MEASUREMENTS OF SECONDARY COSMIC MICROWAVE BACKGROUND ANISOTROPIES WITH THE SOUTH POLE TELESCOPE

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

Published 2010 July 26 © 2010. The American Astronomical Society. All rights reserved.
, , Citation M. Lueker et al 2010 ApJ 719 1045 DOI 10.1088/0004-637X/719/2/1045

0004-637X/719/2/1045

ABSTRACT

We report cosmic microwave background (CMB) power-spectrum measurements from the first 100 deg2 field observed by the South Pole Telescope (SPT) at 150 and 220 GHz. On angular scales where the primary CMB anisotropy is dominant, ℓ ≲ 3000, the SPT power spectrum is consistent with the standard ΛCDM cosmology. On smaller scales, we see strong evidence for a point-source contribution, consistent with a population of dusty, star-forming galaxies. After we mask bright point sources, anisotropy power on angular scales of 3000 < ℓ < 9500 is detected with a signal-to-noise ratio ≳50 at both frequencies. We combine the 150 and 220 GHz data to remove the majority of the point-source power and use the point-source-subtracted spectrum to detect Sunyaev–Zel'dovich (SZ) power at 2.6σ. At ℓ = 3000, the SZ power in the subtracted bandpowers is 4.2 ± 1.5 μK2, which is significantly lower than the power predicted by a fiducial model using WMAP5 cosmological parameters. This discrepancy may suggest that contemporary galaxy cluster models overestimate the thermal pressure of intracluster gas. Alternatively, this result can be interpreted as evidence for lower values of σ8. When combined with an estimate of the kinetic SZ contribution, the measured SZ amplitude shifts σ8 from the primary CMB anisotropy derived constraint of 0.794 ± 0.028 down to 0.773 ± 0.025. The uncertainty in the constraint on σ8 from this analysis is dominated by uncertainties in the theoretical modeling required to predict the amplitude of the SZ power spectrum for a given set of cosmological parameters.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Anisotropy in the cosmic microwave background (CMB) has been well characterized on angular scales larger than a few arcminutes (Jones et al. 2006; Reichardt et al. 2009a; Nolta et al. 2009; Brown et al. 2009), providing stringent tests of cosmological theory and strong evidence for the ΛCDM cosmological model (Hinshaw et al. 2009; Komatsu et al. 2009). The primary CMB anisotropy is a direct record of density fluctuations at the surface of last scattering. On angular scales ≲10', the primary CMB anisotropy is exponentially damped due to photon diffusion in the primordial plasma (Silk 1968); the resulting decline in power with increasing multipole is known as the "damping tail." The anisotropy on very small scales, which is only beginning to be explored experimentally, is instead dominated by foreground emission and secondary distortions to the CMB introduced after the surface of last scattering. This secondary CMB anisotropy is produced by interactions of CMB photons with large-scale structure and can potentially provide sensitive and independent tests of cosmological parameters important to structure formation.

The most significant source of secondary CMB temperature anisotropy on small angular scales is predicted to be the inverse Compton scattering of CMB photons by the hot plasma gravitationally bound to massive, collapsed objects (Sunyaev & Zel'dovich 1972). This is known as the thermal Sunyaev–Zel'dovich (tSZ) effect. The tSZ effect distorts the blackbody spectrum of the primary CMB, creating a decrement in intensity at low frequencies and an increment at high frequencies with a null near 220 GHz. At 150 GHz, the anisotropy induced by the tSZ effect is expected to dominate over the primary CMB fluctuations at multipoles ℓ ≳ 3000. The power contributed to CMB maps by the tSZ effect depends sensitively on the normalization of the matter power spectrum, as parameterized by the rms of the mass distribution on 8 h−1 Mpc scales, σ8.

Galaxy clusters over a wide range of mass and redshift contribute significantly to the tSZ power spectrum. It is challenging to accurately model the low-mass or high-redshift clusters. For instance, lower-mass clusters are more sensitive to non-gravitational heating effects, while we have limited observational data on high-redshift clusters. There is considerable uncertainty as to the expected shape and amplitude of the tSZ power spectrum due to the current lack of knowledge of the properties of intracluster gas in low-mass and high-redshift galaxy clusters.

The scattered CMB photons also obtain a net Doppler shift when ionized matter is moving with respect to the rest frame of the CMB. This so-called kinetic Sunyaev–Zel'dovich (kSZ) effect (Sunyaev & Zeldovich 1980) depends solely on the motion and density of free electrons. In contrast to the tSZ effect, the kSZ effect has contributions from electrons with temperatures as low as 104 K. Therefore higher-redshift epochs, before massive objects finish collapsing, are expected to have relatively larger contributions to the kSZ power. Recent simulations and analytic models also predict a sizable signal from the epoch of the first radiative sources which form ionized regions several tens of Mpc across, within a largely neutral universe. Low-redshift galaxy clusters dominate the power on small angular scales, while high-redshift reionizing regions have their largest relative contribution on angular scales around ℓ = 2000. At 150 GHz, the kinetic effect is expected to amount to tens of percent of the total SZ power.

Only a handful of experiments have had sufficient sensitivity and angular resolution to probe the damping tail of the CMB anisotropy. Early measurements at 30 GHz by CBI (Mason et al. 2003; Bond et al. 2005) reported a >3σ excess above the expected CMB power at multipoles of ℓ>2000. Observations with the BIMA array at 30 GHz (Dawson et al. 2006) also reported a nearly 2σ detection of excess power at ℓ = 5237. However, more recently, the SZA experiment (also observing at 30 GHz) has published an upper limit of 149 μK2 at 95% confidence on excess power at these multipoles (Sharp et al. 2010) in apparent conflict with the previous CBI and BIMA results. For the relatively small patch (0.1 deg2) observed by BIMA, the non-Gaussian nature of the SZ sky means that there is no significant tension between the BIMA and SZA results. The latest CBI measurements (Sievers et al. 2009) include more data, improved radio source removal, and a proper treatment of non-Gaussianity of the SZ sky. These measurements continue to suggest excess power but with a significance of only 1.6σ.

At 150 GHz, the ACBAR (Reichardt et al. 2009a) and QUaD (Friedman et al. 2009) experiments have both measured the damping tail of the primary CMB anisotropy at ℓ < 3000 with high signal-to-noise ratio (S/N). Either with or without the addition of the expected foreground and tSZ contributions, the power measured at the highest multipoles by both experiments is consistent with primary CMB anisotropy. In the last year, the results of 150 GHz observations out to ℓ = 10,000 made with the Bolocam (Sayers et al. 2009) and APEX-SZ (Reichardt et al. 2009b) experiments have been released. These experiments have been used to place upper limits on power above the primary CMB of 1080 μK2 and 105 μK2 respectively at 95% confidence. The constraints on σ8 from these upper limits remain weak, in no small part due to the large, highly non-Gaussian sample variance of the tSZ effect on the small ∼1 deg2 patches of sky observed by Bolocam and APEX-SZ. The cosmic variance of the tSZ effect will be significantly reduced in the on-going ≳100 deg2 surveys being conducted by next-generation experiments such as the Atacama Cosmology Telescope (ACT; Fowler et al. 2007) and the South Pole Telescope (SPT).

In addition to the tSZ and kSZ effects, foreground emission is important on these small angular scales. After bright radio sources are removed, the most significant foreground at 150 and 220 GHz is expected to be a population of unresolved, faint, dusty, star-forming galaxies (DSFGs) with a rest-frame emission spectrum that peaks in the far-infrared. These sources have been studied at higher frequencies close to the peak of their emission spectrum21 (e.g., Holland et al. 1999; Kreysa et al. 1998; Glenn et al. 1998; Viero et al. 2009); however, extrapolating their fluxes to 150 GHz remains uncertain. Adding to the challenge is the expected significant clustering of these sources (Haiman & Knox 2000; Knox et al. 2001; Righi et al. 2008; Sehgal et al. 2010). IR emission from clustered DSFGs was first observed with the Spitzer Space Telescope at 160 μm (Lagache et al. 2007) and more recently this clustering has also been observed at submillimeter wavelengths by the BLAST experiment (Viero et al. 2009). The clustering of these DSFGs is expected to produce anisotropic power at 150 GHz with an angular power spectrum that is similar to that of the SZ effect. However, emission from DSFGs is spectrally separable from the SZ effect and the SZ power spectrum can be recovered by combining information from overlapping maps at 150 and 220 GHz.

In this work, we present measurements by the South Pole Telescope (SPT) which comprise the first significant detections of anisotropy power for ℓ>3000 at 150 and 220 GHz. The SPT has sufficient angular resolution, sensitivity, and sky coverage to produce high-precision measurements of anisotropy over a range of multipoles from ∼100 < ℓ < 9500. However, for the immediate goal of measuring secondary CMB anisotropies, we start with the first bandpower at ℓ = 2000 where primary CMB still dominates the power spectrum. We combine bandpowers from two frequencies to minimize the DSFG contribution and produce the first significant detection of the SZ contribution to the CMB power spectrum. In a companion paper (Hall et al. 2010, hereafter H10), the SPT power spectra are used to place constraints on the amplitude of the Poisson and clustered components of DSFG power. This is the first detection of the clustered DSFG power at mm-wavelengths. We use three results from H10 in setting limits on the residual point-source power in the DSFG-subtracted bandpowers (Section 6.2): an argument that residual clustered components are negligible, a measurement of the Poisson point-source power at 150 GHz, and a limit on the scatter in a spectral index between DSFGs.

We describe the instrument, observations, beams, and calibration strategy in Section 2. The time-ordered data (TOD) filtering and map-making algorithm is outlined in Section 3, along with the procedure to derive bandpowers from maps. The results of tests for systematic errors applied to the SPT data are discussed in Section 4, and the expected astronomical foregrounds are described in Section 5. The bandpowers and cosmological interpretation are given in Section 6.

2. INSTRUMENT AND OBSERVATIONS

The SPT is an off-axis Gregorian telescope with a 10 m diameter primary mirror located at the South Pole. The telescope is optimized to perform high-resolution surveys of low surface brightness sources. The first receiver to be placed on the SPT has the primary goal of identifying a mass-limited sample of galaxy clusters; the description of the first clusters detected in the SZ was reported in Staniszewski et al. (2009, hereafter S09). The receiver is equipped with a 960 element array of superconducting transition edge sensor bolometers. The detectors are split between three frequency bands centered at 95 GHz, 150 GHz, and 220 GHz, allowing the separation of the tSZ effect from the primary CMB anisotropy and foregrounds. Further details of the receiver and telescope can be found in Ruhl et al. (2004), Padin et al. (2008), and Carlstrom et al. (2009).

In this work, we use data at 150 and 220 GHz from one 100 deg2 field centered at R.A. 5h30m, decl. −55° (J2000) observed by SPT in the first half of the 2008 austral winter. Data from SPT's third (95 GHz) frequency band were unusable in 2008. The location of the field was chosen for overlap with the Blanco Cosmology Survey (BCS)22 optical survey and low dust emission. We observed this field for a total of 779 hr, of which 575 hr is used in the analysis after passing data quality cuts. The final map noise is 18 μK arcmin23 at 150 GHz and 40 μK arcmin at 220 GHz. This field accounts for half the sky area observed in 2008 and one-eighth of the total area observed by SPT to date.

The scanning strategy used for these observations involves constant-elevation scans across the 10° wide field. After each scan back and forth across the field, the telescope executes a 0fdg125 step in elevation. A complete set of scans covering the entire field takes approximately 2 hr and we refer to each complete set as an observation. Successive observations use a series of different starting elevations to ensure even coverage of the field. Of the 300 observations used in this analysis, half were performed at an azimuth scanning speed of 0fdg44 s−1 and half at a speed of 0fdg48 s−1.

2.1. Beam Measurements

The SPT beams are measured by combining maps of three sources: Jupiter, Venus, and the brightest point source in the 100 deg2 field. Observations of Jupiter are used to measure a diffuse, low-level sidelobe in the range 15' < r < 40', where r is the radius to the beam center. Although this sidelobe has a low amplitude (−50 dB at r = 30'), it contains approximately 15% of the total beam solid angle. A measurement of this sidelobe is necessary for the cross-calibration with the Wilkinson Microwave Anisotropy Probe (WMAP) described in Section 2.2. The observations of Jupiter show signs of potential nonlinearity in the response of the detectors for r <10'. For this reason we only use the observations of Jupiter to map the sidelobe at r >15'. Observations of Venus are used to measure the beam in the region 4'< r <15'. The angular extent of Jupiter or Venus has a negligible effect on the measurement of the relatively smooth beam features present at these large radii. The brightest point source in the map of the 100 deg2 field is used to measure the beam within a radius of 4'. In this way, the random error in the pointing reconstruction (7'' rms) and its impact on the effective beam are taken into account. The pointing error has a negligible effect on the relatively smooth outer (r >4') region of the beam.

A composite beam map, B(θ, ϕ), is assembled by merging maps of Jupiter, Venus, and the bright point source. The TOD are filtered prior to making these maps, in order to remove large-scale atmospheric noise. Masks with radii of 40', 25', and 5' are applied around the locations of Jupiter, Venus, and the point source, respectively. These masks ensure that the beam measurements are not affected by the filtering.

