For the CMB-S4 Collaboration
Constraining cosmological parameters using the pairwise kinematic Sunyaev-Zel’dovich effect with CMB-S4 and future galaxy cluster surveys
Abstract
We present a forecast of the pairwise kinematic Sunyaev-Zel’dovich (kSZ) measurement that will be achievable with the future CMB-S4 experiment. CMB-S4 is the next stage for ground-based cosmic microwave background experiments, with a planned wide area survey that will observe approximately of the sky. We construct a simulated sample of galaxy clusters that have been optically selected in an LSST-like survey and have spectroscopic redshifts. For this cluster sample, we predict that CMB-S4 will reject the null hypothesis of zero pairwise kSZ signal at . We estimate the effects of systematic uncertainties such as scatter in the mass-richness scaling relation and cluster mis-centering. We find that these effects can reduce the signal-to-noise ratio of the CMB-S4 pairwise kSZ measurement by . We explore the constraining power of the measured kSZ signal in combination with measurements of the galaxy clusters’ thermal SZ emission on two extensions to the standard cosmological model. The first extension allows the dark energy equation of state to vary. We find the CMB-S4 pairwise kSZ measurement yields a modest reduction in the uncertainty on by a factor of 1.36 over the Planck’s 2018 uncertainty. The second extension tests General Relativity by varying the growth index . We find that CMB-S4’s pairwise kSZ measurement will yield a constraint on , and strongly constrain alternative theories of gravity.
I Introduction
The Sunyaev-Zel’dovich (SZ) effect [1, 2] is one of the largest sources of secondary anisotropies in the cosmic microwave background (CMB) and enables powerful probes of astrophysics and cosmology [e.g., 3, 4]. The SZ effect occurs when free electrons, such as those found in the hot intracluster medium of galaxy clusters, Compton scatter CMB photons. The SZ effect can be subdivided into two main subclasses. The first is the thermal SZ (tSZ) effect, which occurs due to an energy transfer from the hot electrons to the CMB photons during inverse Compton scattering. The net effect is to slightly distort the CMB blackbody spectrum by upscattering some CMB photons to higher frequencies. The second is the kinematic SZ (kSZ) effect, which is the Doppler shift induced by the bulk velocity of the free electrons and modifies the observed temperature of the CMB blackbody spectrum. The kSZ effect is the focus of this work.
Observations of the kSZ effect can be used to constrain both cosmological and astrophysical parameters [e.g., 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15]. Measuring the kSZ effect is challenging due to the kSZ effect’s faintness and spectral degeneracy with the CMB temperature fluctuations [16]. However, the kSZ effect is also one of a handful of observables that can directly probe the large-scale velocity field. The pairwise kSZ effect looks at the average relative velocity of two clusters as a function of their physical separation, which on average, these clusters should be falling towards one another due to gravity. This is one of several approaches that can be used to measure the kSZ signal. The first detection of the pairwise kSZ effect [17] was made using high-resolution CMB data from the Atacama Cosmology Telescope (ACT) [18] in conjunction with the Baryon Oscillation Spectroscopic Survey (BOSS) data release 9 spectroscopic galaxy catalog [19]. A subsequent pairwise kSZ analysis was performed using ACT and the Sloan Digital Sky Survey (SDSS) DR15 spectroscopic data set [20], which measured the pairwise kSZ signal to 5.4. This is the highest significance measured so far due to the use of spectroscopic catalogs. Measurements of the kSZ signal with the pairwise statistic have also been carried out using photometric catalogs, notably by the South Pole Telescope [21, 22] and the Dark Energy Survey (DES) [23] collaborations, which reported detections of the pairwise kSZ at the level of around [24, 25]. Other detections of the pairwise kSZ effect with data from the photometric DESI Legacy Imaging Surveys and Planck data have been reported by [26, 27].
Several other methods have been suggested to detect the kSZ signal. Among these are the velocity-weighted stacking approach [28, 29, 30, 31], kSZ tomographic analyses [32, 33], and estimators based on projected fields [34, 35, 36]. Additional strategies include using a velocity reconstruction method [37], angular redshift fluctuations [38], performing cross-correlations with 21 cm and other line-intensity mapping data sets [39, 40], and focusing on local measurements for individual clusters [41]. The variety of techniques explored to detect and analyze the kSZ effect offers the potential for an enhanced understanding of cosmic velocities and structures.
In this work, we forecast the cosmological constraining power of pairwise kSZ measurements with the next-generation CMB-S4 [42] experiment and upcoming large-scale structure surveys, like the Vera C. Rubin Observatory’s Legacy Survey of Space and Time (LSST) [43] and the 4-metre Multi-Object Spectroscopic Telescope (4MOST) [44] with the eROSITA cluster follow-up survey [45]. After near-term stage-3 experiments (like ACT and SPT), CMB-S4 is the next stage for ground-based CMB experiments and will employ both large and small aperture telescopes in order to probe small and large angular scales. In this work, we restrict ourselves to the wide-area survey of CMB-S4, which will cover of the sky after removing the high emission area of our galaxy (See Fig. 1 of [46]). We will use a spectroscopic galaxy cluster catalog, since photometric redshifts dilute the pairwise signal by a factor of , as shown in [25]. We improve the cosmological constraints from the pairwise kSZ measurement by calibrating the optical depth measurements to the tSZ signal of the cluster catalog. Finally, we consider the effects of scatter in the mass-richness scaling relation and the impact of mis-centering biases that can occur when dealing with the measurements from optical galaxy cluster surveys.
II Theoretical Background
II.1 The kinematic Sunyaev-Zel’dovich effect
The kSZ effect yields a small change in the CMB blackbody temperate, , which is proportional to the product of the free electron number density and line-of-sight velocity. The dependence of the kSZ signal on the bulk velocity of the ionized gas in the cluster offers a unique opportunity to probe the cosmological velocity field [47]. The optical depth of a galaxy cluster to CMB photons is low (typically ), so we can use the single-scattering limit. In this limit, the change in the CMB temperature due to the kSZ effect is
(1) |
where is the speed of light, is the Thomson cross section, is the number density of electrons, and is the Thomson optical depth for CMB photons traversing the galaxy cluster [2]. The unaltered CMB temperature is , while is the cluster velocity and a unit vector in the line of sight.
