak & Zaldarriaga 1996) packages.

APPENDIX A: FAST CUBIC ESTIMATORS

We use the following estimators for flocalNL, fequilNL, and bsrc (the amplitude of the point-source bispectrum):

Equation (A1)

Equation (A2)

Equation (A3)

where Fij is the Fisher matrix given by

Equation (A4)

Here, $B_{l_1l_2l_3}^{(i)}$ are theoretically calculated angular bispectra (given below), where i = 1 is used for flocalNL, i = 2 for fequilNL, and i = 3 for bsrc. The denominator of Fij contains the total power spectrum including the CMB signal (CCMBl) and noise (Nl), $\tilde{C}_l\equiv C_l^{\rm CMB}b_l^2+N_l$, and bl is the beam transfer function given in Hill et al. (2009).

While this formula allows one to estimate flocalNL, fequilNL, and bsrc simultaneously, we find that a simultaneous estimation does not change the results significantly. Therefore, we use

Equation (A5)

to simplify the analysis, as well as to make the comparison with the previous work easier. However, we do take into account a potential leakage of the point sources into flocalNL and fequilNL by using the Monte Carlo simulation of point sources, as described later. These Monte Carlo estimates of the bias in flocalNL and fequilNL due to the source contamination roughly agree with contributions from the off-diagonal terms in Equations (A1) and (A2).

Assuming white noise, which is a good approximation at high multipoles in which noise becomes important, one can analytically compute the noise power spectrum as

Equation (A6)

where Ωpix ≡ 4π/Npix is the solid angle per pixel, $M(\hat{\mathbf{n}})$ is the KQ75 mask, fsky = 0.718 is the fraction of sky retained by the KQ75 mask, σ0 is the rms noise per observation, and $N_{\rm obs}(\hat{\mathbf{n}})$ is the number of observations per pixel.

The angular bispectra are given by

Equation (A7)

Equation (A8)

Equation (A9)

where

Equation (A10)

Various functions in $B^{(i)}_{l_1l_2l_3}$ are given by

Equation (A11)

Equation (A12)

Equation (A13)

Equation (A14)

Here, $P_\Phi (k)\propto k^{n_s-4}$ is the primordial power spectrum of Bardeen's curvature perturbations, and gTl(k) is the radiation transfer function that gives the angular power spectrum as Cl = (2/π)∫k2dkPΦ(k)g2Tl(k).

The skewness parameters, Si, are given by (Komatsu et al. 2005; Babich 2005; Creminelli et al. 2006; Yadav et al. 2008)

Equation (A15)

Equation (A16)

Equation (A17)

where w3 is the sum of the weighting function cubed:

Equation (A18)

For a uniform weighting, the weighting function is simply given by the KQ75 mask, that is, $W(\hat{\mathbf{n}})=M(\hat{\mathbf{n}})$, which gives w3 = 4πfsky. For our measurements of flocalNL, fequilNL, and bsrc, we shall use a "combination signal-plus-noise weight," given by

Equation (A19)

where σ2cmb ≡ (1/4π)∑l(2l + 1)Ccmblb2l is the CMB signal variance, σ0 is the rms noise per observation, and $N_{\rm obs}(\hat{\mathbf{n}})$ is the number of observations per pixel. This combination weighting yields a nearly optimal performance for bsrc and fequilNL, whereas it results in a minor improvement in flocalNL over the uniform weighting. The bracket, 〈〉MC, denotes the average over Monte Carlo realizations and "sim" denotes that these are the filtered maps of the Monte Carlo realizations.

The filtered temperature maps, A, B, C, D, and E, are given by

Equation (A20)

Equation (A21)

Equation (A22)

Equation (A23)

Equation (A24)

respectively. Here, lmax is the maximum multipole that we use in the analysis. We vary lmax to see how much the results depend on lmax.

