Articles

A GUIDE TO DESIGNING FUTURE GROUND-BASED COSMIC MICROWAVE BACKGROUND EXPERIMENTS

, , , , , , , and

Published 2014 June 2 © 2014. The American Astronomical Society. All rights reserved.
, , Citation W. L. K. Wu et al 2014 ApJ 788 138 DOI 10.1088/0004-637X/788/2/138

0004-637X/788/2/138

ABSTRACT

In this follow-up work to the high energy physics Community Summer Study 2013 (aka Snowmass), we explore the scientific capabilities of a future Stage IV cosmic microwave background polarization experiment under various assumptions on detector count, resolution, and sky coverage. We use the Fisher matrix technique to calculate the expected uncertainties of cosmological parameters in νΛCDM that are especially relevant to the physics of fundamental interactions, including neutrino masses, effective number of relativistic species, dark energy equation of state, dark matter annihilation, and inflationary parameters. To further chart the landscape of future cosmology probes, we include forecasted results from the baryon acoustic oscillation signal as measured by Dark Energy Spectroscopic Instrument to constrain parameters that would benefit from low redshift information. We find the following best 1σ constraints: σ(Mν) = 15 meV, σ(Neff) = 0.0156, dark energy figure of merit = 303, σ(pann) = 0.00588 × 3 × 10−26 cm3 s−1 GeV−1, σ(ΩK) = 0.00074, σ(ns) = 0.00110, σ(αs) = 0.00145, and σ(r) = 0.00009. We also detail the dependencies of the parameter constraints on detector count, resolution, and sky coverage.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

In the past two decades, cosmic microwave background (CMB) experiments have made great strides in sensitivity improvement. For satellite experiments, there was a factor of 10 improvement in sensitivity from COBE to WMAP and from WMAP to Planck (Abazajian et al. 2013a). These satellites, along with numerous ground-based and balloon-borne experiments as listed on the LAMBDA site,9 collectively moved us from setting upper limits on the temperature anisotropy ($C_\ell ^{TT}$) and the E-mode and B-mode polarization ($C_\ell ^{EE}$ and $C_\ell ^{BB}$) spectra to obtaining precision measurements of the temperature power spectrum (e.g., Story et al. 2013; Planck Collaboration et al. 2013a; Das et al. 2014), shrinking error bars on the E-mode polarization spectrum (e.g., QUIET Collaboration et al. 2012; BICEP1 Collaboration et al. 2013), and detecting gravitational lensing in the B-mode polarization power spectrum (Hanson et al. 2013; The POLARBEAR Collaboration et al. 2013, 2014).

Tremendous opportunities still lie in CMB polarization. CMB polarization maps can be projected into a curl-free component (E modes) and a divergence-free component (B modes). Scalar and tensor perturbations seeded the temperature anisotropies in the CMB. Thomson scattering of the CMB quadrupole anisotropies produces linear polarizations in the outgoing photons. While E modes, like temperature anisotropies, could be sourced by both scalar and tensor perturbations, the only primordial source for B modes is from tensor perturbations, i.e., gravitational waves. Detecting this B-mode polarization signal, which peaks at the degree angular scale, would be a powerful probe of inflation.

B modes at smaller angular scales are generated by gravitational lensing of E modes by large-scale structures (LSSs). The conversion from E modes to B modes provides an exceptionally clean way of reconstructing the lensing potential ϕ and its power spectrum $C_\ell ^{\phi \phi }$ (Okamoto & Hu 2003). The latter is sensitive to the properties of the LSS, making it a probe of the late universe (z < 10; Lewis & Challinor 2006). As a result, it is central to the measurements of parameters sensitive to late universe physics like the total neutrino mass and the dark energy equation of state.

Different techniques have been developed to reconstruct the lensing potential from CMB temperature and polarization maps (Okamoto & Hu 2003; Hirata & Seljak 2003). Recently, experiments have detected gravitational lensing in CMB temperature maps with a high enough significance to measure the lensing potential power spectrum (Das et al. 2011; van Engelen et al. 2012; Planck Collaboration et al. 2013b). Moreover, the first reconstruction through the E-to-B channel has recently been demonstrated (The POLARBEAR Collaboration et al. 2013).

The lensing B-mode signal also acts as a foreground to the inflationary B-mode signal. Specifically, if the tensor-to-scalar ratio r, which parameterizes the amplitude of the tensor perturbation, is smaller than 0.02, the level of inflationary B modes would be lower than the lensing B modes (Kesden et al. 2002; Knox & Song 2002). Algorithms have been developed to remove the lensing-induced B modes (Smith et al. 2012) to study the primordial signal. Therefore, measuring the lensing potential to high precision is fundamental to advancing our understanding of both the early and the late universe.

We envision the Stage IV10 cosmic microwave background polarization experiment (CMB-S4) to be a high-resolution, high-sensitivity experiment for this very purpose. This paper explores how much CMB-S4 could help answer the pressing questions in high energy physics (HEP)—morebroadly, physics of fundamental interactions. In what follows, we outline the main HEP topics that can be addressed by CMB polarization.

Cosmic neutrino background. The relic neutrino number density for each species is 112 cm−3 (Fukugita & Yanagida 2003). This makes neutrinos the second most abundant particle in the universe, after photons. Although neutrinos are massless in the Standard Model, solar and atmospheric neutrino oscillation experiments have shown that they are massive. Assuming normal hierarchy, a lower limit on the total neutrino mass is ∼58 meV. Though minuscule in mass, their large cosmic abundance enables us to observe their cumulative gravitational effect on structure formation: neutrinos suppress structure growth below scales defined by their free streaming distances. As a result, we can measure the total neutrino mass by observing the matter power spectrum, which CMB lensing is particularly sensitive to (Kaplinghat et al. 2003). Another way to study the cosmic neutrino background is to measure the extra number of relativistic species Neff aside from photons at recombination. In the Standard Model, no other particles except neutrinos contribute as extra relativistic species. In this scenario, Neff is predicted to be 3.046. A deviation from this value hints at new physics. The CMB is uniquely sensitive to this photon–baryon–dark matter (DM) interaction at recombination. The current constraints on Neff are $3.30^{+0.54}_{-0.51}$ (at the 95% C.L.) from Planck+WMAP polarization+small-scale CMB data+baryon acoustic oscillation (BAO) measurements, and the current limit on total neutrino mass is <0.23 eV (at the 95% C.L.) from Planck data+BAO data (Planck Collaboration et al. 2013c).

Dark energy. Since the discovery of dark energy (Riess et al. 1998; Perlmutter et al. 1999), its contribution to the energy density of the universe ΩΛ has been determined to the percent level over the past 15 yr. However, its fundamental nature remains mysterious. The three broad possibilities are: (1) vacuum energy manifesting as a cosmological constant; (2) a spatially homogeneous dynamical field; or (3) modified gravity on cosmological scales. These hypotheses can be tested experimentally by measuring the dark energy equation of state w through the clustering of LSS as a function of redshift. So far, measurements are consistent with a cosmological constant. Over the next decade, photometric and spectroscopic surveys such as LSST, Euclid, and Dark Energy Spectroscopic Instrument (DESI) will take these tests to the next level of precision, opening up opportunities for new discoveries (Albrecht et al. 2012). CMB lensing is highly complementary to these surveys, because it provides high-redshift information in the linear regime, see Bartelmann & Schneider (2001). It is especially valuable for searches of early evolution of w when z > 2, which would not be possible with galaxy surveys. In addition, CMB lensing can help optical lensing surveys calibrate their shear bias and boost constraints on cosmological parameters (Pearson & Zahn 2014; Das et al. 2013; Hand et al. 2013). The current constraints of w0 and wa in the w = w0 + wa(1 − a) model are: $w_0 = -1.04^{+0.72}_{-0.69}$ and wa < 1.32 (Planck+WMAP polarization+BAO data), see Planck Collaboration et al. (2013c).

DM annihilation. The gravitational properties of DM have been well constrained by data from a plethora of measurements, e.g., weak (Refregier 2003) and strong (Tyson et al. 1998) lensing, multiwavelength studies of the Bullet Cluster (Clowe et al. 2006), distant supernovae (SN; Riess et al. 1998; Perlmutter et al. 1999), and the CMB (Planck Collaboration et al. 2013c). However, the particle nature of DM remains unknown. There are several types of experiments aiming to understand DM interactions: (1) direct detection experiments, which rely on the nuclear recoil signature of DM–Standard Model particle scattering, (2) indirect detection experiments, which are sensitive to the products of DM decay or annihilation, and (3) collider experiments. For a DM theory and detection review, see Feng (2010).

An alternative observable is the CMB two-point correlation function. When DM annihilates, heat is transferred to the photon–baryon fluid and atoms are ionized and/or excited (Chen & Kamionkowski 2004). Slatyer (2013) performed a detailed study regarding the energy deposition history into the photon–baryon fluid.

DM annihilation also leads to growing ionization fraction perturbations and amplified small-scale cosmological perturbations, leaving an imprint on the CMB bispectrum (Dvorkin et al. 2013). CMB temperature and polarization spectra can constrain the parameter pannf〈σv〉/mDM, where f is the fraction of energy deposited into the plasma, 〈σv〉 is velocity-weighted cross section, and mDM is the mass of the DM particle. Current constraints from WMAP nine-year data and Planck, Atacama Cosmology Telescope (ACT), South Pole Telescope (SPT), BAO, Hubble Space Telescope, and SN data exclude DM masses below 26 GeV at the 2σ level, assuming that all the energy is deposited in the plasma (Madhavacheril et al. 2014). We show in this paper that CMB-S4 will tighten these constraints by a factor of 10.

Inflation. A detection of degree-scale B-mode polarization would help us learn about the physics of inflation. In particular, the amplitude of tensor perturbations is directly related to the energy scale of inflation. In addition, whether r is above or below ∼0.003 determines whether the inflaton field range is sub- or super-Planckian in a broad class of inflationary models. Beyond B modes, small angular scale E-mode polarization, which is less contaminated by foregrounds than the temperature maps, can test the following predictions from slow-roll inflation: (1) the nearly scale invariant primordial power spectrum where ns is close to but not exactly 1, (2) the small running of ns, and (3) the almost flat mean spatial curvature, ΩK ≈ 10−4. The current tightest constraint for ns is 0.9603 ± 0.0073 at the 68% C.L., for dns/dlog k is $-0.014^{+0.016}_{-0.017}$ at the 95% C.L., and for ΩK is $-0.0005^{+0.0065}_{-0.0066}$ at the 95% C.L. (Planck Collaboration et al. 2013c).

We show in this paper, given the range of possible experimental configurations (survey coverage, depth, resolution), how sensitive future ground-based experiments like CMB-S4 will be in probing these areas of new physics. We deliberately extend the experimental configuration space stated in the Snowmass study for CMB-S4 so that we understand the steepness of the improvement in parameter constraints as a function of experimental setups. Furthermore, we detail the forecast methods that are lacking in the Snowmass report.

The paper is organized as follows. First, in Section 2, we describe the methodology and the fiducial cosmology. Sections 36 are assigned to the aforementioned HEP topics and parameters of interest:

  • 1.  
    Section 3: total neutrino mass and extra relativistic species—Mν and Neff;
  • 2.  
    Section 4: dark energy equation of state—w0 and wa;
  • 3.  
    Section 5: DM annihilation—pann;
  • 4.  
    Section 6: inflation—ΩK, ns, αs (running of ns), r, and nt.

For each section above, we present a brief overview of the underlying physical phenomena and their effects on the CMB power spectra, and we present and discuss the results of the forecast. For the tensor-to-scalar ratio r, we have a more thorough method section outlined in Section 6.3.1. Finally, we conclude and summarize the main results in Section 7.

