<<

Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June https://doi.org/10.1088/1538-3873/ab5bfd © 2020. The Astronomical Society of the Pacific. All rights reserved. Printed in the U.S.A.

Data Analysis for Precision 21cm Cosmology Adrian Liu1 and J. Richard Shaw2,3 1 Department of Physics and McGill Space Institute, McGill University, Montreal, QC, H3A 2T8, Canada; [email protected] 2 Department of Physics and Astronomy, University of British Columbia, Vancouver, BC, V6T 1Z1, Canada 3 National Research Council Canada, Herzberg Research Centre for Astronomy and Astrophysics, Dominion Radio Astrophysical Observatory, P.O. Box 248, Penticton, British Columbia, V2A 6J9, Canada Received 2019 July 16; accepted 2019 October 26; published 2020 April 21

Abstract The redshifted 21 cm line is an emerging tool in cosmology, in principle permitting three-dimensional surveys of our universe that reach unprecedentedly large volumes, previously inaccessible length scales, and hitherto unexplored epochs of our cosmic timeline. Large radio telescopes have been constructed for this purpose, and in recent years there has been considerable progress in transforming 21 cm cosmology from a field of considerable theoretical promise to one of observational reality. Increasingly, practitioners in the field are coming to the realization that the success of observational 21 cm cosmology will hinge on software algorithms and analysis pipelines just as much as it does on careful hardware design and telescope construction. This review provides a pedagogical introduction to state-of-the-art ideas in 21 cm data analysis, covering a wide variety of steps in a typical analysis pipeline, from calibration to foreground subtraction to map making to power spectrum estimation to parameter estimation. Key words: dark ages, , first – methods: statistical – techniques: interferometric Online material: color figures

1. Introduction In the next few decades, 21 cm cosmology has the potential to significantly enhance our understanding of our universe. The The last few decades have seen significant progress in basic idea is to take advantage of the 21 cm wavelength spectral observing large portions of our universe using surveys, line that arises from the hyperfine “ flip” transition of Cosmic Microwave Background (CMB) measurements, Lyα neutral hydrogen. By searching for absorption or emission of forest studies, gravitational lensing, galaxy cluster counts, and the redshifted 21 cm line, one has a way to observe neutral now even gravitational wave interferometers. However, in hydrogen, which can in turn be used as a tracer of or as many ways, our universe remains considerably unexplored. For an indirect probe of other properties of our universe such as its example, the CMB and galaxy surveys probe the early and the ionization state or temperature. The 21 cm line of hydrogen is a late universe, respectively, but intermediate portions of our “forbidden” transition, and thus the likelihood of an individual cosmic timeline have not yet been systematically surveyed. As making the transition is low. However, compensating for a result, there exists a gap in direct observations between this is the sheer abundance of hydrogen in our universe, which ∼ ( 400,000 and 1.5 Gyr after the corresponding to a results in an observable signal. In addition, the weakness of the ∼ ∼ ) missing range between z 1100 to z 3 . To be fair, transition means that the transition is optically thin, i.e., 21 cm some information does exist at these intermediate , but are unlikely to be absorbed/emitted more than once as this tends to take the form of individual astronomical objects of they travel to our telescopes. This enables a fully three- ( α interest or in the case of Ly forest studies, just a small dimensional mapping of the spectral line, with the redshift handful of sight lines that extend to the relevant redshifts). providing line of sight distance information. The result will be Moreover, even at redshifts that are nominally covered by a data set with unprecedented reach in volume. surveys, the surveys may be incomplete in a variety of ways. Using the 21 cm line not only enables larger survey volumes, For example, they may not cover the entire sky because of but also pushes state-of-the-art limits in redshift, luminosity, the location of one’s survey instrument. They may also lack and scale. Extremely high redshifts are in principle accessible the resolution to capture details on the finest scales, or the because one only requires neutral hydrogen, which exists prior sensitivity to capture information from anything but the to the formation of the first luminous objects. Moreover, even brightest objects. In general, there is considerable information after luminous objects are formed, only the brightest ones will content locked in our cosmos that has yet to be harnessed. be seen by traditional observations. Observations in 21 cm

1 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw cosmology typically do not resolve such individual bright summary in Section 8, state-of-the-art approaches to dealing objects, instead averaging over many objects that fall inside with the aforementioned challenges are then described in single, relatively large, pixels. The reward of this approach, subsequent sections, with Section 9 discussing calibration, however, is that one is probing the integrated emission over the Section 10 discussing map making, Section 11 discussing entire luminosity function, including photons from objects that power spectrum estimation, and Section 12 discussing fore- would be too dim to detect individually. Without the need to ground mitigation. Section 13 then moves onto parameter resolve individual objects, instrument specifications tend to be extraction and fitting data to theoretical models within the driven by length scales dictated by cosmology. While these context of a power spectrum measurements. Other types of length scales are therefore generally large, it should be noted measurements are described in Section 14. In Section 15 we that the 21 cm line also enables one to access scales that are provide an overview of sky-averaged global signal measure- much finer than other cosmological probes. Much of this ments that have generated much excitement lately because of a is the result of the relative ease in which spectral resolution claimed positive detection. We summarize our conclusions in (and therefore line of sight distance resolution) can be obtained Section 16. Table 1 provides a glossary of useful mathematical in radio astronomy. definitions. Motivated by the aforementioned opportunities, a large number of instruments have been constructed to make a first 2. Science with the 21 cm Line measurement of the cosmological 21 cm signal. The last few 2.1. 21 cm Line Basics years have seen considerable progress toward this goal: detections of the 21 cm line have been made in cross The 21 cm line is the hyperfine transition of atomic correlation with galaxy surveys at z<1, and there has been hydrogen. The parallel alignment of the and a tentative claim of a detection of the sky-averaged 21 cm spins is a slightly higher energy state than the anti-parallel signal at z∼17. In addition, instruments seeking to measure alignment. As an atom transitions from one state to the other, it spatial structures in the 21 cm line across a wide range of emits (or absorbs) a of 21 cm wavelength. To study this redshift are now online and in many cases, have begun setting line we use a quantity called the spin-temperature Ts that upper limits on the signal. Much of the emphasis in the field describes the relative occupancy of the two spin states has therefore shifted over to issues of precision data analysis. In n ⎛ h n ⎞ 1 =-3exp⎜ 21⎟ ,() 1 many ways, 21 cm instruments are as much software telescopes ⎝ ⎠ as they are hardware telescopes—it is becoming increasingly n0 kTbs clear that analysis pipelines capable of meeting exquisite where the factor of three comes from the relative degeneracy of requirements in calibration, data volume, and control of the states, n1 is the number of in the excited hyperfine systematics are just as important as careful hardware design. state, n0 is the number in the ground hyperfine state, h is This paper is a response to the recent emphasis on data ’s constant, kb is Boltzmann’s constant, and n21 » analysis. Our goal is to provide a pedagogical introduction to 1420.406 MHz is the rest frequency of the 21 cm line. The analysis ideas and problems that are currently on the minds of physics of the spin temperature throughout the history of the analysts working in 21 cm cosmology. While we do provide universe is complex; we will cover the relevant parts below. some overview of the theory (Section 2) and instrumentation The key thing to note is that we observe the contrast between (Sections 3–5), these summaries are intended to serve only as the CMB and the spin temperature. Where the spin temperature background for our more thorough discussions of analysis. As is higher than the CMB temperature, Tg, we are emitting such, this paper is complementary to the many excellent photons and see an excess above the CMB temperature; when it previous reviews on 21 cm cosmology, such as Furlanetto et al. is lower than Tg the photons are absorbed from the CMB and (2006), Morales & Wyithe (2010), Pritchard & Loeb (2012), we see a deficit compared to what we expect.4 ( ) Loeb & Furlanetto 2013 . The brightness temperature Tb of the 21 cm line is given by The rest of this paper is organized as follows. Section 2 fl -t (rˆ,z) Tzs(rˆ, )()- Tzg brie y reviews the theoretical foundations of the predicted Teb(rˆ,1n)[=- 21 ] ,2() 21 cm signal. Section 3 then provides a quick introduction to 1 + z 21 cm cosmology from an observational standpoint, qualita- where rˆ is a unit vector that is centered on the observer and tively discussing issues such as instrument design optimization. points in the direction of observation on the sky, and the This is followed by a more technical discussion of inter- observational frequency ν is related to the redshift z via the ferometry in Section 5 after a survey of the current status of the field in Section 4. We provide an overview of many systematic 4 In this paper, we focus exclusively on absorption or emission of the 21 cm effects in Section 6, but leave the chief technical challenge of line relative to the CMB. However, it may also be possible to detect 21 cm — — absorption from a bright, high-redshift radio source (Carilli et al. 2002; 21 cm cosmology that of foreground contaminants to a Furlanetto & Loeb 2002; Xu et al. 2009; mack & Wyithe 2012; Ciardi et al. dedicated description in Section 7. After a brief mid-paper 2013, 2015a, 2015b; Ewall-Wice et al. 2014; Semelin 2016).

2 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw standard relation: phenomena tend to dominate at different redshifts, and in what n follows we provide a quick qualitative description of this. 1.+=z 21 () 3 n

We have additionally defined t21()z as the optical depth across 2.2. Ultra High Redshift: The Dark Ages the 21 cm line at redshift z, which is given by After recombination, though the photons and baryons no 3cA3 xn longer act as a single fluid, scatterings between photons and t (rˆ, z) = 10 HHI ,4() 21 2 residual (and collisions with the rest of the gas) mean 16kbn ()()1 + zdvdrTs 21 that the baryon temperature is coupled to the photon where dvdris the gradient of the proper velocity v along the temperature until z∼300 when these electron-photon scatter- line of sight distance r, xH I is the fraction of hydrogen atoms ings become rare. At this point, the gas starts to cool faster than ÿ 2 that are neutral, nH is their number density, is the reduced the photons ∝(1+z) , and because the collisional coupling ’ = × Planck s constant, c is the speed of light, and A10 2.85 between the gas temperature and the spin temperature remains −15 −1 10 s is the spontaneous emission coefficient of the 21 cm strong, Equation (5) predicts a net 21 cm absorption signal. All line. This optical depth is typically 4% at all regions of of this takes place during the Dark Ages, a pristine interest (Lewis & Challinor 2007). We may therefore Taylor cosmological era where the first luminous sources have yet to expand our expression for Tb to arrive at form (Scott & Rees 1990), rendering the 21 cm line a direct ⎛ ⎞⎡ ⎤ probe of density fluctuations (at least to the extent that 3 ⎛ T ⎞ ⎜ 3cA10 ⎟⎢ xnHHI ⎥ g ) Tb(rˆ, n) = ⎜ ⎟ ⎜1.5- ⎟ () hydrogen traces the overall matter distribution . This era stops ⎝ 16k n 2 ⎠⎣ ()()1 + zdvdr2 ⎦⎝ T ⎠ b 21  s being observationally accessible at z∼30, when collisions between neutral hydrogen atoms become rare enough that the One thing to note is that in the high spin temperature limit the spin degrees of freedom are no longer coupled to the kinetic observed brightness temperature is independent of the spin degrees of freedom, and the spin temperature rises until it temperature itself. This can be understood from the fact that at equilibrates with the CMB temperature. This is illustrated in the high temperatures all spin microstates are equally occupied and middle panel of Figure 1, where we show the global 21 cm thus the observed brightness depends only on the rate of signal. With the CMB temperature and the spin temperature in spontaneous emission rate from the high energy state (A ). 10 equilibrium, the brightness temperature contrast goes to zero Observationally, T can be probed in two ways. One is to b and there is no 21 cm signal to observe. measure the global signal, where T is averaged over all angles b The Dark Ages are of tremendous cosmological interest. The on the sky to produce a single averaged spectrum T : b epoch provides access to a huge number of Fourier modes of the matter density field (Loeb & Zaldarriaga 2004). These TdTbb()nn=Wò (rˆ,.)() 6 fluctuations are graphically depicted in the high redshift regions Another is to try to measure the spatial fluctuations in the full of the top panel of Figure 1. They are a particularly clean probe 21 cm brightness temperature field, either by attempting to of the matter field since this is prior to the formation of the first reconstruct Tb (rˆ, n) or by quantifying the statistical properties luminous sources, thus obviating the need for the modeling of of the fluctuations. For most of this paper, we will be complicated astrophysics. The theoretical modeling effort is in discussing the measurement of spatial fluctuations, but we will fact even simpler than, say, that needed for galaxy surveys, return to the global signal in Section 15. since at these high redshifts, matter fluctuations are in the linear Regardless of the type of observation, one sees from perturbative regime even at very fine scales. This is important Equation (5) that the 21 cm brightness temperature contains a because the spatial fluctuations exist to extremely small scales, rich variety of effects. For example, at certain redshifts Ts is as they are not Silk damped and therefore persist down to the strongly coupled to the baryonic gas temperature (see sections Jeans scale (Tegmark & Zaldarriaga 2009). The result is a vast below). This makes Tb a high-redshift thermometer (albeit a number of modes that can be in principle be used to probe rather indirect one). It is also a probe of the ionization state of fundamental cosmology. The long lever arm in scale, from the hydrogen, via its dependence on xH I. We therefore see that Tb largest structures to the smallest structures, enables constraints is likely to be an excellent probe of many high-redshift on the running of the matter power spectrum’s spectral index astrophysical processes. In addition, it is sensitive to cosmol- (Mao et al. 2008). This would represent an incisive probe ogy, since the distribution of hydrogen (entering via the factor of the inflationary paradigm. Future measurements may also of nH) is driven by the large scale cosmological distribution of be able to detect features in the primordial power spectrum matter, as is the dv/dr velocity term, since it includes peculiar (Chen et al. 2016), or to detect primordial non-Gaussianity velocities in addition to the Hubble flow. Of course, with all of (Muñoz et al. 2015). In addition, the 21 cm line can be used to the aforementioned effects contributing to Tb, it may be probe the existence of relic gravitational waves from inflation, difficult to cleanly probe any of them. Fortunately, different either through their direct effects on large scale structure

3 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Table 1 Glossary of Important Mathematical Quantities

Quantity Meaning/Definition Context

ν21 Rest frequency of 21 cm line Section 2.1 Ts Spin temperature of the 21 cm line Equation (1) Tγ Temperature of the Cosmic Microwave Background Section 2.1

Tb or T Brightness temperature of the 21 cm line Section 2.1 Tb Globally averaged 21 cm brightness temperature Equation (6), Section 15 H0 Hubble parameter today L Dc Line of sight comoving distance Equation (36) z Redshift Equation (3) xH Neutral hydrogen fraction Section 2.1 WH I Normalized neutral hydrogen density Section 2.4 bH I Neutral hydrogen bias Section 2.4 b Baseline vector Section 3.1 u Normalized baseline vector/Fourier dual coordinate to angle in flat-sky approximation Equations(9), (10), (11)

Ap Power primary beam Equation (12), Footnote 11 V Interferometric visibility, cosmological survey volume, or Stokes V (depending on context) Equations (12), (20), Section (5.2) Q, U Stokes Q and U parameters Section 5.2 P(k) Rectilinear power spectrum Equation (16) f Survey volume function, frequency tapering function, or bispectrum phase (depending on context) Equations (17), (175) x ()x Correlation function Equation (22) Δ2 (k) “Dimensionless” power spectrum/variance per logarithmic k bin Equation (23) aℓm Spherical harmonic expansion coefficient Equation (25) Yℓm (rˆ) Spherical harmonic basis functions Equation (25) Cℓ Angular power spectrum Equations (26), (27) k^ Fourier wavevector perpendicular to the line of sight Equation (31) k Fourier wavenumber parallel to the line of sight Equation (31) V Delay transformed visibility Equation (139) Wp Integral of power beam Equation (42)

Wpp Integral of square of power beam Equation (42)

Tsys System temperature Equation (41) E(x, t) Electric field measured at position x at time t Equation (43) e(rˆ, n) Contribution of electric field from direction rˆ and frequency ν Equation (43) A Electric field/voltage beam of an antenna Equation (43) g Instrumental gain Equation (71) B Instrument transfer matrix Equation (79) N Noise covariance matrix of data Equation (84) S Signal covariance matrix of data Section 10.1 C Total covariance matrix of data Equation (116) Ea Quadratic estimator matrix for αth power spectrum bandpower Equation (113) Qa Response of the covariance matrix to the αth power spectrum bandpower Equation (118) W Window function matrix Equations (121), (122), (135) Vab Error covariance between αth and βth power spectrum bandpowers Equations (123) and (125) qˆa Unnormalized power spectrum bandpower estimates Equation (132) pˆa Normalized power spectrum bandpower estimates Equations (113) and (133) M Power spectrum normalization matrix Equation (133) F Fisher information matrix Equation (131)

Note. The “context” column gives Equation references, typically either their defining Equation or their first appearance in the text.