Using the flat-sky approximation, we calculate the Fourier transform of the composite beam map, B(ℓ, ϕ). From this, we compute the azimuthally averaged beam function,

Equation (1)

We note that averaging |B(ℓ, ϕ)|2 instead of B(ℓ, ϕ) would result in a percent-level noise bias in B at very high multipoles due to the presence of noise. The results in this work assume an axially symmetric beam, which is only an approximation for SPT. We simulate the effects of ignoring the asymmetry on the bandpowers and find that the errors introduced by making this assumption are negligible.

Although the measured beam function B is used for the bandpower estimation, an empirical fit is used to quantify the errors on B. B is fit to the empirical model

Equation (2)

There are two components: a main lobe (first term) and a diffuse shelf (second term). The form of the model and the numerical values of the slopes of the exponents were constructed to provide a good fit to the measured B. We note that B150 and B220 are measured and fitted separately. The rms difference between the model and the measured B's is approximately 1%. Two parameters remain free: σb, which describes the width of the main lobe, and a, which sets the relative normalization between the main lobe and the diffuse shelf. These parameters are left free to quantify the uncertainty in B. The uncertainty in the values of these parameters directly translates to an uncertainty in B.

There are a number of factors that limit the accuracy of the measurement of B. These include residual map noise, errors associated with the map-merging process, and spectral differences between the CMB and the sources used to measure the beam. The final uncertainties in the beam model parameters σb and a are constructed as the quadrature sum of the estimated uncertainties due to each of these individual sources of error. To a good approximation, the uncertainties on σb and a can be taken to be Gaussian and uncorrelated.

In practice, a change in the value of a is equivalent to a change in the overall calibration for ℓ>700. After the beam uncertainties are estimated, the uncertainty in a is folded into the estimated uncertainty on the absolute calibration and the parameter a is fixed to the best-fit value of 0.85. The quoted beam uncertainties in the CosmoMC24 (Lewis & Bridle 2002) data files in Section 6.1 include only the uncertainty on σb. Figure 1 shows the measured beam functions for 150 and 220 GHz, along with the 1σ uncertainties in the main lobe beam width σb.

Figure 1.

Figure 1. Average beam functions and uncertainties for SPT. Left axis: the SPT beam function for 150 GHz (red) and 220 GHz (blue). Right axis: the 1σ uncertainties on the beam function for each frequency. The beam uncertainties shown here include only uncertainties on the main lobe beam width, σb, since the uncertainty of the sidelobe amplitude has been subsumed into the calibration uncertainty.

Standard image High-resolution image

2.2. Calibration

The calibration of the SPT data is tied to the superb WMAP5 absolute calibration through a direct comparison of 150 GHz SPT maps with WMAP5 V- and W-band (61 and 94 GHz) maps (Hinshaw et al. 2009) of the same sky regions. Bright radio sources are masked for this comparison. The WMAP5 maps are resampled according to the SPT pointing information and the resulting TOD are passed through the SPT analysis pipeline to capture the effects of TOD filtering. The ratio of the cross-spectra of the filtered WMAP and SPT maps after correcting for the instrumental beams,

Equation (3)

is used to estimate the relative calibration factor between the two experiments. A similar procedure was used to calibrate the Boomerang, ACBAR, and QUaD experiments (Jones et al. 2006; Reichardt et al. 2009a; Brown et al. 2009). The calibration is weighted most heavily to ℓ ∈ [256, 512] and is flat with ℓ ∈ [128, 640]. The largest source of uncertainty is statistical noise in the maps of both experiments. The second largest source of uncertainty is due to the use of a secondary calibration source (RWC38) to bridge the several month gap between the science observations and calibration scans. Dedicated SPT calibration scans of four large fields totaling 1250 deg2 of sky were obtained during 2008. The results for these four fields are combined to achieve an absolute temperature calibration uncertainty of 3.6% at 150 GHz.

The 150 GHz calibration is transferred to 220 GHz through the overlapping coverage of SPT's high S/N maps. We calculate the relative calibration by examining the ratio of the cross-spectra between the 150 and 220 GHz maps to the auto-spectra of the 150 GHz map after correcting for the beam and filtering differences. We estimate the relative calibration uncertainty to be 6.2% and the final absolute calibration uncertainty of the 220 GHz temperature map to be 7.2%.

3. ANALYSIS

In this section, we describe the pipeline used to process the TOD to maps and from maps to bandpowers. Given the small sky area analyzed here, we use the flat-sky approximation. We generate maps in a flat-sky projection and analyze these maps using Fourier transforms. Thus, the discussion of filtering and data-processing techniques refers to particular modes by their corresponding angular wavenumber k in radians. Note that in these units |k| = ℓ.

3.1. Map Making

We generate one map per frequency band for each 2 hr observation of the field. The map-making pipeline used for this analysis is very similar to that used by S09, with some small modifications. An overview of the pipeline is provided below, emphasizing the aspects of data processing and map-making that are most important for understanding the power-spectrum analysis.

3.1.1. Data Selection

Variations in observing conditions and daily receiver setup can affect the performance of individual bolometer channels. The first step in the data processing is to select the set of bolometers with good performance for each individual observation. The criteria for selecting these detectors are primarily based on responsivity (determined through the series of calibrations performed prior to each field observation) and noise performance. Detectors are rejected if they have a low S/N response to a chopped thermal calibration source or atmospheric emission (modulated through a short elevation scan). They are also rejected if their noise power spectrum is heavily contaminated by readout-related line-like features. More details on the bolometer cuts can be found in S09. In addition, bolometers that have a responsivity or inverse-noise-based weight more than a factor of 3 above or below the median for their observing band are rejected. On average, 286 bolometers at 150 GHz (out of 394 possible) passed all cuts for a given observation. At 220 GHz, an average of 161 (out of 254 possible bolometers) passed all cuts.

In a given observation, segments of the TOD corresponding to individual scans may be rejected due to readout or cosmic-ray induced features, anomalous noise, or problems with pointing reconstruction. The scan-by-scan data cuts are the same as those described in S09 and remove approximately 10% of the TOD.

The SPT detectors exhibit some sensitivity to the receiver's pulse-tube cooler. Bolometer noise power spectra are occasionally contaminated by narrow spectral lines corresponding to the pulse-tube frequency and its harmonics. As in S09, we address this by applying a notch filter to the affected data. This filter removes less than 0.4% of the total signal bandwidth.

In addition to the cuts above, we remove a small number of observations that have either incomplete coverage of the field or high noise levels. Observations taken under poor atmospheric conditions are rejected in this way. The application of these cuts removes 36 out of the 336 total observations of the field.

3.1.2. Time-ordered Data Filtering

Let dαj be a measurement of the CMB brightness temperature by detector j at time α. Contributions to dαj are the celestial sky temperature and the atmospheric and instrumental noise. The instrumental noise is largely uncorrelated between detectors. However, the atmospheric signal is highly correlated across the focal plane due to the overlap of individual detector beams as they pass through turbulent layers in the atmosphere.

The signals in the data have been low-pass-filtered by the bolometers' optical time constants. These single-pole bolometer transfer functions are measured by the method described in S09. We deconvolve the transfer functions for each bolometer on a by-scan basis, as the first step in TOD filtering. At the same time, we apply a low-pass filter with a cutoff at 12.5 Hz. In the frequency domain, this filter has the frequency profile:

Equation (4)

The shape of the filter is chosen to be very sharp, yet not induce a significant amount of ringing. This filter removes noise above the Nyquist frequency associated with the final map pixelation. These operations can be represented as a linear operation ${ \bPi ^{\rm fft}}$ on the TOD,

Equation (5)

where a sum is implied over repeated indices.

To remove 1/f noise and atmospheric noise in the scan direction, we project out a 19th-order Legendre polynomial from the TOD for each detector on a scan-by-scan basis. This operation $ {\bPi ^{\rm t}}$ can be described by

Equation (6)

The effect on the signal is similar to having applied a one-dimensional high-pass filter in the scan direction at ℓ ≳ 300.

Significant atmospheric signal remains in the data after this temporal polynomial removal. Because the atmospheric power is primarily common across the entire focal plane, we can exploit the spatial correlations to remove atmosphere without removing fine-scale astronomical signal. The method used in this analysis is to fit for and subtract a plane, a0 + axx + ayy, across all detectors in the detector array at each time sample, where x and y are set by the angular separation of each pixel from the boresight of the telescope. All detectors are equally weighted in this fit. This operation can be represented as a linear operation on all detectors at each time sample α to produce an atmosphere-cleaned data set, d‴

Equation (7)

We mask the brightest point sources in the map before applying the polynomial subtraction and the spatial-mode filtering. The 92 sources that were detected above 5σ in a preliminary 150 GHz map have been masked. A complete discussion of the point sources detected in this field can be found in Vieira et al. (2009, hereafter V09).

3.1.3. Map Making

The data from each bolometer are inverse noise weighted according to the calibrated, pre-filtering detector power spectral density (PSD) in the range 1–3 Hz, corresponding to 1400 < ℓ < 4300. This multipole range covers the angular scales where we expect the most significant detection of the SZ effect. We represent the mapping between time-ordered bolometer samples and celestial positions with the pointing matrix L, which we apply to the cleaned TOD to produce a map m

Equation (8)

${ \bLambda _{\bf w}}$ is a diagonal matrix encapsulating the detector weights. Information on the pointing reconstruction can be found in S09 and Carlstrom et al. (2009). This procedure is repeated for all 300 observations used in this analysis. The mean maps for the 150 GHz and 220 GHz data are shown in Figure 2.

Figure 2.

Figure 2. Maps of the field used in this analysis. The 150 GHz map is shown in the left panel, the 220 GHz data are in the right panel. See Section 3.1 for a description of the map-making pipeline. Near the center of each map the depth is fairly uniform, with a noise rms of roughly 18 μK arcmin at 150 GHz, and 40 μK arcmin at 220 GHz. The apparent apodization on the left and right edges is due to the polynomial filter which has been applied to the TOD. For presentation purposes, the maps have been low-pass filtered at ℓ = 11,000. Point sources have not been masked.

Standard image High-resolution image

3.2. Maps to Bandpowers

We use a pseudo-C method to estimate the bandpowers. In pseudo-C methods, bandpowers are estimated directly from the Fourier transform of the map after correcting for effects such as TOD filtering, beams, and finite sky coverage. We process the data using a cross-spectrum based analysis (Polenta et al. 2005; Tristram et al. 2005) in order to eliminate noise bias. Beam and filtering effects are corrected for according to the formalism in the MASTER algorithm (Hivon et al. 2002). We report the bandpowers in terms of $\mathcal {D}_\ell$, where

Equation (9)

As the first step in our analysis, we calculate the Fourier transform, $\tilde{m}_{\bf {k}}$, of the single-observation maps for each frequency. Each map is apodized using the same window $\textbf {W}$, and thus $\tilde{m}^{(\nu,A)}_{\bf {k}}\equiv \hbox{FFT}(\textbf {W}\textbf {m}^{(\nu,A)})$, where the first superscript, ν, indicates the observing frequency, the second superscript, A, indicates the observation number, and the subscript denotes the angular frequency. Cross-spectra are computed for every map pair and averaged within the appropriate ℓ-bin b:

Equation (10)

Here, Nb is the number of modes in bin b. We use the abbreviation, $D^{\nu _i}\equiv D^{\nu _i\times \nu _i}$, when referring to single-frequency auto-spectra. Recall that |k| = ℓ in the flat-sky approximation.

The FFT of a map, $\widetilde{m}_{\bf k}^{(\nu,A)}$, is linearly related to the FFT of the astronomical sky, aνk. It also includes a noise contribution, n(ν,Ai)k, which is uncorrelated between the 300 observations:

Equation (11)

Here Gk is the two-dimensional amplitude transfer function, which accounts for the TOD filtering as well as the map-based filtering described in Section 3.2.2. $\widetilde{\textbf {W}}_{\bf {k}}$ is the Fourier transform of the apodization mask.

Following the treatment by Hivon et al. (2002), we take the raw spectrum, DABb, to be linearly related to the true spectrum by a transformation, $K_{bb^\prime }$. The K transformation combines the power-spectrum transfer function, Fk, which includes the effects of TOD-based and map-based filtering (Section 3.2.3), the beams, Bk (Section 2.1), and the mode-coupling matrix, $M_{kk^\prime }[\textbf {W}]$ (Section 3.2.1). The mode-coupling matrix accounts for the convolution of the spectrum due to the apodization window, $\textbf {W}$

Equation (12)

Here, Pbk is the re-binning operator (Hivon et al. 2002):

Equation (13)

while the inverse of the re-binning operator is Qkb:

Equation (14)

In the rest of the paper, we will often refer to band-averaged quantities, such as Cb = PbkCk. Cross-spectra between observations that have been corrected for apodization and processing are denoted as

Equation (15)

The final bandpowers are then computed as the average of all cross-spectra,

Equation (16)

We make the simplifying assumption that the noise properties of each observation are statistically equivalent; hence, a uniform weighting is chosen here. The data selection criteria in Section 3.1.1 ensure that these observations target the same region of the sky, have roughly the same number of active bolometers, and have similar noise properties. Therefore, this uniform weighting should be unbiased and only slightly sub-optimal. Simulations suggest that properly weighting the observations would reduce the bandpower uncertainties by 14%.

