Articles

CONSTRAINTS ON PERTURBATIONS TO THE RECOMBINATION HISTORY FROM MEASUREMENTS OF THE COSMIC MICROWAVE BACKGROUND DAMPING TAIL

, , , and

Published 2013 February 1 © 2013. The American Astronomical Society. All rights reserved.
, , Citation M. Farhang et al 2013 ApJ 764 137 DOI 10.1088/0004-637X/764/2/137

0004-637X/764/2/137

ABSTRACT

The primordial cosmic microwave background at small angular scales is sensitive to the ionization and expansion history of the universe around the time of recombination. This dependence has been exploited to constrain the helium abundance and the effective number of relativistic species. Here we focus on allowed ionization fraction trajectories, xe(z), by constraining low-order principal components of perturbations to the standard recombination scenario (xe-eigenmodes) in the circa 2011 South Pole Telescope (SPT), Atacama Cosmology Telescope (ACT), and Wilkinson Microwave Anisotropy Probe seven-year data (WMAP7). Although the trajectories are statistically consistent with the standard recombination, we find that there is a tension similar to that found by varying the helium fraction. As this paper was in press, final SPT and ACT datasets were released and we applied our framework to them: we find that the tension continues, with slightly higher significance, in the new 2012 SPT data, but find no tension with the standard model of recombination in the new 2012 ACT data. We find that the prior probabilities on the eigenamplitudes are substantially influenced by the requirement that xe-trajectories conserve electron number. We propose requiring a sufficient entropy decrease between posterior and prior marginalized distributions be used as an xe-mode selection criterion. We find that in the case of the 2011 SPT/ACT+WMAP7 data only two modes are constrainable, but upcoming ACTPol, Planck, and SPTPol data will be able to test more modes and more precisely address the current tension.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

A primary goal of cosmic microwave background (CMB) experiments after the discovery of the acoustic features in the angular power spectrum was to measure the damping tail. This was achieved by the CBI (Sievers et al. 2003, 2007) and ACBAR (Goldstein et al. 2003; Reichardt et al. 2009) experiments, and spectacularly verified by the Atacama Cosmology Telescope (ACT; e.g., Dunkley et al. 2011) and the South Pole Telescope (SPT; e.g., Keisler et al. 2011). There is even higher precision data on its way from ACT, SPT, and Planck. Initially, the goal was simply to confirm that the damping tail agreed with theoretical predictions based on cosmological parameters determined from the first set of peaks and troughs. Contemporary high-precision data open the opportunity for new constraints of cosmological parameters that specifically influence the damping tail and have little influence on larger scales. The parameters wagged the tail, and now the tail can wag the parameters. Examples of these parameters are the primordial helium abundance Yp and the effective number of relativistic species, Neff. Both experiments have hinted at a deficit of power in the damping tail through the Yp and Neff probes (Keisler et al. 2011; Dunkley et al. 2011; Hou et al. 2011). Other physical possibilities have also been investigated to explain this damping tail tension; these include dark radiation (Smith et al. 2012; Eggers Bjaelde et al. 2012), the annihilation or decay of dark matter particles (Giesen et al. 2012), cosmic strings (Lizarraga et al. 2012), or a high-frequency cosmic gravitational wave background (Sendra & Smith 2012).

CMB anisotropies are suppressed on small scales by shear viscosity and thermal diffusion when the photons and electrons are tightly coupled, and by higher order transport effects as the photons break out from their random-walk Thomson scatterings by electrons to approach free-streaming. The physics of Silk damping (Silk 1968) has been heavily explored, as reviewed in Bond (1996), Hu & White (1997), and Sievers et al. (2003), and is very accurately computed numerically in CMB transport codes (Seljak & Zaldarriaga 1996; Lewis et al. 2000). One can associate a characteristic damping wavenumber kD(z) with the processes, determined by the steep variation in the mean free path (ne(zT)−1 with redshift, where ne(z) is the free-electron number density at redshift z and σT is the Thomson scattering cross section. The overall effect on the CMB power spectrum C can be approximately characterized by an exponentially damped envelope suppressing the baryon acoustic oscillations, encoded in an associated angular multipole damping scale ℓDkD(zdec)χ(zdec) ∼ 1350, where χ(zdec) is the comoving distance to the last scattering redshift zdec ∼ 1088.

Our focus here is how the damping tail is impacted by perturbations of the free-electron fraction, xe(z) = ne/(np + nH i), where np and nH i are the number density of ionized and neutral hydrogen atoms, during cosmological recombination in a model-independent way. This complements model-independent studies of H(z) variations (Samsing et al. 2012). Opening the constraint to the functional form of xe(z) broadens the range of non-standard physics that can be tested, such as energy injection from decaying particles or dark matter annihilation (Chen & Kamionkowski 2004; Padmanabhan & Finkbeiner 2005), which cause a delay of recombination (Peebles et al. 2000).