(Masui & Pen 2010) or through their lensing effects (Book measurements during the Dark Ages provide a good platform et al. 2012). Tests of statistical isotropy and homogeneity may for detecting exotic phenomena beyond the standard models of also be possible (Shiraishi et al. 2016). Small-scale measure- physics and cosmology (see references in Furlanetto ments enable measurements of the neutrino mass (Mao et al. et al. 2019). 2008) and constraints on the existence of warm Measurements during this era are extremely challenging. Not (Loeb & Zaldarriaga 2004). Finally, the cleanliness of only do the extremely long wavelengths (λ≈7to70m) make

4 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Figure 1. A graphical representation of a simulated 21 cm signal from z ~ 90 (during the Dark Ages) to the z ~ 7 (the end of reionization in this particular simulation). The top panel shows the 21 cm brightness temperature contrast with the CMB, displayed as a two-dimensional slice of a three-dimensional volume. Because the line of sight distance for the 21 cm line is obtained via the redshift, the horizontal axis also doubles as a redshift/evolution axis (i.e., the top panel is a picture of a light cone, which is what a real telescope would observe). The middle panel shows the predicted global signal of the 21 cm line, where the spatial information in the vertical direction of the top panel has been averaged over to produce one average brightness temperature per redshift. The bottom panel shows the power spectrum, expressed as Dº23()kkPk ()2p 2, as a function of redshift. This can be thought of as a measure of the variance of spatial fluctuations as a function of spatial wavenumber k (see Equation (23)), with the solid line showing k = 0.1 Mpc-1 and the dotted line showing k = 0.5 Mpc-1). These simulations were generated from the 21cmFAST semi-analytic code (Mesinger et al. 2011;https://github.com/andreimesinger/21cmFAST), and are publicly available athttp:// homepage.sns.it/mesinger/EOS.html as part of the Evolution of 21 cm Structure project (Mesinger et al. 2016). (A color version of this figure is available in the online journal.) simply building an instrument with the required resolution and state of the intergalactic medium (IGM) in a spatially and sensitivity extremely difficult, but as we will discuss later, the temporally non-trivial way. This is reflected in the spatial foregrounds are extremely bright (Sections 7 and 12). fluctuations and redshift evolution of the 21 cm line, which are Additionally, the ionosphere (Section 12.3) is opaque at no longer governed by the relatively simple physics of frequencies lower than a few MHz, and can cause distortions cosmological matter perturbations alone. This is both a defect even at higher frequencies. and an opportunity. It is a defect because the complicated astrophysics of the era (often loosely5 referred to as “Cosmic

2.3. High Redshift: Cosmic Dawn and the Epoch of 5 There is unfortunately no consistent definition of Cosmic Dawn that is Reionization agreed upon in the literature. Some authors use it to refer to a period that began when the first stars formed, and ended with the formation of larger . As the first luminous objects begin to form, the 21 cm line Others define the end of Cosmic Dawn to be when the first galaxies began to systematically reionize the IGM. Yet others consider Cosmic Dawn to be a ceases to directly the dark matter distribution. Radiation broad term that encompasses the entire period from the formation of the first from these objects affects the spin temperature and ionization stars to the end of reionization.

5 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

spin temperature Ts must also be cooler than Tg. The result is an absorption signal that contains information about both the density field and the Lyα flux. As formation continues, the Lyα background eventually becomes sufficiently strong for the coupling between Lyα photons and gas kinetics to be extremely efficient everywhere. Fluctuations from the Lyα background then become negligible. 2. With continued star formation (and eventually galaxy assembly), we enter a period of X-ray heating. High- energy photons from the first luminous objects have a heating effect on the IGM, as these photons cause photoionizations of H I and He I, with the resulting photoelectrons colliding with other . This can result in heating, further ionizations, or atomic excita- tions. The heating contribution raises the gas temperature to be above Tg, which makes the 21 cm signal go from Figure 2. An example 21 cm brightness temperature field at z = 8.1, generated absorption to emission. Because the astrophysical sources using the 21cmFAST semi-analytic code (Mesinger et al. 2011). With the particular astrophysical parameters and models used to generate this that heat the IGM are spatially clustered, this results in simulation, z = 8.1 corresponds to a volume-weighted neutral fraction of spatial fluctuations in the 21 cm signal. Eventually, ∼0.55. Ionized regions are shown in black, with zero brightness temperature however, the entire IGM is heated and the fluctuations since there is no 21 cm transition when there is no neutral hydrogen present. sourced by X-ray heating disappear. fl The spatial uctuations shown here encode a wealth of information about both 3. Cosmic Dawn ends with the Epoch of Reionization astrophysics and cosmology, which 21 cm experiments aim to extract. (EoR), when sustained star formation provides enough (A color version of this figure is available in the online journal.) ionizing photons to systematically ionize the IGM. Reionization does not proceed uniformly because the Dawn”) means that it is hard to use the 21 cm line as a clean, objects responsible for producing the ionizing photons ( ) model-independent probe of fundamental cosmology.6 On the likely galaxies are clustered. This is illustrated in other hand, this very complication makes the 21 cm line a Figure 2, where ionized bubbles form and grow around tremendously promising tool for understanding the nature of regions of high matter density because it is those regions ( the first luminous objects. that contain the most galaxies and therefore produce the ) Figure 1 illustrates a series of three epochs following the most ionizing photons . Since the ionized regions contain 7 Dark Ages: almost no neutral hydrogen, the 21 cm brightness temperature is close to zero there. The intricate pattern α 1. Cosmic Dawn begins with a period of Ly coupling, of ionized versus neutral regions gives rise to strong which runs from z∼20 to z∼12 for the model shown spatial fluctuations in the 21 cm line. These fluctuations in Figure 2. As the first stars are formed, they produce persist until the ionized bubbles are sufficiently large and significant amounts of Lyα flux. This causes the numerous that they overlap, and reionization of the entire Wouthuysen–Field effect, whereby Lyα photon absorp- IGM is complete. tion promotes an electron in a neutral ( from an n=1 state to an n=2 state, only to be followed From the bottom panel of Figure 1 which shows the fl by a decay to a different hyperfine state when the electron uctuation power as a function of redshift; see Section 3.2 for fi ) returns to the n=1 state. This enables Lyα photons to rigorous de nitions we see that each of the three epochs shows fl induce 21 cm spin-flip transitions, and because of the a rise-and-fall pattern: the uctuations increase when a source ( α ) large cross-section of Lyα scattering, this causes the spin of inhomogeneity Ly coupling, X-ray heating, or ionization becomes relevant, peak when (loosely speaking) roughly half temperature Ts to be coupled to the gas temperature. As was the case during the Dark Ages, the gas temperature is of the volume is affected, and fall once the relevant process has uniformly affected our universe. This provides a set of cooler than Tg in this epoch, which in turn means that the distinctive signatures to search for in observational data 6 One exception to this may come from the effect of velocity-induced acoustic oscillations (Dalal et al. 2010; Fialkov et al. 2012; McQuinn & O’Leary 2012; 7 The “ionized” regions shown in Figure 2 are not entirely ionized. At smaller Muñoz 2019a), which have the potential to serve as clean standard rulers at scales (which are unresolved in the simulation), there exist galaxies and neutral high redshifts, thus enabling precision measurements of the Hubble expansion gas clouds that are sufficiently dense to be self-shielded from ionization rate (Muñoz 2019b). (Sobacchi & Mesinger 2014; Watkinson et al. 2015).

6 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

(Lidz et al. 2008; Christian & Loeb 2013). However, it is remaining H I to be found within a range of halo masses 10 13 important to stress that even if our qualitative description of ∼10 to 10 Me. At the lowest mass end halos are not dense Cosmic Dawn is accurate, the dearth of direct IGM observa- enough to effectively self shield, and at the high mass end tions at z>6 mean that theoretical models remain fairly where we are considering clusters, tidal stripping and ram unconstrained, and there still remains a wide landscape of pressure remove gas from within the constituent galaxies possible 21 cm predictions. For instance, the precise timing of (Villaescusa-Navarro et al. 2018). Although the H I that the three Cosmic Dawn epochs can be easily varied, to the remains can be either part of the cold (T  100 K) or warm extent that plausible models can be constructed where some of (T  5000 K) neutral medium, both temperatures are warmer the peaks in fluctuation power merge. As another example, than the CMB temperature within this redshift range and we some models find that if z<8 observations of galaxies can be expect to see the 21 cm line in emission. extrapolated to higher redshifts, the 21 cm signal may in fact This redshift range is the most probed observationally. At very never go into emission, remaining in absorption through the low redshift HIPASS (z<0.03, Barnes et al. 2001) and reionization process (Mirocha et al. 2017). Uncertainties in ALFALFA (z<0.06, Jones et al. 2018) have detected individual feedback prescriptions add to a wide variety of possible extra galactic objects through their 21 cm emission, allowing ( ) theoretical predictions Fialkov et al. 2013 . It is also unclear as constraints on the mass function and abundance of low redshift to what the dominant sources of reionization are. While a fi -4 neutral hydrogen, nding that W~´H I ()3.9 0.6 10 majority of recent models in the literature assume that galaxies (Jones et al. 2018). At higher redshifts cross correlations with ( are primarily responsible for reionization see, e.g., Robertson optical tracers have allowed detection of 21 cm emission. This ) et al. 2010 , it would be fair to say that there is still was initially performed with the Green Bank Telescope considerable room for at least some contribution from quasars combined with DEEP2 and later WiggleZ, which estimated the (Madau & Haardt 2015). Given that there are even more ( ) ∼ hydrogen, abundance times the bias factor bH I at z 0.8 to uncertainties in our understanding of Cosmic Dawn than just be W=b 6.2+2.3 ´ 10-4 (Switzer et al. 2013). the few that we have listed here, current theoretical models HHII -1.5 The measurements described above are all direct measure- should in general be treated as helpful guides rather than ments of 21 cm emission but at higher redshifts we can use ironclad standard models. measurements of Damped Lyman Alpha systems to constrain Fortunately, direct observations of the 21 cm line at the the H I abundance up to z = 4.9, finding W=9.8+2.0 ´ 10-4 redshifts of Cosmic Dawn should significantly enhance our H I -1.8 at z = 4.9 (Crighton et al. 2015). Together this means that understanding. For example, the typical sizes of ionized unlike other epochs there is a well defined target for the bubbles will place constraints on the nature of sources responsible for reionization (McQuinn et al. 2007). A single amplitude of the signal. 21 cm measurement of the midpoint of reionization (performed, As we expect the H I to be hosted in generally quite low perhaps, by locating the peak of ionization fluctuation power) mass halos, and the fraction not bound within haloes to be ( = ) would inform the entire timeline of reionization when small 10% at z 5, smaller at low redshift the neutral combined with CMB and Lyα forest data (Pritchard et al. hydrogen is an excellent tracer of the total mass distribution ( ) 2010). Detailed studies of 21 cm fluctuations over a wide Villaescusa-Navarro et al. 2018 . This makes low redshift variety of spatial scales and redshift will enable constraints on 21 cm observations an excellent probe for cosmology, allowing the high-redshift universe, shedding light on details such as the us to pursue the kinds of science that are currently done by minimum mass for halos to host active star-forming galaxies, or galaxy redshift surveys, but within the radio band. However, the X-ray luminosity of galaxies, among other parameters unlike galaxy redshift surveys we do not need to resolve (Pober et al. 2014; Mesinger et al. 2014; Fialkov et al. 2017; individual galaxies (which is required only to determine their Greig & Mesinger 2017; Kern et al. 2017; Park et al. 2019).In redshift), and can instead map the unresolved emission of all addition to other possible constraints that we have not H I at each frequency of observation, which we can map discussed above, direct 21 cm measurements will also play directly to a redshift. The angular resolution is chosen to match the important role of allowing models of Cosmic Dawn to be scales of scientific interest, and each observed voxel may tested, rather than assumed. contain hundreds to thousands of individual galaxies. This is the idea of Intensity Mapping (Battye et al. 2004; Peterson et al. 2006; Chang et al. 2008; Wyithe & Loeb 2008), hereafter 2.4. Low Redshift: Post-reionization referred to as IM. At low redshift, after the end of reionization, the neutral fraction Observations within this epoch are mapping the large-scale ( hasbeendrivendowntoxH I ~ 2% Planck Collaboration et al. structure of the universe, giving them the ability to do much of 2016b; Villaescusa-Navarro et al. 2018). The gas that remains is the same science as optical galaxy redshift surveys. However, in dense regions that have managed to self-shield against the as they can naturally have wide fields of view and observe all ionizing background. Using simulations as a guide we expect the redshifts in their band simultaneously, IM experiments are

7 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw more readily able to quickly survey very large volumes than Although one only directly measures the inhomogeneity in optical galaxy redshift surveys. And as they do not need to the distribution of H I with 21 cm IM, by looking at the resolve individual galaxies, they can more easily push to high distortions in the observed three-dimensional field caused by redshifts, through the redshift desert (z∼1–3, where optical gravitational lensing, one can in principle infer the total matter spectroscopy is challenging) and beyond (3  z  6). distribution within the volume (Foreman et al. 2018). As the As IM is a very recent technique that is still being developed, lensing displacements are typically small, upcoming 21 cm the initial target within this era are to measure Baryon Acoustic experiments may only be able to see this effect in cross Oscillations (BAO). These are the remnants of primordial correlation with photometric redshift surveys, but it will be sound waves that leave a distinct signature in the correlation extremely powerful for following generations of instruments. structure of matter, and they can be used as a standard ruler, Finally, although most of the science targets we have albeit a statistical one, to measure the expansion history of the outlined above are cosmological, there is tremendous astro- universe (Eisenstein et al. 1998; Seo & Eisenstein 2003). The physical interest in the nature of H I at low redshift. IM by itself BAO signature is quite distinct and thus more robust to gives us direct access to WH I(z), the total amount of neutral systematics, making it an ideal first science target for 21 cm IM hydrogen across redshift, and by using cross-correlations experiments. As a bonus, one can push to ranges in redshifts against optical tracers we can start to obtain information about ( ) not currently probed by existing optical surveys (Peterson et al. which galaxies host the H I Wolz et al. 2017b . 2006; Chang et al. 2008; Wyithe et al. 2008). By constraining the expansion history of our universe we hope to be able to 3. Observational 21 cm Cosmology infer the properties of Dark Energy, in particular its equation of The broad scientific possibilities outlined in Section 2 have state w(z), which can yield clues to a microphysical explana- led to a new generation of instruments that have been custom tion. Such constraints on the equation of state could potentially designed for 21 cm cosmology. These instruments vary in their shed light on recent tensions in measurements of the Hubble detailed specifications, but generally: parameter (Knox & Millea 2020), particularly in the context of 1. Have high sensitivity. The cosmological 21 cm signal is proposed explanations that have non-trivial time evolution of faint across all redshifts, and thus high sensitivity is dark energy at z  1 (e.g., Di Valentino et al. 2016, 2017; essential. This drives telescope designers toward instruments Keeley et al. 2019). with large collecting area, coupled with observational plans As IM is able to survey very large volumes of the universe that call for 1000 hr or more of total integration time. with precise radial distances (unlike photometric redshift 2. Are broadband. A key feature of 21 cm cosmology is the surveys) it is ideal for discovering small statistical effects ability to map our universe in three-dimensions, using where measuring a large number of modes is essential. One redshift information to distinguish between emission target is looking for features in the power spectrum which from different radial distances. Doing so requires broad- fl ( ) might tell us about in ationary physics Chluba 2018 . Another band instruments whose receiving elements are efficient is looking for signatures of new physics by searching for non- over a broad frequency range, as well as backend ( ) Gaussianity in 21 cm data Li & Ma 2017 . electronics that are capable of simultaneously processing The expansion history information that IM can obtain from data over large bandwidths. Note that broad frequency BAO, combined with broadband measurements of the power coverage is helpful not only for achieving the ultimate spectrum shape, plus future CMB measurements gives a potent scientific goals of 21 cm cosmology, but also for combination for probing the contents of the universe. In achieving better calibration and diagnosis of systematics. particular the number of relativistic degrees of freedom and the 3. Are sensitive to the right scales. Spatial fluctuations in the sum of the neutrino masses (Oyama et al. 2016; Obuljen et al. cosmological 21 cm signal are expected to be scientifi- 2018) can be constrained to 20 meV by combining IM with cally interesting over specific ranges of length scales. For other probes. instance, for post-reionization observations the baryon On extremely large scales there are general relativistic acoustic oscillation scale is of particular interest, whereas corrections to the standard observables (Challinor & for reionization observations the typical scale of ionized Lewis 2011), that if observed could be a stringent confirmation bubbles would be a more appropriate scale to target. This of the current cosmological model. To measure these, one affects the physical extent of telescopes, although as we needs to map large volumes of the universe to build up enough discuss later in this section, the reasoning is subtle given samples of these scales (for which IM is ideally suited), and to the ability to probe radial fluctuations by using spectral combine with other probes on similar scales (e.g., photometric information. surveys like the Large Synoptic Survey Telescope) in order to 4. Are stable in time. The faintness of the cosmological remove sample variance effects (Alonso & Ferreira 2015). signal, coupled with the dominating influence of various