The cross-spectrum bandpowers, $D^{\nu _i\times \nu _j,{\it AB}}_b$, generated from the 300 observations are used in conjunction with signal-only Monte Carlo bandpowers to generate empirical covariance matrices, as described in the Appendix. The variance of the Monte Carlo bandpowers is used to estimate the sample variance contribution to the covariance matrix. Meanwhile, we use the variance in the spectra of the real data to estimate the uncertainty due to noise in the maps.

3.2.1. Apodization Mask and Calculation of the Mode-mixing Kernel

Since we have only mapped a fraction of the full sky, the angular power spectra of the maps are convolutions of the true Cs with an ℓ-space, mode-mixing kernel that depends on the map size, apodization, and point-source masking. We calculate this mode-mixing kernel, $M_{kk^\prime }[\textbf {W}]$, following the derivation in Hivon et al. (2002) for the flat-sky case:

Equation (17)

At high-k, elements of the mode coupling kernel depend only on the distance from the diagonal. The normalization of f(x) is ∫0dxf(x) = w2, where $w_n\equiv \left<\textbf {W}^n\right>$ represents the nth moment of the apodization window. Thus, the coupling kernel also serves the purpose of re-normalizing the power spectrum to account for modes lost due to apodization. As we discuss in the Appendix, the coupling kernel also plays an important role in determining the shape and normalization of the covariance matrix.

We avoid areas of the map with sparse or uneven coverage in any single observation. Thus, the apodization window is conservative in its avoidance of the map edges. We also mask 144 point sources detected in the 150 GHz data above a significance of 5σ, which corresponds to a flux of 6.4 mJy. This source list is from a more refined analysis than the preliminary list used in Section 3.1.2; the differences are in sources near the 5σ detection threshold. Each point source is masked by a 2' radius disk, with a Gaussian taper outside this radius. In other words, pixels located at a distance r from the source location receive the weight:

Equation (18)

Many different tapered mask shapes were tested for both efficacy in removing point-source power and noise performance. Given the relatively small area masked by point sources, varying the shape of the point-source mask has little effect on the final spectrum. The effective sky area of the mask is 78 deg2. Simulations have been performed to test whether the application of this mask will bias the inferred power and we see no bias with a 5σ cut. As could be expected, we do observe a mild noise bias when using a more aggressive 3σ point-source mask. The same 5σ mask is used for all maps at both frequencies and for all observations.

3.2.2. Fourier Mode Weighting

The maps produced by the steps described in Section 3.1.2 have anisotropic signal and noise. In particular, the map noise PSD rises steeply at spatial frequencies corresponding to low spatial frequencies in the scan direction (low kx). The covariance of the power estimated in a given ℓ-bin depends on the second power of the map noise PSD for all the Fourier modes in that ℓ-bin. In the presence of either non-uniform noise or signal, applying an optimal mode weighting when calculating the mean bandpower may significantly reduce the final noise covariance matrix of the power-spectrum bandpowers. In the case of SPT, we find that for the purposes of measuring the ℓ ≳ 2000 power spectrum, a simple, uniform selection of modes at kx>1200 is close to optimal, and we apply this mode weighting in calculating the SPT bandpowers.

3.2.3. Transfer Function Estimation

In order to empirically determine the effect of both the TOD-based filtering and the Fourier mode weighting, we calculate a transfer function, Fk, as defined in Hivon et al. (2002). Note that this power-spectrum transfer function is distinct from the amplitude transfer function Gk. Specifically, the transfer function accounts for all map- and TOD-processing not taken into account by the mode-coupling kernel or the beam functions.

We created 300 Monte Carlo sky realizations at 150 and 220 GHz in order to calculate the transfer function of the filtering. The simulations also serve as an input for the covariance matrix estimation. These simulations contain two components: a CMB component and a point-source component. The CMB component is computed for the best-fit WMAP5 lensed ΛCDM model.

The point-source component includes two different populations of dusty galaxies, a low-z population and a high-z population. For each population we generate sources from a Poisson distribution. Sources are generated in bins of flux, S, ranging from 0.01 to 6.75 mJy. This upper limit in flux is close to the 5σ detection threshold in the 150 GHz maps. In each flux bin, the 150 GHz source density, dN/dS, of each population is taken from the model of Negrello et al. (2007). We relate the flux of each source at 220 GHz to its flux at 150 GHz with a power law in intensity, Sν ∝ να. The power-law spectral index, α, for each source is drawn from a normal distribution. We use spectral indices of α = 3 ± 0.5 for the high-z protospheroidal galaxies and α = 2 ± 0.3 for the low-z IRAS-like galaxies. As with any Poisson distribution of point sources, the power spectrum of these maps is white (Cps = constant) and related to the flux cutoff, S0, by

Equation (19)

The power spectra of these simulated point-source maps are well represented by a constant Cps = 1.1 × 10−5 μK2 at 150 GHz and Cps = 6.8 × 10−5 μK2 at 220 GHz.

These simulated maps are smoothed by the appropriate beam. From each map realization, we construct simulated TOD using the pointing information. Each realization of the TOD is processed using the low-pass filter, polynomial removal, and the spatial-mode subtraction described in Section 3.1.2. No time-constant deconvolution is applied, since these realizations of the TOD have not been convolved by the bolometer time constants. The filtered, simulated TOD are then converted into maps according to Equation (8).

We apply the same apodization mask and Fourier mode weighting to these map realizations as is used on the actual data. We then compute the Monte Carlo pseudo-power spectrum, (Dk)MC, for each map. The simulated transfer function is calculated iteratively by comparing the Monte Carlo average, <Dk>MC, to the input theory spectrum, Ctheoryk.

For the single-frequency bandpowers, we start with an initial guess of the transfer function:

Equation (20)

The superscript (0), indicates that this is the first iteration in the transfer function estimates. For this initial guess, we approximate the coupling kernel as largely diagonal. Thus, the factor w2 is the normalization factor required by the apodization window. We then iterate on this estimate using the full mode-coupling kernel:

Equation (21)

We iterate on this estimate five times, although the transfer function has largely converged after the first iteration.

This method may misestimate the transfer function if the simulated spectrum is significantly different from the true power spectrum. The primary CMB anisotropy has been adequately constrained by previous experiments; however, the foreground power spectrum is less well known. We repeated the transfer function estimation using an input power spectrum with twice the nominal point-source power. The resulting transfer function was unchanged at the 1% level, giving confidence that the transfer function estimate is robust.

The transfer function for a multifrequency cross-spectrum is taken to be the geometric mean of the two individual transfer functions:

Equation (22)

This treatment is only strictly correct for isotropic filtering. As a cross-check, we have also computed the cross-spectrum transfer function directly. For the angular multipoles reported in this work, the geometric-mean transfer function estimate is in excellent agreement with the estimate obtained using $\left<D^{\nu _1\times \nu _2}_k\right>_{\rm MC}$ in Equations (20) and (21).

3.2.4. Window Function Estimation

The bandpower window functions, $ {\mathcal W}^b_\ell / \ell$, are defined by the following equation:

Equation (23)

in order to facilitate the comparison of experimental bandpowers, qb, to a theoretical power spectrum, Cth. Following the formalism laid out above, we can write this as

Equation (24)

which implies that

Equation (25)

3.2.5. Frequency-differenced Spectra

We are interested in the power spectra of linear combinations of the 150 and 220 GHz maps designed to remove astronomical foregrounds. One method for obtaining such power spectra would be to directly subtract the maps after correcting the maps for differences in beams or processing. In such a map subtraction, the scaling of the 220 GHz map, x, can be adjusted to optimally remove foregrounds. The differenced maps can then be processed using our standard pipeline to obtain spectra with a reduced foreground contribution.

Equivalently, the differenced spectrum can be generated from the original bandpowers, qib, using a linear spectrum transformation, ξ:

Equation (26)

Here, the index, i, denotes the 150 GHz auto-spectrum, 220 GHz auto-spectrum, or 150 × 220 GHz cross-spectrum. This transformation is computationally fast and takes advantage of the fact that the cross-frequency bandpowers include information on the relative phases of each Fourier component in the map. In this way, the difference spectrum can be represented in terms of the three measured spectra:

Equation (27)

The overall normalization is chosen such that the CMB power is unchanged in the subtracted spectrum. For clarity, we have momentarily avoided here the complications of beams and filtering and have expressed the bandpowers in terms of the Fourier transform of the celestial sky, ak. For a given proportionality constant, x, the values of ξ(x) are

Equation (28)

In order to compute covariances for the subtracted spectrum, one needs to know the correlations between bandpowers measured at different frequencies. Since both frequencies cover the exact same area of sky, the 150 GHz, 220 GHz, and cross-spectra are nearly completely correlated at low ℓ's where the errors are dominated by sample variance. At higher ℓ-ranges the cross-spectrum bandpowers are also partially correlated with both the 150 GHz and 220 GHz bandpowers, since the instrumental contribution to the cross-spectrum variance originates from the 150 GHz and 220 GHz noise. We represent the correlation between ℓ-bins, b and b', measured at frequencies, i and j as follows:

Equation (29)

As before, the superscripts i and j may stand for 150 GHz, 220 GHz, or the cross-spectrum 150 × 220 GHz. The method for computing the multifrequency covariance matrix is discussed in the Appendix. By combining Equations (26) and (29), one can calculate the covariance matrix for the subtracted bandpowers from the multifrequency covariance matrices:

Equation (30)

In this way, we account for correlations between bandpowers of different frequencies when computing the subtracted-spectrum covariance matrix. Ignoring these correlations can lead to a gross overestimate of the uncertainty in the subtracted spectrum. For the SPT bandpowers presented in Section 6.1, neglecting these correlations leads to a 100% overestimate of δC at ℓ = 2000 for x = 0.325.

4. JACKKNIVES

We apply a set of jackknife tests to the SPT data to search for possible systematic errors. In a jackknife test, the data set is divided into two halves, based on features of the data associated with potential sources of systematic error. The two halves are differenced to remove any astronomical signal and the resulting power spectrum is compared to zero. Significant deviations from zero would indicate either a systematic problem or a noise misestimate. We implement the jackknives in the cross-spectrum framework by differencing single pairs of observations and applying the cross-spectrum estimator outlined in Section 3.2 to the set of differenced pairs. In total, we perform 13 different jackknife tests.

Six jackknives are based on the observing parameters, such as time, scan direction, and azimuthal range. The data can be split based on when it was taken to search for systematic changes in the calibration, beams, detector time constants, or any other time variable effect. The "first half–second half" jackknife probes variations on month time scales, while an "even–odd" jackknife differencing every other observation looks for variations on hourly time scales. Results for the "first half–second half" jackknife are shown in the top panel of Figure 3. The data can also be split based on the direction of the scan in a "left–right" jackknife (second panel of Figure 3). We would expect to see residual power here if the detector transfer function had been improperly deconvolved, if the telescope acceleration at turnarounds induces a signal through sky modulation or microphonics, or if the wind direction is important. We observed this field at two scan speeds different by 10%. We check for systematic differences related to the scan speed, such as a mirror wobble, in a "scan speed" jackknife. Side-lobe pickup could potentially introduce spurious signals into the SPT maps from the moon or features on the ground. We test for moon pickup by splitting the data based on whether the moon is above or below the horizon. We test for ground pickup by splitting on the mean azimuth of the observation. To maximize the sensitivity to ground pickup, the azimuthal ranges are selected to be centered on and directly opposite the closest building to the telescope, which is the most likely source of ground signal. The azimuthal jackknife is shown in the third panel of Figure 3.

Figure 3.

Figure 3. Jackknives for the SPT data set at 150 GHz (blue circles) and 220 GHz (black diamonds). For clarity, the 220 GHz jackknives have been shifted to the right by Δℓ = 100. Top panel: bandpowers of the "first half–second half" jackknife compared to the expected errorbars about zero signal. Disagreement with zero would indicate either a noise misestimate or a time-dependent systematic signal. Second panel: power spectrum of the left-going minus right-going difference map. This test yields strong constraints on the accuracy of the detector transfer function deconvolution and on possible directional systematics. Third panel: bandpowers for the difference map when the data is split based on azimuth. Signals fixed in azimuth such as side-lobe pickup from the nearby support building would produce non-zero power. We see no evidence for ground-based pickup. The cumulative probability to exceed the χ2 observed in these three tests is 76% at 150 GHz and 22% at 220 GHz. Bottom panel: the un-differenced SPT power spectra at each frequency for comparison.

Standard image High-resolution image

We also perform jackknives based on four noise and observation-quality statistics of the 150 GHz data. The first is based on the overall rms in the maps, which is affected by atmospheric conditions and detector noise. The second is based on the average raw detector PSD in the range 9–11 Hz, which is a measure of the detector "white" noise level. The third is the rms near ℓ = 3000 where the S/N of the SZ power spectrum is highest. The fourth is based on the number of bolometers active in each observation.

There are also a number of line-like spectral features in the SPT TOD that could potentially affect the power-spectrum bandpowers, and we perform three jackknife tests for sensitivity to these features. Some of these line features appear at harmonics of the receiver pulse-tube frequency and are typically correlated across many bolometer channels. These lines are filtered from the data as described in Section 3.1.2. In addition, some channels exhibit occasional line-like features at other frequencies, which are not filtered in the data processing. We search for residual effects in a jackknife based on the average number of line-like features for all 150 GHz bolometers, as well as a jackknife based on the bandwidth affected by the lines. Finally, we do an additional split using the average number of lines in an observation that appear to be related to the pulse-tube cooler.