Equations (A15) and (A16) involve the integrals over the conformal distances, r. We evaluate these integrals as

Equation (A25)

We use the bispectrum optimization algorithm described in Smith & Zaldarriaga (2006) to compute the optimal weights, (wi)opt, and decide on which quadrature points, ri, to keep. We choose the number of quadrature points such that the bispectrum computed in this way agrees with that from more dense sampling in r to 10−5, which typically gives ∼5 quadrature points for flocalNL and ∼15 points for fequilNL.52

The measurement of these estimators proceeds as follows.

  • 1.  
    Generate the simulated realizations of CMB signal maps, TS, from the input signal power spectrum, Ccmbl, and the beam transfer function, bl. We have generated 300 realizations for the analysis given in this paper.
  • 2.  
    Add random noise, TN, using the rms noise per pixel given by $\sigma _0/\sqrt{N_{\rm obs}(\hat{\mathbf{n}})}$.
  • 3.  
    Add point sources. We use a simplified treatment for the source simulation,
    Equation (A26)
    where Ωpix is the solid angle of pixel, x = hν/(kBTcmb) = 56.80 GHz (for Tcmb = 2.725 K), epsilon is a Poisson random variable with the mean of 〈epsilon〉 = nsrcΩpix, and nsrc is the average number of sources per steradians. This simplified model assumes that there is only one population of sources with a fixed flux, Fsrc, and each source's flux is independent of frequency. We choose nsrc = 0.85 sr−1 and Fsrc = 0.5 Jy, which yields the source power spectrum in the Q band of Cps = 8.7 × 10−3 μK2 sr and the source bispectrum in the Q band of bsrc = 8.7 × 10−5 μK3 sr2, which roughly reproduce the measured values. However, this model does not reproduce the source counts very well. (The source density of nsrc = 0.85 sr−1 at 0.5 Jy is too low.) The main purpose of this phenomenological model is to reproduce the power spectrum and bispectrum—we include point sources in the simulations, in order to take into account the potential effects of the unresolved sources on primordial non-Gaussianity, flocalNL and fequilNL.
  • 4.  
    Coadd them to create the simulated temperature maps, $T(\hat{\mathbf{n}})=T_{S}(\hat{\mathbf{n}})+T_{N}(\hat{\mathbf{n}})+T_{\rm src}(\hat{\mathbf{n}})$.
  • 5.  
    Mask and weight the temperature maps, $T(\hat{\mathbf{n}})\rightarrow \tilde{T}(\hat{\mathbf{n}})=W(\hat{\mathbf{n}})T(\hat{\mathbf{n}})$, where $W(\hat{\mathbf{n}})$ is given by Equation (A19).
  • 6.  
    Remove the monopole and dipole from $\tilde{T}(\hat{\mathbf{n}})$.
  • 7.  
    Compute the harmonic coefficients as
    Equation (A27)
  • 8.  
    Generate the filtered maps, Asim, Bsim, Csim, Dsim, and Esim, and compute the appropriate Monte Carlo averages such as 〈AsimBsimMC. This is the most time-consuming part.
  • 9.  
    Compute the filtered maps from the WMAP data. When we coadd the V- and W-band data, we weight them as TV + W = (TV + 0.9TW)/1.9. The beam transfer function of the coadded map is given by bV + Wl = (bVl + 0.9bWl)/1.9, and σ0/Nobs of the coadded map is given by
    Equation (A28)
  • 10.  
    Compute the skewness parameters, Si, from the filtered WMAP data, and obtain flocalNL, fequilNL, and bsrc, either jointly or separately.
  • 11.  
    Compute these parameters from the simulated realizations as well, and obtain the uncertainties.

For the computations of gTl(k) and generation of Monte Carlo realizations, we have used the maximum-likelihood values of the WMAP 3-year data (the power-law ΛCDM model fit by the WMAP data alone with the Sunyaev–Zel'dovich effect marginalized): Ωb = 0.0414, ΩCDM = 0.1946, ΩΛ = 0.7640, H0 = 73.2 km s-1 Mpc−1, τ = 0.091, and ns = 0.954 (Spergel et al. 2007). These parameters yield the conformal distance to t = 0 as cτ0 = 14.61  Gpc.