The present data are only suggestive of a damping deficit, so discrimination between physical causes is not yet possible, but we develop a framework for considering general modifications and the impact of physical requirements on the posterior distributions. The recombination history is a central quantity in the interpretation of CMB data, and it has evolved significantly over time as more and more effects have been included (Zeldovich et al. 1968; Peebles 1968; Seager et al. 1999), and, with extensive recent work, it may now have nearly converged (Chluba & Thomas 2011; Ali-Haïmoud & Hirata 2011). It is important, however, to test directly whether tensions may exist which suggest that there are missing elements to the recombination story. For example, residual uncertainties in xe(z) directly affect our ability to distinguish between different inflationary scenarios (Hu et al. 1995; Lewis et al. 2006; Rubiño-Martín et al. 2010; Shaw & Chluba 2011).

In this paper we follow the approach of Farhang et al. (2012) to generate eigenmodes for perturbations to the high-redshift ionization history focusing on the epoch of recombination. The method is applied to the seven-year Wilkinson Microwave Anisotropy Probe (WMAP) data combined with the 2011 SPT and ACT data and their recent upgrades, using a Fisher matrix eigenanalysis (Section 2) of finely binned redshift-localized xe-modes. In Farhang et al. (2012) the eigenanalysis was performed for simulated data sets consisting of CMB maps (or their spherical harmonic amplitudes am). Here, the modes are defined from the measured CMB bandpower errors, which leads to a different structure for the Fisher matrix. Further extending Farhang et al. (2012), we jointly treat the foreground nuisance parameters of the ACT, SPT, and WMAP7 experimental data. We use these data to constrain the amplitude of deviations from the standard recombination scenario in Section 3. In Section 4, we investigate the impact of imposing electron number conservation (through xe(z)) on the initial and final probability distributions of the parameters and quantify its effect relative to the information delivered by the data. We conclude with a brief discussion in Section 5.

2. EIGENMODES FOR PERTURBATIONS TO THE HIGH-z IONIZATION HISTORY

2.1. Eigenmode Construction

The functional form of deviations from the standard recombination history can be decomposed into a set of uncorrelated functions ranked by their significance in the data, here the increasing order of their forecasted errors. We construct these functions from the inverse of the Fisher information matrix F for the given combination of data used,

Equation (1)

where the Cb's are the simulated bandpower measurements for a fiducial set of cosmic parameters and for a given experiment, and ${\rm Cov}_{bb^{\prime }}$ is the bandpower covariance matrix for that experiment. The parameters qi include recombination perturbation parameters as well as the standard cosmological parameters and the various nuisance parameters for the experiments and for secondary effects. This form for Fij assumes uniform prior distributions for the parameters, Gaussian likelihoods for the bandpower data points, and that information on the parameters mainly comes from the mean of the bandpowers, Cb, rather than their covariance, Cov. The latter can be verified for the high multipoles of interest by comparing the contributions to the Fisher matrix from the mean and the covariance of data (see, e.g., Tegmark et al. 1997). Gaussianity is a reasonable approximation again for the high multipoles which are of interest in this work (see, e.g., Verde et al. 2003). Further, it is the assumption adopted in the released likelihood codes for SPT and ACT.

We use δln(xe(z)) as the perturbation parameter. xe is the parameter of direct relevance to probe the atomic physics involved at recombination. It is also straightforward to limit its variations within physical ranges. However, it does not affect the CMB anisotropies as closely as, e.g., ne and τ do. Using ne would also decrease possible degeneracies with baryonic matter density. Using the log in the expansion balances the low- and high-z regimes, but the data define the region of dominant impact, namely, around decoupling, rather solidly in the hydrogen recombination regime. These perturbations can be represented interchangeably by any basis that represents their degrees of freedom and does not produce numerical errors (see Farhang et al. 2012 for a discussion on other parameterizations as well as various extended and localized basis functions). Here we have used the cubic M4 spline (see, e.g., Monaghan 2005) representation, and confirmed convergence against an increasing number of basis functions, for the explicit Fisher form, Equation (1), for these experiments.

As mentioned above, the Fisher matrix considered here includes the xe(z) amplitudes, the instrument-dependent nuisance parameters, and the six primary cosmological parameters $(\Omega _{\rm b}h^2,\Omega _{\rm dm}h^2,\theta _{\rm s},\tau , n_{\rm s},\Delta ^2_{\mathcal{R}})$, respectively, describing the physical baryon density, physical dark matter density, the angular size of sound horizon at the last scattering surface, the reionization optical depth, the scalar spectral index, and the curvature fluctuation amplitude. In taking the full Fisher inverse and focusing on sub-blocks of it, we are in effect marginalizing over the parameter directions not in the block in the approximation of a Gaussian posterior.

To analyze multiple experiments with different sets of nuisance parameters, their individual Fisher matrices should be constructed and marginalized to give effective matrices containing only standard cosmological and xe-perturbation parameters that are common among all experiments. The effective Fisher matrices, now with the same dimension, are added to get the total Fisher matrix $\tilde{\bf F}$. The xe-perturbation eigenmodes are the eigenvectors of the perturbation block of ${\tilde{\bf F}}^{-1}$ with their uncertainties estimated from the roots of the corresponding eigenvalues, assuming a Gaussian distribution for the mode amplitudes in the vicinity of their maximum likelihoods. Henceforth, the xe-eigenmodes are unambiguously referred to as the eigenmodes or simply the modes. Details on the Fisher eigenanalysis for multiple experiments with different nuisance parameters are discussed in the Appendix. To generate the bandpowers required in the Fisher matrix construction we modified the publicly available code Camb4 (Lewis et al. 2000) to include the more general ionization histories required for this work.