We calculate the χ2 with respect to zero for each jackknife over the range ℓ ∈ [2000, 10,000] in bins with Δℓ = 500. Some of the tests are highly correlated. For example, we changed scan velocities approximately midway through the observations so splitting the data based on scan velocity is nearly identical to splitting the data between the first and second halves of the season. We calculate a correlation coefficient between the different tests by adding 1/Nobs for each common observation in a half and subtracting 1/Nobs for each distinct observation in a half. This algorithm returns unity for two identical sets and zero for two random sets, as we expect 50% of the observations to be in common for two random selections. The correlation coefficients between the 13 jackknives range from 0 to 0.83, with the maximum correlation being for the previously mentioned scan velocity and first half–second half jackknives. We invert the jackknife correlation matrix, $ {\bf \mathcal {C}}$, and calculate $\chi ^2 = v_{i\alpha } ({\mathcal {C}}^{-1})_{ij} v_{j\alpha }$. Here viα is the ratio of the bandpower over the uncertainty for the ith jackknife and the αth ℓ-bin. The probability of exceeding the measured χ2 for the complete set of 13 jackknives is 77% for the 150 GHz data, 32% for the 220 GHz data, and 57% for the combined set with both frequencies. We thus find no evidence for systematic contaminants in the SPT data set.

5. FOREGROUNDS

The main foregrounds at frequencies near 150 GHz are expected to be galactic dust emission, radio sources, and DSFGs. Note, however, that the SPT field is selected to target one of the cleanest regions on the sky for galactic dust emission, and in the Finkbeiner et al. (1999) model, dust emission is primarily on large angular scales. The contribution for the selected field on arcminute scales is insignificant. The primary foregrounds of consideration for this analysis are radio sources and DSFGs.

Tens of bright radio sources are detected in the SPT maps at >5σ and contribute substantial amounts of power at both 150 and 220 GHz. Information on the fluxes and spectral indices of these and other sources significantly detected in the SPT maps can be found in V09. Without masking, the measured point-source power is Cunmasked = 2.1 × 10−4 μK2 at 150 GHz and Cunmasked = 1.6 × 10−4 μK2 at 220 GHz. These estimates are dominated by the brightest few sources and thus subject to very large sample variance. We mask all sources with 150 GHz fluxes above the 5σ detection threshold, S = 6.4 mJy. By masking these sources we reduce the radio source contribution to the SPT bandpowers by several orders of magnitude. According to the de Zotti et al. (2005) model source counts, after masking bright sources, we expect a residual radio source contribution of Cradio = 8.5 × 10−7 μK2 at 150 GHz.

The point-source masking will remove the SZ contribution from only a few galaxy clusters, leading to negligible reduction of the SZ power. This is because the large majority of radio sources and DSFGs reside outside of SZ clusters. Most of the masked sources are identified as radio sources in V09. Extrapolations from lower frequency observations imply that, at 150 GHz, less than 3% of clusters contain radio source flux exceeding 20% of the tSZ flux decrement (Lin et al. 2009; Sehgal et al. 2010). The masked sources are selected as increments at 150 GHz and therefore have fluxes much greater than 20% of the tSZ of any associated galaxy cluster. The number of clusters masked by the radio source masking should then be much less than 3% and negligible. We also compare the tSZ power spectrum in the Sehgal et al. (2010) simulated sky maps with and without masking >6.4 mJy sources and find the difference to be ≪1%. A small number (six) of the masked sources are identified as DSFGs (V09). Galaxy clusters have a DSFG abundance only 20 times larger than the field (Bai et al. 2007), although they exceed the mass density of the field by a factor of 200 or more. Given the relative rarity of galaxy clusters, it follows that only a small fraction of DSFGs can live in galaxy clusters. Therefore, the number of clusters masked along with the DSFGs should be much smaller than six and negligible.

Both the DSFG and radio source arguments above depend implicitly on the impact of potentially masking ≲10 clusters. As a worst-case study, Shaw et al. (2009) consider the impact of masking the most massive 10 clusters in the field and show that even in this extreme case, the tSZ power spectrum at ℓ = 3000 is reduced by only 11%. Of course, the point-source masking will not select the most massive clusters and is highly unlikely to remove as many as 10 clusters. Hence, the true impact will be significantly less.

After we mask sources above the 5σ threshold, DSFGs are the dominant point-source population in the SPT maps. These sources have been extensively studied at higher frequencies by SCUBA (Holland et al. 1999), MAMBO (Kreysa et al. 1998), Bolocam (Glenn et al. 1998), LABOCA (Siringo et al. 2009), AzTEC (Scott et al. 2008), SCUBA-2 (Holland et al. 2006), and BLAST (Pascale et al. 2008), and there have been some preliminary indications of their contribution in previous small-scale power spectra at 150 GHz (Reichardt et al. 2009a, 2009b). The flux of these galaxies has been observed to scale to higher frequencies as Sν ∝ ν2.4−3.0 (Knox et al. 2004; Greve et al. 2004; Reichardt et al. 2009b), with the exact frequency dependence a function of the dust emissivity, the dust temperature, and the redshift distribution of the galaxy population. This range of spectral indices corresponds to point-source amplitude ratios, δTps220Tps150, of 2.1–2.6 when expressed in units of CMB temperature. The measured spectral index of the DSFGs in the SPT maps is discussed extensively in H10.

In order to obtain an unbiased estimate of the SZ power spectrum, it is essential that we take these sources into account in our fits and modeling. After masking the bright point sources, we significantly detect a Poisson distributed power at 150 GHz of Cps = 7.1 ± 0.5 × 10−6μK2 (H10). This unclustered point-source power climbs with increasing ℓ to be comparable to the SZ effect by ℓ = 2500–3000 and is the dominant astronomical signal in the maps at arcminute scales.

The distribution of DSFGs on the sky is also expected to be clustered, resulting in a significant increase in power at ℓ ≲ 3000. BLAST recently detected this clustered term for DSFGs at 600–1200 GHz (500–250 μm; Viero et al. 2009). Extrapolating the measured clustering to 150 GHz, we expect the clustered contribution to be comparable to the tSZ effect. H10 have analyzed the SPT 150, 220, and 150 × 220 GHz bandpowers presented in Section 6 and have also found that the amplitude of the clustered component is indeed comparable to the tSZ amplitude. Discriminating between clustered DSFGs and the tSZ effect would be extremely difficult for a single-frequency instrument as the angular dependences are very similar. However, the two frequencies used in this analysis allow the spectral separation of these two astronomical signals.

6. RESULTS

In addition to presenting the SPT angular power spectra, we wish to determine the amplitude of the tSZ power spectrum and to use this amplitude to constrain the normalization of the matter power spectrum, σ8. This measurement requires separating the tSZ signal from the other astrophysical signals in our data, which include primary CMB anisotropy (including lensing effects), DSFGs (both Poisson and clustered components), and anisotropy due to the kSZ effect. The primary CMB anisotropy and Poisson point-source component can be separated from a tSZ-like component on account of the distinct angular power spectra of these three signals. However the tSZ, kSZ, and clustered DSFG components are all expected to be roughly flat in $\mathcal {D}_\ell$, and we depend on their distinct frequency dependences to separate them. We use a combination of the two frequency bands to remove the DSFG contribution to the power spectra. We address the remaining degeneracy between the tSZ and kSZ effects by repeating the analysis for a range of assumed kSZ models.

In Section 6.1, we present bandpowers for 150, 220, and 150 × 220 GHz SPT angular power spectra. In Section 6.2, we combine the two frequency bands to generate a set of DSFG-subtracted bandpowers. In Section 6.3, we describe the Monte Carlo Markov chain (MCMC) analysis used to estimate the tSZ power-spectrum amplitude, parameterized as the normalization ASZ of a model template, from the DSFG-subtracted bandpowers. In Section 6.4, we discuss the implications of the measured tSZ power-spectrum amplitude for σ8 and modeling of the tSZ effect.

6.1. Power Spectra

Figure 4 shows the bandpowers we compute by applying the analysis methods described in Section 3 to one 100 deg2 field observed by SPT at 150 and 220 GHz. The bandpowers for the two frequencies and their cross-spectrum are tabulated in Table 1. We combine these multifrequency data as described in Section 3 to produce the "DSFG-subtracted" bandpowers listed in Table 2. This power spectrum is compared to the results from WMAP5, ACBAR, and QUaD in Figure 5. The best-fit model to this combined data set including the primary CMB, kSZ, tSZ, and a Poisson point-source contribution is shown for reference. Due to the normalization of the DSFG-subtracted spectrum, the tSZ amplitude is a factor of 2.2 higher in the subtracted spectrum than at 150 GHz. The primary CMB anisotropy is estimated for a spatially flat, ΛCDM model, which includes gravitational lensing. The bandpower uncertainties are derived from the combination of simulations and the measured intrinsic variations within the SPT data as described in Section 3. The bandpowers can be compared to theory using the associated window functions (Knox 1999). The bandpowers, uncertainties, and window functions may be downloaded from the SPT Web site.25

Figure 4.

Figure 4. SPT 150 GHz (purple circles), 150 × 220 GHz (green diamonds), and 220 GHz (blue triangles) bandpowers. The damping tail of the primary CMB anisotropy is apparent below ℓ = 3000. Above ℓ = 3000, there is a clear excess with an angular dependence consistent with point sources. These sources have low flux (as sources with >6.4 mJy at 150 GHz have been masked) and a rising frequency spectrum, consistent with our expectations for Poisson distributed DSFGs. The point-source population and resulting contributions to anisotropy power are discussed in more detail in H10.

Standard image High-resolution image
Figure 5.

Figure 5. WMAP5 (blue squares), ACBAR (green triangles), QUaD (turquoise diamonds), and SPT (black circles) DSFG-subtracted SPT bandpowers are plotted over the best-fit, lensed ΛCDM cosmological model (dashed red line), best-fit tSZ power-spectrum (solid black line), homogeneous kSZ model (dashed black line), and residual Poisson-distributed point-source contribution (solid orange line). The combined best-fit model is shown by the solid red line. The plotted SPT bandpowers have been multiplied by the best-fit calibration factor of 0.92. Note the entire DSFG-subtracted spectrum has been scaled such that the CMB contribution is calibrated correctly. This scaling means that the tSZ contribution to this spectrum is a factor of 2.2 larger than at 150 GHz. Point sources above 6.4 mJy at 150 GHz have been masked. The patchy kSZ template is also shown for reference (dotted black line).

Standard image High-resolution image

Table 1. Single-frequency Bandpowers

ℓ Range leff 150 GHz 150 × 220 GHz 220 GHz
    q (μK2) σ (μK2) q (μK2) σ (μK2) q (μK2) σ (μK2)
2001–2200 2058 240.5 11.4 269.5 15.3 336.2 30.4
2201–2400 2276 139.4 7.0 155.0 9.8 195.4 20.5
2401–2600 2474 114.3 5.2 119.2 8.1 197.6 18.8
2601–2800 2677 79.0 4.3 114.0 6.7 213.3 19.2
2801–3000 2893 54.1 3.7 77.7 6.1 183.1 19.5
3001–3400 3184 47.0 2.4 76.5 4.2 183.7 13.4
3401–3800 3581 36.9 2.4 75.8 4.3 207.7 15.6
3801–4200 3992 35.0 2.8 82.7 5.0 276.5 18.8
4201–4600 4401 33.9 3.2 87.5 5.9 252.8 21.0
4601–5000 4789 33.6 4.2 99.4 7.3 306.1 24.5
5001–5900 5448 47.1 3.5 135.1 6.3 433.6 21.7
5901–6800 6359 60.5 5.5 158.2 9.4 469.6 33.0
6801–7700 7255 93.5 8.8 195.8 14.5 579.7 46.4
7701–8600 8161 76.8 13.5 221.8 23.6 776.5 73.9
8601–9500 9059 115.6 20.7 299.7 33.7 986.1 105.9

Notes. Band multipole range and weighted value ℓeff, bandpower qB, and uncertainty σB for the 150 GHz auto-spectrum, cross-spectrum, and 220 GHz auto-spectrum of the SPT field. The quoted uncertainties include instrumental noise and the Gaussian sample variance of the primary CMB and the point-source foregrounds. The sample variance of the SZ effect, beam uncertainty, and calibration uncertainty is not included. Beam uncertainties are shown in Figure 1 and calibration uncertainties are quoted in Section 2.2. Point sources above 6.4 mJy at 150 GHz have been masked out in this analysis. This flux cut substantially reduces the contribution of radio sources to the bandpowers, although DSFGs below this threshold contribute significantly to the bandpowers.