II.2 The pairwise kSZ signal
On scales smaller than the homogeneity scale ( Mpc [48]), pairs of galaxy clusters are expected to fall towards one another due to their mutual gravitational pull. This infall will produce a small dipole pattern in the CMB temperature anisotropy at the clusters’ positions due to the kSZ effect (e.g., [49]), which is known as the pairwise kSZ signal. The small pairwise kSZ signal can be detected by stacking many such pairs of clusters. The average pairwise kSZ amplitude for such a stack of clusters at a comoving separation can be related to the mean pairwise velocity of the clusters:
(2) |
where is the average optical depth of the sample. This equation requires two approximations: (1) the internal motion of the cluster’s gas is negligible; and (2) the optical depth and velocity of the clusters are uncorrelated [50]. We adopt a sign convention so that clusters falling towards one another will have a negative relative velocity and negative signal.
The mean pairwise velocity of haloes separated by comoving distance can be analytically modeled in linear theory for a specific cosmology and theory of gravity in terms of the two-point matter correlation function as [51, 52, 53]
(3) |
where is the scale factor, is the Hubble parameter, is the growth rate (with being the linear growth factor), the mass-averaged halo bias, and indicates the average of over a comoving sphere of radius . Fig. 1 compares the theoretical approximation of the pairwise velocity with a measurement obtained from the AGORA simulations, described in Sec. III.
Equations (2) and (3) highlight how measurements of the pairwise kSZ are sensitive to a combination of both cluster astrophysics, through the optical depth and halo bias , and cosmology through the Hubble parameter , the growth rate , and the two-point matter correlation function . In particular, the dependence on the growth rate and matter correlation function makes measurements of the pairwise kSZ signal sensitive to , thus providing complementary information to other cosmological probes such as redshift-space distortions, primarily sensitive to [54]. Therefore, if we break the degeneracy with cluster astrophysics through alternative methods, we can use the pairwise kSZ signal to access information of the velocity field (which on large scales is entirely dependent on the growth of the density field) and probe dark energy and modifications of gravity [53, 55, 56, 57, 58].

II.3 The thermal Sunyaev-Zel’dovich effect
As noted in the introduction, the kSZ effect is the fainter of the two SZ effects associated with a galaxy cluster. The larger signal is the tSZ effect, which is due to inverse Compton scattering between CMB photons and the hot electrons of the intracluster medium [59, 4]. The magnitude of the tSZ effect for a galaxy cluster can be quantified by the Compton- parameter [60], and expressed as an integral over the line of sight:
(4) |
Here is the free electron density, and is the electron temperature. The physical constants are the speed of light , Boltzmann constant, and electron mass . Importantly, both the tSZ and kSZ effects depend on the integrated electron number density, albeit with different weightings (temperature versus velocity). Thus the measured tSZ signal can potentially be used to calibrate the optical depth dependence of the kSZ signal (see [61, 50, 62]). We return to this in Sec. V.2.
III Simulations
III.1 Simulated CMB maps
We use realistic realizations of the millimeter wavelength sky from The Agora: Multi-Component Simulation for Cross-Survey Science [63], which leverages the N-body simulations of the MultiDark Planck 2 Synthetic Skies suite [64]. This simulation is based on Planck’s model [65] with the following cosmological parameters: , , , , .111The cosmological parameters listed are the Hubble parameter, matter density, baryonic matter density, current root mean square (rms) of the linear matter fluctuations on scales of , and the spectral index of the primordial scalar fluctuations, respectively. The simulated skies are generated by pasting astrophysical effects onto the dark matter haloes, obtained through a friends-of-friends algorithm applied to the MultiDark Planck 2 -body simulation [64]. The astrophysical modeling in the simulation has been calibrated using observational data and external hydrodynamical simulations.
Outlined below are the main components of the simulated microwave sky.
-
•
The dark matter density field is used to gravitationally lens the CMB sky.
- •
-
•
The kSZ effect is added in a similar way as the tSZ effect. The same [66] gas profile is used to estimate the electron number density, which is multiplied by the line-of-sight velocity to obtain the kSZ signal from the halo.
-
•
The simulations also include correlated signals from dusty star-forming galaxies, also called the cosmic infrared background (CIB).
-
•
Radio galaxies, which appear as point sources in the simulation maps, have been masked and are not included in the measurements.
We convolve the simulated maps by the expected instrumental beams for each CMB-S4 frequency band and add the expected white and 1/f noise levels to each map. Following [46], the assumed map noise has the form
(5) |
Here is the detector noise level, and are used to estimate the atmospheric noise, and is the spherical harmonic index. The assumed values of these parameters for each of the three relevant frequency maps are listed in Table 1.
Frequency | Beam FWHM | Noise level | 1/f noise |
---|---|---|---|
(GHz) | (′) | () | (, ) |
90 | 2.5 | 2.0 | (2154, 3.5) |
150 | 1.6 | 2.0 | (4364, 3.5) |
220 | 1.1 | 6.9 | (7334, 3.5) |
270 | 1.0 | 16.7 | (7308, 3.5) |
III.2 Mock cluster catalog
Mock cluster catalogs are a key ingredient to the pairwise kSZ forecasts of this work. We expect the cluster catalogs for the CMB-S4 measurement will come from an optical survey, such as LSST or 4MOST. Existing optical cluster finders, such as redMaPPer [68], identify galaxy clusters as overdensities of red sequence galaxies. The number of red sequence galaxies in clusters (often referred to as richness ) is correlated with halo mass and is commonly used as the mass tracer of galaxy clusters.
This approach to cluster detection can suffer from systematic uncertainties, most significantly projection effects due to the chance alignment of structures along the line of sight [69, 70, 71]. Fully simulating these effects is difficult and beyond the scope of this work, as the magnitude is sensitive to galaxy colors, luminosity, and their environmental dependence [70, 72]. However, [73] demonstrates that an approximate treatment, based on counting simulated galaxies within a cylinder along the line of sight, can capture the effect of projection reasonably well. We follow the same cylinder selection method in this work. However, we use a different projection length of Mpc/h since we find this better reproduces the observed number counts as a function of richness in the DES Year 1 data [69].