2.2. Data Sets and Their Eigenmodes

Throughout this work, the SPT (Keisler et al. 2011) and ACT (Dunkley et al. 2011) measurements of the CMB temperature and the seven-year WMAP measurement of CMB temperature and polarization (Larson et al. 2011) are used to study perturbations around the ionization history of the universe during recombination.5 The SPT and ACT data consist of observation of 790 and 296 deg2 of the sky, at 150 and 148 GHz, during 2008–2009 and 2008 seasons, respectively. For simplicity, we neglect non-CMB cosmological constraints.

The left plot in Figure 1 shows the first two ionization eigenmodes generated for the combined SPT+WMAP7 data set. The right and middle plots show the impact of these modes (with an amplitude equal to their 1σ error bars) on the CMB temperature power spectrum and on the Thomson differential visibility g(z) = neσTc(1 + z)−1e−τ, where the Thomson depth to redshift z is τ(z) = ∫neσT/Hdln(1 + z). The first two ACT+WMAP7 modes look very similar, so for definiteness, unless stated otherwise, modes refer to these and lower significance SPT+WMAP7 eigenmodes in this paper. Indeed, it turns out that the first two xe-perturbation modes for Planck+ACTPol data and for a high-resolution cosmic variance limited experiment also have similar shapes (Farhang et al. 2012), with their order reversed in some cases. We now show that the physical significance of the dominant first mode is, not surprisingly, intimately related to basic perturbative features in the differential visibility.

Figure 1.

Figure 1. First two modes (constructed for SPT+WMAP7 data), normalized to have unit norm (left), the corresponding changes in the visibility functions (middle), and the resulting differential changes in the temperature power spectrum C (right). The middle and right plots correspond to perturbations with S/N = 1. The visibility function is defined as g(z) = de−τ/dη, where η is the conformal time and τ is the optical depth to the last scattering surface. The visibility functions have been normalized to the maximum of the fiducial model's visibility, which occurs at zdec = 1088. The width (at 68% and 95% levels) of the visibility function has been marked as error bars about zdec in the figures.

Standard image High-resolution image

2.3. The Damping Physics of the Low-order Recombination Modes

Since the rank-ordered eigenmodes are direct probes of the map from xe-trajectories to C, the data-sensitive top-ranked modes should reflect the most basic C-sensitive recombination effects, namely, through the damping tail, which is intimately related to the sharply peaked differential visibility. We find that the first two SPT+WMAP7 modes confirm this: they largely describe shifts in the decoupling redshift (defined as the peak of g(z), zdec ≈ 1088) and shifts in the decoupling width, σz,dec, the "1σ" spread in g. We find that a +1σ amplitude for the first mode changes the visibility by ∼ − 1.4% in the width and by ∼0.4% in the position of the peak; for the second mode, the width increases by ∼6.2% and the peak by ∼0.8%.

The physical processes that define the structure of the damping tail have been well understood for a long time (for a review, see Bond 1996; Hu & White 1997 and references therein), and were discussed in relation to the experimental unveiling of the damping tail, first by CBI (Sievers et al. 2003) and then by ACBAR (Goldstein et al. 2003). Not surprisingly, the tail is controlled by the Compton scattering rate, neσTc, and the way it runs as the baryon density nb drops, characterized by the local power-law index, pe = 3dln(ne/nb)/dln nb =dln xe/dln(1 + z) (Bond 1996). The basic recombination quantities can be related to pe, which is zero at low and high z, has a maximum of about 12 and is about 9 at zdec for ΛCDM. The peak of the differential visibility g(z) occurs when neσT/H(z) = pe + 2, and the "Gaussian" width of decoupling in ln(1 + z)/ln(1 + zdec) is σz,dec ∼ (pe + 2)−1. Thus for ΛCDM, the Compton time is about one-twentieth of the horizon size at zdec, about one-fifth of the sound crossing time, and the relative width is about 0.09.

Earlier than decoupling, the photons and baryons are so tightly coupled by Thomson scattering that they can be treated as a single fluid with sound speed $c_{\rm s}= (1+R)^{-1/2} c/\sqrt{3}$, lowered by the extra inertia of the baryons, $R \equiv ({3 \bar{\rho }_{\rm b} / 4\bar{\rho }_\gamma })$, a photon+baryon kinematic shear viscosity (4/5)c2s(neσTc)−1 (in a full treatment of Thomson scattering including angular anisotropy and polarization effects), zero bulk viscosity, and thermal conductivity κγ = nbsγ(neσTc)−1, where sγ ∼ 109.8 is the photon entropy per baryon. In this tightly coupled regime, a WKB treatment of the perturbed photon density shows the baryon acoustic oscillations are exponentially damped, ∝exp (− ∫Γ/Hdln a), where the sound wave damping rate relative to the Hubble expansion rate is (see Section C.3.1 in Bond 1996; Kaiser 1983)

Equation (2)

Here a = (1 + z)−1. kcs/Ha multiplies the comoving wavenumber k of the acoustic oscillations by (approximately) the comoving distance sound travels in a Hubble time, cs(Ha)−1. The contribution of thermal diffusion relative to that of shear viscosity is Γdiffvisc ≈ 0.22.