Download table as:  ASCIITypeset image

Table 2. DSFG-subtracted Bandpowers

ℓ Range leff q (μK2) σ (μK2)
2001–2200 2058 221.3 16.9
2201–2400 2276 130.2 11.2
2401–2600 2474 126.5 10.3
2601–2800 2677 60.2 7.7
2801–3000 2893 50.4 8.0
3001–3400 3184 36.6 5.9
3401–3800 3581 21.0 6.5
3801–4200 3992 22.9 8.4
4201–4600 4401 8.1 9.5
4601–5000 4789 3.0 11.3
5001–5900 5448 11.2 10.5
5901–6800 6359 16.0 16.2
6801–7700 7255 60.3 27.9
7701–8600 8161 32.1 42.9
8601–9500 9059 54.7 63.8

Notes. Band multipole range and weighted value ℓeff, bandpower qB, and uncertainty σB for the DSFG-subtracted maps of the SPT field. These bandpowers correspond to a linear combination (see Section 3.2.5) of the 150, 150 × 220, and 220 GHz power spectra, optimized to remove emission from DSFGs below the point-source detection threshold of SPT. Point sources above 6.4 mJy at 150 GHz have been masked out in this analysis. The quoted uncertainties include instrumental noise and Gaussian sample variance of the primary CMB and point-source foregrounds. The sample variance of the SZ effect, beam uncertainty, and calibration uncertainty is not included. Beam and calibration uncertainties are quoted in Sections 2.1 and 2.2 and shown in Figure 6.

Download table as:  ASCIITypeset image

The SPT bandpowers are dominated by the damping tail of the primary CMB anisotropy on angular multipoles 2000 < ℓ < 3000. At these multipoles, the bandpowers are in excellent agreement with the predictions of a ΛCDM model determined from CMB observations on larger angular scales. On smaller scales, the SPT bandpowers provide new information on secondary CMB anisotropies and foregrounds which dominate the primary CMB anisotropy. The SPT data presented here represent the first highly significant detection of power at these frequencies and angular scales where the primary CMB anisotropy is sub-dominant. After masking bright point sources, the total S/N on power in excess of the primary CMB are 55, 55, and 45 at 150, 150 × 220, and 220 GHz, respectively. The majority of the high-ℓ power can be attributed to a Poisson distribution of point sources (likely DSFGs) on the sky. H10 use the bandpowers presented in this work to measure both Poisson and clustered contributions to the measured powers and investigate the implications for the properties of DSFGs. The largest source of secondary CMB anisotropy at 150 GHz is expected to be the SZ effect and we investigate SZ constraints in the following sections.

6.2. DSFG-subtracted Bandpowers

Our immediate goal is to measure the amplitude of the tSZ power spectrum. However, several signals in these maps have similar angular power-spectrum shapes and the single-frequency maps only constrain the sum of the power from these sources. For instance, the 150 GHz data effectively constrain the sum of the tSZ, kSZ, and clustered DSFG power.

As discussed earlier, each of these components has a distinct frequency dependence so a linear combination of SPT's two frequency bands can be constructed (following Section 3.2.5) to minimize any one of them. H10 find significant evidence for a clustered DSFG power contribution to the single-frequency bandpowers listed in Table 1 with an amplitude comparable to that of the tSZ effect. We expect the kSZ effect to be smaller than the tSZ effect at 150 GHz on theoretical grounds. Additionally, due to the frequency dependences of the components, removing the kSZ effect would inflate the relative contribution of clustered DSFGs with respect to the tSZ effect. Therefore, we choose to remove DSFGs from the SPT bandpowers.

For a mean DSFG spectral index, α, the proper weighting ratio, x, to apply to the 220 GHz spectrum for DSFG removal would be

Equation (32)

The spectrum in Table 2 and Figures 5 and 6 is produced with a weighting factor of x = 0.325 corresponding to a mean spectral index of α = 3.6. The contribution from DSFGs can be completely removed only if every galaxy has the same spectral index. However, the comparative closeness of SPT's two frequency bands ensures that power leakage into the difference maps remains small despite some expected scatter in the spectral index of the dusty galaxies. In Section 6.2.1, we motivate this choice of x and discuss predictions for the residual point-source power.

Figure 6.

Figure 6. SPT 150 GHz (purple diamonds) and DSFG-subtracted (black circles) bandpowers overplotted on the best-fit models to the DSFG-subtracted bandpowers. The best-fit, lensed ΛCDM cosmological model for the primary CMB anisotropy is shown by the dashed red line, while the sum of the best-fit ΛCDM model, kSZ, tSZ, and point-source terms is shown by the solid red line. The primary CMB anisotropy alone is a poor fit to the SPT data. The uncertainties on the DSFG-subtracted bandpowers are larger for two reasons. First, the normalization convention inflates the uncertainties by a factor of 1/0.6752 and second, these bandpowers include the more noisy 220 GHz data. Beam and calibration uncertainties are marked by a second blue errorbar for the DSFG-subtracted bandpowers only. Note that the calibration and beam uncertainties are correlated between ℓ-bins. The 150 GHz data have been shifted to the right by Δℓ = 40 for clarity. Point sources above 6.4 mJy at 150 GHz have been masked in this analysis.

Standard image High-resolution image

We assume that there is no tSZ contribution to the 220 GHz data as the 220 GHz band is designed to be centered on the SZ null. Fourier transform spectroscopy measurements of the 220 GHz band pass confirm that the tSZ amplitude in the 220 GHz band will be ≲5% of the 150 GHz amplitude. Any error incurred by subtracting roughly one-third of the 220 GHz amplitude from the 150 GHz data would be less than 3%, far below the present ∼40% statistical uncertainty on ASZ (see Table 3).

Table 3. Constraints on ASZ and σ8

  Primary CMB CMB + ASZ
ASZ  ⋅⋅⋅  0.55 ± 0.21
ASZ (w/homogeneous kSZ)  ⋅⋅⋅  0.42 ± 0.21
ASZ (w/patchy kSZ)  ⋅⋅⋅  0.34 ± 0.21
SZ power at ℓ = 3000    
(tSZ + 0.46 × kSZ)  ⋅⋅⋅  4.2 ± 1.5 μK2
kSZ power at ℓ = 3000    
Homogeneous kSZ   2.0 μK2
Patchy kSZ   3.3 μK2
σ8 (no kSZ) 0.795 ± 0.033 0.778 ± 0.024
σ8 (w/homogeneous kSZ) 0.794 ± 0.028 0.773 ± 0.025
σ8 (w/patchy kSZ) 0.788 ± 0.029 0.770 ± 0.024

Notes. The 1σ constraints on σ8 derived from the DSFG-subtracted analysis of the SPT data, when using the simulations in Shaw et al. (2009) to estimate the non-Gaussian cosmic variance of the tSZ power spectrum. The best-fit value for the amplitude of the tSZ power spectrum is also shown, normalized to unity for a WMAP5 cosmology with σ8 = 0.8. We use the thermal SZ model of Sehgal et al. (2010) as our fiducial template. For this model, ASZ = 1 corresponds to a power of 7.5 μK2 at ℓ = 3000. Results are shown with cosmic variance added in quadrature to the statistical uncertainty; however, the ASZ constraints are dominated by statistical uncertainties. Results are shown for no kSZ effect, for a homogeneous model of the kSZ effect (Sehgal et al. 2010) and the homogeneous model with an additional patchy reionization power contribution (Zahn et al. 2005). Finally, the joint constraint on the combined kSZ/tSZ power is shown under the assumption that the two templates are effectively degenerate. For reference, we also quote the power of the two kSZ models considered. The SPT data constrain the combined amplitude of the SZ contributions, which we quote at ℓ = 3000 where the measurements have the most constraining power. The kSZ and tSZ receive different pre-factors in the frequency-differenced analysis due to their relative spectral dependence.

Download table as:  ASCIITypeset image

It is important to note that the apparent tSZ power in the DSFG-subtracted bandpowers will be a factor of 1/(1 − x)2 = 2.2 higher than at 150 GHz as the differenced bandpowers have been normalized to preserve the amplitude of the primary CMB anisotropy. In this work, we report SZ amplitudes scaled to 153 GHz which is the effective band center of the 150 GHz band for a tSZ spectrum.

6.2.1. Residual Point-source Power

The DSFG-subtracted maps have substantially less power due to both unclustered and clustered point sources as seen in Figure 6. However, we include a Poisson point-source amplitude in all fits since we expect a small fraction of the point-source power to remain in the DSFG-subtracted maps. The best-fit amplitude is consistent with zero and unphysical negative values of the Poisson point-source power are allowed due to noise. To prevent this, we place a prior on the residual Poisson point-source power based on what we know about the observed DSFG population from H10 and radio source population from V09 and de Zotti et al. (2005).

The residual power in the subtracted map due to Poisson-distributed DSFGs, CDSFG, depends on the scatter in spectral indices σα, the accuracy to which the mean spectral index $\bar{\alpha }$ is known, and an estimate of the Poisson DSFG power in the 150 GHz band, Cps,150. For a given combination of these parameters, this residual DSFG power will be

Equation (33)

where ν150 and ν220 are the effective bandcenters of the 150 and 220 GHz bands, and x and xtrue are the assumed and true values of map weighting ratio.

We examine the residual Poisson point-source amplitudes for a broad range of weighting ratios to estimate the optimal x value and the error in that estimate. For each x, we estimate the probability that Cps(x) is less than Cps(x = 0.325) using the MCMC chains described in Section 6.3. The resulting probability distribution is taken to be the likelihood function for xtrue. As shown in Figure 7, there is a broad maximum for x = 0.25 to 0.4 and we adopt the best fit, x = 0.325, for the following results. In the absence of a direct measurement, we place a conservative uniform prior on the scatter in DSFG spectral indices, 0.2 < σα < 0.7, as discussed in H10.

Figure 7.

Figure 7. Probability that the residual point-source power in the DSFG-subtracted map constructed by $\bar{m}_{150} - x \bar{m}_{220}$ is lower than the value at x = 0.325 as a function of x. We can interpret this as the probability that a given value of x is the true value. There is a broad maximum centered at 0.325, which corresponds to a spectral index of 3.6 between the 150 and 220 GHz bands. This is consistent with the ratio of point-source power between single-frequency fits to the 150, 150 × 220, and 220 GHz bandpowers. We estimate the x which minimizes residual point-source power to be 0.325 ± 0.08.

Standard image High-resolution image

Our expectation for the residual radio contribution to the DSFG-subtracted bandpowers, Cradio, is based on the de Zotti et al. (2005) radio source count model. This model is in excellent agreement on the high flux end with the SPT source counts (V09). As discussed in H10, the residual radio source power after masking is expected to be a small fraction of the DSFG power at 150 GHz. However, this small radio contribution may be comparable to the residual DSFG power in the DSFG-subtracted spectrum. The radio source power can be calculated from the integral of S2dN/dS for the de Zotti et al. (2005) counts model from zero to the flux masking threshold of 6.4 mJy. We compute the power these sources contribute to the optimal DSFG-subtracted spectrum to be 3.9 × 10−7 μK2 by assuming an average spectral index of α = −0.5 based on the detected sources in V09. This power level is nearly identical to that predicted by the Sehgal et al. (2010) simulations. To allow for a variation in spectral index as well as uncertainty in the model normalization when extended to lower flux sources, we assign a conservative uncertainty of 50% on the predicted residual radio source power in the prior.

We combine this information to create a prior on the residual point-source power in the DSFG-subtracted maps Cps = CDSFG + Cradio. This prior spans the range Cps∈[3.5, 9.0]×10−7 μK2 at 68% confidence and [1.2, 13.9]×10−7 μK2 at 95% confidence. The best-fit value of the residual Poisson component before applying the prior is Cps(no prior) = (6.2 ± 6.4) × 10−7 μK2 and lies at the middle of our assumed prior range. The upper end of the 95% range is approximately 20% of the best-fit value of the Poisson point-source power in the undifferenced 150 GHz bandpowers. This suggests that we have subtracted over 80% of the point-source power from the 150 GHz spectrum, with the residual point-source power largely from radio sources. Without this prior on the Poisson point-source amplitude, the uncertainty on the ASZ detection presented in the next section would increase by ∼50%.

6.2.2. Residual Clustered Point-source Power

We assume that the contribution of clustered point sources is insignificant in the DSFG-subtracted bandpowers. Using a combination of the SPT bandpowers at 150, 220, and 150 × 220 GHz as well as the DSFG-subtracted spectrum, H10 argue that the residual clustered DSFG component in the DSFG-subtracted bandpowers is less than 0.3 μK2 at 95% confidence. This is several percent of the SZ power spectrum but negligible at the current detection significance of ≲3 σ. We also argue in Section 5 that clustered radio sources are negligible. Therefore, residual power from clustered point sources will not bias SZ constraints from the DSFG-subtracted bandpowers.

The above argument holds if the point sources are uncorrelated with the SZ signal. However, if the clustered term was completely anti-correlated with the SZ signal, the measured SZ power in Table 3 could underestimate the true SZ power by 38%. This is unlikely for two reasons. First, the residual after DSFG subtraction should be uncorrelated as long as the spectral dependence of cluster member DSFGs is similar to the general DSFG population. Second, as argued in Section 5, most DSFGs are not galaxy cluster members. We also look at the correlations between the Sehgal et al. (2010) DSFG and tSZ simulated sky maps. The Sehgal et al. (2010) DSFG model scales the number density of DSFG cluster members linearly with cluster mass; this is a substantially stronger scaling than observed (Bai et al. 2007). Therefore, estimating the cluster–DSFG correlation from Sehgal et al. (2010) should be overly conservative. We calculate the anti-correlation coefficient between the Sehgal et al. (2010) simulated tSZ maps at 148 GHz and a linear combination of the 148 and 219 GHz simulated IR source maps with the same weighting as used for the SPT DSFG-subtracted bandpowers. We find the anti-correlation coefficient between the tSZ effect and total DSFG power to be 21% using the power measured at ℓ = 3000. This highly conservative upper limit of 21% on the anti-correlation implies that the true SZ power is underestimated by less than 4%. Based on these arguments, we assume that correlations between SZ signals and emission from cluster member galaxies are negligible in this analysis.