Within the CMB-S4 footprint, the simulated cluster sample contains clusters in the redshift range , with an optical richness estimate . This corresponds to the redshift and richness range for which the future LSST will provide a complete cluster sample [43]. The photometric redshift estimates from LSST reduce the signal-to-noise ratio by half, as shown in previous works [25]. Therefore, we assume 4MOST will provide the spectroscopic redshifts for the cluster catalog.
IV Analysis Methodology
IV.1 Pairwise kSZ estimator
Following previous works [17, 20, 25], we implement the pairwise kSZ estimator introduced by [74], which has the form
(6) |
This estimator scales the CMB temperature difference at the location of two clusters by a geometrical factor, , to account for the projection of the pair separation onto the line of sight. The temperature difference will depend on the difference between the kSZ signal in the two clusters, which in turn depends on the line of sight projection of the relative velocity between them.
We reconstruct the pairwise kSZ signal in eight equal and linearly spaced bins between comoving pair separation of 10 and 300 Mpc, with a bin width of Mpc.
IV.2 Map filtering and temperature extraction
To improve signal-to-noise, we use a constrained internal linear combination (cILC) approach [75] to combine the simulated multi-frequency maps for CMB-S4. Combining the different frequency bands in this way will lead to a CMB+kSZ map, while minimizing noise and the influence from frequency-dependent foregrounds, and simultaneously explicitly removing the tSZ signal. We combine data from different frequency channels using the cILC technique in harmonic space as
(7) |
where is the signal we want to retrieve from the frequency maps , is the total number of frequency channels, and is the -dependent weights for each frequency channel. These weights are tuned to produce a minimum-variance map along with nulling the contribution of the tSZ signal, where the specific frequency response is known, as
(8) |
where the matrix has dimension and contains the covariance between simulated maps, contains the frequency response vector of the desired signal and the foregrounds, and is used to extract the desired signal as described in [75].
To extract the CMB+kSZ temperature from the cILC combined map for Sec. V.1, we use an aperture photometry filter on the positions of clusters, which is written in map space as
(9) |
where is the characteristic filter scale. The aperture photometry acts as a high pass filter for scales that are larger than the filter scale by subtracting the average temperature in the outer ring from the average temperature inside the disc of radius . In contrast to matched filter techniques used in [24, 25], this approach does not assume a specific model for the cluster profile, but it requires that the cluster is contained within the characteristic filter scale to avoid biases in the temperature estimation .
In this work, we use the radius which is defined as the sphere within which the cluster mass is 500 times the critical density of the Universe at the cluster redshift. This aperture could miss some of the gas in the cluster, thus a model of what fraction is covered within that aperture will be necessary for future analyses. The distribution of in our cluster catalog is shown in Fig. 2.

IV.3 Redshift-dependent foregrounds
Over an extended redshift range, the redshift evolution of foregrounds (like the CIB) can introduce a redshift-dependent bias in the estimated temperatures at the positions of clusters [25]. To mitigate this, we estimate the mean measured temperature as a function of redshift and subtract this mean temperature from the estimated temperatures from the aperture photometry , as
IV.4 Covariance matrix
We estimate the covariance matrix of the binned pairwise kSZ measurement directly from the data using jackknife resampling [76]. The jackknife resampling technique consists of measuring the pairwise kSZ signal by splitting the cluster catalogue into subsamples, removing one of them, and recomputing the pairwise kSZ amplitude from the remaining subsamples. This process is repeated until every subsample has been discarded once from the measurement. Then we estimate the covariance matrix as
(11) |
where is the pairwise kSZ signal in separation bin at jackknife realization , of mean .
The baseline covariance matrix in this work is estimated using the jackknife resampling technique with 1000 subsamples. The off-diagonal elements can be significant, with correlations between adjacent bins.
We have tested the robustness of the covariance estimate to the choice of 1000 subsamples, and have found that the estimated covariance is stable. We show the changes in the diagonals of the covariance matrix between 1000 to 10,000 subsamples in Fig. 3. The diagonals agree within 1.5%, while the eigenvalues of the covariance matrices agree within 8%. We conclude that 1000 subsamples are sufficient to provide a robust estimate of the covariance matrix.

V Measurement forecast
V.1 Pairwise kSZ measurement
We present the expected pairwise kSZ measurement for CMB-S4 in this section. Fig. 4 shows the results for a single redshift bin and Fig. 5 decomposes the contribution from multiple redshift bins. These results have been obtained using the cILC map described in IV.2 and for optically selected clusters with richness .
We fit the measured pairwise kSZ signal to a one-parameter model. We then compute the statistical significance of our measurement. The reported signal-to-noise ratio (SNR) is obtained by fixing the cosmological parameters and obtaining the best-fit and its uncertainty by minimizing the as
(12) |
where is the measured signal, is the covariance matrix estimated using the jackknife method, and is the model computed using Eq.(2). The SNR is then computed with respect to the null-signal as , with .
We forecast that the pairwise kSZ signal will be detected at a significance of across all bins, and a significance of for the multiple redshift bins. These signal-to-noise numbers scale with the number of pairs of clusters, which is a function of the number of clusters in each redshift bin. CMB-S4 in combination with future spectroscopic surveys will detect the pairwise kSZ signal at very high significance.


V.2 Estimating the optical depth from the thermal SZ effect
The ability of the pairwise kSZ signal to constrain the velocity field is weakened by the degeneracy between each cluster’s velocity and optical depth. A possible path to break this degeneracy is to use the observed tSZ signal, which also depends on the product of the number density of free electrons and electron temperature, to independently constrain the clusters’ optical depth. Following [61, 50], we assume that the relationship between the mean integrated Compton-y and mean optical depth can be represented by a power law:
(13) |
where and are parameters to be determined by fitting the simulations. We use noiseless y-maps from the Agora simulations and perform a linear least-squares fit of the relationship between the optical depth and Compton-y parameters measured by applying the aperture photometry filter described in Sec. IV.2 to each halo’s lightcone. To allow for redshift evolution, we perform separate fits for each of the five redshift bins used in the pairwise kSZ forecasts in Fig. 5. The best-fit parameters for each redshift bin are presented in Tab. 2, while Fig. 6 shows the points and model fit for the first and last redshift bins.