To relate this to a WKB damping envelope acting on C, we replace k by ℓ/χdec, where χdec is the comoving distance from us to decoupling, and integrate up to ln adec. The damping scale obtained is

Equation (3)

The dominant first line gives the main dependences, (1 + zdec)1/2(pe + 2)(1 + R)1/2. With pe,dec ≈ 9 and zdec ≈ 1088, $c_{{\rm s,dec}}\approx 0.79 c/\sqrt{3}$ at decoupling, the first line gives ℓD ∼ 1360. The terms in the second and third line are subdominant. The first adds the thermal diffusion contribution to the viscous one, giving a ∼10% decrease; the second from the relativistic matter contribution to H gives a ∼14% rise. With these two, ℓD ∼ 1410. The fourth correction accounts for the decoupling sound speed being about 10% lower than the speed averaged over the z > zdec range; and the last 2.4% correction occurs if we use a sharp integration down to zdec, then stop. The third line terms change ℓD to 1290, but are not there if we just replace ∫Γ/Hdln a by Γ/Hσz,dec.

Of course a full transport treatment taking into account multipoles beyond the three that enter tight coupling (density, velocity, and anisotropic stress) is required to get an accurate damping rate. The phenomenology adopted by Hu & White (1997) estimated damping envelope functions multiplying "undamped" acoustic C's from numerical C-results, fitting them to a form $\exp [-(\ell /\ell _{\rm D})^{m_{\rm D}}]$, similar to the WKB approximation but with a floating slope to allow for a slower falloff reflecting complexities beyond WKB physics (such as the less severe damping associated with fuzziness of last scattering reflected in the g(z) structure cf. the stronger viscous damping; the breakout into higher temperature multipoles in the Thomson-thick to Thomson-thin transition). For ΛCDM parameters, we get ℓD ≈ 1345 and mD ≈ 1.26, in better-than-expected accord with the WKB estimate.

Apart from the residual memory of the acoustic oscillations, the rise in δln C of the first mode seen in Figure 1 (right) conforms to the $(\ell /\ell _{\rm D})^{m_{\rm D}}$ form. The fluctuations in C are dominated by those in ℓD, with less sensitivity to mD. These are related to the fluctuations in the peak position and width (which is in turn related to pe,dec) by Equation (3),

with respect to the fiducial values with subscript f. So we can interpret the first, most significant, mode as primarily due to ℓD variations. Similarly we can understand the sign change in Δg for the first mode as being in response to δzdec.

If we use the same approach for the influence of helium abundance fluctuations, the effect would be the smaller number of hydrogen nuclei near decoupling, suggesting a $\delta C_\ell /C_{{\rm f}\ell } \sim - (\ell /\ell _{\rm fD})^{m_{\rm D}} m_{\rm D} \delta {Y_{\rm p}}/(1-{Y_{\rm p}})$ form, in accord with what we see in Figure 4.

3. CONSTRAINTS FROM CIRCA 2011 ACT, SPT, AND WMAP7 DATA

To search for perturbations in the standard recombination history (Recfast; Seager et al. 1999) with recent recombination corrections included (Chluba & Thomas 2011; Ali-Haïmoud & Hirata 2011), we use the amplitude of the modes introduced in Section 2.2 as a set of new parameters, and estimate them jointly with the six primary and three nuisance parameters. The nuisance parameters follow Keisler et al. (2011) and Dunkley et al. (2011) and are the amplitudes of the Poisson and clustered power from point sources, and a template for the total thermal and kinetic Sunyaev–Zel'dovich power. The shapes of the associated C templates do not look like our xe-modes, and the data can differentiate what is nuisance from what may be standard recombination deviation, albeit with correlations that are fully taken into account in the statistics. This statement holds even though we restricted ourselves to single effective frequency analyses for SPT and ACT. The priors used for the nuisance parameters are taken from multi-band data particular to the flux cut for point-source removal. Unless stated otherwise, throughout the analysis, Yp and Neff are fixed to Yp = 0.2478 (from SPT data for the ΛCDM model and standard BBN; Keisler et al. 2011) and Neff = 3.046 (from the standard model of particle physics; Beringer et al. 2012).

For parameter estimation we use the publicly available code, CosmoMC,6 and modify it to include estimation of the mode amplitudes. We use the versions of CosmoMC adapted for SPT7 and ACT8 data set likelihood functions. We checked that when the mode amplitudes are fixed to zero, the modified CosmoMC recovers the reported SPT and ACT parameter measurements (Keisler et al. 2011; Dunkley et al. 2011). Lensing of the C's has been included throughout this work.

Table 1 summarizes the results when one and two modes are used in the analysis, for SPT+WMAP7 and ACT+WMAP7, respectively, and compares them to the standard six-parameter model. Both experiments find a non-zero (but different) amplitude for the first mode, but they are only of 1.7σ significance and so do not correspond to a detection; they have Δχ2 ∼ 2. ACT+WMAP7 also has a non-zero mean for the second mode, but only at 1.3σ. When the second mode is added to the ACT+WMAP7 analysis, the shift and uncertainty of the first mode change slightly.