2. METHODS AND ASSUMPTIONS

In this section, we describe the Fisher matrix formalism we use in this work to forecast cosmological parameter constraints, spanning over a large grid of experimental inputs. We also lay out the fiducial cosmology that we assume in this work, unless otherwise stated.

2.1. Experiment Input Grid

The exact configuration of CMB-S4 is still being studied. A few examples were given in the Snowmass CF5 Neutrinos report (Abazajian et al. 2013a). Here we extend the experiment input grid (sensitivity, beam size, sky coverage fsky) to understand how constraints for each parameter vary across the experiment design space.

The experimental sensitivity in μK arcmin is derived from the combination of the total number of detectors and the observed fraction of sky. Table 1 presents the sensitivity of an experiment given a total number of detectors Ndet and a sky coverage fsky. We consider Ndet ranging from 104 to 106 and fsky from 1% to 75%. We assume a noise-equivalent temperature (NET) of 350 μK $\sqrt{s}$ per detector (achieved by current generation experiments) and a yield (Y) of 25% that combines both the focal plane yield and the operation efficiency over an observing period of five years (ΔT). To get from NET per detector to the sensitivity s of an experiment, we use

Equation (1)

For each combination of Ndet and fsky, we cover the following beam sizes: 1', 2', 3', and 4'—and in some cases up to 6' and 8'.

Table 1. Table of Experimental Sensitivities s

Ndet/fsky   0.0125 0.025 0.05 0.125 0.25 0.5 0.75
10000   0.75 1.06 1.50 2.37 3.34 4.73 5.79
20000   0.53 0.75 1.06 1.67 2.37 3.34 4.10
50000   0.33 0.47 0.67 1.06 1.50 2.12 2.59
100000   0.24 0.47 0.47 0.75 1.06 1.50 1.83
200000   0.17 0.33 0.33 0.53 0.75 1.06 1.30
500000   0.11 0.21 0.21 0.33 0.47 0.67 0.82
1000000   0.075 0.15 0.15 0.24 0.33 0.47 0.58

Note. s (multiply $\sqrt{2}$ to get polarization sensitivity) given detector count (left column) and sky coverage (top row), in μK arcmin.

Download table as:  ASCIITypeset image

2.2. Fisher Matrix

The Fisher matrix formalism is a simple way to forecast how well future CMB experiments can constrain parameters in the cosmological parameter space of interest. By definition, the Fisher matrix assumes a Gaussian likelihood and may not represent the true likelihood distribution in some cases (Wolz et al. 2012). This approach has been scrutinized and other methods for forecast have been proposed (Perotto et al. 2006; Wolz et al. 2012). However, for a large grid of experimental inputs, it is computationally efficient to arrive at constraints through Fisher information. It is also useful for comparison with previous work, e.g., Albrecht et al. (2006) and Font-Ribera et al. (2013), and within the grid.

Assuming the likelihood function $\mathcal {L}$ of the parameters $\boldsymbol{\theta }$ given data $\boldsymbol {d}$ to be Gaussian, it can be written as

Equation (2)

where C is the covariance matrix of the modeled data. In our case, $\boldsymbol{d}= \lbrace a_{\ell m}^T, a_{\ell m}^E, a_{\ell m}^d \rbrace$ for temperature, E-mode polarization, and lensing-induced deflection. The vector $\boldsymbol{\theta }$ contains the model parameters of the cosmology described in Section 2.4.

The Fisher matrix is constructed from the curvature of the likelihood function at the fiducial values of the parameters:

Equation (3)

where $\boldsymbol{\theta _0}$ is a vector of the model parameters evaluated at their fiducial values, which are chosen to maximize the likelihood.

From Equations (2) and (3), we can write the Fisher components Fij for CMB spectra as

Equation (4)

where ℓ is the multipole of the angular power spectra and

Equation (5)

Similar to Smith et al. (2009), we do not consider a B-mode power spectrum signal in this covariance matrix since we assume that the fields are unlensed and that there are no primordial B modes. We set $C_{\ell }^{Ed}$ to 0 because its effect is small (<1% improvement on constraints in Mν for most cases). The $C_{\ell }^{dd}$ is related to the lensing power spectrum $C_{\ell }^{\phi \phi }$ by $C_{\ell }^{dd} = \ell (\ell +1)C_{\ell }^{\phi \phi }$.

The 1σ uncertainties of each parameter are obtained by marginalizing over the other parameters. For a parameter θi, the marginalized error is given by

Equation (6)

The power spectra include Gaussian noise $N^{XX^{\prime }}_\ell$ defined as

Equation (7)

where s is the total intensity instrumental noise in μK rad, defined in Equation (1), and $\theta ^{\ 2}_{\scriptsize {\rm FWHM}}$ is the full width at half maximum beam size in radians. Note that $s \rightarrow s\times \sqrt{2}$ in the case of XX' = {EE, BB}.

We calculate $C_{\ell }^{dd}$ using the iterative method outlined in Smith et al. (2012), based on Hirata & Seljak (2003) and Okamoto & Hu (2003). We estimate $N^{dd}_\ell$ from lensing reconstruction using $C^{EE}_\ell$ and $C^{BB}_\ell$ only. Figure 1 shows the $N_{\ell }^{dd}$ level for three Ndet at fsky = 0.75 and 1' beam size. We use the EB estimator because it has the lowest noise for the range of sensitivities relevant for CMB-S4 (105Ndet ≲ 106). For reference, the range of the $\ell (\ell +1) N^{dd}_\ell /2\pi$ noise floor for 105 < Ndet < 106 from the EB estimator is between 10−8 and 10−10, while Planck's $\ell (\ell +1) N^{dd}_\ell /2\pi$ noise floor is ∼2 × 10−7 from the 143 GHz and 217 GHz channels (Planck Collaboration et al. 2013b). We note that for a couple of cases with $\mathcal {O} \left(10^4 \right)$ detectors, the TT estimator gives lower reconstruction noise for sensitivity s ≳ 4 μK arcmin and the beam sizes we consider in this paper. To demonstrate the effect of using the TT estimator instead of the EB estimator, the constraint for Mν, which relies most heavily on lensing information, improves from 49 meV to 47 meV for the case with 104 detectors, 1' beam size, and fsky = 0.75. One can also combine the estimators from all reconstruction channels to get a minimum variance $N^{dd}_\ell$ to further improve the constraints if the systematics and foregrounds from the other channels are well understood. The iterative delensing method improves on the quadratic method once the instrumental noise is below the lensing B-mode level (s ∼ 3.5μK arcmin, i.e., polarization noise ∼5 μK arcmin; Seljak & Hirata 2004).

Figure 1.

Figure 1. $N_{\ell }^{dd}$ for three Ndet at fsky = 0.75 and 1' beam size. The deflection angle spectrum $C_{\ell }^{dd}$ is shown in black for reference. The $C_{\ell }^{dd}$ is related to the lensing power spectrum $C_{\ell }^{\phi \phi }$ by $C_{\ell }^{dd} = \ell (\ell +1)C_{\ell }^{\phi \phi }$.

Standard image High-resolution image

2.3. Low-redshift Information

For parameter constraints that could benefit from low-redshift information, we use the BAO signal forecasted from the DESI and put 1% priors on H0, which is achievable in the next decade.

DESI will measure the spectra of approximately 24 million objects in the redshift range 0 < z < 1.9 (we do not include the Lyα forest at z > 1.9 for simplicity). We limit ourselves to use only the DESI BAO signal, because it is the most systematically robust and time-tested measurement—constraints will be limited by statistical errors even at the high level of precision of DESI (Levi et al. 2013).

While constraints from the DESI broadband power spectrum and redshift space distortion measurements will be competitive for many of the models discussed in this work, we do not include them because modeling uncertainties related to galaxy bias make projections for them less certain and we want to focus on the effect of CMB lensing. This way, we will have strong independent measurements from both CMB lensing and optical surveys for these parameters in the coming decade.

Because the BAO signal is completely independent from CMB measurements including the CMB lensing signal, we add the two Fisher matrices,

Equation (8)

when BAO is included in the forecast. The independence holds even when the observing patch overlaps, as one cannot measure the BAO signal from CMB lensing maps. Details of DESI modeling can be found in Font-Ribera et al. (2013). On the practical level, when we use the matrices from that paper, we transformed them to the parameter space of this paper.

To add a 1% H0 prior to the Fisher matrix, we set

Equation (9)

where $F_{H_0 H_0}$ is the diagonal entry for H0 in the Fisher matrix, and H0, fid is the fiducial value of H0.

2.4. Fiducial Cosmology

The fiducial cosmology is a flat νΛCDM universe, with parameter values from Table 2 of the Planck best-fit value (Planck Collaboration et al. 2013c), i.e., Ωch2 = 0.12029, Ωbh2 = 0.022068, As = 2.215 × 10−9 at k0 = 0.05 Mpc−1, ns = 0.9624, τ = 0.0925, and H0 = 67.11 km s−1 Mpc−1. The fiducial Ωνh2 is set at 0.0009, which corresponds to Mν ≃ 85 meV.

Table 2. 1σ Constraints on Neff from CMB, in units of 10−2, for Various Combinations of Beam Sizes, fsky, and Ndet

Beams 1' 2' 3' 4'
104 Detectors        
fsky = 0.25 4.91 5.34 6.02 6.90
fsky = 0.50 4.02 4.36 4.88 5.57
fsky = 0.75 3.60 3.88 4.33 4.93
105 detectors        
fsky = 0.25 3.24 3.54 4.04 4.71
fsky = 0.50 2.57 2.81 3.19 3.72
fsky = 0.75 2.25 2.46 2.79 3.24
106 detectors        
fsky = 0.25 2.32 2.53 2.89 3.41
fsky = 0.50 1.80 1.97 2.25 2.64
fsky = 0.75 1.56 1.70 1.94 2.28

Download table as:  ASCIITypeset image

Unless otherwise stated, for all extensions of the minimal model, we vary all of the above parameters (including neutrino mass) and marginalize over them to obtain the parameter constraints. The parameters investigated in this work are:

  • 1.  
    Mν and Neff, see Section 3;
  • 2.  
    w0 and wa, see Section 4;
  • 3.  
    pann in Section 5;
  • 4.  
    ΩK, ns, αs, r, and nt, see Section 6.

The ℓmax used throughout this paper is 5000 in all the auto- and cross-spectra. Note that high-ℓ polarized radio sources are expected to contribute to the polarization power beyond ℓ ∼ 5000 (Smith et al. 2009). Upper limits on the mean-squared polarization fraction of radio sources and dusty galaxies have been placed by previous studies (Battye et al. 2011; Seiffert et al. 2007). Given knowledge of their millimeter-wave flux density distribution (Vieira et al. 2010; Marriage et al. 2011; Planck Collaboration et al. 2011), we can identify and mask the extragalactic polarized sources with a small data loss. For polarized galactic foregrounds, we can avoid them by scanning patches that are outside of the Milky Way.

Foregrounds in the temperature power spectrum, however, start to dominate over the signal around ℓ = 3000 (Reichardt et al. 2012; Dunkley et al. 2013; Crawford et al. 2014; Planck Collaboration et al. 2013d). The different sources of high-ℓ foregrounds in the temperature are thermal and kinetic Sunyaev–Zel'dovich effects, radio galaxies, and cosmic infrared background. Progress is being made toward an understanding of these foregrounds, as many studies are currently underway (Reichardt et al. 2012; Dunkley et al. 2013; Crawford et al. 2014; Planck Collaboration et al. 2013d). For parameters that rely heavily on high-ℓ information, we make a conservative estimate on parameter constraints by imposing an ℓ = 3000 cut in the temperature spectrum.