8 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

contaminants (see Section 7) means that 21 cm cosmol- and ogy is synonymous with high dynamic range observa- ¥ 22i pu·q tions. To ensure that strong contaminants do not TduTe()q =  ()u ,10 ( ) ò-¥ overwhelm a faint signal, it is crucial that one be able to minimize systematic instrumental effects. A stable with T ()q being the brightness temperature of the sky. A fi instrument helps to suppress the introduction of systema- baseline de ned by two antennas that are physically separated ( ) tics in the first place. by the vector b can then be shown as we do in Section 5 to probe T at 3.1. Single Dish Telescopes or Interferometers? b u = .11() With the aforementioned principles in mind, we examine the l trade offs of various types of radio instruments that have been To be even more quantitatively precise, under the flat-sky ( ) used or proposed for 21 cm cosmology. One such trade off is approximation the visibility V measured by a baseline b is the choice between a single dish telescope or an interferom- given by10 eter.8 In its simplest form, the former consists of a monolithic single dish that focuses radio waves into a feed, and the sky is 22-i plb·q VdTAe()b,,,,12nqnn= ()()qqp () mapped out one pixel at a time. The resolution θdish of the ò resulting map is then roughly determined by the characteristic 11 size D of the dish, where where Ap ()q is known as the primary beam and accounts for the fact that the antennas of an interferometer do not have equal l sensitivity to all parts of the sky. With this extra factor (along qdish = ,7() fi fl ) D with wide eld, non- at sky effects not shown here , one sees from comparing Equations (9) and (12) that a baseline b does and λ is the wavelength of observation. In contrast, an not precisely probe the Fourier mode given by u ==b l interferometer spreads out the total collecting area of a bn c. In particular, the convolution theorem allows us to write telescope into multiple receiving elements, such as multiple 2  ~ small dishes or dipole antennas.9 Signals from pairs of VduTAc()bubu,,,,13nnnn=-ò ()(p ) () antennas are then multiplied together and averaged over a which shows that rather than precisely probing u = bn c, the short period of time. This is the basic correlation operation, visibility probes a smeared out footprint around that mode. which is described in more detail in Section 5.4. The correlated However, if the primary beam is relatively broad, then this data (termed the visibility) from each antenna pair of the footprint is relatively localized, and our basic intuition remains interferometer roughly probes a Fourier mode that corresponds accurate. Under this approximation, each baseline probes a to the characteristic angular scale particular Fourier mode, and considering all the baselines of an l interferometer (i.e., all possible antenna pairings) then provides qint = ,8() b information about multiple Fourier modes of the sky. These Fourier modes can then be Fourier transformed back into the where b the distance between the two antennas that are being image domain to provide a map of the sky. The more unique correlated. It is given the symbol b because a pair of the baselines there are in an interferometer array, the closer our antennas are known as a baseline. More precisely, suppose one map will be to a true image of the sky, and the longer the is considering a small enough patch of the sky that the flat-sky baselines of an array, the higher the resolution of our map. Note approximation holds. We may then define flat-sky coordinates that as the Earth rotates, baseline vectors rotate relative to the ( q º ()qqxy, , and u is as its Fourier dual under a Fourier sky, and thus they rotate through the uv plane i.e., a two- convention where dimensional plane with Cartesian axes given by the two components of u). This is known as rotation synthesis, and ¥ TdTe()u º 22q ()q -i pu·q,9 () enables an interferometer to compile a more complete Fourier- ò-¥ space view of the sky, as illustrated in Figure 3. Of course, this

10 In this paper, we adopt an unusual convention. Typically, instead of the 8 Global signal experiments typically use single elements that are not dishes. brightness temperature T(ν) inside the integral in Equation (12), one has the In this section we are primarily concerned with experiments attempting to map intensity I(ν). We have implicitly assumed that one has already normalized the 22 the spatial fluctuations of the 21 cm signal, leaving global signal experiments to raw data using the Rayleigh–Jeans Law, i.e., I ()nnn= 2 kTb () c, where kb Section 15. is Boltzmann’s constant. 9 11 For simplicity, we will refer to each receiving element of an interferometer To be more precise, Ap is the power primary beam. In the literature it is not as an antenna for the rest of this paper. Unless otherwise stated, our discussions uncommon to see this denoted as A. However, in this paper we reserve A for will apply equally well to arrays of dishes. the electric field beam defined in Section 5.

9 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

zenith. Rather than concentrating all of one’s integration time on a single field, the telescope maps out different parts of the sky as the Earth rotates, surveying a stripe on the celestial sphere. The picture presented in Figure 3 is most suitable for tracking telescopes. Since tracking telescopes focus observa- tional time on a particular field, the coordinates needed to define the (qqxy, ) coordinate system (and by extension the uv plane) can be fixed to this single field. For a drift-scan telescope, on the other hand, the field is constantly changing, and therefore there is not a single uv plane that can be conveniently defined. In particular, once the Earth has rotated through an angle larger than roughly the width of the primary beam, one is essentially looking at a completely different patch of the sky. There are several ways to deal with this. One way is Figure 3. An illustration of rotation synthesis, showing the movement of fi baselines through the uv plane over a full 24 hr day of observations for a to de ne multiple uv planes for different patches of the sky, hypothetical interferometer comprised of 3×2 antennas arranged on a regular with interferometers accumulating uv samplings on one plane square grid. As the Earth rotates, the baselines sample different uv modes, for short integration times before jumping to a different plane although of course in practice it is difficult to get 24 hr of observations. (Cornwell & Perley 1992; Tegmark & Zaldarriaga 2010; Dillon ( Because a baseline b can equally well be considered a baseline -b simply by et al. 2015b). Another solution is to express the sky in a three- reversing the order of one’s antenna pairings), every sampled (u, v) point is accompanied by its conjugate (--uv, ) point. The conjugate points do not dimensional Fourier basis, where interferometric baselines are provide independent information, however, because the sky temperature must sampling a three-dimensional uvw space rather than a uv plane. be a real-valued quantity. Note that we have expressed our uv coordinates in However, in such a formalism the interferometric visibility units of λ, as is conventional. does not map as easily to a Fourier transform of the sky. A third (A color version of this figure is available in the online journal.) solution is to use the m-mode formalism (Shaw et al. 2014).A description of the m-mode formalism is provided in Section 10. view of the sky is still not perfect as there remain unsampled uv But in the current context of rotation synthesis, the essential modes. It is related to the true sky by a point-spread function idea is that because drift-scan telescopes map out a stripe on the known as the synthesized beam, which is given by the Fourier sky, we may describe the entire stripe using cylindrical fi ( transform of the nal uv distribution where a 1 is placed in coordinates, effectively unrolling the stripe into a periodic — — every pixelized cell on the uv plane a uv cell that is Cartesian grid. A new uv space can then be defined as the ) sampled that was used to make the map. Fourier dual to this periodic grid. This is worked out explicitly ’ Although this picture of an interferometer s baselines tracing in Appendix A, where one finds that for a drift-scan telescope, out tracks on the uv plane is a useful one, it has some important baselines do not trace out uv tracks. Despite this subtlety, the limitations. To understand some of these limitations, let us basic picture that we have presented—that at least instanta- distinguish between two types of interferometers: tracking neously, an interferometric visibility probes a Fourier mode of telescopes and drift-scan telescopes. A tracking telescope is the sky—remains a useful bit of intuition. one where antennas are steered to track a particular field in the Having established some basic intuition for interferometry, sky, such that the primary beam is also centered on a particular we now return to the question of whether one should build a coordinate. In addition, each baseline is multiplied by a time- single dish telescope or an interferometer. The answer depends dependent phase factor so that the complex exponential term in on one’s application. For example, if one requires high angular the visibility has zero phase at the center of the chosen field. (In resolution, it is typically advantageous to build an interferom- practice, this phase centring is sometimes done in software eter. Because large dishes are difficult to support mechanically, rather than by the electronic hardware of the telescope, since it a single dish telescope cannot be made arbitrarily large. There can be done after the fact.) In contrast, a drift-scan telescope is are thus practical limitations to the resolution of single dishes. one where the antennas are fixed and always point to the same With interferometers this is much less of an issue, as antenna location on the sky relative to the Earth (i.e., the same altitude elements can be placed extremely far apart. and azimuth).12 This point is often, but not always, the local On the other hand, many interferometers have their antennas laid out in a way that results in holes in their sampling of the uv 12 Of course, a tracking telescope can always be operated in a drift-scan mode, plane. This results in complicated synthesized beams. A single simply by not repointing the telescope. In the discussion that follows, we dish does not suffer from the problem of gaps in the uv plane. continue to use the term “drift-scan telescope,” but the reader should bear in mind that drift scanning should really be thought of as an observing mode While the limited angular resolution of a single dish means that available to all telescopes. maps created from single dish observations do not contain

10 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw information beyond some “radius” from the origin of the uv sufficient sensitivity that each Fourier mode is measured with plane, the modes interior to this cutoff correspond to angular signal-to-noise ratio of order unity or greater. In essence, the scales coarser than the angular resolution of the telescope, requirement for imaging is that the amplitude and phase of each which can be sampled with an appropriate observational mode is reconstructed with high fidelity, enabling Fourier strategy that points the telescope at various parts of the sky. modes to be combined in a way that yields a position space With a fully sampled uv plane (again, up to some cutoff), the image that resembles reality. Images are perhaps the ultimate point-spread function can thus be in principle inverted in one’s data product of a 21 cm cosmology experiment, since they derived maps. This can make it easier to deal with some of the capture the full details of the cosmological signal, including for systematics that we detail in Section 12.1.5 associated with example any non-Gaussian signatures that are present. contaminant emission in our data (such as the foreground Without the sensitivity for imaging, one must resort to emission from our own Galaxy and other astronomical sources; statistical measurements. Here, one takes advantage of the see Section 7). symmetries of the problem to combine individual modes In contrast to foreground systematics, noise systematics can together, thereby increasing signal to noise. For example, one be more difficult to deal with in single dish experiments. With a might use the fact that our universe is rotationally invariant single dish observation, the instrument effectively squares (i.e., statistically isotropic). This means that different Fourier electric field measurements to obtain measurements of power. modes with wave vectors of the same magnitude but different Doing so results in a noise bias, since mean-zero noise in the orientation should—up to statistical fluctuations—carry the pre-squaring data acquires a positive mean after squaring. This same information and be combinable in some way. noise bias must be exquisitely modeled and removed. In A prime example of a statistical measurement would be that contrast, with an interferometer one is correlating (i.e., of the power spectrum. Measuring the power spectrum is the multiplying) data from two different antennas to form a chief focus of most current-generation instruments, and thus baseline, and if the noise in the two antennas are independent much of this paper is devoted to discussing the power (which is frequently a good assumption), then the noise spectrum. Theoretical models of the spatial fluctuations of the contribution in the result will still be mean zero. The lone 21 cm signal typically make predictions for the power exception to this is the baseline b = 0, corresponding to the spectrum, providing a well-defined statistic for testing models. correlation of an antenna with itself. For this zero-length To define the power spectrum, we first define a three- baseline, the correlation is a squaring of the data, and one is dimensional Fourier transform of the sky: essentially treating the antenna as a single dish telescope. The ¥ 3 -ikr· noise bias is therefore present if one chooses to use zero-length TdreT()krº (),14 ( ) ò-¥ baselines. For this reason, interferometric observations often throw out the b = 0 correlation information, and thus where r is a comoving position vector, and k a comoving 13 interferometric images have zero mean, since the u = 0 mode wavevector. The inverse transform is is missing. ¥ 3 dk ikr·  In the past, both types of telescopes have been used for T ()rkº eT().15 ( ) ò-¥ ()2p 3 21 cm cosmology, and a detailed quantitative comparison between the two can be found in Tegmark & Zaldarriaga The power spectrum P()k is then defined via the relation ( ) 2009 . Notable single-dish efforts in the past have included áTT()kk (¢ )*ñ= (2,16pd )3 D ( kk - ¢ ) P () k ( ) observations using the Parkes telescope and the Green Bank D Telescope (see Section 4). These have resulted in detections of with δ being the Dirac delta function, and á...ñ signifying an post-reionization 21 cm fluctuations in cross-correlation with ensemble average, a hypothetical process where one imagines optical galaxy surveys (Chang et al. 2010; Masui et al. 2013; drawing different realizations of our universe from the Anderson et al. 2018a). However, recent efforts have tended to underlying statistical distributions that govern it. favor interferometry, given the flexibility to configure an array While Equation (16) is a formal, rigorous definition of the to have the right amounts of sensitivity on precisely the right power spectrum, it is not particularly useful for understanding scales. Exactly what these requirements are, however, depends how one computes a power spectrum in practice. This is on the type of observations that one is attempting. 13 At this point, we have (confusingly) introduced two different Fourier conventions. Unfortunately, both are standard: the convention with factors of 2π in the complex exponentials (Equations (9) and (10)) is traditionally used in 3.2. From Imaging to Statistical Measurements the radio astronomy literature, whereas the convention with no factors of 2π in the complex exponentials (Equations (14) and (15)) is used in the cosmology Broadly speaking, there are two types of observations that literature. As a general rule of thumb, if a quantity is defined in coordinates one can target when mapping diffuse structures. The first is most suited to describing observations (such as θ or u), it is likely that the radio astronomy convention is being employed; if a quantity is defined in terms of where the end goal is an image. Thinking about imaging in comoving cosmological coordinates, the cosmological convention is almost Fourier space, the process requires a telescope to have certainly being used.