Table 1. The Constraints on Cosmological Parameters with Different Sets of Parameters Used, as Measured by SPT+WMAP7 and ACT+WMAP7

Parameters SPT+WMAP7 ACT+WMAP7
6s + mode 1 + mode 2 6s + mode 1 + mode 2
100Ωbh2 2.221 ± 0.042 2.253 ± 0.046 2.249 ± 0.047 2.219 ± 0.051 2.240 ± 0.050 2.236 ± 0.053
Ωch2 0.1110 ± 0.0048 0.1123 ± 0.0049 0.1118 ± 0.0052 0.1121 ± 0.0052 0.1155 ± 0.0056 0.1121 ± 0.0061
100θs 1.041 ± 0.002 1.041 ± 0.002 1.040 ± 0.003 1.039 ± 0.002 1.039 ± 0.002 1.035 ± 0.004
τ 0.086 ± 0.015 0.089 ± 0.015 0.089 ± 0.015 0.086 ± 0.015 0.089 ± 0.015 0.0875 ± 0.015
ns 0.964 ± 0.011 0.977 ± 0.013 0.975 ± 0.016 0.963 ± 0.013 0.976 ± 0.015 0.960 ± 0.019
$10^9\Delta ^2_{\mathcal{R}}$ 2.43 ± 0.10 2.40 ± 0.10 2.40 ± 0.10 2.45 ± 0.11 2.43 ± 0.11 2.45 ± 0.11
μ1a (0) −0.77 ± 0.46 −0.76 ± 0.47 (0) −1.27 ± 0.74 −1.67 ± 0.86
μ2 (0) (0) −0.39 ± 1.09 (0) (0) −3.5 ± 2.7
σ8(derived) 0.807 ± 0.024 0.825 ± 0.027 0.818 ± 0.032 0.814 ± 0.028 0.841 ± 0.031 0.802 ± 0.040
δzdec/zdecb  ⋅⋅⋅ −0.6% −0.7%  ⋅⋅⋅ −1.0% −1.7%
δσz,decz,decc  ⋅⋅⋅ 1.5% −0.5%  ⋅⋅⋅ 2.6% −14.0%
(|δxe|/xe)maxd  ⋅⋅⋅ 5%(z ∼ 1196) 5%(z ∼ 1039)  ⋅⋅⋅ 8%(z ∼ 1006) 31%(z ∼ 1076)e
Δχ2  ⋅⋅⋅ 2.5 2.5  ⋅⋅⋅ 2.1 2.5

Notes. μ1 and μ2 refer to the amplitudes of the first and second modes. aThe mode amplitudes and errors in this table (and throughout the paper) should be interpreted with respect to the normalized version of the modes as plotted in Figure 1. So, e.g., perturbations with μ1 = 1 correspond to xe changes in the form of mode 1 and with an amplitude exactly as plotted in Figure 1. bRelative change in the redshift of maximum visibility, where zdec = 1088 is the fiducial maximum visibility point. cRelative change in the width of the visibility function. dMaximum relative change in the ionization fraction. The redshift corresponding to this maximum change is also included. eThis large deviation, though looking curious, is not statistically significant. This point is understandable given the relatively large estimated values for μ1 and μ2 and their uncertainties.

Download table as:  ASCIITypeset image

While this paper was in press, improved measurements of the damping tail were released by SPT (referred to as SPT12; Hou et al. 2012) and ACT (referred to as ACT13; Dunkley et al. 2013; Sievers et al. 2013). We present our new results using these recent datasets for the SPT11+WMAP7 modes in Table 2. We verified our SPT12 and ACT12 implementations in COSMOMC by checking that we reproduce the Hou et al. (2012) and Sievers et al. (2013) determinations for the standard six cosmological cases and also for the cases including Yp or the Neff variation. The table shows that the mode amplitudes for SPT12 are highly consistent with those of SPT11, with a slight error decrease, with a mean for the first mode amplitude now 1.9σ away from zero. As our discussion of the similarity of the first mode δCl shape and the perturbed Helium abundance shape illustrates, this is as expected from the new Hou et al. (2012) SPT measurement of Yp = 0.300 ± 0.025 (and Neff = 3.62 ± 0.48). ACT12 reported values of these parameters that were consistent with the unperturbed values of the basic-six parameter case, and as expected we see consistency with zero for the ACT12 mode amplitudes.

Table 2. The Constraints on the First Two Modes (μ1 and μ2), as Measured by SPT12+WMAP7 and ACT12+WMAP7

Parameters SPT12+WMAP7 ACT12+WMAP7
+mode 1 +mode 2 +mode 1 +mode 2
μ1 −0.73 ± 0.38 −0.68 ± 0.39 0.30 ± 0.47 −0.06 ± 0.55
μ2 (0) −1.04 ± 0.83 (0) 1.3 ± 1.1

Download table as:  ASCIITypeset image

4. THE IMPACT OF ELECTRON NUMBER CONSERVATION ON THE POSTERIOR

Our perturbed ionization history is required to satisfy electron conservation through

where xmaxe = 1 + (2Yp/3.97(1 − Yp)) is the maximum total electron fraction, using mHe/mH ≈ 3.97, obtained when helium and hydrogen are fully ionized. When the mode amplitudes are poorly determined by data, the reconstructed xe(z) could break through this bound, which of course we do not allow. Thus, although our starting prior may have been uniform with a wide possible spread in the amplitudes, the true prior distribution can only be determined with a full suite of Monte Carlo calculations restricting the allowed range. The Fisher analysis does not catch this because the amplitudes are supposed to be tiny. They are not in our case for which the data do not have strong discriminatory power so allowed variations in the ionization history can be very broad. Intuitively, if the volume spanned by the prior space is comparable to or smaller than the volume of the likelihood space (at a given significance level), the posterior will be influenced by the physical constraint, and the Fisher matrix analysis will be a poor approximation to the full analysis. Experiments with higher sensitivities will provide information about a larger number of modes before running into this condition (see simulations for Planck+ACTPol-like observations in Farhang et al. 2012).