Redshift | ||
---|---|---|

An important caveat to this approach is that we have calibrated the Compton-y parameter to the integrated number of free electrons, but not the velocity-weighted integral of the number density that shows up in the pairwise kSZ signal. We examine the magnitude of this difference by comparing the mean optical depth of the simulated clusters to the velocity-weighted mean optical depth of the same clusters in each redshift bin in Fig. 7. We find significant offsets between the two quantities, with the mean optical depth differing in each redshift bin by , respectively, relative to the velocity-weighted values (the here is only sample variance and does not include noise). When considering the full sample over all redshifts, the difference is . The discrepancy is likely due to factors like the optically selected catalog containing line-of-sight effects, which the friends-of-friends catalog does not consider. Future work should analyse why this happens and consider approaches to deal with these differences and the magnitude of the biases that would be incurred by using the mean optical depth.

V.3 Cosmological constraints
The pairwise kSZ is a direct measure of the infall velocity between pairs of clusters, and thus the gravitational attraction felt by these clusters. We consider two extensions to that should lead to observable differences in this infall. The first is the model, where the growth index is defined from the growth rate . For general relativity, , but alternate theories of gravity predict other values of [77]. The second extension is , where the dark energy equation of state is allowed to vary from the value of a cosmological constant.
We use a Fisher information formalism to predict expected constraints on cosmological parameters in each model. The covariance between two parameters and , from [57], is given by the Fisher matrix, which is calculated as
(14) |
where is the covariance between two pairwise kSZ bins estimated from the data in Sec. IV.4.
We determine which parameters have a higher effect on the pairwise kSZ signal by computing their log-derivatives with respect to the parameter being
(15) |
The log-derivatives give an estimate of how much each parameter contributes to the pairwise kSZ signal, i.e., the higher the log-derivative, the higher the plausible constraints on that parameter. We show these derivatives in Tab. 3, where we can see that the main cosmological parameters that contribute to the signal are , and .
Parameters | |||||||||
---|---|---|---|---|---|---|---|---|---|
1.00 | 0.64 | 0.21 | 0.18 | 0.09 | 0.04 | 0.02 |
We show the combined constraints of the main parameters in Fig. 8(a) for the and cosmologies. The plotted constraints include the pairwise kSZ measurement in all redshift bins, with a prior on the cluster optical depths. These constraints involve adding different redshift bins, using the mean optical depth values and uncertainties from the -map as a prior, and adding Planck’s 2018 covariance from the TTTEEE low+low and MCMC chains [65] as a prior to the cosmological parameters. We present the and cosmological parameters in Tab. 4.
In the model, the growth index covers the entire prior range of . Adding the pairwise kSZ measurement tightly constrains , with an uncertainty of . The pairwise kSZ data do not appreciably improve constraints on the other parameters, with the largest fractional improvement being only 4% for . There are mild degeneracies between and both and , suggesting that some improvement might be obtained by combining the data with external probes of large-scale structure or the expansion rate. On the other hand, since Planck leaves the posterior of the same as the prior, we obtain a 96% improvement. For the cosmology, we can see how the pairwise kSZ improves on all the fiducial values from Planck. , and are improved by 38%, 46%, and 42%, respectively.
Due to higher redshift uncertainties and/or higher CMB map noise levels, current Stage-3 measurements [20, 25] only detect the pairwise kSZ measurement at , compared to for CMB-S4. As a result, the Stage-3 measurements are largely unable to constrain cosmological models. For instance, using the Stage-3 uncertainties we estimate a constraint of at , which is 25 times larger than the forecast for CMB-S4. Thus, while Stage-3 measurements can provide initial insights, the advanced capabilities of CMB-S4 are crucial for achieving the precision needed to significantly constrain non-standard cosmological models.
We also observe a correlation between and all the cosmological parameters, which conveys the importance of properly constraining through techniques like the Compton values of the clusters. We explore the full range of cosmological parameters of the that are not as well constrained by the pairwise kSZ in Appendix A.


Parameter | Fiducial | Improvement over Planck | ||
: | ||||
67.77 | 0.61 | 0.57 (0.60) | 6% (2%) | |
0.818 | 12% (5%) | |||
: | ||||
67.77 | 0.61 | 0.60 (0.60) | 2% (2%) | |
0.818 | 4% (4%) | |||
0.57 | 0.57 | 0.02 (0.03) | 96% (93%) | |
: | ||||
67.77 | 8.94 | 5.51 (6.79) | 38% (24%) | |
0.818 | 46% (28%) | |||
-1 | 0.26 | 0.15 (0.19) | 42% (27%) |
V.4 Systematics
In this section, we assess the impact of systematic effects for the pairwise kSZ measurement. We consider the following sources.
-
•
Mass scatter: In order to match the mass range from simulations, where the masses of clusters are known, to the optical cluster catalog that is selected in richness, we need a good understanding of how to obtain an accurate representation of the mass range under analysis. This is of particular interest because the analytical model in Eq. 3, which is used to infer the optical depth of clusters, depends on the mass range of interest and changing the typical cluster mass could significantly bias this results. In this work, we have selected the simulation sample using the relation given in [78]. However, we need to take into account that these are estimated, and therefore a scatter in the cluster mass of the optical data can occur. To model this scatter, we draw mass errors from a normal distribution with width , which is an estimation of the scatter for a Gaussian error model in our optically selected catalog. We then use these errors to compute the pairwise kSZ signal from the simulations, obtaining an average decrease of the signal detection of , which corresponds to about a reduction on the signal strength.
-
•
Mis-centering: The measured pairwise kSZ signal can be diluted due to the fact that the position of clusters estimated from the optical survey catalog might not coincide with the location of the cluster kSZ signal. This mis-centering has a larger impact on clusters that are not fully relaxed or are merging, where the potential minimum is not located on the brightest cluster galaxy. To model the mis-centering, we draw RA and DEC errors from a normal distribution with width . This value corresponds to an uncertainty of Mpc at the mean redshift of the cluster catalog . This uncertainty is higher than the mean mis-centering of the optical catalog in comparison with high-resolution X-ray measurements [79]. This error can dilute the signal up to , which corresponds to a decrease in the significance of the measurement.