11 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw because it assumes that one is able to take a Fourier transform where Nk is the number of k voxels that fall have magnitude of T ()r that is infinite in extent. If we imagine a finite (but between k−Δk/2 and k+Δk/2, where Δk is an analyst- otherwise still ideal and noiseless) survey, the Fourier modes defined bin width that is narrow enough to capture any wiggles that we measure are not the true T()k modes, but instead are in the power spectrum. This formula provides a more intuitive given by view of what a power spectrum measures: it is a measurement ¥ of the variance of fluctuations in a field (in our case, T) as a obs TdreT ()krr= ò 3 -ikr· ()f () function of lengthscale, or more precisely, k. Another way to -¥ think of the power spectrum is to show (using Equation (16)) ¥ 3 dk 15 = T()(kkk¢ f - ¢ ),17 ( ) that it is the Fourier transform of the correlation function ò 3 -¥ ()2p x ()xrrxºáTT () ( - ) ñ, i.e., where in the last equality we made use of the convolution dk3 theorem, and f()r is a function that equals one within the x()xk= Pe()-ikx· ,22 ( ) survey region, and zero outside. If we now compute the ò ()2p 3 absolute magnitude squared and take the ensemble average of the result, we obtain thus illustrating that the power spectrum is a measure of position space correlations in our data, but expressed in Fourier ¥ 3 3 obs 2 dk1 dk2 space. áñ=á∣()∣TkkkTT()12 ( )* ñ ò-¥ ()()22pp3 3 In Figure 4, we show some example 21 cm power spectra from 6<z<10. Instead of plotting P(k), we have opted to ´-ff()()kk12* kk - ¥ 3 instead plot dk1 2 =-P()∣(kkkf )∣.18 ( ) 3 ò 3 11 k -¥ ()2p Dº2()k Pk ().23 ( ) 2p2 If the survey volume is large, then f is a relatively narrowly peaked function in comparison to a reasonably smoothly This quantity is frequently seen in the literature, and can be motivated by the following argument. Consider the variance of varying P()k . The latter can then be factored out of the integral, fi and invoking Parseval’s theorem then gives a zero-mean random temperature eld at some location. We can write this as áñ∣()∣Tobs k 2 P()k » ,19() ⎛ ¥ dk3 ⎞ V áñ=T 2 ()rk⎜ eTikr· ()⎟ ⎝ò-¥ ()2p 3 ⎠ where ⎛ ¥ dq3 ⎞* ´ ⎜ eTiqr· ()q ⎟ 32 3 ò 3 Vdr==òòff()rr dr () (20 ) ⎝ -¥ ()2p ⎠ ¥ dk3 dq3 is the volume of the survey. Now, with real data one cannot =áñeTTi()·kqr- ()kq ()* perform an ensemble average. However, if one is willing to ò-¥ ()()22pp3 3 ¥¥3 make the assumption of isotropy, then the statistical properties dk 2 ( ==DPk() dln k () k , ( 24 ) of our Fourier modes including the mean of their squared òò-¥ ()2p 3 0 magnitudes, i.e., their variances) should only depend on the ( ) wavenumber k º ∣∣k , and not on the orientation of k.14 This where in the penultimate equality we used Equation 16 and Δ2( ) allows one to replace the ensemble average with an average assumed isotropy. From this, we see that the quantity k over direction. If we imagine dividing Fourier space into may be thought of as the contribution to the position space discrete voxels at discrete spacings of k, then we can write this variance per logarithmic bin in k. It is often termed the average as dimensionless power spectrum, but in the case of 21 cm cosmology this may be somewhat of a misnomer, since it has ∣()∣Tobs k 2 åkÎk dimensions of temperature squared. Pk()» ,21() While power spectra may not contain the richness of an NVk image, they do capture much of the relevant physics. For

14 example, one sees from Figure 4 that as reionization proceeds, This is not true when one includes the effect of redshift space distortions, ( k) ( where peculiar velocities are confused for the Hubble flow and lead to an power moves from small scales high to large scales small incorrect mapping of observed redshifts to radial distances. Since this sort of mis-mapping only happens radially, it causes a measured power spectrum to 15 Confusingly, in the nomenclature of statistics, the correlation function as we depend on the direction of k, even if the true underlying power spectrum is have defined it here would be more properly termed a covariance function. isotropic and only depends on k. See Section 15.1.1 for an example of an However, because such terminology is standard in the cosmology literature, we expression for P(k) that includes redshift space distortions. will continue to use it.

12 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

where Yℓm denotes a standard spherical harmonic basis function and aℓm its expansion coefficient. In analogy to Equation (16), we may define the angular power spectrum Cℓ as

áñ=aaℓm ℓm*¢¢ ddℓℓ¢ mm¢ C ℓ,26() and the analog to Equation (21) is ℓ 1 2 Cℓ = å ∣∣aℓm .27 () 21ℓ + mℓ=- The angular power spectrum has been the workhorse statistic for decades in the CMB community, since the CMB resides on a two-dimensional surface on the sky. It has been less commonly used in 21 cm cosmology because of the three- dimensional nature of 21 cm surveys. One option would be to create a separate angular power spectrum for each frequency. However, since each frequency is treated in a way that is Figure 4. Example 21 cm power spectra from the reionization epoch, when the independent of all others, this method is unable to capture global neutral fraction xH I decreases from xH I ~ 1 to xH I ~ 0. These were generated using 21cmFAST (Mesinger et al. 2011), and for the particular set of valuable information on correlations in the 21 cm data along theoretical parameters chosen, most of reionization happens over the redshift the line of sight. An alternative is to compute all possible range 6<z<10. Although power spectra do not contain the full information angular cross-power spectra between frequency channels, i.e., content of an image when non-Gaussianities are present, they do capture many important features of the underlying physics. We note that these curves are 1 +ℓ Cℓ ()nn, ¢ º å aaℓm ()() nℓm* n¢ ,28 () intended to be for reference only, as there are considerable modeling 21ℓ + uncertainties both within models (due to uncertain values of free parameters) mℓ=- and between models. where aℓm (ν) is the spherical harmonic expansion coefficient (A color version of this figure is available in the online journal.) for mode Yℓm of the map at frequency ν (Santos et al. 2005; Datta et al. 2007; Bharadwaj et al. 2019). This is essentially a hybrid statistic, being a harmonic space quantity in the angular k) as the characteristic size of bubbles increases,16 until the directions but a correlation function in the frequency/line of overall global neutral fraction has decreased sufficiently for the sight direction. An alternative to Cℓ (ν, ν′) that retains its power to drop to zero on all scales. Power spectrum spherical geometry in the angular direction but is expressed in measurements can be used to constrain the underlying harmonic space in all three directions is the spherical Fourier- parameters of a model, and in Section 13 we outline how (in Bessel power spectrum (Liu et al. 2016). To compute it, one general) one can go about performing this data analysis step. decomposes the sky into spherical Bessel functions in the radial Power spectra can also be used to select between models, and direction and spherical harmonics in the angular direction to we discuss this in Section 13.3. obtain In addition to the basic power spectrum that we have defined fi ¥ above, there exist variations that can be useful in speci c data 2 32 Tkℓm ()º drrjkrYℓ ( )ℓm* (rrˆ)() T ,29 ( ) analysis contexts. If one is dealing not with a three-dimensional p ò0 volume of data but instead with observations on the surface of a where jℓ is the ℓth order spherical Bessel function of the first sphere, it is more appropriate to construct an angular power kind. The spherical Fourier-Bessel power spectrum Sℓ(k) is then spectrum. Here, one decomposes maps into spherical harmonic computed as modes rather than Fourier modes, so that17 ℓ 1 2 ¥ +ℓ Skℓ ()µ å ∣Tkℓm ()∣,30 ( ) TYa(rrˆ)(= åå ℓmˆ)() ℓm,25 21ℓ + mℓ=- ℓmℓ=0 =- where for clarity we have omitted constants of proportionality. These constants depend on one’s survey volume, just like with 16 This may be a reasonable story for explaining the trends seen in the the analogous expression for the power spectrum in evolution of the power spectrum, but it is not one that is rigorous. See ( ) Furlanetto & Oh (2016) for a considerably more well-defined and subtle view Equation 21 . If the constants are included, then one can of how ionized bubbles grow during reionization. show that averaging Sℓ(k) over all ℓ yields P(k). 17 For simplicity, here we assume that one has surveyed the entire sky. A full Although Sℓ(k) ultimately just gives P(k) when averaged over discussion of the more general case for incomplete sky coverage is a little more ℓ involved than it was for rectilinear power spectrum, and so we omit it here. , it is extremely useful as an intermediate data product because Interested readers may wish to consult resources such as Alonso et al. (2019). it explicitly indexes data by ℓ. In other words, Sℓ(k) explicitly

13 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw picks out angular fluctuations, allowing one to focus on only have made the same approximation as in Section 3.1, where we those fluctuations that occur on a particular angular scale. This assumed that Ap was reasonably broad, and have defined is extremely useful because 21 cm experiments survey the sky in starkly different ways in the angular and radial directions— TddeT()u,,.34hqnº ò 22-+i phn(·)u q ()q n () fluctuations in the angular direction are probed by different pointings of a single-dish telescope or by an interferometer’s In going from Equations (13) to (33) we have neglected the  sampling of the uv plane, whereas fluctuations in the radial frequency dependence of Ap, evaluating the function at some ν direction are probed by analyzing data at different frequencies central frequency 0. The intrinsic frequency dependence of the  (i.e., by looking at the redshift of the 21 cm line). Systematics primary beam (i.e., the second argument of Ap) can be in one’s data will therefore often have telltale signatures in a minimized in hardware, and this is often a chief consideration harmonic space that separates out angular from radial in designing one’s antenna elements. The dependence arising fi ( fluctuations. This is partially accomplished by Sℓ(k), which from the bn c term in the rst argument due to the fact that a fi separates out the angular fluctuations withℓ. But Sℓ(k) retains a baseline of xed physical length samples different angular dependence on k, which describes fluctuations by wavenumber, scales at different frequencies) may be a reasonable approx- regardless of orientation. A cleaner separation can be obtained imation for very short baselines, and we will take advantage of in the flat-sky limit, where the analogous quantity to Sℓ(k) is the this in Section 11.2. However, the relaxation of this assumption cylindrical power spectrum P(k⊥, k).Todefine P(k⊥, k),we has profound consequences for avoiding the foreground note in the narrow-field, flat-sky limit, the radial direction contaminants that we will describe in Section 7. We will essentially points along a particular Cartesian direction. This explore this in detail in Section 12.1.5. allows us to designate coordinates along this direction as r, For now, however, we can combine our last two equations to and those along the two directions perpendicular to the line of give sight as r^. The corresponding Fourier dual coordinates k⊥ and ~ fi VddeT()b,,.35hqnµ 22-+icphn(·b q n0 ) ()q n () k are then de ned implicitly via ò ¥  2 -+ikr(·kr^^  ) ( ) Tk()kr^,,,31º ò drdre^ Tr ()^  () This is very similar to Equation 31 except that there the -¥ relation is written in terms of cosmological coordinates such as and the cylindrical power spectrum is then r^ and r, whereas here our expression is in terms of observational coordinates q and ν. To relate these two sets of 1 ~obs 2 Pk()^, k» åå ∣()∣()Tkk^,. 32 coordinates, we begin with the line of sight (radial) comoving NV kk^,  kk^^ÎÎkk distance Dc to redshift z (Hogg 1999) Just as Equation (21) represents an averaging of power z c dz¢ 3 spectrum estimates over spherical shells in Fourier space to Dc º ò ;136Ez()ºWm ( + z ) +WL ( ) H0 0 Ez()¢ yield P(k), Equation (32) tells us to form P(k⊥, k) by averaging ± over pairs of rings of radius k⊥ located at k, with each ring where c is the speed of light, H0 is the Hubble parameter, Ωm is Ω containing Nkk^,  independent k modes. the normalized matter density, Λ is the normalized dark energy density, and z is the redshift of observation, which is 3.3. Designing an Interferometer: Achieving Sensitivity to related to the observational frequency ν via Equation (3). the Right Modes Under the small-angle approximation, Dc relates q to r^ via

The cylindrical power spectrum provides a way to separate r^ = Dc q.37() fluctuations in the maps into fluctuations along the line of sight ν and fluctuations perpendicular to the line of sight. This can be For relating to r, one might be tempted to simply use the particularly illuminating for considering the optimal design of expression for Dc, since Dc gives a radial distance as a function an interferometer. Our strategy for this is to compare our of frequency. However, when computing power spectra from ’ interferometer’s measurement equation, Equation (13), to the data, one typically splits up the full bandwidth of one s Fourier modes that we wish to measure. If we take the Fourier instrument into relatively small bandwidth chunks. This is done transform of Equation (13) along the line of sight, we end up in order to avoid the effects of cosmological evolution. Thus, with the relevant relation is not the one between frequency and total radial distance Dc, but the local relation (defined relative to the ~ 2  ~ VduTAc()bubu,,hhnn=-ò ()(p 00 , ) middle of a sub-band) between changes in frequency Δν and changes in distance Δr. This is given by » Tc()bnh0 ,, () 33 2 η ν ( ¶Dcc ()1 + z where is the Fourier dual to although see Section 11.2 for a D=r D=-n Dn.38() discussion of the subtleties regarding this definition). Here, we ¶n H021n Ez()

14 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

If we now assume that both the frequency and the radial distance are measured relative to values appropriate to the approximate center of one’s observational band, we may replace Δr and Δν with r and ν, respectively, in a bit of notational convenience. Additionally, we are free to flip the direction of one of our Cartesian axes to rid ourselves of the minus sign in Equation (38). Inserting Equations (37) and (38), sans minus sign, into Equation (35) then gives ~ VdrT()b,,hnµ ò 3 ()q ⎡ ⎤ ⎛ HEznn() br· ⎞ ´-exp⎢ i 2p⎜ 021 r + ^ 0 ⎟⎥,39() 2  ⎣ ⎝ cz()1 + cDc ⎠⎦ and comparing this to Equation (31) reveals that 2pnb 2 pn HEz() k ==0210; k h.40() ^  2 cDc cz()1 + We therefore see that an interferometer’s baseline distribution determines which angular wavenumbers k^ it is able to access. Figure 5. A qualitative picture of the region of Fourier space that are accessible In particular, the finest angular scales (largest k^^º ∣∣k ) are to radio interferometers. The largest scale (lowest k⊥ wavenumber) angular limited by the longest baselines. The largest angular scales Fourier modes are limited by the extent of the survey region. The smallest ( ) smallest k^ values are limited by either the shortest baselines, scales are limited by one’s array configuration, since higher k⊥ are probed by or if the interferometer is extremely compact (e.g., with nearly longer baselines, and the number of baselines eventually thins out as one goes touching elements), the survey area. In the line of sight to longer baselines, even for arrays with a large number of elements. In the line ’ fi ( ) of sight direction, high k modes are limited by one s spectral resolution. The direction, the nest modes which reside at high k are limited ( ) ’ low k large scale modes are dominated by cosmic variance, the radial extent by the spectral resolution of one s instrument. The coarsest of one’s survey (i.e., one’s instrumental bandwidth), and foregrounds (see modes (at low k) are limited by the bandwidth over which one Sections 7 and 12). In addition, the inherent chromaticity of interferometers is able to collect data. These limits are illustrated in Figure 5, results in a further leakage of foregrounds to higher k, leading to the along with a “wedge” signature from foreground contaminants foreground wedge (see Section 12.1.5). The remaining region in Fourier space that is discussed in detail in Section 12.1.5. Of course, within is sometimes termed the Epoch of Reionization window, and is a promising region in which to pursue a first detection of the 21 cm power spectrum. the accessible region (often denoted the Epoch of Reionization (A color version of this figure is available in the online journal.) window18) the observations still contain noise that can only be reduced by averaging over independent measurements, for instance by averaging over an extended period of time. We distribution of k^ modes will inevitably push sensitivity to discuss noise in more detail in Section 5.4. higher k values. However, since we are primarily interested in To recap, the k modes are accessed via an interferometer’s ^  a three-dimensional mapping, there is another way to access spectral information while the pattern of k coverage is ^ small-scale information. Even with a single short baseline that determined by the baseline distribution. This leads to some probes low values of k⊥ (large angular scales), one can access instrument designs for 21 cm cosmology that may seem small-scale information along the line of sight direction. Said counterintuitive at first. Consider two qualitatively different differently, if one assumes statistical isotropy, then the power scenarios. One where the interferometer is designed to be as 2212 compact as possible, with many short baselines, and another spectrum depends only on k º+()kk^  , and one can where the interferometer has a more spread out configuration. If access a wide variety of length scales despite only sampling a one were conducting a two-dimensional mapping of the sky at very narrow range of k⊥. a particular frequency, the latter configuration would have its With two qualitatively different ways to access the same Fourier modes, it is natural to ask if one strategy is preferable sensitivity spread out over a greater variety of k^ modes (or equivalently, uv modes). This also enables one to reach modes over the other. In some scenarios, one does not have a choice. with higher k, i.e., finer spatial scales, since a more spread out For instance, suppose one were interested in measuring fine spatial scales (high k) at very high redshifts (e.g., the Dark 18 The origin of this name is historic, because the geometry of Figure 5 was Ages) using a hypothetical futuristic interferometer. Such a elucidated in the reionization literature. However, the qualitative structure of the plot applies at all redshifts, even if the precise locations of the various measurement will necessarily require reaching high k values by boundaries are different in detail. accessing high k modes. To see this, consider Figure 6, where