Figure 2 shows the marginalized one-dimensional (1D) distributions of the amplitudes of the first three modes, μ1 to μ3, in an analysis with three modes included, and for two experimental setups: the posterior distributions of the SPT+WMAP7 case (solid black lines) and the prior-only simulations (dotted black lines) by ignoring the likelihood, i.e., assuming infinite errors in the data. The overplotted blue lines (in the left plot only) correspond to a case with only one mode being included in the analysis. Note that the prior distributions of the first and second modes are skewed toward negative values for the case with three modes in the analysis (black lines). However, the prior distribution for the first-mode-only case (the quite symmetric blue dashed curve) shows that the μ1 measurement is not prior-driven.

Figure 2.

Figure 2. Black lines: marginalized 1D-prior (dotted lines) and posterior probabilities (solid lines) with SPT+WMAP7 data for the first three modes, in an analysis where six standard parameters, three nuisance parameters, and the first three xe-perturbation modes were used. Blue lines, left plot: the prior (dashed line) and posterior (solid line) distributions for μ1, similar to black lines, but with only one mode included in the analysis. Note that the solid black and blue curves coincide in the left plot.

Standard image High-resolution image

The very narrow posterior distributions of μ1 and μ2 relative to their priors illustrate the power of the ACT/SPT data in constraining these parameters, although they are not found to be significantly different from zero. For μ3, on the other hand, the comparable widths of the prior and posterior distributions imply that the data set under consideration hardly contains more information about this parameter than the limits set by electron conservation. The measured errors on the fourth and higher modes differ significantly from their Fisher forecasts, not even keeping their ranking. The insensitivity of the data to higher modes explains why we have limited our study to the first two modes.

We can quantify the impact of the prior by measuring the Shannon entropy decrease in the measurement of n-parameters, q = {qi}, associated with the transition from the prior distribution pi to the posterior distribution pf when data are added,

Equation (4)

For us the relevant qi's are the amplitudes of the first n modes.

Although the full calculation can be made, we have found that a Gaussian approximation works reasonably well, and does not have the numerical challenges associated with an accurate full calculation. The posterior pf is closer to Gaussian than is the prior pi. With an nD-Gaussian distribution with zero mean and covariance matrices C, S = 1/2ln det(C) + nln(2π)/2 + 1/2〈qTC−1q〉. The last term is one-half of the mean χ2 associated with the measurement, hence is n/2 since 〈qTC−1q〉 = n. The entropy difference is then the ratio of mean log-volumes of the parameter space in question, namely,

Equation (5)

where Ci,f are the prior and posterior parameter covariance matrices. We checked the Gaussian approximation by comparing Equation (5) with estimates of the integral form Equation (4) of the entropies. The integral was calculated from the nearest neighbor entropy estimate (Singh et al. 2003). This non-parametric entropy estimation method is based on the distribution of the nearest neighbor distance of the samples, here the MCMC chain outputs, and is used for parameter spaces with more than one dimension. We found that the results from Gaussian approximation agree well with those from full integration, and are less noisy as the dimensionality of the parameter space increases. Using the determinant ratio to measure the level of improvement with improved data is familiar as a figure of merit (see, e.g., Mortonson et al. 2010). Although we have found for our application for deciding which modes to include that Equation (5) is adequate, Equation (4) is the better expression for a more accurate figure of merit (Farhang et al. 2011).

As n increases, the data add less information about the parameters relative to the prior. Thus, the difference between successive R's gradually decreases. By adding new parameters, the volumes of the posterior and prior spaces change by a similar prior-dominated factor. In the limit of very large n, R(n) → constant. This is shown in Figure 3 where we compare R(n) for different data sets and various numbers of modes included, n. For this plot, the modes of each curve are the eigenmodes constructed for the corresponding experiment. For ACT+WMAP7 and SPT+WMAP7, the difference between one and two parameters is greater than the difference between other subsequent modes. This shows that these data sets are much more informative about the second mode than the higher modes, which are entering the prior-dominated regime. This difference between the first two modes and the higher ones is also evident from Figure 2. With higher precision data sets, we expect the transition from likelihood to prior-dominance to happen at a higher mode number. This prior-likelihood dominance transition hints to a natural criterion for mode-hierarchy truncation. However, one should note that choosing a quantitative mode selection criterion can be rather subjective and not necessarily applicable to all data sets.

Figure 3.

Figure 3. Increase in the information content of the measured modes delivered by data relative to the volume of parameter space allowed by electron conservation, for different number of modes included in the analysis and various experimental cases. The modes in each case are constructed for the corresponding data set.