3. COSMIC NEUTRINO BACKGROUND

3.1. Overview

The standard hot big bang model predicts a relic sea of neutrinos—the cosmic neutrino background, whose density is the second highest of all species. Neutrinos decoupled from the primordial plasma at a temperature of T ∼ 1 MeV (Lesgourgues & Pastor 2006) but maintained temperature equilibrium with photons as the universe expanded. Once the temperature of the photons dropped below the mass of the electrons, electron–positron annihilation transferred heat to the photons. Using entropy conservation, and assuming neutrinos had completely decoupled from the plasma by that time, we can relate the temperatures of neutrinos and photons by

Equation (10)

where the factor of 4/11 comes from the effective degrees of freedom of positrons, electrons, and photons before and after electron–positron annihilation is complete (Steigman 2002).

To account for the radiation-like behavior that neutrinos have in the primordial plasma, it is conventional to parameterize the relativistic energy density ρR as a function of photon energy density ργ and Neff, the effective number of relativistic species, as

Equation (11)

where the factor of 7/8 accounts for the fermionic degrees of freedom of neutrinos. Given three neutrinos in the Standard Model of particle physics, one would naively expect Neff = 3. However, QED corrections in the primordial plasma and spectral distortion due to electron–positron annihilation raise the value of Neff to 3.046 (Mangano et al. 2005). Because of the generic parameterization, Neff encapsulates radiation-like behavior from any other relativistic species as well. For example, sterile neutrinos, a candidate for nonstandard radiation content, could change Neff on the basis of their masses and mixing angles with active neutrinos (Fuller et al. 2011). Therefore, tight constraints on Neff from the CMB and LSS, along with big bang nucleosynthesis (BBN), would be essential to either confirm the Standard Cosmological Model or discover new physics relating to extra relativistic species.

Total neutrino mass, Mν, contributes to the critical density of the universe as (Mangano et al. 2005)

Equation (12)

Main Effects of Mν and Neff on CMB Spectra:

  • 1.  
    Suppression on small angular scale lensing power spectrum for nonzero Mν. Neutrinos' large thermal velocities allow them to free stream on scales smaller than ∼(Tν/mν) × (1/H) (Abazajian et al. 2011), where mν is the mass of an individual neutrino and H is the Hubble expansion rate, instead of falling into gravitational wells. As a result, structure formation below this scale is suppressed (Kaplinghat et al. 2003). Figure 2 shows the effect of neutrino free streaming on the lensing potential power spectrum $C_\ell ^{\phi \phi }$ for various total neutrino masses.
  • 2.  
    Increasing Neff increases Silk damping. The main effects of Neff on the CMB temperature and E-mode polarization power spectra is increased Silk damping at small scales (Hou et al. 2013).
Figure 2.

Figure 2. Ratio of the lensing potential power spectrum for different total neutrino masses to a massless neutrino case: the heavier the neutrinos, the more suppressed the potential is.

Standard image High-resolution image

The current constraint on Neff is $3.30^{+0.54}_{-0.51}$ (95% C.L.) from Planck+WMAP polarization+small-scale CMB+BAO data, and the current limit on total neutrino mass is Mν < 0.23 eV (95% C.L.) from Planck+BAO data (Planck Collaboration et al. 2013c). From solar and atmospheric neutrino oscillation experiments, we know the mass differences between each mass eigenstate satisfy

Equation (13)

From these equalities, the lower bound for the sum of neutrino masses is ∼58 meV for the normal hierarchy, where the smallest mass is set to zero, and ≳ 100 meV for the inverted hierarchy, in which case m3 is set to have the smallest mass (Lesgourgues & Pastor 2006).

The next qualitatively interesting result for neutrino mass through cosmological probes will be to detect the total neutrino mass with significance and potentially distinguish normal from inverted hierarchies if the smallest mass is <100 meV. In addition, a precise measurement of Neff would determine whether the three active neutrinos scenario accurately describes the thermal content of the early universe. An order of magnitude improvement from current constraints is likely to be needed to achieve this goal.

Experiments like Katrin11 directly probe the effective electron neutrino mass. Coupled with neutrino mixing angle measurements like Noνa,12 T2K,13 or Double Chooz,14 one can pin down the masses of each neutrino species. Agreement will confirm our understanding of cosmology and particle physics, and disagreement will initiate search for new physics.

3.2. Results and Discussion

3.2.1. Neff Forecast

Table 2 presents a sample of Neff constraints given different beam sizes, sky coverage, and detector counts in an experiment. The best constraint we get in this grid is σ(Neff) = 0.016 from the 106 detectors, 1' beam, fsky = 0.75 case (equivalent sensitivity is 0.58 μK arcmin), which distinguishes 3.046 from 3 at the 3σ level. We observe that increasing Ndet from 104 to 105 improves constraints in Neff by about 35%, while increasing Ndet from 105 to 106 improves constraints in Neff by about 30%. Increasing fsky also improves the constraints even though the sensitivity of the experiment decreases. The constraints improve by about 10% for each arcminute decrease in beam size. Out of all the parameters studied in this work, Neff improves most with decreasing beam size. This is partly due to the high constraining power on Neff from high ℓ multipoles in the $C_\ell ^{TT}$ and $C_\ell ^{EE}$ spectra, which are sensitive to effects from Silk damping. Figure 3 shows the trend of Neff constraints for increasing fsky, Ndet, and beam sizes.

Figure 3.

Figure 3. Constraints for σ(Neff) as a function of the number of detectors and sky fraction for 1'–4' beam sizes.

Standard image High-resolution image

We know that for multipoles ℓ > 3000, the temperature power spectrum is also contaminated by foregrounds. For this reason, we run the forecast up to ℓ = 3000 in $C_\ell ^{TT}$, while keeping the other spectra the same. In this case, we get the Neff 1σ constraint at 0.021 with 106 detectors, fsky = 0.75, and 1' beam size. The set of experiments that could constrain Neff to better than 0.025 shrinks to {Ndet, beam, fsky} = {106, 0.75, 1'}, {5 × 105, 0.75, 1'}, {2 × 105, 0.75, 1'}, {106, 0.5, 1'}, {106, 0.75, 2'}, and {5 × 105, 0.75, 2'}.

Relationship with relic helium abundance YP. In our analysis, we set the relic helium abundance YP = 0.248, assuming that its value is measured externally. However, if we assume $N_{\rm eff}^{\rm BBN}$ is the same as $N_{\rm eff}^{\rm CMB}$, we can improve the constraints of Neff at recombination by varying the value of YP self-consistently for each $N_{\rm eff}^{\rm CMB}$ (Nν in the expression below) and Ωbh2 using the relation (Simha & Steigman 2008)

Equation (14)

where η10 = 273.9 ΩBh2, S = (1 + 7ΔNν/43)1/2, and ΔNν = Nν − 3. With the imposed relation, we improve the constraints over the entire parameter space by 15%–20%. In particular, the best constraint for Neff is 0.013, in the range of experimental configurations we consider in this work.

On the other hand, if $N_{\rm eff}^{\rm BBN}$ is different from $N_{\rm eff}^{\rm CMB}$ because of new physics, we can use CMB to constrain Neff and YP independently. In this case, the constraints on Neff degrade from those listed in Table 2 by about a factor of two, and the constraints on YP is on the order of 10−3.

3.2.2. Mν Forecast

Table 3 presents a sample of Mν constraints given different beam sizes, sky coverage, and detector counts in an experiment. The best constraint we get for CMB only is Mν = 34.1 meV from the 106 detector (0.58 μK arcmin equivalent), fsky = 0.75, 1' beam case. This constraint improves to 15.1 meV when BAO is added. The top row of Figure 4 shows the trend of how neutrino constraints vary as a function of detector count, sky fraction, and beam size.

Figure 4.

Figure 4. 1σ constraints on the total neutrino mass for various detector numbers and observed sky fraction in units of meV. The top two panels show constraints from CMB for 1'–4' beams. The bottom panels show constraints from CMB+BAO with the same beams in the CMB experiments. "CMB" includes lensing. We see that data from BAO push the constraints from CMB with lensing to a lower floor and a wide range of experimental configurations can obtain sub-20 meV constraints on the sum of neutrino masses.

Standard image High-resolution image

Table 3. 1σ Constraints on Mν, in units of meV, from CMB and from CMB+BAO

  CMB CMB+BAO
1' 2' 3' 4' 1' 2' 3' 4'
104 Detectors                
fsky = 0.25 71.7 72.8 74.4 76.6 22.8 23.0 23.4 23.9
fsky = 0.50 54.7 55.7 57.2 59.2 20.6 20.9 21.3 21.9
fsky = 0.75 48.1 49.0 50.5 52.5 20.1 20.4 20.9 21.5
105 detectors                
fsky = 0.25 63.3 65.2 66.7 68.3 19.7 19.8 19.9 20.1
fsky = 0.50 46.4 47.2 48.2 49.4 16.9 17.0 17.1 17.2
fsky = 0.75 38.5 39.2 40.0 41.0 15.7 15.8 15.9 16.0
106 detectors                
fsky = 0.25 54.9 58.1 62.2 64.7 19.1 19.2 19.3 19.4
fsky = 0.50 40.8 42.7 45.1 46.5 16.4 16.4 16.5 16.6
fsky = 0.75 34.1 35.7 37.2 38.3 15.1 15.2 15.3 15.3

Note. "CMB" includes lensing.

Download table as:  ASCIITypeset image

For the levels of sensitivity achieved with Ndet ⩾ 105, constraints on Mν are sample variance limited as opposed to sensitivity limited. We see from Figure 4 that as we increase detector counts, the constraints on Mν plateau. For CMB alone, increasing Ndet from 104 to 105 improves constraints for fsky = 0.5 cases by about 15%, while increasing Ndet from 105 to 106 improves constraints for the same cases by 6%–12% (smaller beam sizes give more improvements). In contrast, with CMB and BAO increasing Ndet from 104 to 105 improves constraints for fsky = 0.5 cases by about 20%, while increasing Ndet from 105 to 106 improves constraints by only 3%. This shows that increasing Ndet beyond 105 does not improve the constraints on Mν much once BAO is included. Decreasing beam sizes help the CMB-only cases more when Ndet is high, e.g., in the 106 Ndet case, the improvement is almost 10% for each decreased arcminute. However, this effect is washed away when BAO is added. For the lower Ndet regime, the improvement over decreasing beam size is a few percent.

CMB lensing significantly improves the constraints on Mν obtained using primary spectra alone. Figure 5 compares the constraints on Mν with and without using CMB lensing from an experiment with 105 detectors and 4' beam. This is because the lensing spectrum breaks the degeneracy that neutrino mass has with Ωch2 in the primary spectra. Adding BAO information further breaks degeneracies as seen in the right column of Table 3. The bottom row of Figure 4 shows that a wide range of experimental configurations can lead to tight neutrino mass constraints. Specifically, constraints go below 25 meV for fsky ⩾ 0.125 and for any number of detectors and beam sizes. These constraints are similar to those estimated by Font-Ribera et al. (2013) for Planck CMB+redshift-space distortions+DESI BAO and/or lensing measurements based on galaxies, e.g., DES/LSST.

Figure 5.

Figure 5. Constraints on total neutrino mass across all sky fractions, comparing CMB without lensing to CMB with lensing for experiments having 105 detectors and 1' or 4' beams. The constraints on Mν are greatly improved by CMB lensing. This general trend is observed across all experimental configurations when the lensing power spectrum is added.