APPENDIX B: AXION

In this Appendix, we derive relations among the tensor-to-scalar ratio r, the axion mass density Ωah2, the entropy-to-curvature perturbation ratio α, the phase of the Pecci–Quinn field θa, and the axion decay constant fa.

Let us write the expectation value of the complex Pecci–Quinn field, ψPQ, as

Equation (B1)

where fa is the axion decay constant and θa is the phase. Quantum fluctuations during inflation generate fluctuations in the phase, δθa, as

Equation (B2)

As the number density of axions scales as the phase squared, na ∝ θ2a, the mass density fluctuation is given by

Equation (B3)

As the energy density of axions was negligible during inflation, the axion density perturbation, δρaa, would produce the isocurvature perturbation. While radiation (including photons) is generated by decay of inflaton fields, (some of) dark matter is in the form of axions whose generation is independent of photons; thus, the entropy perturbation between photons and axions would be generated. We assume that axions were not in thermal equilibrium with photons in the subsequent evolution of the universe.

The entropy perturbations and curvature perturbations are given, respectively, by

Equation (B4)

where Ωa ⩽ Ωc is the axion mass density, Hk is the expansion rate during inflation at which the wavenumber k went outside of the horizon, ϕk is the value of inflaton at the same time, $\epsilon \equiv -\dot{H}/H\approx \big(M_{\rm pl}^2/2\big)(V^{\prime }/V)^2$ is the usual slow-roll parameter, V(ϕ) is the inflaton potential, and $M_{\rm pl}=1/\sqrt{8\pi G}$ is the reduced Planck mass. We have used the slow-roll approximation, $\dot{\phi }\approx -V^{\prime }/(3H)$ and H2V/(3M2pl).

Here, let us comment on our choice of m = 1, which makes $k^3P_{\cal S}\propto k^{m-1}$ independent of k. Since $k^3P_{\cal S}\propto H_k^2\propto k^{-2\epsilon }$, where $\epsilon =-\dot{H}/H^2$, this choice corresponds to having a very small slow-roll parameter, epsilon ≪ 1. This is consistent with our limit on the curvature power spectrum, ns = 1 + 6epsilon − 4η ≃ 1 − 4η < 1, where η is another slow-roll parameter. As the current limit is 1 − ns ≈ 4η ≃ 0.04, our approximation, m = 1, is valid for epsilon < 0.01. It should be straightforward to extend our analysis to the case in which m ≠ 1.

By dividing $P_{\cal S}(k)$ by $P_{\cal R}(k)$, we find the entropy-to-curvature perturbation ratio for axions, α0(k), as

Equation (B5)

At this point, it is clear that one cannot solve this constraint uniquely for any of epsilon, fa, or θa.

In order to break the degeneracy, we use the axion mass density (Kawasaki & Sekiguchi 2008, and references therein)

Equation (B6)

where γ is a dilution factor, representing the amount by which the axion density could have been diluted by a late-time entropy production between the QCD phase transition at ∼200 MeV and the epoch of nucleosynthesis at ∼1 Mpc.

Combining equation (B5) and (B6) to eliminate the phase, θa, and using the relation between the tensor-to-scalar ratio r and the slow-roll parameter epsilon, r = 16epsilon, we find

Equation (B7)

Alternatively, we can eliminate the axion decay constant, fa, to obtain

Equation (B8)

This is Equation (48).

APPENDIX C: EQUATION OF STATE OF DARK ENERGY: A NEW PARAMETRIZED FORM

In this Appendix, we describe the models of dark energy that we explore in Section 5.4. Our goal is to obtain a sensible form of time-dependent dark energy equation of state, w(a). One of the most commonly used form of w(a) is a linear form (Chevallier & Polarski 2001; Linder 2003)

Equation (C1)

where w0 and wa parametrize the present-day value of w and the first derivative. However, this form cannot be adopted as it is when one uses the CMB data to constrain w(a). Since this form is basically the leading-order term of a Taylor series expansion, the value of w(a) can become unreasonably too large or too small when extrapolated to the decoupling epoch at z* ≃ 1090 (or a* ≃ 9.17 × 10−4), and thus one cannot extract meaningful constraints on the quantities, such as w0 and wa, that are defined at the present epoch.