Standard image High-resolution image

The mode selection criterion introduced here is much stronger than the Occam's razor argument developed in Farhang et al. (2012), where the truncation of the mode hierarchy was based on the change in the information content as more ordered modes were added. Here, we have used the posterior information that the modes have in excess of the prior, a change of perspective motivated by the weak constraints from current data sets. The analysis enters a prior-dominated regime beyond just one or two modes and there is no need to consider the more sophisticated Farhang et al. (2012) criteria.

5. DISCUSSION

In this work we studied how allowing for some freedom in the recombination history gives a better fit to the damping tail of the CMB power spectrum as measured by SPT and ACT, compared to the primary six-parameter model. The red solid line in Figure 4 shows the relatively small shift between the best-fit SPT+WMAP7 C's, one with the basic six parameters fixed at their best-fit values with the xe-perturbations on, and the other with the basic six parameters fixed at their unperturbed values. That is because the non-zero μ1 is accompanied by compensation in the values of the standard parameters, most significantly shifting ns and Ωbh2 (see Figure 5) to give a small net ΔC. (Although the modes are marginalized over standard parameters, they are generally correlated with them.) When the basic six parameters are set to their xe-unperturbed values, and only μ1 varies, the red dashed line is obtained. Apart from the oscillation difference, the shape and value look rather like that for Yp variation for the reasons discussed in Section 2.3.

Figure 4.

Figure 4. Solid red line corresponds to the difference between the best-fit C's for the standard case and the case with the first mode included, measured by SPT+WMAP7. That is, the two cases have different background cosmology (as measured by data) as well as different values for μ11 = 0 and μ1 = −0.77). The dashed and dotted curves show the response to changes in Yp, Neff, and the first mode. In these cases, the six standard parameters are fixed for all models, while Yp = 0.296, Neff = 3.898, and μ1 = −0.77 have been chosen for their corresponding curves.

Standard image High-resolution image
Figure 5.

Figure 5. Marginalized 68% and 95% ns–Ωbh2 contours for various sets of parameters being included in the analysis. The 6s contours represent the standard model with six parameters. Other cases each have one extra parameter, being Yp, Neff, and μ1. Note that these extended models favor a slightly higher value of ns compared to the standard case.

Standard image High-resolution image

The goodness of fit with μ1 added to the analysis is comparable to cases where the recombination history is assumed to be perfectly known (see Chluba & Thomas 2011; Ali-Haïmoud & Hirata 2011 for the most recent calculations) and instead either Yp or Neff are allowed to vary (Keisler et al. 2011; Dunkley et al. 2011). (The corresponding best-fit C difference between these two cases with the standard six-parameter case is not shown here but is similar to the red solid line for the μ1 case.) As mentioned, including Yp has a similar impact on the standard parameter measurements to that of μ1, as illustrated in Figure 5, where ns and Ωbh2 have the most significant shifts. Neff, on the other hand, behaves differently, with a different C-shape, and it also has a large impact on Ωdmh2 and H0 (see Keisler et al. 2011; Dunkley et al. 2011).

It is also noteworthy that the measurements of σ8 (the amplitude of linear matter fluctuations at z = 0 on scales of 8h−1Mpc) for the model with Neff for ACT data (σ8 = 0.906 ± 0.059; Dunkley et al. 2011) and for SPT data (σ8 = 0.859 ± 0.043; Keisler et al. 2011) are currently slightly disfavored by σ8 inferred from clusters (σ8m/0.25)0.47 = 0.813 ± 0.013 ± 0.024; here the second set of errors is systematic and due to the uncertainty in cluster masses; Vikhlinin et al. 2009). Including μ1 and especially μ2 in the analysis, with fixed Neff = 3.046, brings σ8 toward lower values (see Table 1), consistent with external data sets.

The first principal component xe-mode, Yp, and Neff are strongly correlated due to their similar effects on the damping tail of the power spectrum, as illustrated in Figure 4. The correlation coefficients of μ1 with the other two are corr(μ1, Yp) = 0.95 and corr(μ1, Neff) = 0.72, based on results from CosmoMC for SPT+WMAP7 bandpower data. Due to this partial degeneracy, varying more than one of these parameters simultaneously would significantly increase their uncertainties. The improvement in errors forecasted for near-future data and the addition of polarization information will modestly reduce the degeneracy of μ1 with Yp and Neff. For example, for a Planck+ACTPol-like scenario the correlation coefficients for an ℓ-by-ℓ analysis from CosmoMC are forecasted to be corr(μ1, Yp) = 0.77 and corr(μ1, Neff) = 0.60. (An ℓ-by-ℓ analysis is more sensitive to acoustic oscillation phase information than data with wide bandpowers.)

The perturbative xe(z)-eigenmodes are, by definition, those that are most constrained by the CMB. The perturbations in xe(z) induced by varying Yp and Neff, are shown in Figure 6. They look very different than our most significant mode. What is interesting is the response in differential visibility. The first mode and the Yp-induced perturbations look somewhat similar, although the differences are important. In C the dominant damping tail behavior of ΔC is even closer, although the details of the peak-trough oscillations about it differ. The full Yp-induced perturbation involves a coherent sum over many eigenmodes, but the data are mostly trying to constrain its component in the first mode, as Figure 4 shows. The low-order mode amplitudes are therefore a way to efficiently transfer information from the CMB into constraints on ionization perturbations. Alternative high-z ionization history models due to specific physical effects (and priors they may impose) could then be differentiated in, e.g., the μ1 − μ2 plane.