Standard image High-resolution image

4. DARK ENERGY EQUATION OF STATE

4.1. Overview

The observed acceleration in the expansion of the universe is currently explained phenomenologically by dark energy. In this section, we focus on the constraints for a homogeneous dark energy with the equation of state parameterized as

Equation (15)

allowing for a time-varying dark energy field. In this parameterization, a cosmological constant would be the same as dark energy in its functional form when the equation of state w(a) is constant and set to −1, i.e., w0 = −1 and wa = 0. Two questions ensue: (1) how well could future experiments differentiate a cosmological constant from dark energy? and (2) how does dark energy evolve, if it does? This later question is the first step to capturing the much broader class of possible dark energy models. Once we precisely measure the equation of state, we will have a much better sense of directions in the theoretical modeling.

The effects of dark energy on cosmological observations are encapsulated in the Hubble parameter $H \equiv \dot{a}/{a}$, where a = a(t) is the scale factor. The Friedmann equation relates the expansion rate of the universe with the energy densities of the contents in it:

Equation (16)

where Ωr, ΩM, ΩDE, and ΩK are the density of radiation, matter, dark energy, and curvature, respectively, and w(a) is the specific parameterization of the equation of state for dark energy.

The Hubble expansion rate H also governs the growth of linear perturbations, δ, in structure formation:

Equation (17)

Primordial CMB fluctuations constrain dark energy parameters through the effects H have on them at recombination, i.e., z ≃ 1100—acoustic peaks positions are sensitive to the value of w (Huterer & Turner 2001). If w < −1, both $C_\ell ^{TT}$ and $C_\ell ^{EE}$ would shift toward higher ℓ, and if w > −1 they would shift toward lower ℓ. The wa parameter gives the same effects as w0 but at a smaller amplitude. One gets the same qualitative feature when H0 (value of H(z) at z = 0) is decreased or increased, respectively. Therefore, when we only use information from the CMB (z ≃ 1100), dark energy equation of state parameters are degenerate with geometrical parameters, like H0 and ΩK.

To break this degeneracy, low redshift probes like weak lensing (CMB and optical), BAO, and H0 measurements are essential. Here we focus on CMB lensing in relation to dark energy. CMB lensing is sensitive to structures at various redshifts (the lenses) and the CMB (the source), see Hanson et al. (2010). Because they are both a function of H(z), we can observe the effects of dark energy through them. The effect of the equation of state w on the lensing power spectrum is an overall enhancement or suppression in power across all scales with a minor scale dependence. Because CMB lensing probes redshifts higher than optical surveys, it is particularly useful for studying structure formation beyond z > 2—a direct observatory of the early history of dark energy. Specifically, if effects of a nonstandard dark energy manifest at high z (e.g., z ∼ 5), then structure formation could be significantly suppressed, which CMB lensing can uniquely constrain.

4.2. Results and Discussion

As discussed in the previous section, to constrain the dark energy equation of state, it is essential to have observations from various redshifts and multiple probes. In this section, we present the constraints on w0, wa, and the figure of merit (FoM) as defined by the Dark Energy Task Force (DETF; Albrecht et al. 2006), from the CMB, with and without H0 prior and with and without BAO measurements from DESI.

In addition to constraints on w0 and wa, and the DETF FoM, we quote a constraint on wp at ap, where ap is the value of the scale factor a when the uncertainties of w(a) are minimized (Albrecht et al. 2006). To be explicit, FoM = 1/(σ(wa)σ(wp)). Table 10 in the Appendix lists the constraints from CMB, CMB+BAO, and CMB+BAO+1% H0 prior. CMB includes CMB lensing. The best FoM is 303. As defined earlier, this Fisher matrix parameter space includes massive neutrinos. For reference to the original DETF FoM, we include the constraints with fixed sum of neutrino mass in Table 11 in the Appendix. In this scenario, the best FoM is 576.

In contrast to the rest of the models investigated in this work, the DETF FoM is sensitivity limited. For example, for CMB+BAO+H0 prior 1' beam cases, increasing Ndet from 104 to 105 improves the FoM by 10%–19%, while increasing Ndet from 105 to 106 improves the FoM by 21%–29% (bigger fsky gives more improvement). For CMB alone and a 1' beam, increasing Ndet from 104 to 105 and from 105 to 106 improves the FoM by more than double, respectively. Improvements from decreasing beam sizes become more noticeable at high Ndet, and the relative improvement per arcminute decreases as the beam size decreases. The improvement ranges from a few percent to tens of percent. While fsky does not drive the improvement of the FoM as much, the FoM does improve for increasing fsky for all cases in the table though the sensitivity of the experiment decreases.

Figure 6 shows the trend of how the FoM improves with increasing detector counts and sky fraction. We note that BAO and H0 prior improve the FoM from using CMB alone from a factor of several to an order of magnitude. Also, with BAO and H0 prior, increasing the detector count does not improve the FoM much at fsky smaller than 0.2.

Figure 6.

Figure 6. Dark Energy Task Force (DETF) figure of merit (FoM; neutrino mass included in Fisher space) as a function of detector number counts and sky fraction for 1'–4' beam sizes. The top row shows FoM from CMB. The middle row shows FoM from CMB+BAO. The bottom row shows FoM from CMB+BAO+1% H0 prior. "CMB" includes lensing. The obvious trend is that the more detectors and sky coverage, and the smaller the beam, the better the DETF FoM is. However, the case of a CMB+BAO+H0 prior combination shows that different beams have very close performances.

Standard image High-resolution image

Figure 7 illustrates how much a CMB-S4 experiment improves on the constraints on w0 and wa over Planck including DESI BAO and marginalizing over neutrino masses. The FoM from Planck (no lensing)+DESI BAO is 22 and from CMB-S4+DESI BAO is 141. In this figure, the assumed CMB-S4 configuration has 500,000 detectors, 3' beam, and covers half of the sky. The improvement comes mainly from the tight constraints one gets for neutrino masses from CMB-S4 and DESI BAO.

Figure 7.

Figure 7. Error ellipses for w0 and wa comparing CMB-S4 and Planck given BAO information from DESI and marginalizing over Mν. The FoM from Planck (no lensing)+DESI BAO is 22 and from CMB-S4+DESI BAO is 141. This CMB-S4 experiment has 500,000 detectors, 3' beam, and covers half of the sky.

Standard image High-resolution image

To further constraint dark energy parameters, adding extra probes like SNe and optical weak lensing, as suggested by the DETF report, will certainly help.

5. DARK MATTER ANNIHILATION

5.1. Overview

If DM is a weakly interacting massive particle and it is a thermal relic, then its self-annihilation cross section can be determined by its relic density today through the Boltzmann equation (Scherrer & Turner 1986; Gondolo & Gelmini 1991; Steigman et al. 2012).

Depending on the model, DM particles can annihilate into gauge bosons, charged leptons, neutrinos, hadrons, or more exotic states. These annihilation products then decay or interact with the photon–baryon fluid and produce electrons, positrons, protons, photons, and neutrinos. Neutrinos do not further interact with the photon–baryon fluid, but they interact gravitationally and their effects can be observed through the lensing of the CMB. A proton's energy deposition to the fluid is inefficient because of their high penetration length. Hence, the main channels for energy injection are through electrons, positrons, and photons. At high energies, positrons lose energy through the same mechanisms as electrons. High energy electrons lose energy through inverse Compton scattering of the CMB photons, while low energy electrons lose energy through collisional heating, excitation, and ionization. Photons lose energy through photoionization, Compton scattering, pair-production off nuclei and atoms, photon–photon scattering, and pair-production through CMB photons (Zdziarski & Svensson 1989). The rate of energy release per unit volume by a self-annihilating DM particle is given by (Galli et al. 2009)

Equation (18)

where ρc is the critical density of the universe today, ΩDM is the DM density, f is the energy deposition efficiency factor, 〈σv〉 is the velocity-weighted annihilation cross section, and mDM is the mass of the DM particle, assumed to be a Majorana particle in this work.

Because of these processes, the photon–baryon plasma is heated and the ionization fraction is modified. This leads to modifications of the recombination history (Padmanabhan & Finkbeiner 2005; Galli et al. 2009) and, consequently, of the CMB spectra. For details on the energy injection processes, see Slatyer et al. (2009). The energy injection due to DM annihilation broadens the surface of last scattering but does not slow recombination (Padmanabhan & Finkbeiner 2005). The extra scattering of photons at redshift z ≲ 1000 damps power in the CMB temperature and polarization fluctuations at small angular scales (ℓ ≳ 100) and adds power in the E-mode polarization signal at large scales (ℓ ≲ 100). The "screening" effect at ℓ ≳ 100 goes as an exponential suppression factor Ce−2ΔτC, where Δτ is the excess optical depth due to DM annihilation. This exponential factor is partially degenerate with the amplitude of the scalar perturbation power spectrum As, and polarization data help to break this partial degeneracy.

The CMB is sensitive to the parameter pann, defined as

Equation (19)

In this work, we take a thermal s-wave annihilation cross section 〈σv〉 = 3 × 10−26 cm3 s−1 and f = 1, and we will present the 95% C.L. upper limit of the DM particle mass, mDM, that a CMB Stage IV experiment could reach. We note that 〈σv〉 could vary for different ranges of mDM (Steigman et al. 2012) and that the value of f is a function of redshift and the interaction of annihilation products (Slatyer 2013). The reader can easily scale our results for any other models considered. If, instead, DM has a p-wave annihilation cross section, with a v2 dependence, we do not expect any relevant bound from the CMB, since velocities at DM freeze out and recombination can be orders of magnitude different.

In particular, we expect most of the constraining power of CMB-S4 to come from its polarization spectrum $C_\ell ^{EE}$ and the cross spectrum $C_\ell ^{TE}$. Figure 8 illustrates the percentage deviation from the fiducial cosmology in $C_\ell ^{EE}$ at the 95% C.L. value of pann that Planck and CMB-S4 would be able to differentiate, respectively. The CMB-S4 configuration chosen for this figure has 106 detectors, fsky = 0.75, and 1' beam size.

Figure 8.

Figure 8. Percentage relative deviation of $C_\ell ^{EE}$ from fiducial cosmology when DM annihilation is taken into account. The Planck and CMB-S4 lines illustrate the power of each experiment to differentiate a model with DM annihilation from the fiducial cosmology at the 95% C.L.

Standard image High-resolution image

5.2. Results and Discussion

In the range of sensitivities and beam sizes considered, we found that the main factor that improves the limit in mDM is sky coverage fsky. This means that the constraints are largely sample variance limited.

Figure 9 illustrates how steep the dependence is on fsky (it goes as ${\sim} f_{{\rm sky}}^{1/2}$) and the mild dependence on detector number and beam size. It also shows how much more many of the possible CMB-S4 configurations improve over limits from Planck.

Figure 9.

Figure 9. Ninety five percent C.L. upper limit for mDM in GeV as a function of fsky for a few experimental configurations. The blue (black) lines correspond to 106 (104) detectors, and the solid (dash) lines correspond to 1' (4') beams. The dash-dotted line shows the limit expected from Planck. While it matters more to have a small beam when there are a lot of detectors, the dependence is small compared with the dependence on the sky fraction. Therefore, to have a tighter constraint on mDM, we need as much sky coverage as possible.

Standard image High-resolution image

At perfect energy deposition f = 1, Planck could exclude mDM < 114 GeV at 95% C.L. for 〈σv〉 = 3 × 10−26 cm3 s−1, when including polarization and lensing information. Any CMB-S4 configuration with beams smaller than 4', minimum of 104 detector, covering half the sky could exclude mDM < 190 GeV at the 95% C.L. or better. In particular, the best exclusion limit considered here is 255 GeV with 106 detectors, 1' beam size, and fsky = 0.75. It remains exciting to see significant improvement on mDM limits from Planck.