15 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

1/Nbl t, since the power spectrum goes as the square of V. This stands in contrast to an interferometer with a spread out configuration that has little redundancy. In such an inter- ferometer, different baselines will tend to sample different uv (and thus different k^) modes. Each of these modes contains different information about the sky, and this information cannot be coherently added. In other words, each baseline measures V()b, h at different b, and so one cannot directly average their measurements. The best one can do is to take advantage of statistical isotropy to average together statistics like the power spectrum, as suggested in Equations (21) and (32). Thus, in this regime the power spectrum error (instead of the V error) is what scales as 1 Ntbl , which is a slower scaling than what a

Figure 6. Maximum k⊥ (as a function of redshift) accessible to interferometers redundant array offers. with baselines of various lengths, compared to the maximum k accessible to an In the high signal-to-noise regime, however, one’s errors are interferometer with 10 kHz spectral resolution. At very high redshifts, the only dominated by cosmic variance. This is an error contribution practical way to measure Fourier modes with high wavenumber k is to measure that arises because the cosmological field that we seek to small scale modes along the line of sight (i.e., those with high k). (A color version of this figure is available in the online journal.) measure is random, and thus there is no guarantee that the amplitude of a particular Fourier mode is representative of the true underlying variance (i.e., the power spectrum) from which we plot the relations given in Equation (40). One sees that the is it is drawn. Though it is by definition rare, it is entirely cosmological scalings are such that to measure Fourier modes possible for a particular realization of a mode to lie in the of any appreciable wavenumber in the perpendicular direction extreme tails of a probability distribution. In the low signal-to- requires impractically long baselines. This is especially true noise regime this is not a concern, as the contribution to the when one accounts for the fact that measuring ultra high error budget from measurement uncertainties is larger than that redshift signals will require extremely high sensitivity, so one from cosmic variance. Once a particular mode has been needs a very large number of long baselines, and not just the measured to a signal-to-noise ratio of order unity, however, it is small handful that are often used in very long baseline no longer beneficial to continue integrating on this mode interferometry. (Tegmark 1997a). Instead, one gains by measuring other modes At lower redshifts where the baseline lengths are not that can be averaged together incoherently to reduce cosmic prohibitive, one’s strategy should be informed by the signal- variance. This favors arrays that are less compact and have to-noise ratio. In the low signal-to-noise regime one is limited fewer repeated redundant baselines. by sensitivity, and the optimal strategy is to observe with a Similar arguments can be made regarding the observational compact interferometer with antennas that are as closely packed strategy of a telescope. Just as rare fluctuations may cause a as possible. The reason for this is that compact interferometers particular Fourier mode to be unrepresentative of the under- will typically contain many identical copies of the same lying statistics of the sky—and therefore result in a misleading baselines (which therefore sample precisely the same k⊥ estimate of the power spectrum—the same can be true if one modes), especially if the antennas are placed on a regular grid samples only a small (possibly unrepresentative) patch of the (see Section 3.4). Moreover, rotation synthesis causes baselines sky. Given the finite field of view of a telescope’s primary to rotate into one another on the uv plane, further increasing beam, it is therefore necessary to consider the trade-off of this redundancy. This effect is particularly pronounced for short baselines, as a shorter “radius” in uv space causes baselines to whether one should concentrate observations on a small part of ’ spend more integration time in each uv cell. Thus, compact the sky or if one should spread out one s integration time over a ( ) interferometers are ones that concentrate their integration time larger portion of the sky Trott 2014 . This optimization is into a small number of Fourier modes. They measure a few precisely analogous to the optimization problem for inter- modes to high precision, rather than a large number of modes at ferometer layout: one should integrate on a particular patch of low signal to noise. the sky until the patch has been measured to unity signal to Concentrating on a select handful of modes results in very noise, at which point one should switch to another patch ( ) high sensitivity to the power spectrum. This is because Nbl Tegmark 1997a . And just as with our discussion of baselines integrating for time t on the same Fourier mode interferometer layout, the optimization of observational results in a coherent averaging of V, with the error averaging strategy has consequences for instrument design. Given the down as 1 Ntbl . The power spectrum error then goes as rotation of the Earth, the continuous observation of a small

16 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw patch of sky requires a tracking telescope. On the other hand, if Thyagarajan et al. 2017). However, this comes at the cost of one desires broad sky coverage, a drift-scan telescope suffices. carrying around extra zeros at grid points where there are no extra antennas, preventing an exact ()NNantlog ant scaling. Still, the flexibility of arbitrary antenna placement is enticing, 3.4. Costs and Systematics in Instrument Design and early prototypes of next-generation hardware correlators Our discussion so far has focused mostly on minimizing (whether based on regular antenna grids or not) have shown uncertainties in our observations, be they uncertainties from the promising initial results (Beardsley et al. 2017; Kent et al. measurements themselves or from cosmic variance. This 2019). motivated compact, regular interferometers in the low signal- In closing, we note also that redundant, regular arrays have to-noise regime, and extended interferometers with many the advantage of enabling non-traditional calibration strategies. unique baselines in the high signal-to-noise regime. The former Essentially, regular arrays have multiple copies of the same is also more appropriate for statistical measurements like baseline, and so while there may be Nant()N ant - 1 2 baselines measurements of the power spectrum, while the latter is more in total, the number of unique baselines is considerably smaller. appropriate for imaging. Since identical baselines should yield identical measurements However, in practice one must contend with issues of cost up to noise fluctuations if one’s array is correctly calibrated, and instrument stability: an ideal instrument from a sensitivity one can demand this property of mathematical consistency in standpoint may be prohibitively expensive, or may have one’s data analysis and work backwards to solve for an array’s various practical problems that result in serious systematic calibration parameters. We explore the problem of calibration errors. As an example, many current-generation 21 cm and provide context for why it is necessary in Section 9. telescopes (see Section 4) are drift-scan telescopes even though they are in a pre-detection low signal-to-noise regime. This 3.5. A Recipe for Estimating Sensitivity design choice obviates the need for a telescope with moving As a summary of our discussion on interferometer design, parts, and is therefore commonly made to increase instrument we provide a basic recipe for computing the errors on a power stability in an effort to minimize the possibility of systematics. spectrum measurement. For a tracking telescope, the recipe is ( fi ) Another example of what may at rst glance be considered as follows: a strange choice is the proposal of large, regular interferometric arrays for next-generation arrays that will presumably be deep 1. Given the antenna locations of an interferometer, into the high signal-to-noise regime. Under pure signal-to-noise compute the uv coordinates of all baselines. considerations, one would expect an irregular array that is more 2. Segment a uv plane into discrete cells, each of which has suited to imaging. However, regular arrays drive down the roughly the same width as Ap. Roughly speaking, this hardware cost of producing visibilities from an interferometer. width is the inverse of the instantaneous field of view As we shall see in Section 5.3, a baseline’s measured visibility of an antenna. Define a three-dimensional uvh space, is computed by cross-multiplying the measured voltages from extending the uv plane into a third dimension of Fourier the baseline’s constituent antennas. With Nant antennas one can space. Each cell in this dimension should have an extent form Nant()N ant - 1 2 pairs of antennas, and this is how many of 1/B, where B is the frequency bandwidth of the baselines worth of visibilities one must compute. Since this observation. 2 goes as Nant, the computational cost goes up rapidly with the 3. Simulate the movement of baselines through the uv cells size of one’s array. over a sidereal day of observations (or over however Futuristic array designs can circumvent this computational much time the telescope is being operated per sidereal bottleneck in a number of ways. With a regular array, one can day). Record how much time tu each baseline spends in in principle reduce the computational cost to ()NNantlog ant . each uv cell. To understand this, suppose the voltages measured by each 4. Assuming a roughly constant noise contribution across antenna were laid out on a regular grid, where the grid points frequency channels, the uncertainty (standard deviation) correspond to the locations of the antennas. Cross-multiplying from measurement noise for a given uvη cell is given by every voltage with every other voltage is then equivalent to Dc2 ()1 + z2 W2 T2 convolving the gridded function of voltages with itself. This c p sys ,41() operation is equivalent to computing a series of Fourier n21HEz 0 () Wpp 2tu transforms, thanks to the convolution theorem, and is therefore where realizable in ()NNantlog ant time using the Fast Fourier Transform (FFT) algorithm (Tegmark & Zaldarriaga 2 WºppppòòdA W(rrˆ)(;42 W º dA W p ˆ)() 2009, 2010). In fact, if the voltages are gridded onto a grid finer than one’s antenna spacing, even irregular antenna layouts and Tsys is the system temperature, which is the sum of can take advantage of FFT speedups (Morales 2011; the sky brightness temperature and the receiver

17 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

temperature (see Section 5.4). This is the uncertainty for between two samples in uvη space. A proper treatment one baseline. If multiple baselines spend time in a given therefore requires intentionally over-resolving the space while uv cell, the integration times of each baseline must be taking into account error covariances (rather than just error added together before being inserted in Equation (41). variances) between tiny uvη cells (Liu et al. 2014a). Despite The above expressions are derived in Morales (2005), these limitations, the recipe outlined above is reasonable for McQuinn et al. (2006), Parsons et al. (2012a)(albeit with approximate forecasting work, and is implemented (with different notations), but see Appendix B of Parsons et al. several convenient additions) by publicly available packages ( ) (2014) for a discussion of the subtleties in defining Ωp such as 21cmSense Pober et al. 2013b, 2014 . and Ωpp, which were missed by earlier works. 5. Divide the uncertainty computed above by the total 4. Current Status of Observations number of sidereal days tdays of observation. Note that the -1 -12 The last decade has seen a large increase in experimental standard deviation scales as tdays rather than tdays because 21 cm activity. A variety of instruments have been used to visibility measurements repeat every day and thus can be place increasingly stringent limits on the 21 cm power coherently averaged prior to the squaring step of forming spectrum (in the case of experiments targeting Cosmic Dawn a power spectrum, as discussed in Section 3.3. and reionization) as well as to detect spatial fluctuations in 6. Associate each uvh cell with its location in k space using cross-correlations with traditional galaxy surveys (in the case of ( ) ( ) Equations 11 and 40 . post-reionization experiments). Additionally, there has been a 7. For high signal-to-noise measurements it is necessary to recent tentative detection of the sky-averaged (“global”) 21 cm account for cosmic variance which contributes a standard signal at z∼17. We will return to a discussion of this result in ( ) deviation of P()k i.e., the cosmological signal itself to Section 15. every uvh cell. This is added to the instrumental For now, we review the instruments that have been used in uncertainty. attempts to measure the spatially fluctuating 21 cm signal. Each 8. Assuming that different cells in the 3D k space are instrument is optimized in subtly different ways, which has averaged together with an inverse variance weighting to implications not only for observational capabilities and ( ) obtain Pk( ^, k) or P k , the corresponding errors can be strategies, but also for analysis methods: combined in inverse quadrature. In other words, if e()k is ( ) the standard deviation for a k-space cell, then the final 1. The Giant Metrewave Radio Telescope GMRT . GMRT averaged error bars are given by (åe--212) , where the is a general-purpose low-frequency interferometer located sum is over all the k-space cells that contribute to a in India approximately 10 km east of the town of Narayangaon. It consists of thirty steerable dishes, each particular bin in (k⊥, k) or k. 45 m in diameter. Observations can be made around six There are several caveats to this (reasonably simple) recipe. frequency bands, with the lowest one centered around The first is that it assumes a tracking telescope. For a drift-scan 50 MHz and the highest around 1420 MHz (Kapahi & telescope, one should simulate the movement of baselines (in Ananthakrishnan 1995). Being a general-purpose instru- Step 3 of our recipe) for only the amount of time it takes for ment, its layout is not optimized for sensitivity to the Earth rotation to move through one primary beamwidth. The 21 cm power spectrum. Nonetheless, the GMRT Epoch primary beamwidth defines the size of a single patch of the sky of Reionization (GMRT-EoR) project was the first to that one might consider to be an independent observation. The provide upper limits at redshifts relevant to reionization observation of Npatch patches of the sky can then be accounted (Pen et al. 2009; Paciga et al. 2011, 2013). The analysis for by dividing the final errors by 1 Npatch . Of course, scaling pipeline used to derive these limits employed the pulsar the errors in this way is an approximation. For a rigorous gating calibration technique described in Section 9.1, and treatment, one should discard the notion of observing discrete the principal component-based methods for radio fre- patches and treat the full, curved sky in a single calculation, quency interference and astrophysical foreground mitiga- perhaps using a more suitable basis such as a spherical Fourier- tion described in Sections 12.4 and 12.1.3, respectively. Bessel basis (Liu et al. 2016) or the m-mode techniques of 2. The Murchison Widefield Array (MWA). The MWA is a Appendix A. custom-built interferometer located in the Shire of The other big assumption that our recipe makes is that the Murchison in the Western Australian desert. It is coherence scale of noise on the uv plane is given by the width designed for a wide variety of science cases (including of Ap, and that the coherence scale in η is given by 1/B. Milky Way science, extragalactic studies, physics, Measurements smaller than these scales are assumed to be and radio transients), but with Cosmic Dawn and Epoch perfectly correlated, while measurements spaced farther apart of Reionization studies as key motivators (Bowman et al. than these scales are assumed to be entirely uncorrelated. In 2013). Each element of the interferometer consists of a practice, noise correlations fall off smoothly with the distance square 4×4 grid of crossed dipole antennas (forming a

18 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

tile) whose signals are combined electronically. By elements centered in the Netherlands complemented by a changing the relative phases with which the signals from set of remote international elements (van Haarlem et al. each dipole are combined in a process known as 2013). Each interferometric element is known as a beamforming, each tile can be electronically “pointed” station, and is in fact made of two different types of at different parts of the sky. Each tile then serves as a antennas: a set of high-band antennas (HBAs) cover the single element in a 256-element interferometer operating 110–250 MHz frequency range and a set of low-band between 70 and 300 MHz (Tingay et al. 2013). At any antennas (LBAs) cover the 10–90 MHz range. The given instant, 30 MHz of bandwidth and 128 elements number of antennas in each station varies from location can be correlated. True to its broad science case, the tile to location, but the stations in the Netherlands (which layout of the MWA is a hybrid between a spread out non- give rise to the shortest baselines and therefore are the regular configuration and two closely packed, hexagonal ones with greatest sensitivity to cosmology) each have 96 grids of 36 tiles each (Wayth et al. 2018). The MWA has LBAs and 48 HBAs. Whereas each LBA is an individual published a number of upper limits over a broad range of dipole antenna, each HBA is in fact a tile of 16 antennas redshifts in recent years ranging from Cosmic Dawn (in a similar fashion to the MWA) tied together by an redshifts to reionization redshifts (Dillon et al. 2014, analog beamformer. At each station, the outputs from the 2015a; Beardsley et al. 2016; Ewall-Wice et al. 2016b; LBAs and the HBAs are digitized and then digitally Jacobs et al. 2016; Trott et al. 2016, 2019a, 2019b; Barry beamformed before the data from all the stations are sent et al. 2019). MWA analyses tend to involve multiple to a central correlator located at the University of pipelines for validation purposes (Jacobs et al. 2016), but Groningen. LOFAR is a multi-purpose observatory that most involve some combination of sky-based calibration accommodates a wide variety of key science projects, off of point source catalogs (Section 9.1), a hybrid including deep extragalactic surveys, transient phenom- approach for foreground mitigation involving the sub- ena, cosmic rays, solar science, cosmic magnetism, and traction of bright point sources and foreground avoidance cosmology. It has recently placed upper limits on both (Section 12.1.5), and map-making-based power spectrum Cosmic Dawn (Gehlot et al. 2019) and the EoR (Patil estimation methods (whether these maps are traditional et al. 2017). The analysis pipelines used to derive these images or in Fourier space; Section 11.1). Planned limits employed sophisticated direction-dependent cali- instrument upgrades will target improvements to the bration schemes (Section 9.4), and use either Generalized electronic hardware, enabling simultaneous correlation of Morphological Component Analysis (Section 12.1.3) or all 256 tiles over a larger instantaneous bandwidth Gaussian Process Regression (Section 12.1.2) for fore- (Beardsley et al. 2019). ground mitigation. 3. The Donald C. Backer Precision Array for Probing the 5. The Green Bank Telescope (GBT). GBT is a large 100 m Epoch of Reionization (PAPER). PAPER is an inter- single-dish telescope located in Green Bank, West ferometer that is optimized for EoR power spectrum Virginia (Prestage et al. 2009). Operating between measurements. As such, it embraced the idea of 100 MHz and 115 GHz, it is a general purpose instru- maximizing power spectrum sensitivity by arranging ment. It holds the distinction of having made a detection 112 dual-polarization dipole antennas in a regular of 21 cm fluctuations at z∼0.53 to 1.12 via cross- rectangular grid (Parsons et al. 2012a), complemented correlations with the DEEP2 optical galaxy survey by 16 antennas that served as outriggers to provide longer (Chang et al. 2010) and at z∼0.8 via cross-correlations baselines. Operating between 100 and 200 MHz, PAPER with WiggleZ (Masui et al. 2013). Additionally, it has has published a number of upper limits on the EoR power placed upper-limits on the 21 cm only auto spectrum and spectrum (Parsons et al. 2010, 2014; Pober et al. 2013a; used the combination of cross- and auto-correlation ) Ω ( Ali et al. 2015; Jacobs et al. 2015 , but has now been measurements to place constraints on H IbH I Switzer decommissioned, with its site in the South African Karoo et al. 2013). The data used for these results were desert now being used for the Hydrogen Epoch of calibrated using a combination of known bright sources Reionization Array (HERA; see below). PAPER’s and data obtained from injecting signals from a noise analysis approach uses the redundant baseline calibration diode into the antenna. Map-making was performed using algorithm described in Section 9.2 and bypasses map- the maximum likelihood estimator described in making in favor of direct power spectrum estimation Section 10.1, and power spectrum estimation was using visibilities via the delay spectrum approach of performed from the resulting maps. Foregrounds were Section 11.2, couched in the language of the quadratic suppressed by projecting out principal component spectra estimator framework (Section 11). (Section 12.1.3), with special care taken to compute 4. The LOw Frequency Array (LOFAR). LOFAR is a low- transfer functions to correct for the possible subtraction of frequency radio interferometer with a dense core of cosmological signal (Section 12.5).