To avoid this problem, yet to keep a close contact with the previous work in the literature, we shall consider an alternative parametrized form. Our idea is the following: we wish to keep the form given by Equation (C1) at low redshifts, lower than some transition redshift, ztrans. However, we demand that w(a) approach −1 at higher redshifts, z > ztrans. This form of w(a), therefore, has the following property: at early times, before the transition redshift, ztrans, dark energy was just like a cosmological constant, and thus the dark energy density was nearly constant, that is, ρde(z > ztrans) ≈ constant. Then, dark energy began to become dynamical at zztrans, with the equation of state given by the conventional linear form, Equation (C1).

Some of the properties of our form of w(a) are similar to those of "thawing models," (Caldwell & Linder 2005) in which a scalar field was moving very slowly initially, giving w(a) ≈ −1 at early times, and then began to move faster toward low redshifts, causing w(a) to deviate more and more from −1 at low redshifts. Our parametrization can describe a more general class of models than single scalar field models, as it allows for w to go below −1. However, models that are based upon a single scalar field cannot have w < −1 (e.g., Hu 2005). The "Forever regular" parametrization, explored in Wang & Tegmark (2004), also approaches a constant density at early times, if the late-time equation of state is w < −1. The "Kink model" explored in Bassett et al. (2002, 2004) and Corasaniti et al. (2004) also extrapolates a constant equation of state at early times to a different constant equation of state at late times. Our parametrization is more general than theirs, as their form only allows for a constant equation of state at late times.

We wish to find a smooth interpolation between wearly = −1 and wlate = w0 + (1 − a)wa. We begin by writing

Equation (C2)

where atrans = 1/(1 + ztrans), and the function f(x) goes to zero for x ≪ 1 and to unity for x ≫ 1. Here, $\tilde{w}(a)$ is the form of w at low redshifts. Any function that has this property is adequate for f(x). We choose

Equation (C3)

which gives the desired form of the equation of state of dark energy,

Equation (C4)

where

Equation (C5)

One nice property of this form is that it allows one to obtain a closed, analytical form of the effective equation of state, weff(a), which gives the evolution of dark energy density, $\rho _{\rm de}(a)=\rho (0)a^{-3[1+w_{\rm eff}(a)]}$:

Equation (C6)

This property allows one to compute the expansion rate, H(a) (Equation (7)), and hence the distance (Equation (2)), easily.

Finally, we use the present-day value of w, w0w(z = 0), and the first derivative, w' ≡ dw/dz|z = 0, as free parameters. They are related to $\tilde{w}_0$ and $\tilde{w}_a$ as

Equation (C7)

Equation (C8)

The inverse relations are

Equation (C9)

Equation (C10)

In the limit of very early transition, atrans ≪ 1, one finds $w_0\approx \tilde{w}_0$ and $w^{\prime }\approx \tilde{w}_a$, as expected. This completes the description of our form of w(a).

Figure 20 shows the evolution of dark energy density, $\rho _{\rm de}(z)=\rho (0)(1+z)^{3[1+w_{\rm eff}(a)]}$, the equation of state, w(z), and the effective equation of state, weff(z), computed from Equation (C4) and (C6). We choose w0 = −1.1 and w' = 1, which are close to the best-fitting values that we found in Section 5.4 (see Figure 14). We show three curves for the transition redshifts of ztrans = 0.5, 2, and 10. We find that the form of w(z) that we have derived achieves our goal: w(z) approaches −1 and the dark energy density tends to a constant value at high redshifts, giving sensible results at the decoupling epoch. The dark energy density is totally subdominant compared to the matter density at high redshifts, which is also desirable.