6. INFLATION

6.1. Overview

The theory of cosmic inflation was proposed to solve the missing monopole problem, the flatness problem, and the horizon problem from the standard big bang theory (Guth 1981). Inflation invokes a period of rapid expansion of the universe before the standard big bang expansion phase. Because of the rapid expansion, any relics would be diluted to a point where they would be extremely rare in our observable universe. Similarly, the curvature of the universe could be diluted as well. For the horizon problem, modes that would not have been in causal contact in the standard big bang theory were in fact once in causal contact in the inflationary framework before their horizon exit (Dodelson 2003). Outside the horizon, their amplitudes were frozen. Therefore, after they reentered the horizon, modes that would not have been in causal contact if not for inflation appear to have equilibrated with each other. Not only does inflation solve these problems, it sets the initial conditions for LSS, thus providing a physical foundation for the observed fluctuations in the CMB and LSS. During inflation, quantum fluctuations were stretched and became classical perturbations. Scalar perturbation of the metric seeded the formation of LSS. Tensor perturbation of the metric created primordial gravitational waves. The observables caused by these perturbations provide definite signatures to confirm or falsify the theory of inflation. To study them, we parameterize the perturbation spectra using a power law in wavenumber k, as follows.

The scalar spectrum is

Equation (20)

where As is the amplitude of the scalar spectrum, ns is the scalar spectral index, k* is the pivot scale, and αs is the running of ns. The tensor spectrum can be written as

Equation (21)

where At is the amplitude of the tensor spectrum, and nt is the tensor spectral index.

We can measure As, ns, αs, At, and nt using the CMB. One of the most sought-after parameters, the tensor-to-scalar ratio rPt/Ps, tells us the energy scale of inflation. Specifically,

Equation (22)

Therefore, a value of r larger than 0.01 would confirm that inflation happened at an energy scale comparable to that at the Grand Unified Theory, at which the strong, weak, and electromagnetic forces are unified. The CMB provides a unique window to the HEP that cannot be attained in terrestrial accelerator experiments.

The value of r is also directly related to how big the field excursion Δϕ is from when fluctuations seen in the CMB were created to the end of inflation. Given the number of e-folds Ne, in single-field slow-roll and single-field Dirac-Born-Infeld inflation (Silverstein & Tong 2004; Lyth 1997), we can write (Lyth 1997)

Equation (23)

where Mpl is the Planck mass, r = Pt/Ps, $P_t \propto H^2/M_{{\rm pl}}^2$, and $P_s \propto H^4/\dot{\phi }^2$. For Ne ∼ 50, a Planck mass field range corresponds to r ∼ 0.003.15 (Similarly, for Ne ∼ 60, Δϕ = Mpl implies r ∼ 0.002.) For super-Planckian field excursion, Δϕ > Mpl implies that the inflation field is sensitive to an infinite series of operators of arbitrary dimensions (Baumann et al. 2009). In order for a large-field inflation model to be UV complete, effective field theory requires a shift symmetry in the field, which protects the flatness of the potential over a large field range. Differentiating between sub-Planckian and super-Planckian field excursion during inflation would rule out different classes of inflationary models.

Here we use the simplest class of models—single-field slow-roll inflation—to illustrate how observations can falsify inflationary models. The conditions for slow-roll require the inflationary potential to be flat enough such that the slow-roll parameters epsilon and η are small. They are defined as follows:

Equation (24)

Equation (25)

where V is the inflaton potential, and the subscript, ϕ denotes the partial derivative with respect to ϕ.

We can write ns, αs, nt, and r, to leading order, in terms of the slow-roll parameters:

Equation (26)

Equation (27)

Equation (28)

Equation (29)

where $\xi ^2 = M^4_{{\rm pl}} V_{,\phi } V_{,\phi \phi \phi }/V^2$.

Therefore, for each specific single-field slow-roll model, there exists a unique set of predictions on each of these parameters. This allows us to rule out models by measuring the values of the parameters.

Inflation predicts an almost but not exactly flat universe, where |ΩK| is on the order of 10−5 (Planck Collaboration et al. 2013e; Abazajian et al. 2013b). This small spatial curvature is a consequence of the curvature fluctuations of the large-scale modes at the horizon. A statistically significant deviation from the expectation of inflation for the curvature would give us information on the process of inflation. For example, if a significant departure from flatness is measured, it can mean that inflation was not slow-rolling when perturbations of the scales just larger than our observable universe exited the inflationary horizon.

The current best constraints for these parameters are from the first release of Planck data (Planck Collaboration et al. 2013e). It is measured that ns = 0.9603 ± 0.0073 (and ns = 0.9629 ± 0.0057 when combined with BAO), αs = −0.0134 ± 0.0090, and $\Omega _K = -0.058^{+0.046}_{-0.026}$ with Planck and WMAP polarization data (and ΩK = −0.0004 ± 0.00036 when combined with BAO). An upper bound for r is set at r < 0.11 at 95% C.L.

To get to the next qualitatively significant level of constraints for these parameters or confirm null results, we need to at least constrain r to 0.0006 at 1σ. This will confirm a discovery for large-field inflation in the case where r ≳ 0.003 if Ne ∼ 50. Alternatively, a null result at this level will rule out large field inflation. For curvature ΩK, it would be very interesting to get 1σ constraints at the level of 10−5. While ns is constrained to a level where we confirm inflation predictions, more precise constraints will be needed to rule out models. As for αs, most inflationary models predict it to be undetectable, so any detection would be interesting.

6.2. Results: ΩK, ns, αs

In this section, we present the 1σ constraints of ΩK, ns, and αs that are close to getting to evidence (3σ) and/or discovery (5σ) regimes for the grid of experimental setups. The treatment for constraining the tensor-to-scalar ratio r is different from that for the other three parameters, so we devote a separate section for r.

6.2.1. ΩK Forecast

Table 4 lists the constraints for ΩK using CMB combined with BAO measurements and 1% H0 prior. CMB experimental inputs listed have 104, 105, and 106 detectors, 1'–4' beams, and a fixed sky fraction of 0.75.

Table 4. ΩK 1σ Constraints in Units of 10−3 for Five Combinations of CMB with DESI BAO and 1% H0 Prior at fsky = 0.75

  1' 2' 3' 4'
104 Detectors        
CMB (no lens) 8.03 8.42 9.02 9.77
CMB 5.29 5.41 5.55 5.70
CMB+H0 1.30 1.31 1.33 1.36
CMB+BAO 0.94 0.96 0.98 1.01
CMB+BAO+H0 0.93 0.95 0.97 1.00
105 detectors        
CMB (no lens) 6.26 6.61 7.09 7.71
CMB 4.24 4.31 4.39 4.52
CMB+H0 1.21 1.23 1.24 1.26
CMB+BAO 0.84 0.86 0.88 0.90
CMB+BAO+H0 0.84 0.85 0.87 0.89
106 detectors        
CMB (no lens) 4.89 5.38 5.96 6.47
CMB 3.37 3.65 3.84 4.03
CMB+H0 1.13 1.16 1.19 1.22
CMB+BAO 0.75 0.79 0.83 0.85
CMB+BAO+H0 0.74 0.78 0.82 0.84

Download table as:  ASCIITypeset image

Constraints on ΩK are sample variance limited in the sensitivity range we consider for CMB-S4 in this study. Increasing Ndet by an order of magnitude in the CMB only improves ΩK constraints by about 20%, while adding BAO and/or H0 improves the constraints by factor of about five. Decreasing beam sizes improve the constraints at the percent level.

The effects of the curvature density ΩK on the CMB temperature and E-mode polarization power spectra are degenerate with those from H0 (Efstathiou & Bond 1999; Hu et al. 2006), because a larger correlation angle between hot spots could be due to both the curvature of space and the surface of last scattering being closer to us. Measuring ΩK at multiple redshift slices could break this degeneracy. Therefore, we will use multiple probes—CMB lensing, BAO, and H0 priors—for constraining ΩK in this section.

CMB lensing provides an overall factor of ∼2 improvement from CMB (TT, EE, TE) for the smaller beam cases and about a factor of three improvement for the 8' beam. When BAO information is added to CMB lensing, there is a factor of 5–10 improvement across the whole grid. This is because the BAO signal is orthogonal to the CMB in ΩK and H0 space, as illustrated in Figure 10. We expect the Hubble parameter to be constrained to about or better than 1% in the coming decades. The ΩK constraints improve by about a factor of three to five from CMB lensing given the 1% H0 prior. Figure 11 shows the trends of the constraints in ΩK for CMB, CMB+1% H0 prior, and CMB+BAO, respectively (CMB includes lensing).

Figure 10.

Figure 10. 1σ (dash line) and 2σ (solid line) constraint ellipses of ΩK vs. H0. This plot shows how having the BAO handle breaks degeneracies between the two parameters.

Standard image High-resolution image
Figure 11.

Figure 11. 1σ constraints of ΩK× 10−3 in the plane of beam sizes and detector counts for fixed fsky = 0.75. We show constraints from CMB, CMB+1% H0 prior, and CMB+BAO from top to bottom panel. "CMB" includes lensing. BAO provides about an order of magnitude improvement in the whole space by breaking degeneracies, better than adding a 1% H0 prior to CMB data.

Standard image High-resolution image

From Table 4, we note that the 1σ constraints on ΩK is 0.00075 from CMB+DESI BAO, which is very close to the regime where we want to be able to constrain ΩK. To further shed light on the mean spatial curvature of our universe, 21 cm mapping (Mao et al. 2008), QSOs, and the Lyα forest will also be useful (Font-Ribera et al. 2013).

6.2.2. The ns, αs Forecast

Table 5 summarizes the constraints for ns, the scalar spectral index, for experiments with 104, 105, and 106 detectors and 1'–4' beams, covering 25%–75% of the sky, lensing included. We can also see that the constraints are similar across beam sizes and number of detectors.

Table 5. The ns 1σ Constraints in Units of 10−3 from CMB and from CMB+BAO

  CMB CMB+BAO
1' 2' 3' 4' 1' 2' 3' 4'
104 Detectors                
fsky = 0.25 2.91 2.94 2.98 3.04 2.19 2.23 2.29 2.36
fsky = 0.50 2.11 2.13 2.16 2.21 1.64 1.67 1.71 1.75
fsky = 0.75 1.76 1.77 1.80 1.83 1.39 1.42 1.45 1.48
105 detectors                
fsky = 0.25 2.66 2.73 2.80 2.86 1.93 1.98 2.04 2.12
fsky = 0.50 1.94 1.97 2.01 2.06 1.44 1.47 1.51 1.56
fsky = 0.75 1.60 1.63 1.66 1.70 1.22 1.24 1.28 1.32
106 detectors                
fsky = 0.25 2.38 2.48 2.62 2.73 1.70 1.76 1.83 1.92
fsky = 0.50 1.75 1.81 1.90 1.96 1.28 1.32 1.37 1.43
fsky = 0.75 1.45 1.51 1.57 1.61 1.10 1.12 1.16 1.20

Note. "CMB" includes lensing.

Download table as:  ASCIITypeset image

The best constraint for ns is 0.00145 from Table 5 for 106 detectors, 1' beam, and 75% fsky. For reference, the constraint from TT, EE, and TE spectra alone is 0.00151. This is a factor of five improvement from the current Planck best constraint (0.0073). We note that the gain over going to a larger sky area plateaus once fsky hits 0.3, and adding BAO information helps by about 30%. For ns, the major input that changes the constraints is fsky. In the CMB only case, going from 104 to 105 Ndet and from 105 to 106 Ndet improves the constraints by 5%–10%, respectively. The improvement is better for smaller beam cases. The improvement from decreasing beam sizes per arcminute is on the percent level.