Separately, one might worry about correlations between radio sources and SZ clusters. Radio sources could suppress the tSZ signal by "filling in" the tSZ decrements. However, the work by Lin et al. (2009) and Sehgal et al. (2010) shows that the number of cluster-correlated radio sources is expected to be small. For instance, we can examine this correlation in the simulated tSZ and radio source maps produced by Sehgal et al. (2010). We look at the power spectrum at ℓ = 3000 for the tSZ map, the tSZ+radio-source map, and the radio source only map after masking sources above 6.4 mJy. We find an anti-correlation coefficient of 2.3% for the two components. Given the expected radio source power level of ∼0.6 μK2 at ℓ = 3000 (Section 6.2.1), radio source–cluster correlations should not affect the results in this work.

6.3. Markov Chain Analysis

The DSFG-subtracted bandpowers presented in Section 6.2 detect at high significance a combination of the primary CMB anisotropy, secondary SZ anisotropies, and residual point sources. In this section, we use an MCMC analysis to separate these three components and to produce an unbiased measurement of the tSZ power-spectrum amplitude.

6.3.1. Elements of the MCMC Analysis

We fit the DSFG-subtracted bandpowers to a model including the lensed primary CMB anisotropy, secondary tSZ and kSZ anisotropies, and a residual Poisson point-source term. We use the standard, six-parameter, spatially flat, lensed ΛCDM cosmological model to predict the primary CMB temperature anisotropy. The six parameters are the baryon density Ωb, the density of cold dark matter Ωc, the optical depth to recombination τ, the angular scale of the peaks Θ, the amplitude of the primordial density fluctuations ln [1010As], and the scalar spectral index ns. To fit the high-ℓ power, we extend the basic, six-parameter model with two additional parameters: the amplitude of a tSZ power-spectrum template, ASZ, and a constant, Cps, such as would be produced by a Poisson distribution of point sources on the sky. We also explore the potential impact of the kSZ effect on these parameters by using three different kSZ models.

Gravitational lensing of CMB anisotropy by large-scale structure tends to increase the power at small angular scales, with the potential to influence an SZ power-spectrum measurement. The calculation of lensed CMB spectra out to ℓ = 10,000 proved prohibitively expensive in computational time. We avoid this computational limitation by calculating the lensing contribution for the best-fit cosmological model and adding this estimated lensing contribution to the unlensed CMB power spectrum calculated at each step in the Markov chain. Given the small allowed range of Ωm with current CMB data, we predict that using the fixed lensing contribution will misestimate the actual lensing by less than 30%. We have checked this assertion on a sampling of parameter sets drawn from the chains. The lensing contribution to the high-ℓ spectrum is ∼1.5 μK2, while the modeled tSZ spectrum for the same best-fit WMAP5 cosmology averages ∼8.6 μK2 near ℓ = 3000 where SPT has the highest S/N on the SZ spectrum. In the differenced spectra used to derive the ASZ constraints (see Section 6.2), the ratio of lensing to the tSZ effect is suppressed by a factor of (1 − x)2 = 0.46. This reduction occurs because the subtracted spectrum is normalized such that the CMB power is unaffected, though the SZ power is enhanced. As a result, we expect possible lensing misestimates to introduce a negligible error of ≲3% on the ASZ constraints. Of course, this error will be larger for smaller values of ASZ.

The tSZ template we use as a fiducial model is based on simulations by Sehgal et al. (2010).26 The simulations are for a WMAP5 cosmology with σ8 = 0.80 and Ωbh = 0.0312 and an observing frequency of 148 GHz. We rescale the template to 153 GHz, the effective center frequency of the SPT 150 GHz band for signals with a tSZ spectrum.

For a limited number of cases, we also compare the results of the Sehgal numerical tSZ template to those of the Komatsu & Seljak analytic template with WMAP5 parameters27 (Komatsu & Seljak 2002) and the numerical template by Shaw et al. (2009). The primary difference between the Shaw and Sehgal simulations is the value of the energy feedback parameter in the Bode et al. (2007) intracluster gas model. The Shaw simulations have the feedback parameter reduced to 60% of the Sehgal value. Increasing the feedback parameter causes the gas distribution in clusters to "puff-out" or inflate; in low-mass clusters the gas may become unbound altogether. The overall effect is to reduce the predicted tSZ power spectrum, especially at small angular scales. We limit our analysis of specific tSZ models to these three models.

An independent measurement of the kSZ spectrum is outside the scope of this work, however, it is necessary to take the kSZ effect into account when inferring the tSZ amplitude from the data. We consider three kSZ cases based on two published kSZ models. The kSZ effect is assumed to be zero in the first case ("no kSZ"). As a second, intermediate case, we use the model by Sehgal et al. (2010, referred to as "homogeneous kSZ"). This model includes kSZ contributions from a homogeneous reionization scenario, but does not include the additional kSZ power produced by patchy reionization scenarios. We take this case to be the fiducial kSZ model. We include an estimate of patchy reionization in the third kSZ case. For the patchy reionization phase, we use the "brief history" model B from Zahn et al. (2005), re-calculated for WMAP5 best-fit cosmological parameters. The sum of the homogeneous kSZ model and the patchy contribution will be referred to as the "patchy kSZ" model. We do not scale these templates for different cosmological parameter sets as we expect the kSZ theoretical uncertainty to be at least as large as the cosmological dependence. Finally, we include a residual Poisson point-source component in all chains. The construction of the point-source prior is outlined in Section 6.2.1.

Previous CMB experiments have produced exceptional constraints on the primary CMB anisotropy and we use the bandpowers from WMAP5 (Dunkley et al. 2009), ACBAR (Reichardt et al. 2009a), and QUaD (Brown et al. 2009) at ℓ < 2200 in all parameter fitting. We refer to this collection along with the SPT DSFG-subtracted bandpowers as the "CMBall" data set. It is important to note that the two-parameter extension to the ΛCDM model for point sources and the tSZ effect is only restricted to the SPT bandpowers. This restriction is imposed for two reasons. First, the point-source contributions to each experiment may be different due to the different frequencies and flux cuts for masking sources. Arguably, the point-source power is likely to be similar in the 150 GHz SPT, ACBAR, and QUaD results, but we would be unable to use frequency information to discriminate between the SZ effect and clustered DSFGs. Second, the primary CMB is dominant below ℓ ∼ 3000 and the other experiments lack sufficient statistical weight at high-ℓ to improve upon SPT's measurement of SZ effect and point-source power.

Parameter estimation is performed by MCMC sampling of the full multi-dimensional parameter space using an extension of the CosmoMC package (Lewis & Bridle 2002). We include the code extension produced by the QUaD collaboration (Brown et al. 2009) to handle uncertainties on a non-Gaussian beam in CosmoMC. CMB power spectra for a given parameter set are calculated with CAMB (Lewis et al. 2000). We use the WMAP5 likelihood code publicly available at http://lambda.gsfc.nasa.gov. After the burn-in period, each set of four chains is run until the largest eigenvalue of the Gelman–Rubin test is smaller than 0.0005. Wide uniform priors are used on all six parameters of the ΛCDM model. A weak prior on the age of the universe (t0 ∈ [10, 20] Gyr) and Hubble constant (h ∈ [0.4, 1]) is included in all chains, but should not affect the results. We use a uniform prior on ASZ over a wide range from −1 to 10 times the value expected for σ8 = 0.80.

6.3.2. Constraints on SZ Amplitude

We fit for the normalization factor of a fixed tSZ template and, in Table 3, report both this template-specific normalization, ASZ, and the total inferred SZ-power at ℓ = 3000, near the multipole with maximum tSZ detection significance. This estimate of the SZ power includes both tSZ and kSZ terms, and is included to facilitate comparison with other SZ models. We expect both the thermal and the kinetic SZ spectra to vary slowly with angular multipole.

The chains are run for three different assumptions about the kSZ effect: the no kSZ, homogeneous kSZ, and patchy kSZ models described above. In each case, we produce MCMC chains with a fixed kSZ amplitude. The χ2 of the 15 SPT bandpowers is between 16.3 and 16.4 for the best fits of the three kSZ cases considered; there is essentially no impact on the quality of the fit. The ASZ and power constraints for each case are listed in Table 3 and plotted in Figure 8. Shifting the frequency-differencing factor, x, by ±0.05 changes the resultant best-fit ASZ by less than 0.5σ.

Figure 8.

Figure 8. One-dimensional marginalized ASZ constraints from the SPT DSFG-subtracted bandpowers. Three kSZ cases are considered: no kSZ effect (dashed line), the homogeneous kSZ model (solid line), and the homogeneous model plus a patchy reionization term (dotted line). These models are described more fully in Section 6.3.1. Top axis: the corresponding tSZ power at ℓ = 3000 for reference. The no-kSZ curve (dashed line) can be interpreted as a constraint on the sum of DtSZ3000 + 0.46 × DkSZ.

Standard image High-resolution image

The bandpower uncertainties in Table 2 do not include the sample variance of the tSZ effect, and we must convolve the ASZ distribution in the chains with an estimate of the sample variance to find the true ASZ likelihood function. The sample variance of the tSZ effect is estimated using the simulations of Shaw et al. (2009) re-run with the same intracluster gas model parameters as Sehgal et al. (2010). The simulation consists of 300 map realizations of the size of the SPT sky patch. These are constructed from a base sample of 40 independent maps by separating the components of each map into eight redshift bins (between 0 ⩽ z ⩽ 3) and shuffling these bins between maps to generate a larger sample. We take the spectrum of each realization and find the best-fit ASZ amplitude after weighting ℓ-bins based on the SPT bandpower uncertainties. We use the distribution of these amplitudes to map out the likelihood function for the tSZ power (see Figure 9). Due to the number of independent realizations, we have limited ability to resolve the tail of the likelihood function. Fortunately, the non-Gaussianity is small for such a large sky area and the distribution of ln(ASZ) is fit well by a Gaussian with a 12% width. We use this Gaussian fit as an estimate of the full likelihood surface. Small deviations from the true description of the tSZ sample variance will not impact the final results as the sample variance is small compared to both the statistical uncertainties and the assumed 50% model uncertainty. The uncertainties on ASZ are essentially unchanged by the inclusion of the tSZ sample variance. The sample variance and model uncertainty are shown in Figure 9.

Figure 9.

Figure 9. Sample variance and assumed theoretical uncertainty on the tSZ amplitude. The histogram shows the sample variance of ASZ/〈ASZ〉, where 〈ASZ〉 is the mean value measured over a sample of 300 simulated thermal SZ maps (see Section 6.3.2 for details). The overlying dashed black line shows the lognormal fit to the distribution, i.e., a Gaussian fit to ln(ASZ)), with σln A = 0.12. The solid black line is the assumed 50% theoretical uncertainty on ASZ, which we model as a Gaussian distribution in ln(ASZ).

Standard image High-resolution image

As discussed earlier, the DSFG-subtracted bandpowers are sensitive to a linear combination of the tSZ and kSZ effects. We expect analysis of 2009 and later SPT data which include 95 GHz data to be able to separate the two SZ effects. The linear combination is not a simple sum, as the frequency differencing used to produce the DSFG-subtracted spectrum suppresses the kSZ relative to the tSZ by a factor of (1 − x)2 = 0.46. This factor is uncertain at the 15% level due to the relative calibration uncertainty between the bands. The SPT data detect the combined SZ effect at 2.6σ with tSZ + 0.46 × kSZ = 4.2 ±  1.5 μK2 at ℓ = 3000. Using this combined constraint implicitly assumes that the tSZ and kSZ templates are perfectly degenerate, which is a good assumption for the current data quality and templates used in this work.

We can compare the power detected with SPT to that reported by the CBI Collaboration (Sievers et al. 2009). We use the best-fit normalization of a Komatsu & Seljak template for WMAP5 parameters to compare directly the results of the two experiments. There are sub-percent differences in the assumed values of σ8 and Ωbh between the template we adopted here and that used in Sievers et al. (2009), which would change the amplitude of the template by ∼1% for an assumed scaling of σ78bh)2. This effect is negligible. We find the best-fit normalization of the WMAP5 Komatsu & Seljak model to be 0.37 ± 0.17 for the SPT data under the homogeneous kSZ scenario. This is 2.4σ below the best-fit CBI normalization of 3.5 ± 1.3. The model includes the frequency dependence of the tSZ effect. The smaller SPT bandpowers suggest that the CBI excess power may be produced by foregrounds with a frequency dependence falling more steeply than the SZ effect such as radio sources.

6.4. Implications of the ASZ Measurement

The best-fit normalization for the fiducial tSZ spectrum, ASZ, is significantly lower than unity. The cosmological parameters assumed when generating this template may be slightly different than the best-fit models, so we scale the template to match the cosmologies explored by the Markov chain. When these scalings are taken into account, the measured values of ASZ are still low. The tSZ power spectrum depends on the details of how the baryon intracluster gas populates dark matter halos and on cosmology through the number density of these halos. The paucity of tSZ power may reflect an overestimate of the intracluster gas pressure by the fiducial model and so we compare the template to other SZ models. At the same time, this low value of ASZ favors a shift in the derived cosmological parameters, particularly σ8.