Figure 20.

Figure 20. Evolution of dark energy for w0 = −1.1 and w' = 1, and various transition redshifts, ztrans = 0.5, 2, and 10, above which w(z) approaches −1. (Top left) Evolution of the dark energy to matter density ratio as a function of z. Note that the vertical axis has been multiplied by 106. (Top right) Evolution of the dark energy density relative to the dark energy density at present. The dark energy density was nearly constant at high redshifts above ztrans; thus, these models can describe the "thawing models" (Caldwell & Linder 2005), in which dark energy was nearly constant at early times, and had become dynamical at lower redshifts. (Bottom left) Evolution of the equation of state, w(z) = Pde(z)/ρde(z). By construction of the model, w(z) approaches −1 beyond ztrans. (Bottom right) Evolution of the effective equation of state, weff(z), which determines the evolution of dark energy density as $\rho _{\rm de}(z)=\rho _{\rm de}(0)(1+z)^{3[1+w_{\rm eff}(z)]}$.

Standard image High-resolution image

The constraints that we have obtained for w0 and w' are not sensitive to the exact values of ztrans (see the right panel of Figure 14). This is because all of the curves shown in Figure 20 are very similar at z ≲ 1, where the BAO and SN data are currently available.

APPENDIX D: COMPARISON OF SN COMPILATIONS AND EFFECTS OF SYSTEMATIC ERRORS

In Table 14, we show the ΛCDM parameters derived from WMAP+BAO+SN, where we use various SN compilations: "Union" for the latest union compilation (Kowalski et al. 2008), "Union+Sys.Err." for the union compilation with systematic errors included, "Davis" for the previous compilation by Davis et al. (2007), and "Alternative" for the compilation that we used in the original version of this paper (version 1 of arXiv:0803.0547).

Table 14. Comparison of ΛCDM Parameters from WMAP+BAO+SN with Various SN Compilations

Class Parameter Uniona Union+Sys. Err.b Davisc Alternatived
Primary 100Ωbh2 2.267+0.058−0.059 2.267 ± 0.059 2.270 ± 0.060 2.265 ± 0.059
  Ωch2 0.1131 ± 0.0034 0.1134+0.0036−0.0037 0.1121 ± 0.0035 0.1143 ± 0.0034
  ΩΛ 0.726 ± 0.015 0.725 ± 0.016 0.732+0.014−0.015 0.721 ± 0.015
  ns 0.960 ± 0.013 0.960 ± 0.013 0.962 ± 0.013 0.960+0.014−0.013
  τ 0.084 ± 0.016 0.085 ± 0.016 0.085 ± 0.016 0.084 ± 0.016
  $\Delta ^2_{\cal R}(k_0^{\rm e})$ (2.445 ± 0.096) × 10−9 (2.447+0.096−0.095) × 10−9 (2.429+0.096−0.095) × 10−9 (2.457+0.092−0.093) × 10−9
Derived σ8 0.812 ± 0.026 0.813+0.026−0.027 0.807 ± 0.027 0.817 ± 0.026
  H0 70.5 ± 1.3 km s−1 Mpc−1 70.4 ± 1.4 km s−1 Mpc−1 70.9 ± 1.3 km s−1 Mpc−1 70.1 ± 1.3 km s−1 Mpc−1
  Ωb 0.0456 ± 0.0015 0.0458 ± 0.0016 0.0451+0.0016−0.0015 0.0462 ± 0.0015
  Ωc 0.228 ± 0.013 0.229+0.014−0.015 0.223 ± 0.013 0.233 ± 0.013
  Ωmh2 0.1358+0.0037−0.0036 0.1361+0.0038−0.0039 0.1348 ± 0.0038 0.1369 ± 0.0037
  zfreion 10.9 ± 1.4 10.9 ± 1.4 10.9 ± 1.4 10.8 ± 1.4
  tg0 13.72 ± 0.12 Gyr 13.72 ± 0.12 Gyr 13.71 ± 0.12 Gyr 13.73 ± 0.12 Gyr