To account for these effects, we introduce an amplitude parameter to the pairwise kSZ measurement as
(16) |
This amplitude will constrain the amount of dilution that occurs to the pairwise kSZ measurement due to mass scatter and mis-centering of the clusters. We estimate this amplitude by calculating the ratio of the pairwise kSZ bins with respect to ones obtained after the selection effects are added to the simulations. Following [80], we estimate the uncertainty as a normal approximation to the distribution of the ratio of two random variables, obtaining .
If we include in the Fisher formalism, using the previously obtained value as the central value, we obtain the constraints as shown in Fig. 8(a) for , and Fig. 8(b) for . We observe that is highly correlated with all the parameters, but particularly with and . This decreases the improvements coming from the pairwise kSZ, as shown in the numbers inside the parenthesis in Tab. 4. This decrease is worse for cosmology, seeing a reduction of around for all the parameters. Meanwhile, for , there is only a 3% reduction on the constraining power of . Thus, the selection effects can decrease the effectiveness of the pairwise kSZ measurement and have to be taken into account.
VI Conclusions
We have presented forecasts for the pairwise kSZ measurement that will be achievable through the combination of the CMB-S4 survey and an LSST-like galaxy cluster catalog with spectroscopic redshifts. The data combination should reject the null hypothesis of no pairwise kSZ signal at across all redshifts, and at a significance of for each of five equal-width redshift bins covering the range . CMB-S4 will be able to detect the pairwise kSZ signal at very high significance.
To break the degeneracy between cosmological parameters and mean optical depth of the clusters, we independently estimated the latter using the Compton- map. We used a power-law form of to parameterize the relationship between the mean Compton- signal and , and calibrated this relation using noiseless simulations. Using this calibrated relationship, the estimated with the -map differs relative to the values obtained with the pairwise kSZ for the different redshift bins. Meanwhile, for the full redshift range, the values disagree with each other at a level. This indicates that we need to better model the relationship between the mean Compton-y value and with improved simulations that can tell us the optical depth of optically selected clusters. We leave this to future work.
However, despite the disagreement, we have still shown how the pairwise kSZ signal can constrain cosmological parameters. Using the Fisher formalism, we produced constraints for two models: (a) ; and (b). We found that is highly correlated with and . This means that the better we can model and constrain , the better our constraints on modifications to cosmology will be. For , we can measure to , which should constrain alternatives to general relativity [81]. Current Stage-3 measurements are limited by high redshift uncertainties and noise levels, achieving only a modest detection of the pairwise kSZ effect and providing constraints on cosmological parameters like that are substantially less precise than those anticipated from CMB-S4. For , we can improve on the Planck cosmological parameters constraints approximately , showing that including pairwise kSZ measurements with CMB power spectrum will help constrain alternatives to .
We have also tested how systematics could influence the measurement of the pairwise kSZ signal. We tested the impact due to the scatter in the richness-mass scaling relation and the cluster mis-centering. These effects combined can dilute the signal up to 20% of the true amplitude value. To account for these effects, we have introduced an amplitude parameter to the pairwise kSZ measurement. We have estimated this parameter to be , by estimating the average of the ratios of the signal with and without the effects included. Including this additional parameter reduces the constraining power of the pairwise kSZ by 20%.
Overall, our results show that the pairwise kSZ signal using the future CMB-S4 experiment and an LSST-like cluster catalog with spectroscopic redshifts can be a powerful tool to investigate alternatives to cosmology, as well as shed some light on the gravitational properties of the Universe at large scales.
Acknowledgments
We thank Daan Meerburg and Bruce Partridge for providing useful comments on the manuscript.
SR acknowledges support by the Illinois Survey Science Fellowship from the Center for AstroPhysical Surveys at the National Center for Supercomputing Applications.
CS acknowledges support from the Agencia Nacional de Investigación y Desarrollo (ANID) through Basal project FB210003.
The Melbourne group acknowledges support from the Australian Research Council’s Discovery Projects scheme (DP200101068). ES acknowledges support from the David Lachlan Hay Memorial Fund.
CMB-S4 is supported by the Director, Office of Science, Office of High Energy Physics of the U.S. Department of Energy under contract No. DEAC02-05CH11231; by the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility under the same contract; by the Divisions of Physics and Astronomical Sciences and the Office of Polar Programs of the U.S. National Science Foundation under Mid-Scale Research Infrastructure award OPP-1935892; and by the NSF Cooperative Agreement award: AST-2240374. Considerable additional support is provided by the many CMB-S4 team members and their institutions.
This research used resources of the National Energy Research Scientific Computing Center, a DOE Office of Science User Facility supported by the Office of Science of the U.S. Department of Energy under contract No. DE-AC02-05CH11231. This research was supported by The University of Melbourne’s Research Computing Services and the Petascale Campus Initiative.
References
- Sunyaev and Zel’dovich [1970a] R. A. Sunyaev and Y. B. Zel’dovich, Ap&SS 7, 3 (1970a), URL http://adsabs.harvard.edu/cgi-bin/nph-bib_query?bibcode=1970Ap%26SS...7....3S&db_key=AST.
- Sunyaev and Zel’dovich [1980] R. Sunyaev and Y. Zel’dovich, ARAA 18, 537 (1980).
- Birkinshaw [1999] M. Birkinshaw, Physics Reports 310, 97 (1999).
- Carlstrom et al. [2002] J. E. Carlstrom, G. P. Holder, and E. D. Reese, ARA&A 40, 643 (2002).
- Lahav et al. [1991] O. Lahav, M. J. Rees, P. B. Lilje, and J. R. Primack, MNRAS 251, 128 (1991).
- Haehnelt and Tegmark [1996] M. G. Haehnelt and M. Tegmark, MNRAS 279, 545+ (1996), URL http://adsabs.harvard.edu/cgi-bin/nph-bib_query?bibcode=1996MNRAS.279..545H&db_key=AST.