Table 6 summarizes the constraints for αs, the running of ns, for experiments with 105 and 106 detectors and 1'–4' beams, covering 25%–75% of the sky, lensing included. The best constraint for αs is 0.00160 (no lensing), 0.00146 (with lensing), and 0.00145 (lensing+BAO). Compared with the current best 1σ constraints from Planck+WMAP polarization+BAO of 0.009, future CMB experiments alone can give a factor of five improvement.

Table 6. The αs 1σ Constraints in Units of 10−3 from CMB and from CMB+BAO

  CMB CMB+BAO
1' 2' 3' 4' 1' 2' 3' 4'
104 Detectors                
fsky = 0.25 3.40 3.51 3.69 3.92 3.40 3.51 3.68 3.91
fsky = 0.50 2.58 2.66 2.79 2.96 2.58 2.66 2.78 2.95
fsky = 0.75 2.20 2.26 2.37 2.51 2.20 2.26 2.37 2.51
105 detectors                
fsky = 0.25 2.79 2.92 3.09 3.31 2.78 2.91 3.08 3.30
fsky = 0.50 2.09 2.17 2.29 2.45 2.08 2.17 2.29 2.44
fsky = 0.75 1.76 1.83 1.93 2.06 1.76 1.82 1.92 2.05
106 detectors                
fsky = 0.25 2.32 2.46 2.65 2.87 2.27 2.42 2.63 2.86
fsky = 0.50 1.75 1.83 1.96 2.12 1.72 1.82 1.96 2.11
fsky = 0.75 1.47 1.54 1.65 1.77 1.45 1.53 1.64 1.77

Notes. "CMB" includes lensing. BAO measurements improve constraints in α by a few percent. The improvement is more significant for small sky fractions and small beam size scenarios.

Download table as:  ASCIITypeset image

Similar to ns, once fsky reaches 0.3, the improvement on the constraints for αs over adding sky area is quite flat. This is particularly true when lensing is included. We also note that the sensitivities and the beam sizes of the experiments do not have as big of an impact on the constraints as the sky area does. So again, fsky is the key input. For αs, going from 104 to 105 Ndet improves the constraints by 16%–20% and from 105 to 106 Ndet by 13%–17%. Like ns, the improvement is higher for smaller beams. With that noted, lensing has a larger impact on the constraints when beam sizes are smaller and when the sky area is smaller.

6.3. Tensor-to-scalar Ratio r

Most of the sensitivity on r comes from the B-mode polarization, which is already known to be much smaller in power than the E modes and the temperature anisotropies. Since tensor modes are so far undetected, we set r = 0 to be the fiducial value and study the 1σ uncertainty on r for various experimental configurations. In this "discovery" phase, it is initially more advantageous to improve sensitivity on a few spatial modes (Jaffe et al. 2000). When lensing starts to dominate (r ∼ 0.01), high-resolution B-mode data can be used to reconstruct the lensing field with high fidelity for delensing. In this regime, there are complicated trade-offs between sky coverage, delensing residual, and polarized foregrounds. As a result, the optimized survey for r can be significantly different from the lensing survey, which is important for constraining the sum of the neutrino mass and the dark energy equation of state. We use this separate section to study the strategies for the tensor search, taking delensing and foregrounds into account. This deep tensor survey is assumed to have an equal observing time as the lensing survey, with enough frequency coverage to achieve the target foreground residual. If r is large (>0.01), CMB-S4 can be a powerful tool for a detailed characterization of the tensor perturbations. In this section, we also look at how well r and nt could be constrained for larger values of r. Another potentially interesting goal is to measure the three-point functions of B modes. However, this is beyond the scope of this paper.

6.3.1. Method

We use a smaller range of sky fractions when producing constraints for r, while maintaining the range of detector numbers and beam sizes. The grid of net sensitivities on r is given in Table 7. The range of fsky is chosen to observe the so-called recombination bump at multipole ℓ ∼ 100. The inflationary tensor signal at very large angular scales (the "reionization bump" at ℓ ∼ 8) can in principle exceed lensing even for low values of r (r < 10−3). It is buried deep in galactic foregrounds but could in principle be recoverable with frequency-component separation. We acknowledge this exciting possibility but choose not to include the reionization bump in the forecast for simplicity.

Table 7. Table of Experimental Sensitivities s Given Detector Count and Sky Coverage, in μK arcmin

Ndet/fsky 0.0004 0.0006 0.0025 0.01 0.0225 0.04 0.25
10000 0.13 0.17 0.33 0.67 1.00 1.34 3.34
20000 0.09 0.12 0.24 0.47 0.71 0.95 2.37
50000 0.06 0.07 0.15 0.30 0.45 0.60 1.50
100000 0.04 0.05 0.11 0.21 0.32 0.42 1.06
200000 0.03 0.04 0.07 0.15 0.22 0.30 0.75
500000 0.02 0.02 0.05 0.09 0.14 0.19 0.47
1000000 0.013 0.02 0.03 0.07 0.10 0.13 0.33

Note. This grid is used for investigating the constraints for the tensor-to-scalar ratio r.

Download table as:  ASCIITypeset image

At r = 0.01, the magnitude of the primordial gravitational wave B-mode power is comparable to lensing at ℓ = 100. Lensing-induced B modes have the same frequency dependence as the CMB and cannot be distinguished by multi-frequency observations. The lensing contamination can be debiased (i.e., subtracted in power spectrum space) in the same way instrumental noise is removed in temperature power spectrum measurements when the expected lensing power is well known. Alternatively, the lensing deflection field can be reconstructed from arcminute-scale B-mode measurements and the expected lensing contamination to degree-scale B modes predicted and subtracted from the observed B-mode map. The quadratic estimators (Hu & Okamoto 2002) and maximum-likelihood estimators (Hirata & Seljak 2003) are techniques that have been developed to delens CMB maps. Here we use the method outlined in Smith et al. (2012) to forecast the residual noise level after iterative delensing, which converges to the maximum-likelihood solution.

Polarized synchrotron and thermal dust constitute the major sources of astrophysical foregrounds in the measurement of the B-mode power spectrum. In this study, we use the Planck Sky Model (Delabrouille et al. 2013) to estimate the level of polarized foreground we would have given that we observe the cleanest patches of the stated sky area. We compute the BB spectrum from polarized dust and synchrotron at 95 GHz, which is amongst the cleanest frequencies for small sky areas (Dunkley et al. 2009) and expect them to be cleaned to 1%–10% of the observed foreground BB power using existing component separation techniques (Planck Collaboration et al. 2013f). For synchrotron, we use a curved power law for emission with the reference frequency set at 20 GHz and curvature at −0.3. We use the model developed in Miville-Deschênes et al. (2008) for the synchrotron emission index. For polarized dust, we employ the galactic polarization model from the same paper (Miville-Deschênes et al. 2008), where the thermal dust emission in intensity is based on Model 7 of Finkbeiner et al. (1999), with the mean polarization fraction set to 5%. We then incorporate the cleaned foreground level as a noise term in the Fisher forecast.

The effect of tensor modes on the temperature and E-mode polarization power spectra is small compared to that on the B-mode power spectrum. So we can forecast constraints on r using only the B-mode spectrum to obtain almost all of the constraining power. In the B-mode spectrum, given perfect delensing and aside from the B-power enhancement at low-ℓ by τ during reionization, the constraint on r is independent of the rest of the νΛCDM parameter space. Therefore, in the forecast for r, losing the extra constraining power from τ, the Fisher matrix Fij in Equation (3) can be reduced to

Equation (30)

with

Equation (31)

where $C^{B_{{\rm tens}}}$ is the tensor contribution of the BB power spectrum, $N^{BB}_\ell$ is defined in Equation (7), $N^{{\rm fg}}_\ell$ is noise from foreground, and $N^{{\rm res}}_\ell$ is the BB residual after delensing the B-mode power spectrum using the fast algorithm developed in Appendices A and B of Smith et al. (2012).

In the scenario that delensing is imperfect, the constraints on r do depend on our knowledge of other parameters that determine the amplitude and shape of the lensing power spectrum. This is because the uncertainty in the lensing power spectrum cascades into the uncertainty in $N^{{\rm res}}_\ell$. Given a CMB-S4 like experiment with 5 × 105 detectors, 3' beam, and fsky = 0.5, we estimate a 0.5% uncertainty (95% C.L.) in the level of BB lensing amplitude averaged over ℓ range of 2–3000. If the power of $N^{{\rm res}}_\ell$ was 10% of the lensing BB power, the uncertainty in $N^{{\rm res}}_\ell$ would be 0.05%, which is small compared with other terms.

Since we take the fiducial r to be 0, the constraints in r inform us how well (in σ) we can differentiate r ≠ 0 from r = 0. For the specific experimental cases considered in this work, a comparison of Equation (30) with the formalism developed in Errard et al. (2011) shows a relative difference of ⩽1%.

6.3.2. Results and Discussion

Table 8 presents the 1σ constraints for r for a range of detector counts, sky fractions, beam sizes, and foreground residuals. The main result is that almost all of these experiments can constrain r at 1σ to 0.0006 or better if the foreground residual is at 1% level. A few 106 Ndet cases at 10% foreground residuals can also reach this level of constraints. This means that many of these configurations can certainly determine whether the field range is sub-Planckian or super-Planckian, and it will drive theoretical research on models of inflation. More importantly, CMB-S4 will be able to measure and characterize the B-mode power spectrum with high significance if r is large (r > 0.01).

Table 8. 1σ Constraints on r, in Units of 10−3

  1% Foreground 10% Foreground
1' 2' 3' 4' 1' 2' 3' 4'
104 Detectors                
fsky = 0.0025 0.50 0.53 0.59 0.66 1.76 1.82 1.91 2.03
fsky = 0.04 0.52 0.54 0.56 0.58 1.19 1.21 1.24 1.28
fsky = 0.25 0.72 0.72 0.73 0.74 1.46 1.47 1.48 1.50
105 detectors                
fsky = 0.0025 0.23 0.24 0.28 0.32 1.20 1.25 1.34 1.45
fsky = 0.04 0.18 0.19 0.20 0.22 0.66 0.68 0.71 0.74
fsky = 0.25 0.22 0.22 0.23 0.24 0.72 0.73 0.75 0.77
106 detectors                
fsky = 0.0025 0.15 0.16 0.18 0.21 0.96 1.01 1.08 1.19
fsky = 0.04 0.09 0.10 0.11 0.12 0.48 0.49 0.52 0.56
fsky = 0.25 0.10 0.11 0.12 0.13 0.51 0.52 0.54 0.56

Notes. For 1% foreground residuals, almost all the constraints are better than 0.6 × 10−3 in the wide range of experimental configurations we considered. For 10% foregrounds, the cases that have 106 detectors and fsky > 0.04 can get better than 0.6 × 10−3 1σ constraints on r.

Download table as:  ASCIITypeset image

Figure 12 shows the constraints on r as a function of detector count and sky coverage for 1'–4' beams with 1% and 10% foreground residuals in power. The constraints with 1% foreground residuals are three to six times better than those with 10% foregrounds. Under a wide range of experimental parameters, we find a broad minimum in σ(r) to lie between 0.01 < fsky < 0.1. This is because delensing requires high enough experimental sensitivity, which one loses when scanning a larger sky area. Even for the most ambitious configuration we studied (106 detectors, 1' beam), we find the optimal survey area to be around 1000 deg2.