Notes. aCompilation by Kowalski et al. (2008) without the systematic errors included. bCompilation by Kowalski et al. (2008) with the systematic errors included. cCompilation by Davis et al. (2007). dCompilation used in the original version of this paper (version 1 of arXiv:0803.0547). ek0 = 0.002 Mpc−1. $\Delta ^2_{\cal R}(k)=k^3P_{\cal R}(k)/(2\pi ^2)$ (Equation (15)). f"Redshift of reionization," if the universe was reionized instantaneously from the neutral state to the fully ionized state at zreion. gThe present-day age of the universe.

Download table as:  ASCIITypeset image

For the "Alternative" compilation, we have combined measurements from the Hubble Space Telescope (HST; Riess et al. 2004, 2007), the SNLS (Astier et al. 2006), and the Equation of State: ESSENCE survey (Wood-Vasey et al. 2007), as well as some nearby Type Ia SNe. In the "Davis" and "Alternative" compilations, different light curve fitters were used for the SN data taken by different groups, and thus these compilations were not as optimal as the union compilation, for which the same SALT fitter was used for all the SNe samples. Moreover, the union compilation is the largest of all. For this reason, we have decided to update all the cosmological parameters using the union compilation.

Nevertheless, we find that all of these compilations yield similar results: the mean values shift no more than ∼0.5σ.

The effects of the systematic errors in the Type Ia SN data on the ΛCDM parameters are also very small for WMAP+BAO+SN; however, the effects on the dark energy parameters, w0 and w', turn out to be significant. In Figure 21, we show the two-dimensional joint constraint on w0 and w' from WMAP+BAO+SN+BBN (also see Section 5.4.2) with the systematic errors included. Comparing this with Figure 15, where the systematic errors are ignored, we find that the constraints on w0 and w' weaken significantly: we find w0 = −1.00 ± 0.19 and w' = 0.11 ± 0.70 with the systematic errors included, whereas w0 = −1.04 ± 0.13 and w' = 0.24 ± 0.55 without the systematic errors.

Figure 21.

Figure 21. The same as Figure 15, but the systematic errors in the Type Ia SN data are included.

Standard image High-resolution image

Footnotes

  • WMAP is the result of a partnership between Princeton University and NASA's Goddard Space Flight Center. Scientific guidance is provided by the WMAP Science Team.

  • 15 

    Here, "TB" refers to the power spectrum of a cross-correlation between the temperature and B-mode polarization, while "EB" refers to a correlation between the E-mode and B-mode polarization.

  • 16 

    This quantity, Aps, is the value of the power spectrum, Cl, from unresolved point sources in the Q band, in units of the antenna temperature. To convert this value to the thermodynamic units, use Cps = 1.089Aps (Nolta et al. 2009).

  • 17 

    For our limits on the residual polarized foreground contamination, see Dunkley et al. (2008).

  • 18 

    In Percival et al. (2007), the authors used a different notation for the drag redshift, z*, instead of zd. We have confirmed that they have used equation (6) of Eisenstein & Hu (1998) for rs, which makes an explicit use of the drag redshift (W. Percival 2008, private communication).

  • 19 

    We use a Gaussian prior on $A=D_V(z=0.35)\sqrt{\Omega _mH_0^2}/(0.35c)= 0.469(n_s/0.98)^{-0.35}\pm 0.017$.

  • 20 

    See, for example, Watanabe & Komatsu (2006) for the spectrum of the primordial gravitational waves itself.

  • 21 

    We have performed a similar, but different, analysis in Section 6.2 of Page et al. (2007). In this paper, we include both the scalar and tensor contributions to EE, whereas in Page et al. (2007), we have ignored the tensor contribution to EE and found a somewhat tighter limit, r < 4.5 (95% CL), from the low-l polarization data. This is because, when the tensor contribution was ignored, the EE polarization could still be used to fix τ, whereas in our case, r and τ are fully degenerate when r ≳ 1 (see Figure 3), as the EE is also dominated by the tensor contribution for such a high value of r.