Even a ∼2.5σ detection of ASZ will produce cosmologically interesting constraints on σ8, as we expect ASZ to scale strongly with σ8 and less strongly with the baryon density. The scaling is approximately σγ8bh)2 where 7 < γ < 9, depending on the exact cosmology (Komatsu & Kitayama 1999; Komatsu & Seljak 2002). We explore the σ8 dependence under the Press–Schechter halo model and find that this relationship steepens with the currently favored lower values of σ8. Sampling cosmological parameter values from the WMAP5 MCMC chains, we find that the modeled amplitude of the tSZ power spectrum varies roughly as σβ8, with β ≈ 11. This sampling reflects the degeneracies between parameters, and so the value of this exponent, β, includes the degeneracy between σ8 and Ωmh. Constraining the amplitude of the tSZ effect offers an independent measurement of σ8 that can be compared to measurements based on primary CMB anisotropy or large-scale structure. Such comparisons test our understanding of the physical processes involved in structure formation.

For each point in the MCMC chain, we calculate the predicted ASZ value from the six basic cosmological parameters. For this calculation, we use the mass function of Jenkins et al. (2001) to determine the abundance of galaxy clusters of a given mass. We then use the mass–concentration relation of Duffy et al. (2008) to determine the dark matter halo properties, and the gas model used in Komatsu & Seljak (2002) to estimate the tSZ signal for each halo according to its mass. In order to convert this analytic tSZ spectrum into an amplitude, we take an ℓ-weighted average designed to match the relative weights each multipole receives in the real tSZ fits. This amplitude is normalized to unity for the cosmological parameters assumed in the fiducial tSZ model. We also allowed the mass–concentration index to vary with cosmology by appropriate scaling of the characteristic mass M*, but this was found to be a negligible effect within the explored range in parameters.

At each point in the chain, the measured tSZ amplitude is compared to the predicted tSZ amplitude to construct a tSZ scaling factor, ASZ/AtheorySZ. This procedure will account for any correlations between the measured ASZ parameter and the six ΛCDM parameters in a self-consistent fashion, although we do not see evidence for such correlations in the current data. The distribution of scaling factors versus σ8 is illustrated by the black contours in Figure 10.

Figure 10.

Figure 10. Two-dimensional likelihood contours at 68% and 95% confidence for σ8 vs. the tSZ scaling factor, ASZ/AtheorySZ, derived from the SPT DSFG-subtracted bandpowers. For each point in the Markov chain, the tSZ scaling factor compares the ASZ value fit to the SPT data to the AtheorySZ value predicted for that point's ΛCDM model parameters (see Section 6.4). The black contours show the likelihood surface for the CMBall data set. We observe no dependence between ASZ and the six parameters of the ΛCDM model. The tilt toward higher scaling factors at lower σ8 is expected since the predicted AtheorySZ depends steeply on the value of σ8. The black contours also do not account for the cosmic variance of ASZ; without cosmic variance or uncertainty in modeling the tSZ power, the tSZ scaling factor would be constrained to be exactly unity (solid orange line). The red shaded regions about unity illustrate the uncertainty we assume for the tSZ scaling factor due to theoretical uncertainty and cosmic variance (also see Figure 9). This uncertainty is modeled as a lognormal distribution. The measured value of the tSZ scaling relation, including the theoretical uncertainty and sample variance in the model, is used to importance sample the Markov chain and obtain the likelihood surface marked by the blue contours. Left panel: likelihood surfaces assuming no kSZ contribution. Right panel: likelihood surfaces assuming the patchy kSZ model. The constraints for the homogeneous kSZ model will lie between the results for these two cases.

Standard image High-resolution image

In general, the tSZ scaling factors are less than unity. These low scaling factors suggest either an overestimate of the tSZ effect or lower values of σ8. Models predicting larger kSZ or tSZ effects lead to lower scaling factors. However, the results cannot be purely explained by an overestimate of the kSZ effect since this tension persists in the no-kSZ case. Alternatively, the scaling factors may indicate that the explored range in cosmological parameters is systematically overestimating the rms of the mass distribution. For instance, as we see in Figure 10, points of the chain with lower values of σ8 have scaling factors closer to unity.

The first interpretation of the low tSZ scaling factor is that the Sehgal tSZ template overestimates the tSZ power spectrum. There is currently some degree of uncertainty in the expected shape and amplitude of the tSZ power spectrum as predicted by analytic models or hydrodynamical simulations. One reason for this is that cosmological simulations of the intracluster medium have only recently begun to investigate in detail the impact of radiative cooling, non-gravitational heating sources (such as AGNs), and possible regulatory mechanisms between them. The computational expense of running hydrodynamical simulations with sufficient resolution to resolve small-scale processes (such as star formation) while encompassing a large enough volume to adequately sample the halo mass function is prohibitive to a detailed analysis of the predicted SZ power spectrum. In order to accurately predict the tSZ power spectrum, it is especially important to correctly model the gas temperature and density distribution in low-mass (M < 2 × 1014h−1M) and high-redshift (z>1) clusters, which contribute significantly to the power spectrum at the angular scales where SPT is most sensitive (Komatsu & Seljak 2002).

In Figure 11, we plot the tSZ power spectrum derived from several different simulations (note that all curves have been normalized to the fiducial cosmology). The thick black solid line shows the base template, obtained from maps generated by Sehgal et al. (2010). The black dot-dashed line shows the power spectrum obtained from the simulations of Shaw et al. (2009), which have a lower energy feedback parameter than the Sehgal et al. (2010) simulations. Increasing the amount of feedback energy has the effect of inflating the gas distribution and suppressing power at small angular scales. The red solid and dot-dashed lines show the power spectrum from maps constructed from a 240 h−1 Mpc box simulation run using the Eulerian hydrodynamics code ART (Kravtsov et al. 2005; D. Rudd 2009, private communication) and from the MareNostrum simulation (O. Zahn et al. 2010, in preparation), respectively. Both of these simulations were run in the adiabatic regime, i.e., no cooling, star formation, or feedback. The gas distribution is more centrally concentrated than in the Shaw or Sehgal models, producing tSZ power spectra with significantly more power at small angular scales and less at larger angular scales. The blue solid line shows the analytic template predicted by the Komatsu & Seljak (2002) halo model calculation. The histogram shows the SPT sensitivity to the thermal SZ signal in each ℓ band (with arbitrary normalization).

Figure 11.

Figure 11. Comparison of the tSZ power spectrum (at 153 GHz) as predicted by numerical simulations and halo model calculations. Note that all curves have been renormalized to the fiducial cosmology. The thick, black, solid line shows the base template obtained from maps generated by Sehgal et al. (2010). The black, dot-dashed line shows the modeled spectrum obtained from maps generated by applying the semi-analytic model for intracluster gas of Bode et al. (2007) to the halos identified in an N-body lightcone simulation (as described in Shaw et al. 2009). The red, solid line shows the results from maps constructed from an adiabatic simulation produced using the Eulerian hydrodynamical code ART (Kravtsov et al. 2005), while the red, dot-dashed line shows results from maps made from the MareNostrum simulations (O. Zahn et al. 2010, in preparation). The blue, solid line shows the predictions of the Komatsu & Seljak (2002) halo model calculation. The histogram shows the SPT sensitivity to the thermal SZ signal in each ℓ-band (with arbitrary normalization) The gray band illustrates the 68% confidence interval of our theoretical uncertainty (see Figure 9).

Standard image High-resolution image

The current data are in tension with even the high-feedback simulations for the CMB-derived best-fit cosmological parameter set. Even with the kSZ effect set to zero, the tSZ scaling factor is only 0.55 ± 0.21 of what is predicted for the fiducial WMAP5 cosmology. Meanwhile, the tSZ scaling factors are 0.42 ± 0.21 for the homogeneous kSZ model and 0.34 ± 0.21 for the patchy kSZ model. The tension grows worse for the medium-feedback simulations by Shaw et al. (2009). Although the Bode et al. (2007) model is calibrated to reproduce observed X-ray scaling relations for high-mass, low-redshift clusters, it may significantly overestimate the contribution of low-mass or high-redshift clusters (for which there are few direct X-ray observations with which to compare).

As mentioned above, an alternate interpretation of the low tSZ scaling factor is that the CMBall parameter chains without SZ constraints overestimate σ8. In order to constrain σ8 based on the measurement of ASZ, we need to construct the likelihood of observing ASZ for a given cosmological parameter set. At a minimum, this likelihood would reflect the 12% uncertainty due to sample variance of the tSZ effect. However given the large range of tSZ predictions, we add an additional theory uncertainty in quadrature with the sample variance. The gray region in Figure 11 shows the 1σ region encompassed by a lognormal distribution of width σAsz = 0.5 around the fiducial model (black solid line). In the range where SPT is most sensitive (2000 ⩽ ℓ ⩽ 6000), all the predicted curves lie within this region. In order to account for the range of predicted power spectra, we adopt a 50% theory uncertainty on the value of ln(Asz) in the likelihood calculation. The resulting σ8 constraints are dominated by the theory uncertainty. With this prior, we construct a new chain from the original parameter space through importance sampling. The regions preferred by this prior are shown in Figure 10, as are the results of the new Markov chain.

Constraints on σ8 with and without including the SPT ASZ measurements are shown in Table 3 and Figure 12. Under the assumption of the homogeneous kSZ model and a 50% theoretical uncertainty in the amplitude of the tSZ power spectrum, the addition of the SPT data slightly tightens the constraint on σ8, while reducing the central value from σ8 = 0.794 ± 0.028 to 0.773 ± 0.025. The uncertainty in the resulting constraint on σ8 is dominated by the large theoretical uncertainty in the tSZ amplitude. If instead we assume that the fiducial tSZ model is perfectly accurate and do not account for model uncertainty, the uncertainty on σ8 is reduced by 30% and the preferred value is significantly reduced to σ8 = 0.746 ± 0.017 for the homogenous kSZ model. Despite the fact that the adopted template is the lowest of the tSZ models shown in Figure 11, the Sehgal et al. (2010) template, the value of σ8 inferred by the SPT data using this model is lower than that favored by WMAP. Additional SPT data will soon determine if the apparent tension between σ8 inferred from SZ and primordial CMB measurements is robust. In any case, improving our theoretical understanding of both the kSZ and tSZ power spectra is essential for fully realizing the potential of SZ power-spectrum measurements to constrain cosmological parameters such as σ8.

Figure 12.

Figure 12. One-dimensional marginalized σ8 constraints with and without including the SPT DSFG-subtracted bandpowers for three kSZ cases. The black lines denote the σ8 constraints without SPT, while the red lines include SPT's bandpowers. Constraints with the patchy kSZ template are shown with a dotted line. The results when including only the homogeneous kSZ model are shown with the solid lines and the results for no kSZ effect are shown with dashed lines. The SPT data tighten the σ8 constraint in all three cases.

Standard image High-resolution image

7. CONCLUSIONS

We have presented the first CMB temperature anisotropy power-spectrum results from the SPT experiment. Table 1 contains bandpowers based on simultaneous observations at 150 and 220 GHz of a single 100 deg2 field for three months during the austral winter of 2008. A total of 286k detector hours of data was collected during these observations, resulting in a final map depth of 18 μK arcmin at 150 GHz and 40 μK arcmin at 220 GHz. The SPT 150 GHz temperature maps are calibrated to an accuracy of 3.6% through a direct comparison of CMB temperature anisotropy as observed by SPT and WMAP5. This absolute temperature calibration is extended to 220 GHz with a 7.2% uncertainty by an internal comparison of the primary CMB anisotropy measured by SPT at 150 and 220 GHz.

The SPT bandpowers include angular multipoles from ℓ = 2000 to 9500. This broad angular range covers the transition from where the primary CMB anisotropy dominates the measured power to where secondary CMB anisotropies and foregrounds become dominant. The primary anisotropy observed by SPT is consistent with ΛCDM cosmological models based on lower-ℓ CMB data. At ℓ>3000, the SPT bandpowers are used to detect an excess of power above the primary anisotropies at ∼50σ at both 150 and 220 GHz after masking bright (>6.4 mJy) sources.

Most of this power can be explained by a population of faint DSFGs, the properties of which are discussed in H10. We expect the DSFG population to contain a clustered component, which will produce a power term of comparable size to the tSZ effect at 150 GHz. This clustered component would bias the measurement of power from secondary CMB anisotropies in single-frequency maps. However, we use SPT's multifrequency data to construct the DSFG-subtracted bandpowers shown in Figure 6. This frequency differencing reduces the clustered point-source power to negligible levels and significantly reduces the Poisson power.

Secondary CMB anisotropies, consisting of a linear combination of the kSZ and tSZ effects, are detected at 2.6σ in an analysis of the DSFG-subtracted bandpowers. Assuming there is no kSZ contribution, this equates to 4.2 ± 1.5 μK2 of tSZ power at ℓ = 3000. This is substantially lower than the 8.5 μK2 predicted by the homogeneous kSZ and fiducial tSZ models using WMAP5 cosmological parameters. The lower than expected value of the measured SZ power may point to either lower values of σ8 or an overestimate of the modeled SZ power spectrum. Adopting a conservative prior on the uncertainty in the model predictions for the SZ amplitude, we combine the measurement of the tSZ power spectrum with CMB data on larger angular scales to improve the constraint on the CMB-derived matter power-spectrum normalization, σ8. This results in a decrease in the preferred value to σ8 = 0.773 ± 0.025. A major source of uncertainty in this analysis is due to the current theoretical uncertainty on the amplitudes and shapes of the thermal and kinetic SZ effect power spectra. In the absence of this theoretical uncertainty, the constraint tightens about a still lower value of σ8 = 0.746 ± 0.017. The tension between the predicted and measured SZ amplitudes decreases for models that predict less kSZ or tSZ power, for instance models with shorter reionization epochs or higher energy feedback in clusters. However, the tension remains at some level for all the models we considered. It is our hope that the progress in observational efforts represented by this work will lead to further work in improving theoretical models for secondary CMB anisotropies.