19 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

6. The Parkes Radio Telescope. Parkes is a general-purpose diameter (DeBoer et al. 2017). The dishes are arranged in 64 m single dish telescope (Staveley-Smith et al. 1996) a compact configuration that is essentially an “exploded” operating from 1230 to 1530 MHz. It is located in New hexagon: the configuration would be a close-packed South Wales, Australia. In a similar way to GBT, it has hexagon of touching dishes except that the hexagon is been used to measure 21 cm fluctuations in cross- split into three parallelograms that are displaced from correlation. Combining Parkes data with the 2dF galaxy each other by sub-dish spacings (Dillon & Parsons 2016). survey has enabled positive detections in the range This produces sub-aperture sampling in the baseline 0.057<z<0.098 (Anderson et al. 2018a). Concep- distribution, reducing grating lobes. In addition to 320 tually, the analysis pipelines used for the analyses were dishes in the exploded hexagon configuration, there are very similar to those used for the GBT results. 30 outrigger dishes to provide long baselines for better 7. The Owens Valley Long Wavelength Array (OVRO- angular resolution. All the elements (including the LWA). OVRO-LWA consists of 288 dipoles, with 251 of outriggers) are placed in a such a way that the entire these dipoles located within a 200 m diameter compact array can be calibrated using the redundant calibration core and the remaining 32 dipoles located farther away to method described in Section 9.2. HERA is a telescope provide baselines of up to 1.5 km in length. The elements with no moving parts. It surveys the sky as a drift-scan in the core are arranged in a pseudo-random fashion, telescope that is always pointed at zenith. With a custom- enabling a better point-spread function for imaging. This designed broadband feed, HERA is in principle capable makes OVRO-LWA a powerful instrument for transient of observing between 50 and 250 MHz. It is designed to science, for instance in its searches for radio emission have sufficient sensitivity to make a high-significance from -ray bursts (Anderson et al. 2018b), gravita- detection of the 21 cm power spectrum both during tional wave events (Callister et al. 2019), or from reionization (Pober et al. 2014; Liu & Parsons 2016) and (Anderson & Hallinan 2017). With observa- cosmic dawn (Ewall-Wice et al. 2016c; Kern et al. 2017). tions spanning an instantaneous bandwidth from 27 MHz 2. The Canadian Hydrogen Intensity Mapping Experiment to 85 MHz, OVRO-LWA has also produced new results (CHIME). CHIME is located at the Dominion Radio for 21 cm cosmology, including a detailed set of low- Astrophysical Observatory in British Columbia, Canada. frequency foreground maps (Eastwood et al. 2018) and a It is comprised of four large cylindrical reflectors, each upper limit on the 21 cm power spectrum at z≈18.4 measuring 20 m×100 m (with the parabolic curvature (Eastwood et al. 2019). The pipelines for these results being along the shorter dimension, hence the visual lean heavily on the m-mode formalism for map-making impression of a cylinder). The long axes of the cylinders (Section 10.5), the Karhunen–Loève mode projection are oriented north–south, and since it is these axes that are technique for foreground suppression (Section 12.1.3), flat and open, the instantaneous field of view is large and quadratic estimator-based power spectrum estimation (∼100°) in the decl. direction (Bandura et al. 2014).In (Section 11). the east–west direction CHIME has a smaller instanta- neous field of view, but because the telescope operates as Thus far, there has yet to be a detection of the 21 cm auto- a drift-scan telescope, over time the entire range of R.A. power spectrum (i.e., not in cross-correlation with other values is covered. In each cylinder are 256 dual- probes), although upper limits have become increasingly polarization antennas that serve as feeds (Newburgh stringent over the years. The experiments listed above were et al. 2014; Hincks et al. 2015). Operating between 400 chosen because they have all published such limits, which are and 800 MHz, CHIME is designed to make precision illustrated in Figure 7. To be sure, there have been setbacks, measurements of the Baryon Acoustic Oscillation scale including revisions and retractions of incorrect upper limits, from z = 0.8 to 2.5 (Shaw et al. 2015). However, in and we discuss this topic in Section 12.5. On the more recent years alternate processing backends have been optimistic side, however, current instruments (e.g., the MWA) installed on the telescope to enable pulsar monitoring and are being upgraded and new instruments are being built. These searches for Fast Radio Bursts (FRBs; CHIME/FRB new instruments not only have the sensitivity (at least in Collaboration et al. 2018). The FRB searches, in principle) to make high-significance detections, but also to particular, have revolutionized FRB studies with CHIME diagnose possible systematics with new levels of precision: discovering an enormous number of new FRBs, including 1. The Hydrogen Epoch of Reionization Array (HERA). some down to 400 MHz (CHIME/FRB Collaboration HERA occupies the same South African Karoo desert site et al. 2019a, 2019b). that was formerly occupied by PAPER. When complete, 3. The Hydrogen Intensity and Real-time Analysis eXperi- it will consist of 350 dishes, each of which has a 14 m ment (HIRAX). HIRAX is similar to CHIME in its dual

20 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Figure 7. A summary of current upper limits on EoR power spectrum measurements. The limits are expressed as Dº23()kkPk ()2p 2values evaluated at the k bin that gives the most competitive (i.e., lowest) upper limit for each experiment. Thus, different points on the plot come from different k values. A direct comparison between them is thus unfair in principle, and in practice is somewhat reasonable only because the theoretical predictions for Δ2(k) are often fairly flat in the range 0.1 h Mpc−1<k<1 h Mpc−1 (see Figure 4). (Only the LOFAR points fall outside this range.) Also shown is a fiducial theoretical model from the 21cmFAST semi- analytic code (Mesinger et al. 2011). (A color version of this figure is available in the online journal.)

science goals (21 cm cosmology from z∼0.8 to 2.5 plus forecasted to make precision cosmological measurements transient surveys), but is located in South Africa. With (Xu et al. 2015). access to the southern sky, HIRAX is particularly 5. The BAO from Integrated Neutral Gas Observations well-positioned to take advantage of cross-correlation (BINGO) experiment. Located in South America, opportunities, given the large number of galaxy BINGO aims to make measurements in the 0.12< surveys, CMB experiments, and future non-H I inten- z<0.48 range (960–1260 MHz). As such, it is quite sity mapping surveys with footprints in the south. (See complementary to CHIME, Tianlai, and HIRAX, provid- Sections 12.1.7, 14.1,and14.2 for the power and ing large scale structure measurements at a different set of limitations of cross correlations). When construction is redshifts. From a technology development standpoint it is complete, HIRAX will consist of a square 32×32 grid also complementary, as it is not an interferometer. Instead of 1024 dishes operating from 400 the 800 MHz. With it is a single-dish experiment with a two-mirror crossed- 6 m dishes, its instantaneous field of view (∼6°) is Dragone design that is outfitted with a 50-element feed smaller than that of CHIME. However, the HIRAX horn array at the plane (Battye et al. 2013; dishes can be repointed (by hand), enabling large sky Wuensche & the BINGO Collaboration 2019). The coverage. Each pointing maps out a stripe in decl., and instrument has an instantaneous field of view of ∼15° with a repointing every 150 days over four years, and operates in a drift-scan mode to map out a stripe in HIRAX can cover a quarter of the sky with sufficient the sky. depth for high-precision cosmological measurements 6. The Square Kilometre Array (SKA). The SKA is the (Newburgh et al. 2016). largest effort of all upcoming telescopes. It will consist of 4. The Tianlai experiment. Tianlai is again a post-reioniza- two separate telescopes, SKA-low and SKA-mid. SKA- tion 21 cm intensity mapping experiment, but with a low (50–350 MHz) will be located in the Western slightly larger frequency range (400–1420 MHz or Australian desert, while SKA-mid (350 MHz–15.3 GHz) 0<z<2.5) than CHIME or HIRAX. It is located in will be located in the South African desert. The former the Xinjiang Autonomous Region, China. Its final design will be comprised of over ∼100,000 antennas grouped is yet to be decided, with the Tianlai team building two into stations, while the latter will be comprised of 197 separate pathfinder experiments: one is a set of three dishes. To achieve a small, compact synthesized beam cylindrical reflectors (each measuring 15 m×40 m and without strong grating lobes, the elements will be outfitted with 96 feeds) and the other is a set of 16 dishes arranged in a pseudo-random configuration with a (each 6 m wide). The dishes are arranged in two maximum baseline of 65 km for SKA-low and 150 km concentric circles at radii of 8.8 and 17.6 m around a for SKA-mid. While many details regarding the SKA are central dish. Like CHIME and HIRAX, Tianlai is still to be determined, it is generally expected to have an

21 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

expansive science case, including the EoR (Koopmans direction of the vector indicating polarization direction) coming et al. 2015), post-reionization cosmology (Maartens et al. from each direction rˆ and frequency ν 2015), pulsars (Kramer & Stappers 2015), transients dtEr()=Wε (ˆ,,43nn)() e2pnit dd (Fender et al. 2015), galaxy formation (Prandoni & Seymour 2015), magnetism (Johnston-Hollitt et al. 2015), such that the total electric field that could be measured at time t planet formation (Hoare et al. 2015), and individually is the integral over all angles and frequencies.19 The electric detected H I galaxies (Staveley-Smith & Oosterloo 2015). field at a general point x simply adds a direction dependent time delay relative to the reference point While we have endeavoured to be reasonably complete in our list of experimental efforts, there will inevitably have been dtdtEx(),44=- E ( x · rˆ c)() some omissions. For the list of upcoming experiments, we =We(rˆ,,45nn)()edd2pnit(·-xrˆ c) focused on those that are funded (or are reasonably likely to ( ) secure funding) and also list 21 cm cosmology as one of their which encodes the fact that as we move toward or away from ( main science objectives. A more complete list (at least for post- the source of the emission, we receive the emission sooner or ) reionization experiments), including telescopes that could in later . principle be used for 21 cm cosmology (even if they were not In radio astronomy when we talk about an antenna we can fi designed for it), can be found in Bull et al. (2015). mean many things; here we de ne it in terms of its operational While carefully designed hardware is crucial for achieving behavior. That is that an antenna is the part of the system which the high levels of precision necessary for 21 cm cosmology, maps the emission on the sky into the measured signal. fl equally important are the analysis choices that one makes. In a Physically this may consist of many components: a re ector sense, 21 cm experiments are software telescopes. Having that focuses the radiation onto a dipole, the dipole itself, fi summarized the general state of the field, the rest of this paper ampli ers, cables among others. These components have many focuses on 21 cm data analysis in detail, starting with a deeper complex interactions within and between signal chains. We fi discussion of interferometry. abstract over this complexity and simply de ne the operation of the antenna labeled i (located at position xi on the ground) as the response of the measured output si to emission from the 5. Fundamentals of Interferometry sky. That is In Section 3 we introduced the basics of interferometry, dsii= anAr(ˆ,,,46)· d Ex ( i t ) ( ) defining the key observable, the visibility, in the commonly used flat-sky, unpolarized limit. While largely adequate for where the antenna response Ai is more commonly called its 20 α current instruments, the next generation of survey interferom- primary beam or often just its beam, and is an arbitrary eters have large instantaneous fields of view, and need to excise scaling we will set later to simplify the notation. Note that as fi both polarized and unpolarized foregrounds with high fidelity. the electric eld is a vector quantity, the primary beam must be To appreciate when these approximations are appropriate and a vector response and the two multiplied with a dot product to when they are not, we need to go back to the start and look at give a scalar voltage output. Provided the properties of the interferometry with no approximations. antenna do not depend on time themselves, i.e., the antenna is a In this section we go through the fundamentals of stable part of the instrument, the beam depends only on the ( ) interferometry, starting from the electromagnetism and statis- frequency a consequence of the Fourier convolution theorem tics to explain: what are we measuring? Why do we measure it? and the direction of the incoming radiation, which we have And how do we do it in practice? However, later sections can used above. be mostly understood using just the results of Section 3, and so The beam is a function which, for every position on the sky, the hurried reader may wish to jump forward to Section 7. and at every frequency, describes the response of the instrument. That response is given as a complex vector, with the vector being defined as transverse to the direction vector rˆ 5.1. Antenna Response (i.e., it is defined in the tangent space to the sphere in that In radio astronomy we make coherent measurements, 19 meaning that an individual antenna directly measures a It is worth noting that although the sign choice within the exponential Equation (43) is arbitrary, this choice fixes the sign of all other Fourier weighted linear sum of the electric field in the surrounding exponentials within interferometry. We use the standard convention in radio volume. The antenna generates a voltage proportional to this astronomy and electrical engineering that the transform from frequency to time has a positive sign in the exponential. signal that we can read out. If we consider only emission in the 20 Do not confuse this with the primary beam Ap defined in Section 3, which far field, this is equivalent to measuring a direction-weighted was a power primary beam, and a scalar quantity, unlike the vector electric summation of incoming electromagnetic plane waves. More field/voltage primary beam that we are denoting as A here. As we will see, the fi former roughly goes as the square of the latter. Confusingly, both quantities are precisely, we will model the electric eld at a reference point as often referred to simply as the “primary beam,” and it is necessary to made up of small vector contributions ε(rˆ, n) (with the understand the context in which the term is being applied.