  • 22 

    See Polnarev et al. (2008); Miller et al. (2007) for a way to constrain r from the TE power spectrum alone.

  • 23 

    This is the one-dimensional marginalized 95% limit. From the joint two-dimensional marginalized distribution of ns and r, we find r < 0.27 (95% CL) at ns = 0.99. See Figure 2.

  • 24 

    This is the one-dimensional marginalized 95% limit. From the joint two-dimensional marginalized distribution of ns and r, we find ns < 1.007 (95% CL) at r = 0.2. See Figure 2.

  • 25 

    For recent surveys of inflation models in light of the WMAP 3-year data, see Alabidi & Lyth (2006a), Kinney et al. (2006) and Martin & Ringeval (2006).

  • 26 

    This classification scheme is similar to, but different from, the most widely used one, which is based upon the field value (small-field, large-field, hybrid) (Dodelson et al. 1997; Kinney 1998).

  • 27 

    These choices are used to sample the space of positive curvature models. Realistic potentials may be much more complicated: see, for example, Destri et al. (2008) for the WMAP 3-year limits on trinomial potentials. Also, the classification scheme based upon derivatives of potentials sheds little light on the models with noncanonical kinetic terms such as k-inflation (Armendariz-Picon et al. 1999; Garriga & Mukhanov 1999), ghost inflation (Arkani-Hamed et al. 2004), Dirac–Born–Infeld (DBI) inflation (Silverstein & Tong 2004; Alishahiha et al. 2004), or infrared-DBI (IR-DBI) inflation (Chen 2005b, 2005a), as the tilt, ns, also depends on the derivative of the effective speed of sound of a scalar field (for recent constraints on this class of models from the WMAP 3-year data, see Bean et al. 2007a, 2008; Lorenz et al. 2008).

  • 28 

    The positive sign case, V(ϕ) ∝ 1 + cos(ϕ/f), belongs to a negative curvature model when ϕ/f ≪ 1. See Savage et al. (2006) for constraints on this class of models from the WMAP 3-year data.

  • 29 

    In the language of Section 3.4 of Peiris et al. (2003), the models 1 and 2 belong to "small positive curvature models," and model 3 to "large positive curvature models" for $\tilde{\phi }\ll 1$, "small positive curvature models" for $\tilde{\phi }\gg 1$, and "intermediate positive curvature models" for $\tilde{\phi }\sim 1$.

  • 30 

    To simplify our discussion, we ignore the dark energy contribution, and assume that the universe is dominated matter at the present epoch. This leads to a small error in the estimated lower bound on Ntot.

  • 31 

    For simplicity, we assume that reheating occurred as soon as inflation ended.

  • 32 

    Note that flocalNL can be related to the quantities discussed in earlier, pioneering work: −Φ3/2 (Gangui et al. 1994), −Ainfl/2 (Wang & Kamionkowski 2000), and −α (Verde et al. 2000).

  • 33 

    For a pedagogical introduction to the bispectrum (three-point function) and trispectrum (four-point function) and various topics on non-Gaussianity, see Komatsu (2001) and Bartolo et al. (2004).

  • 34 

    Since the angular bispectrum is the harmonic transform of the angular three-point function, it forms a triangle in the harmonic space. While there are many possible triangles, the "squeezed triangles," in which the two wave vectors are long and one is short, are most sensitive to flocalNL (Babich et al. 2004).

  • 35 

    Previously, the Kp0 mask was defined by the K-band map, which contains CMB as well as the foreground emission. By cutting bright pixels in the K-band map, it could be possible to cut also the bright CMB pixels, introducing the negative skewness in the distribution of CMB. Since we did not include isolated "islands" on the high Galactic latitudes, some of which could be bright CMB spots, in the final mask when we defined the Kp0 mask, the skewness bias mentioned above should not be as large as one would expect, if any. Nevertheless, with the new definition of mask, the masked maps are free from this type of bias. For more details on the definition of the mask, see Gold et al. (2009).