The SPT power-spectrum results presented here use only a small fraction of the complete data set. The SPT is continuing to take data and has already observed 800 deg2 to similar depth at 150 and 220 GHz to the 100 deg2 used in this work, and 600 deg2 of this area also has 95 GHz coverage. The complete SPT survey is expected to cover over 2000 deg2 with these three frequencies. The complete data set promises a >10σ detection of the tSZ power spectrum, and the addition of the third (95 GHz) frequency band will allow us to separate the kSZ effect from DSFGs and the tSZ effect. Measuring the kSZ power spectrum will open a new window onto the reionization history of the universe. The complete SPT data set will also produce a catalog of hundreds of SZ-selected clusters, high S/N images of known clusters, a direct detection of gravitational lensing of the CMB, and the widest study to-date on the resolved and unresolved source populations at these frequencies.

The SPT team gratefully acknowledges the contributions to the design and construction of the telescope by S. Busetti, E. Chauvin, T. Hughes, P. Huntley, and E. Nichols and his team of iron workers. We also thank the National Science Foundation (NSF) Office of Polar Programs, the United States Antarctic Program, and the Raytheon Polar Services Company for their support of the project. We are grateful for professional support from the staff of the South Pole station. We thank T. M. Lanting, J. Leong, A. Loehr, W. Lu, M. Runyan, D. Schwan, M. Sharp, and C. Greer for their early contributions to the SPT project, J. Joseph and C. Vu for their contributions to the electronics, and P. Ralph for his useful discussions and insights.

The SPT is supported by the National Science Foundation through grants ANT-0638937 and ANT-0130612. Partial support is also provided by the NSF Physics Frontier Center grant PHY-0114422 to the Kavli Institute of Cosmological Physics at the University of Chicago, the Kavli Foundation, and the Gordon and Betty Moore Foundation. The McGill group acknowledges funding from the National Sciences and Engineering Research Council of Canada, the Quebec Fonds de recherche sur la nature et les technologies, and the Canadian Institute for Advanced Research. The following individuals acknowledge additional support: K.K.S. and B.A.B. from a KICP Fellowship, O.Z. from a Berkeley Center for Cosmological Physics Fellowship, J.J.M. from a Fermi Fellowship, Z.S. from a GAAN Fellowship, and A.T.L. from the Miller Institute for Basic Research in Science, University of California Berkeley. N.W.H. acknowledges support from an Alfred P. Sloan Research Fellowship.

This research used resources of the National Energy Research Scientific Computing Center, which is supported by the Office of Science of the U.S. Department of Energy under Contract no. DE-AC02-05CH11231. Some of the results in this paper have been derived using the HEALPix (Górski et al. 2005) package. We acknowledge the use of the Legacy Archive for Microwave Background Data Analysis (LAMBDA). Support for LAMBDA is provided by the NASA Office of Space Science.

APPENDIX: BANDPOWER COVARIANCE MATRIX ESTIMATION

The bandpower covariance matrix includes both signal and noise contributions. The signal covariance is calculated from simulations. The noise covariance is estimated from the data. We calculate the variance of the mean power spectrum using the variance of cross-spectra between independent real maps. With 300 independent observations of the same field, these maps are sufficient to generate an accurate estimate of the covariance matrix. Several details of our approach are motivated by the analytical treatments of Tristram et al. (2005) and Polenta et al. (2005), and we first review the analytical estimate of the covariance matrix before discussing the estimator.

A.1. Analytical Considerations

Following Tristram et al. (2005), we represent the expected covariance between two cross-spectra as Ξ:

Equation (A1)

Equation (A2)

Our goal is to express this covariance in terms of the noise and signal in the maps, and to compute the magnitude of the diagonal elements, as well as the correlation between bandpowers. As the first step, the central term can be rewritten as

Equation (A3)

To simplify Equation (A3), Tristram et al. (2005) and Polenta et al. (2005) make the following assumptions.

  • 1.  
    Fluctuations in the map, i.e., from CMB anisotropies, confusion limited point sources and noise, are well described by a Gaussian random field.
  • 2.  
    The beams and filtering applied to the data are isotropic; EkGkBk depends only on $|\textbf {k}|$.
  • 3.  
    The instrumental noise is isotropic.
  • 4.  
    The power spectrum, Ck, is smoothly varying with k, and changes little over scales comparable to the width of the mode coupling matrix.

Using the assumption that Ck and |Ek| do not vary much over small changes in $\textbf {k}$, these products become

Equation (A4)

Here CABk is shorthand for $C_k+\frac{N^{A}_k}{B_k^2}\delta _{\it AB}$, the expected cross-spectrum between two unfiltered, perfectly beam-corrected—though noisy—maps. The additional term is the noise bias that exists in the map auto-spectrum. Assuming isotropic beams and filtering, we can combine Equations (A3) and (A4) to obtain a relatively simple expression:

Equation (A5)

Combining Equations (A3) and (A5) yields

Equation (A6)

To obtain a simplified expression for the magnitude of the diagonal elements of the covariance matrix, one typically assumes the mode-coupling matrix is nearly diagonal: $M_{kk^\prime }[\textbf {W}]\approx w_2 \delta _{kk^\prime }$. In this approximation $\left(K[\textbf {W}]^{-1}\right)_{bb^\prime }\approx w_2^{-1}E_b^{-2}\delta _{bb^\prime }$ and $M_{kk^\prime }[\textbf {W}^2]\approx w_4 \delta _{kk^\prime }$ and thus the covariance of any two cross spectra is

Equation (A7)

where νb is the effective number of independent k-modes in each ℓ-band. For isotropic filtering, $\nu ^{\rm {iso}}_b=\frac{N_b w_2^2}{w_4}$.

One subtlety in estimating the covariance is the fact that although the noise in each map is independent, each map has the same sky coverage. Hence the signal in all maps is correlated. The correct estimator must take this correlation into account. Under the simplifying assumption that all maps are statistically equivalent, the correlation between two cross-spectra depends only on whether the spectra have a map in common (e.g., the cross-spectrum D12b is more strongly correlated to the spectrum D23b than D34b). Comparing the covariance of two cross-spectra taken from four different observations:

Equation (A8)

to the covariance of a pair of cross-spectra with a common map,

Equation (A9)

to the covariance of a pair of cross-spectra with two maps in common,

Equation (A10)

The degree of correlation is also ℓ dependent since it depends on the relative signal versus noise power in the maps. In the high-ℓ regime, where noise dominates the power in an individual map, all cross-spectra are nearly independent. Conversely all cross-spectra are nearly completely correlated at low-ℓ, where the primary CMB anisotropy overwhelms the noise.

Given the assumption of statistical equivalence, we can then compute the expected variance of the mean spectrum based on these variance estimates. This is equal to the correlation between any particular cross-spectrum and the mean,

Equation (A11)

This estimate of the variation of the mean spectrum agrees with the uncertainty estimates given in Polenta et al. (2005).

It should be noted that the noise and filtering of the SPT data are anisotropic. Equation (A7) can therefore be used only as a guideline. In the case of anisotropic filtering or anisotropic beams, the number of independent modes per bin will be typically smaller than νisob, since anisotropic filtering will weight different k-space modes unevenly. For example, the k-space mask completely eliminates all modes with kx < 1200. The variance in each ℓ-bin increases with fewer independent modes. However, even if we account for the effective number of independent modes in an ℓ-bin, Equation (A7) does not account for the anisotropic nature of the atmospheric noise contribution.

A.2. The Empirical Covariance Estimator

The existing analytic treatments are not directly applicable to the SPT data due to anisotropies in the noise and filters. By the nature of SPT's scan strategy, atmospheric fluctuations preferentially contaminate low kx modes. Likewise the filters intended to remove these fluctuations preferentially remove low kx modes. Instead, we have designed an empirical estimator which reproduces the analytical results when applied to isotropic data, while accurately accounting for the increased uncertainty due to the noise and filtering anisotropies in the actual data.

The noise covariance matrix estimate is divided into two parts, a signal contribution obtained from the Monte Carlo simulations described in Section 3.2.3 and a noise contribution obtained from real single-observation maps:

Equation (A12)

The signal contribution is straightforward to estimate with an approach similar to the MASTER power-spectrum error estimator. We use the signal only simulations to obtain an empirical estimate of the sample variance:

Equation (A13)

Note that here $\Delta x\equiv x - \overline{x}$ is defined with respect to the sample mean. Since the simulations include only CMB realizations and point sources in the confusion limit, the simulated signals are essentially Gaussian. Therefore, we expect the usual sample variance contribution:

Equation (A14)

As before, νb is the effective number of independent Fourier modes in each ℓ-band.

The noise contribution is computed from the cross-spectra of single-observation maps. We use the following estimator for the noise contribution:

Equation (A15)

Here f(nobs) is a correction due to the finite number of realizations. In the limit of many observations this function asymptotes to unity; we use 300 observations so this term can be ignored. The first term can be identified as the sample variance of the cross-spectra. The second term accounts for the additional correlations between cross-spectra with a common map.

We can now calculate the expectation value for the noise component of the covariance estimator defined in Equation (A15):

Equation (A16)

This is combined with the signal, or cosmic variance, component in Equation (A13) to get the expectation value of the estimator:

Equation (A17)

This agrees with the analytic estimate (Equation (A11)) for the variance of the mean.

A.3. Multifrequency Cross-covariances

A multifrequency data set requires an estimate of both the covariance of the each individual set of bandpowers (i.e., both the single-frequency bandpowers and the cross-frequency bandpowers) and the cross-covariance between these sets. We naturally expect signal correlations between different sets of bandpower due to the fact that all three sets of bandpowers reported here are derived from the same patch of sky. However we also expect noise correlations between the 150 GHz × 220 GHz cross-spectrum bandpowers and each set of single-frequency bandpowers since the noise uncertainty in the cross-spectrum is entirely due to noise in the 150 GHz and 220 GHz data. Thus, we compute the cross-covariance matrices, $ {\textbf {C}_{bb^\prime }}^{(i,j)}$, where i and j denote one of the three sets of bandpowers: 150 GHz, 220 GHz, and 150 GHz × 220 GHz:

Equation (A18)

A.4. Treatment of Off-diagonal Elements

Given the finite number of simulations and data maps, we expect some statistical uncertainty in the covariance estimate, particularly in the off-diagonal elements. Such uncertainty is not unique to the estimation technique described here, rather it is expected for any covariance estimate which is computed from a finite number of realizations. We expect the covariance estimates to be Wishart distributed, with nobs = 300 degrees of freedom. A given covariance element, $\textbf {C}_{ij}$ has a statistical variance of

Equation (A19)

For diagonal elements we expect a standard deviation of $\sqrt{2/n_{\rm obs}}=1/\sqrt{150}=8.1\%$. In addition, there is a statistical uncertainty on the apparent correlation between two bins. If we assume that the true correlation between bins is small, then the standard deviation of the apparent correlation between two bins is $\sqrt{1/n_{\rm obs}}$ or 5.7%. For the choice of bin size, the statistical error on the correlation of any two bins is much larger than the expected correlation (i.e., the fractional error on the apparent correlation estimates is greater than 100% even for adjacent bins). For bins that are widely separated, these false bin–bin correlations may skew model fitting. Therefore, we "condition" the published covariance matrices in order to reduce this statistical uncertainty on the covariance matrices.

From Equation (A5), we see that the shape of the correlation matrix (i.e., the relative size of the on-diagonal to off-diagonal covariance elements as a function of bin separation) is determined by the apodization window through the quadratic mode-coupling matrix, $M[\textbf {W}]_{kk}$. For the ℓ range considered in this work, the off-diagonal elements of this matrix depend only on the distance from the diagonal, |kk'|. Therefore, we condition the estimated covariance matrix, $\widehat{\textbf {C}}_{kk^\prime }$ by first computing the corresponding correlation matrix, and then averaging all off-diagonal elements of a fixed separation from the diagonal:

Equation (A20)

The bandpowers reported in Tables 1 and 2 are obtained by first computing power spectra and covariance matrices for a bin width of Δℓ = 100 with a total of 80 preliminary bins. This covariance matrix was then conditioned according to Equation (A20) before averaging the bandpowers and covariance matrix into the final bands.

Equation (A5) is based on the assumption that the filtering is isotropic. In order to test the validity of this equation for the anisotropic filtering, we perform 10,000 simple Monte Carlo simulations. In each simulation, a white noise realization is subjected to a simplified, though similarly anisotropic, version of the filtering scheme. The variance of the resultant spectra is computed and compared to Equation (A5). Though the apparent correlations between all bins exhibit the expected 1% scatter, the correlations between neighboring bins are consistent with Equation (A5).

Footnotes

Please wait… references are loading.
10.1088/0004-637X/719/2/1045