Figure 12.

Figure 12. 1σ constraints of r, in units of 10−3, as a function of detector number and sky fraction. The beam sizes of the experiments shown from left to right go from 1'–4'. The top and bottom row show constraints with 1% and 10% foreground residuals in power, respectively. This set of plots highlights the difference foreground removal makes in the constraints for r. We see that almost all experimental configurations with 1% foreground can differentiate sub-Planckian and super-Planckian inflation. We find a broad minimum in r constraints between 0.01 < fsky < 0.1 regardless of foreground.

Standard image High-resolution image

We compared iterative and quadratic delensing methods and learned that iterative delensing helps constrain r a lot more than quadratic delensing at a small sky area compared with a large sky area. For example, it improves the constraints at smaller patches by a factor of 6–20 for fsky = 0.01 for 105 detectors depending on what beam sizes the experiment has. However, the improvement is only a factor of less than two for fsky = 0.25. This is important at the discovery phase and when we do not observe large patches.

Beam size matters less when the sky area is large because having a larger sky area provides more modes that debias the result regardless of how well delensing is done. On the other hand, for a smaller sky fraction, one relies more heavily on delensing to provide competitive constraints, thus a smaller beam is advantageous. However, this improvement from delensing using a small beam gets greatly washed out if the foreground level is high, especially when the experiment has high sensitivity. For example, at fsky = 0.04, for different beam sizes and detector numbers, at least a factor of two and at most an order of magnitude of constraining power on r is lost when the foreground residual is 10% and not 0.25% (a very foreground clean case).

The nt and r constraints if r > 0.01. In the scenario that r is bigger than 0.01, the B-mode peak around ℓ of 100 is higher than the lensing B-mode power. We can therefore constrain r and nt without delensing. We run the forecast using only $C^{BB}_{\ell }$ for the ℓ range of 10–500 for experiments with a grid shown in Table 1 and 4' beam. The reason for this ℓ range is to exclude the reionization bump at ℓ < 10 and noise above ℓ > 500. Beam size is not critical here because we are not delensing. We picked a few fiducial r values greater than 0.02 to illustrate the trend. The fiducial nt value is nt = 0, i.e., we are not imposing the consistency relation between r and nt valid in single-field slow-roll inflation. Table 9 presents the constraints for r and nt for several fiducial r values.

Table 9. 1σ Constraints for r and nt for Various rfid for Combinations of Ndet and fsky at 4' Beam Size

fsky σ(r) σ(nt)
0.25 0.50 0.75 0.25 0.50 0.75
rfid = 0.2            
104Ndet 0.030 0.026 0.025 0.09 0.08 0.07
105Ndet 0.021 0.016 0.014 0.07 0.05 0.04
106Ndet 0.020 0.015 0.012 0.06 0.05 0.04
rfid = 0.1            
104Ndet 0.022 0.020 0.019 0.13 0.11 0.11
105Ndet 0.016 0.012 0.010 0.10 0.07 0.06
106Ndet 0.015 0.011 0.009 0.09 0.07 0.05
rfid = 0.05            
104Ndet 0.016 0.015 0.014 0.18 0.16 0.16
105Ndet 0.012 0.009 0.007 0.14 0.10 0.08
106Ndet 0.011 0.008 0.007 0.13 0.09 0.08
rfid = 0.02            
104Ndet 0.012 0.011 0.011 0.31 0.29 0.28
105Ndet 0.008 0.006 0.005 0.22 0.16 0.14
106Ndet 0.008 0.005 0.004 0.21 0.15 0.12

Notes. Constraints are derived without delensing and counting the lensing $C^{BB}_\ell$ as one of the noise terms. Foreground contributions are not included.

Download table as:  ASCIITypeset image

We observe that for the few high fiducial r values listed, we can have 3σ–5σ constraints without delensing in the grid we consider in this work. We can also constrain nt at the 10% level for many rfid ≳ 0.1 cases. However, with this approach, nt cannot be measured to a precision that allows us to verify the consistency relation between r and nt in single-field slow-roll inflation.

After submitting this manuscript, the BICEP2 team released the first measurement of B mode at degree scales and reported r = 0.2 (BICEP2 Collaboration et al. 2014). With this new information of a high r value, we note that many configurations of CMB-S4 can constrain nt to better than 10% without delensing and information from the reionization bump (ℓ < 10). This does not, however, include the contribution of foregrounds on the B-mode power spectrum. Projecting forward, better information on polarized foregrounds will be essential for us to extract the primordial B-mode power spectrum.

7. CONCLUSION

In this work, we forecasted how well a highly capable next-generation ground-based CMB experiment can constrain cosmological parameters of fundamental physics relevant to both HEP and cosmology. We forecast for a range of experimental inputs—104–106 detectors, 1'–4' beams (6' and 8' in some cases), and 1%–75% sky fraction—in order to see how the constraints for each parameters vary with these inputs.

We detailed in Section 2 the methods used to estimate the performance of a given CMB experimental design. We presented our results in Sections 36. Here we quote the range of constraints each parameter falls in for experiments with 104–106 detectors, 1'–4' beam, and 25%–75% fsky as illustrations and summarize the CMB-only parameter improvement dependence on Ndet, beam size, and fsky.

  • 1.  
    0.0156 ⩽ σ(Neff) ⩽ 0.0690 (CMB). Increasing Ndet from 104 to 105 improves σ(Neff) by about 30%, the same for going from 105 to 106; decreasing beam size improves σ(Neff) by ∼10% per arcminute; σ(Neff) is not sample variance limited.
  • 2.  
    15 ⩽ σ(Mν) ⩽ 24 [meV] (CMB+BAO). Increasing Ndet from 104 to 105 improves σ(Mν) by 10%–20% (a smaller beam size gives a better improvement with Ndet), while increasing Ndet from 105 to 106 improves σ(Mν) by 5%–15%; decreasing beam size improves σ(Mν) at percent levels per arcminute; σ(Mν) is sample variance limited.
  • 3.  
    164 ⩽ DETF-FoM ⩽ 303 (CMB+BAO+H0). Increasing Ndet from 104 to 105 improves the FoM by more than factor of two, the same for going from 105 to 106; decreasing beam size improves the FoM by a few percent to tens of percent, depending on the configuration; the FoM is not sample variance limited.
  • 4.  
    0.00588 ⩽ σ(pann) ⩽ 0.0110 [3 × 10−26 cm3 s−1 GeV−1] (CMB). Increasing Ndet from 104 to 105 improves σ(pann) by about 4%, the same for going from 105 to 106; decreasing beam size improves σ(pann) by ≲1%; σ(pann) is sample variance limited.
  • 5.  
    0.00074 ⩽ σ(ΩK) ⩽ 0.0014 (CMB+BAO+H0). Increasing Ndet from 104 to 105 improves σ(ΩK) by about 20%, while increasing Ndet from 105 to 106 improves σ(ΩK) by 10%–20% (a smaller beam size gives a better improvement with Ndet); decreasing beam size improves σ(ΩK) at percent levels per arcminute; σ(ΩK) is sample variance limited.
  • 6.  
    0.00110 ⩽ σ(ns) ⩽ 0.00236 (CMB+BAO). Increasing Ndet from 104 to 105 improves σ(ns) by ∼5%–10% (a smaller beam size gives a better improvement with Ndet), the same for going from 105 to 106; decreasing beam size improves σ(ns) at percent levels per arcminute; σ(ns) is sample variance limited.
  • 7.  
    0.00145 ⩽ σ(αs) ⩽ 0.00330 (CMB+BAO). Increasing Ndet from 104 to 105 improves σ(αs) by 16%–20% and by 13%–17% going from 105 to 106 (a smaller beam size gives a better improvement with Ndet); decreasing beam size improves σ(αs) at percent levels per arcminute; σ(αs) is sample variance limited.
  • 8.  
    0.00009 ⩽ σ(r) ⩽ 0.00203 for 1% and 10% foreground residual. σ(r) is foreground limited; when the foreground is high, the optimal fsky shifts higher; decreasing beam size improves σ(r) at percent levels per arcminute (the slope of this trend with beam increases with lower sky coverage).

Detailed constraints in specific cases can be read off from tables listed in each section.

Besides learning the approximate ranges of how well these parameters can be constrained with CMB-S4, we also learn how the constraints improve as functions of Ndet, beam size, and fsky.

  • 1.  
    For all parameters, except r, increasing fsky always improves the constraints even though the overall sensitivity of the experiment decreases.
  • 2.  
    For all parameters, except those related to the dark energy equation of state, going from 105 to 106 detectors yields the same or less percentage improvement on the constraints than going from 104 to 105 detectors. The improvements range from a few percent to tens of percent. For Mν, the improvement beyond 105 detectors is marginal when the BAO signal is added.
  • 3.  
    The dependence on beam size is quite mild—constraints on Neff improve by about 10% per arcminute decrease, while for all other parameters it is around a few percent improvement per arcminute decrease.

We envision CMB-S4 to be a powerful next-generation ground-based CMB polarization experiment with high resolution and high sensitivity. With CMB-S4, we showed that most constraints on cosmological parameters are sample variance limited. Combining these data sets with spaceborne observations will allow access to a larger sky fraction, thus further improving the constraints on sample variance limited parameters.

This work and collaboration were encouraged by the preparation for the Snowmass Community Summer Studies 2013. We thank Bradford Benson, John Carlstrom, Tom Crawford, Daniel Eisenstein, Daniel Green, Ryan Keisler, John Kovac, Lloyd Knox, Eric Linder, Toshiya Namikawa, Peter Redl, Leonardo Senatore, Neelima Sehgal, and Kendrick Smith for useful discussions. W.L.K.W. thanks Olivier Doré for assistance with the Fisher matrix code.

We acknowledge the use of the PSM, developed by the Component Separation Working Group (WG2) of the Planck Collaboration. We also used CAMB and an implementation of iterative delensing developed by Wei-Hsieng Teng. C.D. was supported by the National Science Foundation grant number AST-0807444, NSF grant number PHY-088855425, and the Raymond and Beverly Sackler Funds. C.L.K. acknowledges the support of an Alfred P. Sloan Research Fellowship and an NSF Faculty Early Career Development (CAREER) Award (award number: 1056465). A.T.L. acknowledges support from the U.S. Department of Energy Office of Science. O.Z. acknowledges support by National Science Foundation through grants ANT-0638937 and ANT-0130612.

APPENDIX: DETF CONSTRAINTS

Tables 10 and 11 provide the 1σ constraints for dark energy parameters with and without fixing neutrino masses for various detector count, sky coverage, and beam sizes.