  • 36 

    The uncertainty for lmax > 500 is slightly larger than that for lmax = 500 due to a small suboptimality of the estimator of flocalNL (Yadav et al. 2008).

  • 37 

    A more general relation is $\delta \rho _x/\dot{\rho }_x=\delta \rho _y/\dot{\rho }_y$, where x and y refer to some energy components. Using the energy conservation equation, $\dot{\rho }_x=-3H(1+w_x)\rho _x$ (where wx is the equation of state for the component x), one can recover Equation (36), as wr = 1/3 and wm = 0. For a recent discussion on this topic, see, for example, Weinberg (2003).

  • 38 

    This variable, B, is the same as B used in Gordon & Lewis (2003), including the sign convention.

  • 39 

    Note that the sign convention of flocalNL inLyth & Rodriguez (2005) is such that flocalNL,WMAP = −flocalNL, theirs.

  • 40 

    See Lepora (1998); Klinkhamer (2000); Adam & Klinkhamer (2001) for studies on a spacelike pα, including its signatures in CMB.

  • 41 

    The 68% limit is w = −1.15+0.21−0.22 (WMAP+BAO; Ωk = 0).

  • 42 

    The 68% limit is w = −0.977+0.065−0.064 (WMAP+SN; Ωk = 0).

  • 43 

    The 68% limit is w = −0.992+0.061−0.062 (WMAP+BAO+SN; Ωk = 0).

  • 44 

    For the WMAP+BAO limit, there is a long degenerate valley with a significant volume at w < −1. Models anywhere in this valley are good fits to both datasets. It is dangerous to marginalize over these degenerate parameters as conclusions are very sensitive to the choice and the form of priors.

  • 45 

    The 68% limits are Ωk = −0.0049+0.0066−0.0064 and w = −1.006+0.067−0.068 (WMAP+BAO+SN).

  • 46 

    To obtain the WMAP+BAO contours in the right panel of Figure 12, we have re-weighted the WMAP+BAO data in the middle panel of Figure 12 by the likelihood ratio of L(Eisenstein'sBAO)/L(Percival'sBAO). As a result, the contours do not extend to w ∼ 0; however, the contours would extend more to w ∼ 0 if we ran a MCMC from the beginning with the Eisenstein et al. BAO.

  • 47 

    The 68% intervals are w0 = −1.06 ± 0.14 and w' = 0.36 ± 0.62 (WMAP+BAO+SN; Ωk = 0).

  • 48 

    The 68% intervals are w0 = −1.04 ± 0.13 and w' = 0.24 ± 0.55 (WMAP+BAO+SN+BBN; Ωk = 0).

  • 49 
  • 50 

    While the current cosmological data are not yet sensitive to the mass of individual neutrino species, that is, the mass hierarchy, this situation may change in the future, with high-z galaxy redshift surveys or weak lensing surveys (Takada et al. 2006; Slosar 2006; Hannestad & Wong 2007; Kitching et al. 2008; Abdalla & Rawlings 2007).

  • 51 

    This is similar to what happens to the curvature constraint from the CMB data alone. A positive curvature model, Ωk < 0, shifts the acoustic peaks to lower multipoles; however, this shift can be absorbed by a reduction in the value of H0. As a result, a closed universe with Ωk ∼ −0.3 and ΩΛ ∼ 0 is still a good fit, if Hubble's constant is as low as H0 ∼ 30 km s-1 Mpc−1 (Spergel et al. 2007).

  • 52 

    Note that Smith & Zaldarriaga (2006) used 10−6 as a criterion, which gives more quadrature points to evaluate. We find that 10−5 is sufficient for the size of statistical and systematic errors in the current measurements.

Please wait… references are loading.
10.1088/0067-0049/180/2/330