- Battaglia et al. [2013] N. Battaglia, A. Natarajan, H. Trac, R. Cen, and A. Loeb, Astrophys. J. 776, 83 (2013), eprint 1211.2832.
- Ma and Zhao [2014] Y.-Z. Ma and G.-B. Zhao, Phys. Lett. B 735, 402 (2014), eprint 1309.1163.
- Bianchini and Silvestri [2016] F. Bianchini and A. Silvestri, Physical Review D 93 (2016), ISSN 2470-0029, URL http://dx.doi.org/10.1103/PhysRevD.93.064026.
- Alonso et al. [2016] D. Alonso, T. Louis, P. Bull, and P. G. Ferreira, Phys. Rev. D 94, 043522 (2016), eprint 1604.01382.
- Smith and Ferraro [2017] K. M. Smith and S. Ferraro, Physical Review Letters 119, 021301 (2017).
- Battaglia et al. [2017] N. Battaglia, S. Ferraro, E. Schaan, and D. N. Spergel, J. of Cosm. & Astropart. Phys. 2017, 040 (2017), eprint 1705.05881.
- Alvarez et al. [2021] M. A. Alvarez, S. Ferraro, J. C. Hill, R. Hložek, and M. Ikape, Phys. Rev. D 103, 063518 (2021), eprint 2006.06594.
- Raghunathan and Omori [2023] S. Raghunathan and Y. Omori, Astrophys. J. 954, 83 (2023), eprint 2304.09166.
- Raghunathan et al. [2024] S. Raghunathan, P. A. R. Ade, A. J. Anderson, et al., arXiv e-prints arXiv:2403.02337 (2024), eprint 2403.02337.
- Shaw et al. [2012] L. D. Shaw, D. H. Rudd, and D. Nagai, Astrophys. J. 756, 15 (2012), eprint 1109.0553.
- Hand et al. [2012] N. Hand, G. E. Addison, E. Aubourg, N. Battaglia, E. S. Battistelli, D. Bizyaev, J. R. Bond, H. Brewington, J. Brinkmann, B. R. Brown, et al., Physical Review Letters 109, 041101 (2012), eprint 1203.4219.
- Swetz et al. [2011] D. S. Swetz, P. A. R. Ade, M. Amiri, J. W. Appel, E. S. Battistelli, B. Burger, J. Chervenak, M. J. Devlin, S. R. Dicker, W. B. Doriese, et al., ApJS 194, 41 (2011), eprint 1007.0290.
- Dawson et al. [2013] K. S. Dawson, D. J. Schlegel, C. P. Ahn, S. F. Anderson, É. Aubourg, S. Bailey, R. H. Barkhouser, J. E. Bautista, A. Beifiori, A. A. Berlind, et al., AJ 145, 10 (2013), eprint 1208.0022.
- Calafut et al. [2021] V. Calafut, P. Gallardo, E. Vavagiakis, S. Amodeo, S. Aiola, J. Austermann, N. Battaglia, E. Battistelli, J. Beall, R. Bean, et al., Physical Review D 104 (2021), ISSN 2470-0029, URL http://dx.doi.org/10.1103/PhysRevD.104.043502.
- Padin et al. [2008] S. Padin, Z. Staniszewski, R. Keisler, M. Joy, A. A. Stark, P. A. R. Ade, K. A. Aird, B. A. Benson, L. E. Bleem, J. E. Carlstrom, et al., Appl. Opt. 47, 4418 (2008), URL http://adsabs.harvard.edu/abs/2008ApOpt..47.4418P.
- Carlstrom et al. [2011] J. E. Carlstrom, P. A. R. Ade, K. A. Aird, B. A. Benson, L. E. Bleem, S. Busetti, C. L. Chang, E. Chauvin, H.-M. Cho, T. M. Crawford, et al., PASP 123, 568 (2011), eprint 0907.4445.
- Flaugher et al. [2015] B. Flaugher, H. T. Diehl, K. Honscheid, T. M. C. Abbott, O. Alvarez, R. Angstadt, J. T. Annis, M. Antonik, O. Ballester, L. Beaufore, et al., AJ 150, 150 (2015), eprint 1504.02900.
- Soergel et al. [2016] B. Soergel, S. Flender, K. T. Story, L. Bleem, T. Giannantonio, G. Efstathiou, E. Rykoff, B. A. Benson, T. Crawford, S. Dodelson, et al., MNRAS 461, 3172 (2016), eprint 1603.03904.
- Schiappucci et al. [2023] E. Schiappucci, F. Bianchini, et al. (SPT-3G, DES), Phys. Rev. D 107, 042004 (2023), eprint 2207.11937.
- Chen et al. [2021] Z. Chen, P. Zhang, X. Yang, and Y. Zheng, Monthly Notices of the Royal Astronomical Society 510, 5916–5928 (2021), ISSN 1365-2966, URL http://dx.doi.org/10.1093/mnras/stab3604.
- Li et al. [2024] S. Li, Y. Zheng, Z. Chen, H. Xu, and X. Yang, ApJS 271, 30 (2024), eprint 2401.03507.
- Li et al. [2014] M. Li, R. E. Angulo, S. D. M. White, and J. Jasche, MNRAS 443, 2311 (2014), eprint 1404.0007.
- Schaan et al. [2016] E. Schaan, S. Ferraro, M. Vargas-Magaña, K. M. Smith, S. Ho, S. Aiola, N. Battaglia, J. R. Bond, F. De Bernardis, E. Calabrese, et al., Phys. Rev. D 93, 082002 (2016), eprint 1510.06442.
- Schaan et al. [2021] E. Schaan, S. Ferraro, S. Amodeo, N. Battaglia, S. Aiola, J. E. Austermann, J. A. Beall, R. Bean, D. T. Becker, R. J. Bond, et al., Phys. Rev. D 103, 063513 (2021), eprint 2009.05557.
- Hadzhiyska et al. [2024] B. Hadzhiyska, S. Ferraro, B. Ried Guachalla, et al., arXiv e-prints arXiv:2407.07152 (2024), eprint 2407.07152.