Table 10. Dark Energy Parameter Constraints (Neutrino Mass Marginalized)

  CMB CMB+BAO CMB+BAO+1% H0 Prior
FoM σ(w0) σ(wa) σ(wp) σ(ΩK) FoM σ(w0) σ(wa) σ(wp) σ(ΩK) FoM σ(w0) σ(wa) σ(wp) σ(ΩK)
104 Detectors                              
0.25, 1' 8 0.26 1.2 0.10 12.0 86 0.12 0.39 0.030 1.3 170 0.069 0.26 0.023 1.3
0.50, 1' 12 0.21 0.97 0.086 9.4 100 0.11 0.36 0.028 1.1 187 0.067 0.24 0.022 1.0
0.75, 1' 15 0.18 0.84 0.078 7.9 108 0.11 0.34 0.027 1.0 196 0.065 0.24 0.021 0.9
0.25, 2' 7 0.28 1.3 0.11 12.4 84 0.13 0.40 0.030 1.4 169 0.069 0.26 0.023 1.3
0.50, 2' 11 0.22 1.0 0.093 9.7 97 0.12 0.37 0.028 1.1 185 0.067 0.25 0.022 1.0
0.75, 2' 14 0.19 0.88 0.084 8.2 104 0.11 0.35 0.027 1.0 193 0.066 0.24 0.022 0.9
0.25, 3' 6 0.29 1.3 0.12 12.8 81 0.13 0.40 0.031 1.4 166 0.070 0.26 0.023 1.3
0.50, 3' 9 0.23 1.1 0.10 10.1 93 0.12 0.38 0.029 1.1 181 0.068 0.25 0.022 1.0
0.75, 3' 12 0.21 0.93 0.093 8.5 99 0.12 0.36 0.028 1.0 189 0.066 0.24 0.022 0.9
0.25, 4' 5 0.32 1.4 0.14 13.3 78 0.13 0.41 0.031 1.4 164 0.070 0.26 0.023 1.4
0.50, 4' 8 0.25 1.1 0.12 10.5 88 0.13 0.39 0.029 1.1 178 0.068 0.25 0.023 1.1
0.75, 4' 10 0.22 0.98 0.11 8.8 94 0.12 0.38 0.028 1.0 185 0.067 0.24 0.022 1.0
105 detectors                              
0.25, 1' 18 0.18 0.80 0.069 9.3 106 0.11 0.34 0.028 1.2 187 0.066 0.24 0.022 1.2
0.50, 1' 28 0.14 0.66 0.054 7.2 132 0.093 0.30 0.025 1.0 214 0.062 0.23 0.020 0.9
0.75, 1' 37 0.13 0.57 0.047 6.0 153 0.085 0.28 0.024 0.8 234 0.060 0.22 0.020 0.8
0.25, 2' 15 0.19 0.88 0.074 9.9 101 0.11 0.35 0.028 1.3 183 0.067 0.25 0.022 1.2
0.50, 2' 25 0.15 0.70 0.058 7.4 126 0.097 0.31 0.026 1.0 209 0.063 0.23 0.021 0.9
0.75, 2' 33 0.13 0.60 0.050 6.1 144 0.089 0.29 0.024 0.9 227 0.061 0.22 0.020 0.8
0.25, 3' 13 0.21 0.97 0.081 10.3 96 0.12 0.36 0.029 1.3 179 0.067 0.25 0.022 1.2
0.50, 3' 21 0.16 0.76 0.064 7.7 118 0.10 0.32 0.026 1.1 203 0.064 0.23 0.021 1.0
0.75, 3' 28 0.14 0.64 0.056 6.3 134 0.094 0.30 0.025 1.0 219 0.062 0.22 0.020 0.8
0.25, 4' 10 0.24 1.1 0.092 10.7 90 0.12 0.38 0.029 1.3 175 0.068 0.25 0.023 1.3
0.50, 4' 17 0.18 0.82 0.073 8.0 110 0.11 0.34 0.027 1.0 197 0.065 0.24 0.021 1.0
0.75, 4' 23 0.15 0.69 0.064 6.5 124 0.10 0.31 0.026 0.9 211 0.063 0.23 0.021 0.9
106 detectors                              
0.25, 1' 50 0.10 0.38 0.052 6.3 146 0.085 0.27 0.026 1.0 227 0.058 0.21 0.021 1.0
0.50, 1' 74 0.085 0.34 0.040 5.0 190 0.072 0.23 0.023 0.8 270 0.054 0.20 0.019 0.8
0.75, 1' 92 0.077 0.32 0.034 4.3 225 0.065 0.21 0.021 0.7 303 0.051 0.19 0.018 0.7
0.25, 2' 40 0.11 0.45 0.056 7.0 133 0.090 0.29 0.026 1.1 214 0.060 0.22 0.021 1.1
0.50, 2' 59 0.096 0.40 0.042 5.5 173 0.077 0.25 0.023 0.9 253 0.056 0.20 0.019 0.9
0.75, 2' 75 0.086 0.37 0.036 4.8 204 0.069 0.23 0.021 0.8 283 0.053 0.19 0.018 0.8
0.25, 3' 30 0.13 0.55 0.061 8.0 120 0.097 0.31 0.027 1.2 202 0.063 0.23 0.021 1.2
0.50, 3' 45 0.11 0.48 0.047 6.2 154 0.083 0.27 0.024 0.9 235 0.059 0.21 0.020 0.9
0.75, 3' 58 0.099 0.43 0.040 5.2 182 0.075 0.25 0.022 0.8 261 0.056 0.20 0.019 0.8
0.25, 4' 22 0.16 0.67 0.069 8.8 109 0.10 0.33 0.028 1.3 191 0.065 0.24 0.022 1.2
0.50, 4' 34 0.13 0.56 0.053 6.7 138 0.090 0.29 0.025 1.0 221 0.061 0.22 0.020 0.9
0.75, 4' 44 0.11 0.50 0.045 5.5 161 0.082 0.27 0.023 0.9 243 0.058 0.21 0.019 0.8

Notes. Table 10 shows the constraints for w0, wa, wp, and ΩK, and the FoM for various experimental setups using CMB (with lensing) and adding external information like DESI BAO and a 1% H0 prior. The first column contains (fsky, beam size) parameters. The constraints for ΩK are in units of 10−3.

Download table as:  ASCIITypeset image

Table 11. Dark Energy Parameter Constraints (Fixed Neutrino Mass)

  CMB CMB+BAO CMB+BAO+1% H0 Prior
FoM σ(w0) σ(wa) σ(wp) σ(ΩK) FoM σ(w0) σ(wa) σ(wp) σ(ΩK) FoM σ(w0) σ(wa) σ(wp) σ(ΩK)
104 Detectors                              
0.25, 1' 9 0.26 1.2 0.096 6.1 160 0.10 0.28 0.022 1.2 299 0.059 0.18 0.018 1.1
0.50, 1' 13 0.21 0.94 0.080 4.9 188 0.092 0.25 0.021 1.1 330 0.056 0.17 0.018 1.0
0.75, 1' 17 0.18 0.83 0.073 4.3 206 0.087 0.23 0.021 1.0 347 0.055 0.17 0.017 0.9
0.25, 2' 8 0.27 1.2 0.10 6.3 155 0.11 0.29 0.022 1.3 295 0.059 0.18 0.018 1.1
0.50, 2' 12 0.22 0.98 0.086 5.1 182 0.095 0.26 0.022 1.1 324 0.057 0.17 0.018 1.0
0.75, 2' 15 0.19 0.86 0.078 4.6 198 0.090 0.24 0.021 1.0 340 0.055 0.17 0.018 0.9
0.25, 3' 7 0.29 1.3 0.11 6.7 150 0.11 0.30 0.023 1.3 290 0.060 0.19 0.019 1.1
0.50, 3' 10 0.23 1.0 0.096 5.4 174 0.098 0.26 0.022 1.1 316 0.057 0.18 0.018 1.0
0.75, 3' 13 0.21 0.92 0.087 4.8 188 0.093 0.25 0.021 1.0 331 0.056 0.17 0.018 0.9
0.25, 4' 6 0.32 1.4 0.13 7.0 144 0.11 0.30 0.023 1.3 286 0.060 0.19 0.019 1.1
0.50, 4' 8 0.25 1.1 0.11 5.7 165 0.10 0.27 0.022 1.1 309 0.058 0.18 0.018 1.0
0.75, 4' 10 0.22 0.97 0.098 5.1 177 0.097 0.26 0.022 1.0 320 0.057 0.17 0.018 0.9
105 detectors                              
0.25, 1' 20 0.18 0.77 0.064 4.5 201 0.088 0.24 0.021 1.1 339 0.055 0.17 0.017 1.0
0.50, 1' 31 0.14 0.63 0.051 3.5 262 0.074 0.20 0.019 1.0 402 0.051 0.15 0.016 0.9
0.75, 1' 41 0.12 0.55 0.044 3.0 312 0.067 0.18 0.018 0.8 455 0.048 0.14 0.016 0.8
0.25, 2' 17 0.19 0.85 0.069 4.8 191 0.091 0.25 0.021 1.2 330 0.056 0.17 0.018 1.1
0.50, 2' 27 0.15 0.68 0.054 3.7 249 0.077 0.21 0.019 1.0 390 0.052 0.15 0.017 0.9
0.75, 2' 36 0.13 0.58 0.047 3.1 295 0.069 0.18 0.018 0.9 439 0.049 0.14 0.016 0.8
0.25, 3' 14 0.21 0.94 0.076 5.1 181 0.095 0.26 0.021 1.2 321 0.057 0.17 0.018 1.1
0.50, 3' 23 0.17 0.73 0.060 3.9 232 0.081 0.22 0.020 1.0 376 0.053 0.16 0.017 0.9
0.75, 3' 30 0.14 0.63 0.052 3.3 274 0.073 0.19 0.019 0.9 422 0.050 0.15 0.016 0.8
0.25, 4' 11 0.24 1.0 0.087 5.4 171 0.10 0.27 0.022 1.2 313 0.058 0.18 0.018 1.1
0.50, 4' 19 0.18 0.79 0.068 4.1 216 0.086 0.23 0.020 1.0 363 0.054 0.16 0.017 0.9
0.75, 4' 25 0.15 0.67 0.059 3.4 253 0.078 0.20 0.019 0.9 404 0.052 0.15 0.016 0.9
106 detectors                              
0.25, 1' 53 0.10 0.38 0.049 2.9 260 0.073 0.20 0.019 1.0 398 0.051 0.15 0.017 0.9
0.50, 1' 79 0.085 0.34 0.037 2.3 357 0.060 0.16 0.017 0.8 495 0.046 0.13 0.015 0.8
0.75, 1' 100 0.077 0.31 0.032 2.0 437 0.053 0.14 0.016 0.7 576 0.042 0.12 0.014 0.7
0.25, 2' 43 0.11 0.44 0.053 3.1 241 0.077 0.21 0.020 1.1 378 0.052 0.16 0.017 1.0
0.50, 2' 64 0.096 0.39 0.040 2.5 327 0.064 0.17 0.018 0.9 465 0.047 0.14 0.015 0.8
0.75, 2' 82 0.086 0.36 0.034 2.2 399 0.056 0.15 0.017 0.8 537 0.044 0.13 0.015 0.8
0.25, 3' 32 0.13 0.54 0.058 3.6 218 0.083 0.23 0.020 1.1 355 0.054 0.16 0.017 1.0
0.50, 3' 49 0.11 0.46 0.044 2.9 293 0.068 0.18 0.018 0.9 431 0.049 0.15 0.016 0.9
0.75, 3' 63 0.099 0.42 0.038 2.5 357 0.061 0.16 0.017 0.8 496 0.046 0.13 0.015 0.8
0.25, 4' 24 0.16 0.65 0.065 4.0 199 0.089 0.24 0.021 1.2 337 0.055 0.17 0.018 1.1
0.50, 4' 37 0.13 0.55 0.050 3.1 264 0.074 0.20 0.019 1.0 405 0.051 0.15 0.016 0.9
0.75, 4' 49 0.11 0.48 0.043 2.7 320 0.065 0.17 0.018 0.9 462 0.048 0.14 0.015 0.8

Notes. Table 11 shows the constraints for w0, wa, wp, and ΩK, and the FoM for various experimental setups using CMB (with lensing) and adding external information like DESI BAO and a 1% H0 prior. The first column contains (fsky, beam size) parameters. The constraints for ΩK are in units of 10−3. Same as Table 10, except that Mν is fixed as in the DETF Fisher matrix.

Download table as:  ASCIITypeset image

Footnotes

Please wait… references are loading.
10.1088/0004-637X/788/2/138