Figure 6.

Figure 6. Relative xe and differential visibility changes due to relative infinitesimal changes in Yp and Neff (with other parameters fixed). The first mode has been added in the background (the gray dotted lines) to aid visual comparison. It has been normalized to be comparable to Yp changes.

Standard image High-resolution image

There are several alternative effects that can cause modifications to the CMB power spectrum in the damping tail. These include possible modifications to the physics of recombination, dark radiation (e.g., Archidiacono et al. 2011 and references therein), changes in the fine structure constant α (Kaplinghat et al. 1999; Battye et al. 2001; Scóccola et al. 2008), high-frequency cosmic gravitational wave background (Smith et al. 2006), dark matter annihilation, and particle decay (Chen & Kamionkowski 2004; Padmanabhan & Finkbeiner 2005; Zhang et al. 2006, 2007; Hütsi et al. 2009, 2011; Galli et al. 2009, 2011; Giesen et al. 2012). Deviations from standard recombination may be differentiated or corroborated by non-CMB measurements. Apart from invoking additional physical processes to explain damping tail measurements, it is important to note that the tail is sensitive to experimental issues, such as the instrumental beam, point sources in the maps, and detector time constants.

As shown in Table 2, we have now found that the damping tail tension in the SPT11 and ACT11 data of Table 2 persists in the full 2500 deg2 SPT data reported in Story et al. (2012) and Hou et al. (2012), but not in the three-season ACT12 data. The Planck data, with its much larger sky coverage, should be able to address whether or not tension exists. Our forecasts for the Planck 2.5-year data show that the current SPT11 best-fit amplitude of the first mode, μ1 = −0.77 (which is only 1.7σ now), could be detected at more than 10σ.

If such a deviation were found, it would be well beyond the levels of the standard recombination corrections which have been discussed extensively over the past several years (see Dubrovich & Grachev 2005; Chluba & Sunyaev 2006; Kholupenko & Ivanchik 2006; Switzer & Hirata 2008; Hirata 2008; Grin & Hirata 2010; Ali-Haïmoud & Hirata 2010; and additional references) and would have to indicate a new and possibly non-standard process at work. These physical mechanisms, with their high level of nuance and theoretical concordance, result in a modification to recombination which is 3.5 times smaller than the 1σ errors found here (Dunkley et al. 2011). In the far future, measurements of the cosmological recombination radiation from hydrogen and helium (see Sunyaev & Chluba 2009 for an overview) may provide another way to investigate this question and break some of the expected degeneracies.

If a detection of deviation is made with high significance, further explorations could be done with the same sort of analysis as that given here, but with modes weighted toward different redshift regimes. Indeed, although we have focused here on just the recombination epoch, viewing recombination, and reionization as a connected xe-trajectory has appeal, since CMB data (though at large rather than small angular scales) inform the latter. In these extended studies one needs to explore other possible variables to linearly expand in rather than δln(xe(z)), as discussed in Farhang et al. (2012). The merit of xe(z) expansions is that one can weight them to concentrate on specific recombination regions, e.g., at higher z where helium recombines, or at lower z as xe-freezeout is approached. Ultimately showing in a model-independent way that the allowed xe-trajectories do not compromise our determination of cosmological parameters would further increase our confidence in conclusions drawn from CMB data sets.

This work was accomplished by support from NSERC and the Canadian Institute for Advanced Research.

APPENDIX: FISHER ANALYSIS

The goal of this work is to search for deviations from the standard ionization scenario at high redshifts, around the epoch of recombination. For this purpose, we search for the perturbation patterns in xe best constrained by data (see Farhang et al. 2012 for more details). The most constrained perturbation parameters are the eigenvectors of the xe-perturbation block of the inverse of Fisher information matrix. The Fisher matrix for each data set under consideration is

where the q's represent any of the standard, the xe-perturbation, the secondary, or experimental nuisance parameters. The bandpowers Cb are

where the window functions, Wb's, are specific to the experiment and ${\rm Cov}_{bb^{\prime }}=\langle \delta C_b \delta C_{b^{\prime }}\rangle$ is their covariance matrix. The derivatives are calculated at the fiducial qi's. If multiple experiments are used, the total Fisher matrix is the sum of the individual Fisher matrices constructed for each experiment—if they are statistically independent—and marginalized over their nuisance, experiment-dependent parameters. To make this marginalized Fisher matrix, we divide F into blocks of nuisance parameters (represented by n) and the cosmologically interesting parameters, represented by y. We then have

where F|m is the Fisher matrix (including only cosmologically interesting parameters) marginalized over nuisance parameters. The individual F|m's should be added to get the total Fisher matrix. The xe-perturbation eigenmodes are the eigenvectors of the perturbation block of the total Fisher matrix after it has been marginalized over the standard parameters, similar to the above marginalization.

Another approach is to calculate Fisher matrices that include all nuisance parameters. Then, for those experiments that provide no constraint on a set of nuisance parameters, set those matrix entries to zero, sum the matrices over experiments and marginalize over nuisance and standard cosmic parameters.

Footnotes

Please wait… references are loading.
10.1088/0004-637X/764/2/137