- Ho et al. [2009] S. Ho, S. Dedeo, and D. Spergel, arXiv e-prints arXiv:0903.2845 (2009), eprint 0903.2845.
- Smith et al. [2018] K. M. Smith, M. S. Madhavacheril, M. Münchmeyer, S. Ferraro, U. Giri, and M. C. Johnson, arXiv e-prints arXiv:1810.13423 (2018), eprint 1810.13423.
- Hill et al. [2016] J. C. Hill, S. Ferraro, N. Battaglia, J. Liu, and D. N. Spergel, Physical Review Letters 117 (2016), URL https://doi.org/10.1103%2Fphysrevlett.117.051301.
- Kusiak et al. [2021] A. Kusiak, B. Bolliet, S. Ferraro, J. C. Hill, and A. Krolewski, Phys. Rev. D 104, 043518 (2021), eprint 2102.01068.
- Patki et al. [2023] R. Patki, N. Battaglia, and S. Ferraro, Phys. Rev. D 108, 043507 (2023), eprint 2306.03127.
- Deutsch et al. [2018] A.-S. Deutsch, E. Dimastrogiovanni, M. C. Johnson, M. Münchmeyer, and A. Terrana, Phys. Rev. D 98, 123501 (2018), eprint 1707.08129.
- Chaves-Montero et al. [2020] J. Chaves-Montero, C. Hernández-Monteagudo, R. E. Angulo, and J. D. Emberson, Monthly Notices of the Royal Astronomical Society 503, 1798–1814 (2020), ISSN 1365-2966, URL http://dx.doi.org/10.1093/mnras/staa3782.
- Sato-Polito et al. [2021] G. Sato-Polito, J. L. Bernal, K. K. Boddy, and M. Kamionkowski, Phys. Rev. D 103, 083519 (2021), eprint 2011.08193.
- Li et al. [2019] D. Li, H.-M. Zhu, and U.-L. Pen, Phys. Rev. D 100, 023517 (2019), eprint 1811.05012.
- Sayers et al. [2013] J. Sayers, T. Mroczkowski, M. Zemcov, P. M. Korngut, J. Bock, E. Bulbul, N. G. Czakon, E. Egami, S. R. Golwala, P. M. Koch, et al., Astrophys. J. 778, 52 (2013), eprint 1312.3680.
- Abazajian et al. [2019] K. Abazajian et al., arXiv e-prints arXiv:1907.04473 (2019), eprint 1907.04473.
- Ivezić et al. [2019] Ž. Ivezić, S. M. Kahn, J. A. Tyson, B. Abel, E. Acosta, R. Allsman, D. Alonso, Y. AlSayyad, S. F. Anderson, J. Andrew, et al., Astrophys. J. 873, 111 (2019), eprint 0805.2366.
- de Jong et al. [2019] R. S. de Jong et al., The Messenger 175, 3 (2019), eprint 1903.02464.
- Finoguenov et al. [2019] A. Finoguenov, A. Merloni, J. Comparat, K. Nandra, M. Salvato, E. Tempel, A. Raichoor, J. Richard, J. P. Kneib, A. Pillepich, et al., The Messenger 175, 39 (2019), eprint 1903.02471.
- Raghunathan et al. [2022] S. Raghunathan, N. Whitehorn, M. A. Alvarez, H. Aung, N. Battaglia, G. P. Holder, D. Nagai, E. Pierpaoli, C. L. Reichardt, and J. D. Vieira, Astrophys. J. 926, 172 (2022), eprint 2107.10250.
- Bernardeau et al. [2002] F. Bernardeau, S. Colombi, E. Gaztañaga, and R. Scoccimarro, Phys. Rep. 367, 1 (2002), eprint astro-ph/0112551.
- Yadav et al. [2010] J. K. Yadav, J. S. Bagla, and N. Khandai, MNRAS 405, 2009 (2010), eprint 1001.0617.
- Diaferio et al. [2000] A. Diaferio, R. A. Sunyaev, and A. Nusser, ApJL 533, L71 (2000), eprint astro-ph/9912117.
- Soergel et al. [2018] B. Soergel, A. Saro, T. Giannantonio, G. Efstathiou, and K. Dolag, Monthly Notices of the Royal Astronomical Society 478, 5320 (2018), URL https://doi.org/10.1093%2Fmnras%2Fsty1324.
- Juszkiewicz et al. [1999] R. Juszkiewicz, V. Springel, and R. Durrer, Astrophys. J. Lett. 518, L25 (1999), eprint astro-ph/9812387.
- Sheth et al. [2001] R. K. Sheth, A. Diaferio, L. Hui, and R. Scoccimarro, Mon. Not. Roy. Astron. Soc. 326, 463 (2001), eprint astro-ph/0010137.
- Bhattacharya and Kosowsky [2008] S. Bhattacharya and A. Kosowsky, Phys. Rev. D 77, 083004 (2008), eprint 0712.0034.
- Percival and White [2009] W. J. Percival and M. White, MNRAS 393, 297 (2009), eprint 0808.0003.
- Kosowsky and Bhattacharya [2009] A. Kosowsky and S. Bhattacharya, Phys. Rev. D 80, 062003 (2009), eprint 0907.4202.
- Keisler and Schmidt [2013] R. Keisler and F. Schmidt, ApJL 765, L32 (2013), eprint 1211.0668.
- Mueller et al. [2015] E.-M. Mueller, F. de Bernardis, R. Bean, and M. D. Niemack, Astrophys. J. 808, 47 (2015), eprint 1408.6248.
- Ma et al. [2015] Y.-Z. Ma, M. Li, and P. He, Astron. Astrophys. 583, A52 (2015), eprint 1509.06413.
- Sunyaev and Zel’dovich [1970b] R. A. Sunyaev and Y. B. Zel’dovich, Comments on Astrophysics and Space Physics 2, 66 (1970b).
- Sunyaev and Zel’dovich [1972] R. A. Sunyaev and Y. B. Zel’dovich, Comments on Astrophysics and Space Physics 4, 173 (1972).