22 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw direction). For a physical interpretation of the beam we can sourced from the sum of many independent systems that have think of the amplitude of the vector as telling us how sensitive no time/phase correlations between themselves, for example we are in each direction on the sky; the orientation of the vector electrons spiralling in a magnetic field emitting synchrotron tells us what orientation of the incoming electric field we are radiation that have random initial velocities and positions. sensitive to (i.e., the polarization); finally, the complex phase However, radio emission is often polarized. That is, there are tells us about any relative time delay with which we receive the non-zero correlations between the orthogonal polarization emission. states. Expressing both of these facts mathematically we have In integral form we can write the output signal si in terms of 4kZB 0 the beam and emission from the sky as ᢢñ=-¢-¢enena (rrˆ,,)(* ˆ )(Tab rrrˆ,,nd)(ˆˆ)( dn n ) b l2 2pnit(·-xrˆ c) ()48 stii()=Wann∬ Ar (ˆ,,)·ε ( rˆ )() ei d d n .47

where Z0 is the impedance of free space and Tab is the Though the antenna beam Ari (ˆ, n) is easy to define, actually coherency matrix describing the polarization of the emission measuring it from data, or calculating it using simulations is which, as we have defined here, has units of temperature. The very difficult and is one of the major challenges facing 21 cm polarization of radiation is usually described in terms of the cosmology. This is discussed more in Section 6.2. four Stokes parameters each describing how one component of the polarised emission varies over the sky and with frequency. 5.2. Sky Emission The four components are: T (rˆ, n) which describes the total To understand how polarization affects our measurements intensity of the emission; Q(rˆ, n) and U (rˆ, n) which give the we must first define a reference frame on the sky. As linear polarization; and V (rˆ, n) which gives the intensity of electromagnetic radiation has no component in the direction circular polarization. In terms of these the coherency matrix is fi of propagation, the electric eld lies in the two-dimensional 1 ⎛ TQUiV++⎞ plane perpendicular to the direction vector rˆ.Defining an Tab(rˆ, n)()= ⎜ ⎟.49 2 ⎝UiVTQ--⎠ orthonormal basis in this plane eˆa, the electric field vector can a be decomposed into orthogonal polarization states ea = eˆ · ε, There are very few astrophysical sources of Stokes V with the beam decomposing similarly. As we are considering polarization, and those that exist are typically from fast wide-field instruments we need to use a basis which works on transients like pulsars, and so generally we can assume that the full sky and so we follow the cosmological convention and Stokes V is zero. We should also remember, that though the ˆ ˆ use the spherical polar unit vectors eˆa = {qf, }, with celestial underlying electric field is random with no spatial or spectral north aligned along the polar axis. In this convention the vector correlations, the same is not true for its statistics, i.e., the ˆ ˆ q points south, while the vector f points to the east. elements of the coherency matrix Tab can be correlated across First, we should remember that the electric field from the sky frequencies and angles (as we will see in Section 7). (that is ε) is a random signal and so is described statistically. As almost all astrophysical radiation is produced by a large 5.3. Interferometry in TheoryK ensemble of independent emitters (electrons, atoms, mole- fi ) As the underlying electric eld is a Gaussian process, the cules... by invoking the central limit theorem we can see that coherency matrix contains all the information about the the distribution of the observed electric field is Gaussian21 ( ) emission from the sky. Our goal with interferometry is to try with zero mean , and so its statistical properties are encoded and capture as much information about this as we can. entirely within its two-point statistics áenena (rrˆ,,)(*b ˆ¢¢ñ) ( ) First, because the emission is spectrally incoherent, it will be Nityananda 2003 . If we know that function we know more useful to describe our observations in frequencies rather everything of interest about the emission from the sky; than times. This means that we usually think of our however, it is generically a complicated function of multiple measurements from the antennas as a function of frequency, positions and frequencies. and so will use their Fourier transforms si(ν), which written Fortunately, on this front we receive some help from our explicitly are universe: while our detector is coherent, in radio astronomy the emission from the sky (with rare exception) is spectrally and -2pnit sstedtii()n = () spatially incoherent. That means that there is no correlation ò between emission from different parts of the sky or at different a -2pnicxri·ˆ =WanenAedi (rrˆ,,)(a ˆ )() 50 frequencies. Physically this occurs when the emission is ò

21 where we have implicitly summed over a using an Einstein Even searches for non-Gaussian distributed electric fields from highly ( ) coherent emitters such as masers (Evans et al. 1972; Dinh-v-Trung 2009) and summation convention for indices a and shortly b , which we pulsars (Smits et al. 2003) have proved negative. will use for the rest of this section.

23 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

As the emission from the sky is Gaussian distributed, and the 3. ∼1s—on timescales longer than a second the change in measurement by the antenna is a linear process, this means that orientation of the telescope due to rotation of the Earth is the signal si(ν) is also Gaussian distributed. As all the significant compared to its resolution. information about a Gaussian process is captured in the two Put together, these give us a practical way to build an point statistics, this means that for a fixed set of antennas all the interferometer. Between the first two timescales we can information is contained in the covariance matrix of the signals transform our data from the time domain into the frequency from the set of antennas. These elements of this covariance domain to give ourselves distinct frequency channels. This matrix are the visibilities, the key quantities in radio performs the operation described in Equation (50) and is interferometry. The elements of the matrix are the cross performed by the F-engine. Between the second and third correlations between two feeds, and for a pair i and j are timescales the samples from each frequency channel should be statistically equivalent, but independent, which allows us to Sss()nnn=á ()* () ñ ij i j evaluate the ensemble average of Equation (51) by integrating a b* -2piurij·ˆ in the domain. This is performed by the X-engine. =Wò AAi (rrrˆ,,,nnn)(j ˆ )( Tab ˆ )() ed 51 5.4.1. Analog Chain where we define the baseline vector bij=-xx i j and uijº b ij l, which is the separation between the antennas in The purpose of the analog chain is to transform the electric 2 12 fi wavelengths. We have also set the value of al= ()4kZB 0 eld coming into the telescope into a signal which can be to eliminate the prefactor found in (48). Here Sij specifically digitized. At radio frequencies many analog components refers to the part of signal coming from the sky; in the next introduce noise into the system, typically by attenuating the section we will look at the contribution of noise. signal we care about and introducing a small amount of thermal Equation (51) is the fundamental equation of interferometry, noise in its place. The primary goal of an analog chain is to describing how emission on sky generates a signal in our data, minimize this effect. and includes exactly all curved sky and polarization effects. Its There are many stages to the analog chain, the first that we resemblance to Equation (12) is apparent, and we will make the have already discussed is the antenna itself (which may also be connection more closely in Section 6.3. paired with a reflector), that receives the electric field and turns it into a voltage. At this point the signal is extremely weak and 5.4. Kand in Practice must be quickly amplified so as to not introduce noise. This is performed by a Low Noise Amplifier (LNA) which amplifies Above we have presented the idealised picture of an the signal by many orders of magnitude, at the expense of interferometer, but to appreciate fully the challenges of introducing some noise of its own. The LNA is designed to building one and analyzing its output we need to delve into amplify the signal to well above the thermal noise level in the how they work in practice. An interferometer can be divided rest of the system such that any further attenuation does not into three distinct parts, the analog chain and two digital introduce any extra noise. systems, the F-engine and the X-engine.22 We discuss these At this point there are two more roles for the analog chain: below. One of the key things to appreciate in how an interferometer 1. Removing any unwanted frequencies, i.e., frequencies works is the separation of the various timescales involved. above the maximum frequency we are interested in and There are three distinct timescales we need to consider: below the minimum, which is important for digitisation (see the following section). 1. ∼1nsto10ns—for 21 cm cosmology we are interested 2. Transporting the signal to the location of the digital in frequencies of around 100–1000 MHz, so we are system. Although serving a trivial purpose, this particular probing fluctuations in the electric field on roughly stage can be the source of several systematics (see nanosecond timescales. Section 6). 2. ∼10 μs—the smallest spectral structures we are interested in are around 100 kHz which means the longest At the end of the analog chain we have amplified, and fluctuations in the electric field we are interested in filtered the signal all the while introducing noise into the are ∼10 μs. system. We write the final quantity passed onto the digital stage as 22 In this work we discuss only FX correlators which are dominant in low- frequency radio astronomy. However, we should note the existence of XF xtii()=+ st () nt i (),52 ( ) correlators (also known as lag correlators) which perform the frequency transform last. This may have some advantages in terms of spectral line where we have rescaled such that this output is proportional to sensitivity and correcting quantisation effects, but requires more computation fi for interferometers with a large number of antennas. See Thompson et al. the input signal. This moves all the effects of ampli cation and (2017) for a detailed comparison. filtering into the final noise term which we write as ni(t).

24 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

5.4.2. F-engine chunks being transformed. The frequency channel shape is controlled by The job of the F-engine is to take the time stream data and turn it into distinct frequency channels (hence the “F”), however, before that can happen the data must be digitized. wwaiat˜(npn )=-Då [ ]exp ( 2 ) , ( 55 ) This first step is performed by an Analog–Digital Converter a (ADC) which produces a discrete series of quantized time the Discrete Fourier Transform of the window function w[a] samples from the amplified input. The quantization reduces the which is typically of width ∼1/T. To choose the window precision of the input and is typically treated as an extra source function we need to trade-off sensitivity against localizing the ( ) ( ) of noise Thompson et al. 2017 that can be absorbed into ni t , response (i.e., keeping w˜()n compact around ν=0), which is although modern correlators typically operate at high enough important for mitigating RFI (see Section 12.4). In practice bit-depth that this contribution is sub-percent (Mena-Parra et al. most interferometers use Polyphase Filter Banks (Price 2016) 2018). The rate at which the input is sampled is controlled by rather than a Short-Time Fourier Transform which combines the range of frequencies we are interested in. For the simplest multiple time chunks (called taps). This gives even better scenario of standard Nyquist sampling the input must be localization of the spectral response, but requires more ν > ν sampled at sample 2 max where frequencies above the computational power, and introduces slight correlations maximum were filtered out earlier in the signal chain to between spectra close in time. For modern instrumentation prevent aliasing of higher frequency signals. However, usually the channelization is typically performed by Field Program- ν we do not care about frequencies below some min, in that case mable Gate Arrays (FPGAs), integrated circuits that can be we can reduce the sampling rate by either: programmed at the hardware level (Bandura et al. 2016a; ) 1. Mixing—using a local oscillator and a mixer, we can shift Hickish et al. 2016 . the frequency range of interest down toward zero, and we 5.4.3. X-engine then only need to sample at νsample>2 (νmax−νmin). 2. Alias sampling—if the maximum and minimum frequen- The role of the X-engine is to take the frequency channelised cies are both multiples of some interval Δν it is possible data and cross correlate all inputs. However, the output of the to alias sample, where we deliberately sample at less than F-engine is a time series of spectra for each individual input, the Nyquist frequency, aliasing our signal down to lower but for the cross correlation we need to have access to the data frequencies without the use of a mixer. In this case we for all inputs at the same frequency. This requires a large need only sample at νsample>2Δν. transpose of the data set from frequency-ordered to input- ordered. This is a significant bottleneck for large interferom- This gives the shortest timescale we are concerned with: eters, and usually requires a combination of fast mesh networks the length of an individual sample ∼1/ν . We will denote sample and large pools of memory to perform (Lutomirski et al. 2011; the discrete samples taken at some time tat=Das a Bandura et al. 2016b). xt[]= xt (). ia ia After the transpose, all that remains is to calculate the The next stage in the digital chain is channelization, where visibilities. This is done by exploiting the separation of we turn the time samples into frequency channels. The simplest timescales and calculating the sample covariance of the way to perform this is to make use of the the Short-Time frequency channels. By doing this we are forming an estimate Fourier Transform. In this process we group the timestream of the underlying, true covariance. As it is an estimator, for the into chunks of N samples, each starting at samples t and of A moment we write the visibilities as Vˆ given by length T=NΔt, apply a window function w[a] to account for ij the non-periodic nature of the chunk and then use an FFT to ν = / ˆ 1 turn the data into frequency channels centered at b b T Vij[]nnn b = å xtxt ˜[ibA;;,56 ]˜[*j bA ] () M A ⎛ 2piab ⎞ ⎜⎟ though for notational convenience we will just use V in xt˜[ibAn ;exp ]=-å xtwaiAa [+ ] [ ]⎝ ⎠ (53 ) ij a N subsequent sections. A typical spectral resolution of an interferometer is around 100 kHz, meaning we get a new 2pnitA μ =-ò xtw˜(iAnnnn;,54 ) ˜( b ) ed ( ) spectrum every 10 s. If we average these up to the timescale on which we expect the signal to change by Earth rotation, this 5 where we have inserted the definition of xi[tA+a] in terms of its implies that we sum over M  10 spectra. Computationally continuous Fourier transform to show how the measured this step is extremely challenging for large interferometers, frequency channels relate to the underlying continuous requiring ∼ΔνN2 operations per second, but is generally easy frequency signal. As expected, to gain higher spectral to parallelize. This is usually performed on FPGAs or GPUs resolution we simply need to increase the length of the time (Kocz et al. 2014; Denman et al. 2015).

25 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

The quantity Vˆij[n b] can be thought of as an estimator for the further analysis it must be understood better than the 21 cm underlying “true” visibilities23 with expectation signal to astrophysical foreground ratio (i.e., the S/F ratio),to avoid leaking foregrounds into our data. This varies depending ¯ 2 VwSNdij[nnnnnn b ]=-ò ∣˜( b )∣[ ij () + ij ()],57 ( ) on where on the sky we are looking and at what frequency band, but is often around 10−4. where the noise covariance Nij()nnn=ánn i ()*j () ñ. The covar- iance of the estimator Vˆ is ij 6.1. Gain Variations ˆˆ ˆˆ* Cov(VVij , kl)(= Cov VV ij , lk ) Converting the signal arriving at the antenna into one which 1 can be digitized is the job of the analog chain (Section 5.4.1) = VV¯¯ik lj.58() M and requires a variety of components which amplify, filter and Equation (58) gives an exact and compact description of the transport the signal. The effect of the analog chain can be (ν) noise of an interferometer. However, we are often in a regime described by a complex frequency-dependent gain g that multiplies each of the received signals. (particularly for low redshift 21 cm observations) where Sij = Generally we can think of the amplitude as describing how Nij, and as we can often assume the noise is uncorrelated i fi much the signal has been amplified by the chain, and the phase between feeds Nij » Trecv dij,we nd the standard result that the noise on an interferometric visibility is describes how it has been been delayed by components along the chain. These are free functions of frequency, though ij TTrecv recv amplitudes are often smooth, and the phase behavior may be Var(Vˆij)()= , 59 M dominated by a single time delay for all frequencies. We should with distinct visibilities being uncorrelated. note that absolute phases are typically meaningless, so the phase typically refers to how much one input or frequency has 6. Instrumental Systematics been delayed relative to the others. Learning what these gains are, and thus how to relate our received signal to the true signal Modern radio telescopes are complex instruments. Using our from the sky is the process of calibration that we discuss in data to map the H I distribution in the universe requires that we Section 9. first understand how our data is generated by emission from the Gains are typically multiplicative scalars. That is for some sky. This requirement is compounded by the presence of true signal s(ν) the observed voltage xgs()nnn=+ ()() noise extremely bright astrophysical emission (relative to the 21 cm where the overall gain g(ν) is the product of all the individual signal), that mean we must control and understand our component gains, e.g., gg()nnnn= filter () g cable () g amplifier (). instrument well enough to prevent proliferation of these Unfortunately, the gain for each signal chain is time foreground modes. dependent. There are many possible sources for this, but the Below we talk about the various types of systematics effects primary source is the thermal environment of the analog chain, generated by radio interferometers. For each of these instru- two examples of this are: mental properties there is some fraction that we can understand fi fi precisely, and some part that we cannot determine. The parts 1. The level of ampli cation of an ampli er depends on its ∼ we understand can in principle be added to our analysis, and temperature. This is typically at the 0.5% per K level, ∼ their effects removed in an optimal manner, though in practice and so over the course of a day we may expect 5% fi this may be computationally challenging. Even if we can variations in the ampli cation along any signal chain ( ) optimally treat the systematic, we may suffer a loss of Davis 2012 . sensitivity, one which can only be rectified by changing the 2. Cables cause both variable attenuation of the signal as the instrument itself. The parts we do not understand are far more material properties change with temperature, and phase dangerous, as they lead to uncorrected mode proliferation, and variations due to thermal expansion of the cable thus bias in our results. The key challenge of any 21 cm increasing the propagation time along it. observation program is to devise analysis methods which However this is made even more problematic by the fact that minimise exposure to undetermined instrumental effects, and, different components experience slightly different thermal constrain these instrumental unknowns such that they do not environments (are they exposed to the Sun? The ? Are appreciably bias our results. they heated internally?), and also have varying susceptibilities As a rule of thumb for thinking about instrumental to the temperature they experience. This means the variations in systematics, when a systematic needs to be understood for the gain of each signal chain are time variable, (partially) fi 23 independent, and dif cult to predict. Technically the estimator has a complex Wishart distribution, that is The effect of these gain variations is to modify the ([])(MVijn b~  C V¯ ij [] b, M), but for typical interferometers M is large enough that we can treat its statistics as Gaussian. measurements we make. Known gains are trivial to correct,