- Battaglia [2016] N. Battaglia, Journal of Cosmology and Astroparticle Physics 2016, 058–058 (2016), ISSN 1475-7516, URL http://dx.doi.org/10.1088/1475-7516/2016/08/058.
- Vavagiakis et al. [2021] E. Vavagiakis, P. Gallardo, V. Calafut, S. Amodeo, S. Aiola, J. Austermann, N. Battaglia, E. Battistelli, J. Beall, R. Bean, et al., Physical Review D 104 (2021), URL https://doi.org/10.1103%2Fphysrevd.104.043503.
- Omori [2024] Y. Omori, MNRAS 530, 5030 (2024), eprint 2212.07420.
- Klypin et al. [2016] A. Klypin, G. Yepes, S. Gottlöber, F. Prada, and S. Heß, MNRAS 457, 4340 (2016), eprint 1411.4001.
- Planck Collaboration et al. [2020] Planck Collaboration, N. Aghanim, Y. Akrami, M. Ashdown, J. Aumont, C. Baccigalupi, M. Ballardini, A. J. Banday, R. B. Barreiro, N. Bartolo, et al., A&A 641, A6 (2020), eprint 1807.06209.
- Mead et al. [2020] A. J. Mead, T. Tröster, C. Heymans, L. Van Waerbeke, and I. G. McCarthy, Astronomy & Astrophysics 641, A130 (2020), ISSN 1432-0746, URL http://dx.doi.org/10.1051/0004-6361/202038308.
- McCarthy et al. [2016] I. G. McCarthy, J. Schaye, S. Bird, and A. M. C. Le Brun, Monthly Notices of the Royal Astronomical Society 465, 2936–2965 (2016), ISSN 1365-2966, URL http://dx.doi.org/10.1093/mnras/stw2792.
- Rykoff et al. [2014] E. S. Rykoff, E. Rozo, M. T. Busha, C. E. Cunha, A. Finoguenov, A. Evrard, J. Hao, B. P. Koester, A. Leauthaud, B. Nord, et al., Astrophys. J. 785, 104 (2014), eprint 1303.3562.
- Abbott et al. [2020] T. M. C. Abbott, M. Aguena, A. Alarcon, S. Allam, S. Allen, J. Annis, S. Avila, D. Bacon, K. Bechtol, A. Bermeo, et al., Phys. Rev. D 102, 023509 (2020), eprint 2002.11124.
- To et al. [2021] C.-H. To, E. Krause, E. Rozo, H.-Y. Wu, D. Gruen, J. DeRose, E. Rykoff, R. H. Wechsler, M. Becker, M. Costanzi, et al., MNRAS 502, 4093 (2021), eprint 2008.10757.
- Sunayama et al. [2020] T. Sunayama, Y. Park, M. Takada, Y. Kobayashi, T. Nishimichi, T. Kurita, S. More, M. Oguri, and K. Osato, MNRAS 496, 4468 (2020), eprint 2002.03867.
- To et al. [2024] C.-H. To, J. DeRose, R. H. Wechsler, E. Rykoff, H.-Y. Wu, S. Adhikari, E. Krause, E. Rozo, and D. H. Weinberg, Astrophys. J. 961, 59 (2024), eprint 2303.12104.
- Wu et al. [2022] H.-Y. Wu et al. (DES), Mon. Not. Roy. Astron. Soc. 515, 4471 (2022), eprint 2203.05416.
- Ferreira et al. [1999] P. G. Ferreira, R. Juszkiewicz, H. A. Feldman, M. Davis, and A. H. Jaffe, The Astrophysical Journal 515, L1 (1999), URL https://doi.org/10.1086/311959.
- Remazeilles et al. [2011] M. Remazeilles, J. Delabrouille, and J.-F. Cardoso, MNRAS 418, 467 (2011), eprint 1103.1166.
- Escoffier et al. [2016] S. Escoffier, M. C. Cousinou, A. Tilquin, A. Pisani, A. Aguichine, S. de la Torre, A. Ealet, W. Gillard, and E. Jullo, arXiv e-prints arXiv:1606.00233 (2016), eprint 1606.00233.
- Huterer et al. [2015] D. Huterer et al., Astropart. Phys. 63, 23 (2015), eprint 1309.5385.
- McClintock et al. [2019] T. McClintock, T. N. Varga, D. Gruen, E. Rozo, E. S. Rykoff, T. Shin, P. Melchior, J. DeRose, S. Seitz, et al., MNRAS 482, 1352 (2019), eprint 1805.00039.
- Rozo and Rykoff [2014] E. Rozo and E. S. Rykoff, Astrophys. J. 783, 80 (2014), eprint 1303.3373.
- Díaz-Francés and Rubio [2013] E. Díaz-Francés and F. Rubio, Statistical Papers 54, 309 (2013), URL https://ideas.repec.org/a/spr/stpapr/v54y2013i2p309-323.html.
- Avila et al. [2022] F. Avila, A. Bernui, A. Bonilla, and R. C. Nunes, Eur. Phys. J. C 82, 594 (2022), eprint 2201.07829.
Appendix A Redshift dependence of the entire cosmological parameter space
Here we show the effects of all the cosmological parameters that we considered for the pairwise kSZ analysis. This is a cosmological model of + + + , where we want to observe also the redshift dependent constraints that can be achieved by using a redshift bin separation of 0.2. The redshift dependence is shown in Fig. 9, where we did not include priors from Planck on the parameters, allowing us to see the constraining power of purely the pairwise kSZ at a binned redshift level. The full combined constraints are presented in Fig. 9, where Planck’s covariance from the low+low+BAO were used as priors for the main + parameters, and the uncertainty on that comes from the y-map was used as a prior for this parameter. We show the constrained parameters in Tab. 5, where we observe how most parameters are not changed at all by adding the pairwise kSZ, but can still be constrained at the level.
Parameter | Fiducial | Improvement over Planck | ||
---|---|---|---|---|
67.77 | 1.44 | 1.42 | 1% | |
0.048 | ||||
0.259 | ||||
0.818 | 2% | |||
0.96 | - | |||
0.57 | 0.57 | 0.04 | 93% | |
-1 | 0.34 | 0.34 | - | |
0 | 0.06 | 0.06 | - |