26 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw as we can divide each signal by its known gains. However, are other techniques (see Section 12.1.5) which avoid unknown gains limit in the degree to which we can combine combining baselines, and though this has lower requirements different baselines and frequencies to make maps and remove on the knowledge of the primary beams, it does place stronger foregrounds without biasing our results. requirements on their properties, in particular their frequency It is often useful to divide the gains up into three distinct smoothness and polarization purity (see the next section). terms for each signal chain An additional systematic effect introduced by primary beams are feed to feed variations, which can introduce non- redundancy on otherwise redundant arrays. This can cause an gtiii(nnndn,1,60 )=+ gband () g ()( gt ( )) ( ) indirect systematic effect whereby the use of redundant calibration techniques (see Section 9) can mix the beam non- where redundancy into incorrect gain solutions (Orosz et al. 2019). 1. gband(ν) is the bandpass of the telescope. A real, time- independent function describing the sensitivity of the whole telescope to each frequency. 6.3. Polarization Leakage 2. g (ν) is a time-independent function, that encodes the i With rare exceptions such as the Ooty Wide Field Array input dependent gain applied to the instrument. (Subrahmanya et al. 2017), almost all radio interferometers are 3. δg(ν, t) describes time variation of the gains of each constructed from pairs of co-located antennas with orthogonal channel. feed polarizations, generally dipoles oriented at right angles. One further complexity is that gains are not always scalar This gives us sensitivity to the polarization of emission on the quantities. We may have linear mixing between signal paths, sky. However, this sensitivity is imperfect, and can lead to a generally referred to as cross talk, that must be represented by a fundamental entangling of our data where we cannot complex matrix rather than a scalar. This can happen anywhere completely separate out the various polarized components. that the signal paths are electrically “close,” for instance Although within intensity mapping we are generally interested between polarizations on an antenna (Hamaker et al. 1996),or only in the total intensity of the 21 cm emission (i.e., Stokes I), between inputs on the same analog-to-digital-converter (ADC) to remove foregrounds we need to rely on their spectral chip. This does not fundamentally alter the picture above, but smoothness (as we will see in Sections 7 and 12) and although does make correcting for it harder. this is a very good approximation in total intensity they may have significant spectral structure within their polarized 6.2. Primary Beams emission (see Section 7). If we cannot disentangle the polarized To fully understand the measurements we are making we components we may misestimate the true total intensity, must understand the primary beams of our antennas Ai (rˆ, n), introducing spurious spectral structure and hindering our that describe the spatial and frequency sensitivity of the ability to remove foregrounds. antennas in the telescope. These directly appear in the To see how this entanglement can occur, let us look at measurement Equation (51). constructing the instrumental Stokes I visibility for some Beams are typically frequency dependent, with complicated baseline. For our telescope, at every location we have two spatial structure and polarization response (Ewall-Wice et al. orthogonal dipoles (labeled X and Y), which have nearly 2016a; Neben et al. 2016; Cianciara et al. 2017; Patra et al. orthogonal response on the sky around the center of the beam 2018), and while some of these effects can be minimised (this is typical of most telescopes). To construct the Stokes I fi through speci c design decisions, they cannot be removed visibility VT we form entirely. Additionally, the beam function can vary significantly / ( from feed to feed due to manufacturing installation errors e.g., 1 deformations in reflectors, non-identical antennas) and posi- VVVTXXYY=+[] ()61 2 tional effects (e.g., whether the antenna is at the edge of the array or the center). 1 a b a b =+[(AAX rrˆ)(X**ˆ)( AAY rrˆ)(Y ˆ)] As we will discuss later (Section 12), different analysis 2 ò -2piruˆ· techniques place different restrictions on the knowledge and ´WTeab(rˆ)() d.62 properties of the primary beams. Some technique will internally make linear combinations of different visibilities in order to Let us write the primary beam of each of our feeds as a deconvolve instrument effects, and these are typically very complex scalar weight multiplied by a unit vector giving the a a sensitive to our knowledge of the primary beams, requiring a direction of response. That is AX (rrˆ)(= AxX ˆ) ˆ and similarly sensitivity similar to the S/F level (see Section 12.1.3). There for the Y polarization. Note that xˆa may have complex

27 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw components. Using this we can rewrite Equation (61) as More generally, we can see that because the polarization leakage terms DA(rˆ) and m(rˆ) are functions of angular position ¯2 -2piruˆ· VATedT =Wò (rrˆ)(ˆ) on the sky, we can never construct a linear combination of the polarizations for a visibility that measures only the unpolarized 1 2 ab + ò ((AxxX rˆ) ˆˆ* sky. Polarization leakage can only be corrected by combining 2 visibilities from different baselines together in a way which 22ab - piruˆ· +WAyyPeY (rrˆ) ˆˆ*)(ab ˆ)() d 63 allows us to discriminate between different angular positions. where we have defined the polarization averaged beam ¯2 22 A (rrrˆ)((=+AAXYˆ)(ˆ)) 2, and Pab (rˆ) is the polarized part 6.4. Noise Systematics of the coherency matrix, i.e., Although instrumental noise is typically thought of as a “ ” ⎛ QUiV+ ⎞ random error, and in a sense easy to deal with, in a real life Pab(rˆ,.64n)()= ⎜ ⎟ radio interferometer, there are still effects which must be ⎝UiVQ--⎠ understood. The first term of Equation (63) is equivalent to the usual One complication arises from the use of closed packed arrays unpolarized formalism, whereas the second term describes the to maximise sensitivity (Section 3.3). This leads to many leakage of the polarized components into this. antennas which are close to each other, and may be able to To gain more insight into the source of this leakage, let us interact directly with each other. In this scenario instrumental look at the effect of small deviations from a perfect system with noise which is generated within one signal chain can be identical beam amplitudes (AX = AY) and orthogonal real broadcast out of its antenna and received by a neighbor. This polarization orientations across the beam. In this case we can gives a source of noise which is correlated between antennas, 2 22 leading to the visibilities having a bias due to the correlated treat the difference in amplitudes DAAA(rrrˆ)(=-XYˆ)(ˆ) and noise. This effect tends to diminish as the feed separation m(rxyˆ) ==+ˆ · ˆ* mmri( rˆ)(i rˆ) as small parameters, and expand Equation (63) to linear order. Using this the leakage increases, but can be problematic for the shortest baselines that are sensitive to the largest scales. of polarized intensity into instrumental total intensity VPT is One approach that can work to remove this contribution is to 1 -22piababruˆ· take advantage of the fact that the sky changes over the day and VeAxxyyPT =D-[(rˆ)(ˆˆ ˆˆ) 2 ò remove an appropriate average of each visibility in order to 2 ab* ab ++AxyyxPd¯ (rrˆ)(mm (ˆ)**ˆˆ ( rˆ) ˆˆ )]ab ( rˆ) W. remove the noise bias. However, this is only as good as the ()65 noise properties are stable: like the gains discussed in Section 6.1 the noise level is sensitive to the changing If the axes of our polarization basis are aligned with respect to environment of each analog chain and may have substantial the zeroth order antenna polarization directions xˆ0 and yˆ0 this variation. Additionally, subtracting the average will also excise can be written as the average signal, and this is important to account for (see Section 12.5). VAQAU=D22rrˆ ˆ +m rrˆ ¯ ˆ rˆ PT ò [()()(r ) ( )() ¯ 22- piruˆ· 6.5. A Summary of Interferometry +Wmi (rrˆ)AV(ˆ)( rˆ)] e d (66 ) Summarizing the last few sections, we see that individual showing that in this basis, leakage of feed oriented Q into I is antenna elements of an interferometer directly sample incident caused by a mismatch in the beam amplitudes whereas leakage electric fields from the sky, and when pairs of measurements of U is caused by real non-orthogonality in the beam responses. from different antennas are correlated with one another Stokes V can leak into instrumental I whenever there is a delay (forming a baseline of two antennas), one obtains a visibility. (i.e., non zero phase) between the overlap of the two The visibility is the fundamental quantity probed by an polarisations, though as the sky has very little Stokes V individual baseline of an interferometer. Ignoring the compli- polarization, this term is small. The reverse is also true, in that cations of polarization leakage, the Stokes I visibility is given leakage from I into V is expected to be small, although in by practice some leakage is often seen due to calibration errors (Kohn et al. 2019). Assuming no such errors, however, Stokes ¯2 2piruˆ· V measurements are therefore often used as a proxy for noise VATedT =Wò (rrˆ)(ˆ)(),67 levels in an observation, provided one can further make the (fairly reasonable) assumption that there are relatively few where we have replaced the Stokes I intensity with the intrinsically circularly polarized sources on the sky (Patil et al. temperature units used in Section 3, assuming that any relevant 2017; Gehlot et al. 2018). multiplicative constants have been applied. Under the flat sky

28 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw approximation, we have where ν is the frequency, αsyn≈2.8 is the spectral index, Δαsyn≈0.1 is the running of this spectral index, and ν0 is an 22 rˆ =--»()()()qqxy,,1 qxy q qqxy ,,1 68 (arbitrary) pivot frequency that is degenerate with the overall amplitude of emission in a particular pixel (Wang et al. 2006). θ to linear order, where qx and y are angular coordinates on the In addition to synchrotron radiation, there are several other sky defined relative to some arbitrary axes centered on an important sources of foregrounds. These include free–free origin known as the phase center. This then gives (up to some emission, bright radio point sources, and unresolved point arbitrary constant phase terms) sources, with dust emission generally negligible at the low frequencies of interest for 21 cm cosmology (Tegmark et al. ¯2 22piq·u^ VATedT = ò ()qq ()q,69 ( ) 2000; Planck Collaboration et al. 2016a, 2018b). While many of these emission mechanisms may be subdominant to Galactic where u^ is the projection of u onto a plane perpendicular to the synchrotron radiation at the relevant frequencies, they are fi vector pointing to the phase center. If we de ne the phase nonetheless important contaminants to consider, as they can ( ) center to be the zenith, then this is precisely Equation 12 ,up still be much brighter than the cosmological signal. For fi ¯2 to the de nition Ap º A to relate the power beam to the reference, experiments targeting the EoR are aiming to measure electric field beams. We have thus returned to our starting point cosmological signals on the order of ∼10s of mK, while those in our discussion of interferometry, and the most important targeting lower redshifts seek to measure a ∼0.1 mK signal. At piece of intuition continues to be that a visibility is roughly a higher frequencies, the latter has dimmer foregrounds, but in measure of a Fourier mode of the sky. However, we are now both cases, the foreground-to-signal ratios are a formidable equipped with a better appreciation for the approximations that ∼105 in temperature. are required for this to be true, as well as some of the In some respects, 21 cm observations are similar to systematic challenges with 21 cm cosmology instrumentation. observations of the CMB in that the removal of foregrounds These will all inform the analysis methods that we discuss later is a prerequisite for precision measurements. However, the in this paper. foreground challenge in 21 cm is different from that of the CMB in two important ways. First, the aforementioned 7. Foregrounds foreground-to-signal ratio is much worse for 21 cm cosmology. One of the most formidable obstacles to a successful With the CMB, the cosmological signal is in fact the dominant measurement of the 21 cm signal is the issue of contamination source of emission far away from the Galactic plane (e.g., at the by astrophysical foregrounds and how they interact with one’s Galactic poles). Thus, exquisite foreground removal is needed instruments. These foregrounds consist of all sources of radio only for precision measurements and the desire to map large emission within our observing band except for the cosmolo- portions of the sky, and not for a simple detection. With 21 cm gical 21 cm signal of interest. Of course, these signals may be cosmology, foregrounds dominate in all parts of the sky, and of significant scientific interest in their own right—for instance, robust removal strategies are required even for a detection-level data from many 21 cm instruments can be used to map the measurement. Galaxy (e.g., Wayth et al. 2015; Hardcastle et al. 2016; Hurley- The second crucial difference between foreground removal Walker et al. 2017; Su et al. 2017a, 2017b), which is currently in the CMB and 21 cm cosmology comes from the fact that the not well-surveyed at low frequencies—but, for our present former is mostly a measurement of angular anisotropies, while purposes, any emission that is not the 21 cm signal can be the latter aims to make three-dimensional tomographic maps. considered a foreground contaminant, and must be suppressed Although CMB observations are beginning to target spectral or removed. distortions from a pure blackbody spectrum (e.g., with thermal The mitigation of foregrounds is not an optional exercise. Sunyaev–Zel’dovich measurements and future experimental This is clear from a simple consideration of how bright the proposals for measuring μ-ory-type distortions; e.g., foregrounds are. In Figure 8, we show a map of Galactic radio Chluba 2018), most CMB constraints to date have relied on emission at 408 MHz. At such frequencies, the emission is measurements of angular anisotropies. When measuring dominated by Galactic synchrotron radiation, and a glance at angular anisotropies, the different frequency channels of a the typical brightness temperature reveals that foreground CMB observation simply provide redundant information: the contaminants enter at ∼100s of K. Toward lower frequencies, cosmological signal follows a blackbody spectrum while the synchrotron emission gets brighter, and it is a reasonable foregrounds do not. Thus, extracting the cosmological signal approximation to model the spectral dependence as amounts to discarding non-blackbody contributions to the measurement, and the redundancy allows (in principle) a perfect separation of foregrounds from the cosmological signal ⎛ n ⎞--Daannsyn synln() 0 T ()n µ ⎜ ⎟ ,70 ( ) provided. This redundancy is lost in 21 cm cosmology because ⎝ ⎠ n0 every frequency channel of an observation carries unique

29 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Figure 8. The 408 MHz Haslam map (Haslam et al. 1981, 1982; Remazeilles et al. 2015) of diffuse synchrotron radio emission in our galaxy. Synchrotron emission is a dominant source of foreground contamination for cosmological 21 cm measurements. The emission is clearly brightest in the galactic plane, but the brightness of the foregrounds (∼10s to 100s of Kelvin at 408 MHz and stronger at lower frequencies) is much greater than the cosmological 21 cm signal everywhere on the sky. (A color version of this figure is available in the online journal.) cosmological information, since they each correspond to a provide a quick qualitative discussion of how one might go different redshift. Thus, unlike with the CMB, there is no about sequestering foregrounds. One popular proposal is to mathematical method to solve for the cosmological signal self- consider the spectral properties of foregrounds. As one sees consistently from the data. With N frequency channels, N from Equation (70), the spectrum of Galactic synchrotron pieces of cosmological information to solve for, and (up to) N emission is a relatively smooth function of frequency. Other foreground degrees of freedom to simultaneously constrain, sources of foregrounds are expected to behave in a reasonably one has too many variables for too few measurements. similar way (Tegmark et al. 2000; Planck Collaboration et al. To successfully eliminate foreground contaminants, then, 2016a, 2018b). Decomposing spectral data into modes that requires reducing the number of variables that one needs to vary rapidly with frequency and modes that vary slowly with solve for. One solution, in principle, is to simply eliminate the frequency might then allow foregrounds (which dominate the need to solve for any of the foreground degrees of freedom by latter) to be separated from the cosmological signal (which having exquisitely precise models of the foregrounds that can dominate the former). The foreground modes are then simply be directly subtracted from the raw data. Essentially, one would discarded, while the cosmological modes are retained for then be able to devote all N frequency channels to solving for N further analysis. cosmological degrees of freedom. In practice, it is impossible The use of spectral smoothness to enact foreground to write down an a priori foreground model that is precise to the separation is a promising approach, and indeed, it is the 105 dynamic range between foregrounds and cosmological central assumption of most foreground mitigation methods signal. It thus follows that one cannot hope to constrain all N described in Section 12. However, there are two key caveats modes of the cosmological signal; instead, there will in general that are important to bear in mind. First, note that formulae be some modes where foregrounds will dominate, with the such as Equation (70) are merely phenomenological fits, and at cosmological signal irretrievably buried within the bright the high precision necessary for 21 cm cosmology, there may be contaminants. Foreground mitigation is therefore tantamount non-smooth components to foreground emission. This is certainly to identifying some way in which the foregrounds can be a possibility that one should consider, particularly since future robustly sequestered to a limited set of modes. data sets may provide evidence suggesting this. But for now, In Section 12, we provide a detailed discussion of strategies empirical models suggest that foregrounds are relatively simple that have been proposed for foreground mitigation. Here, we both spectrally and spatially (de Oliveira-Costa et al. 2008;

30 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Liu & Tegmark 2012; Zheng et al. 2017a). In addition, studies of 8. Interlude: From Our Universe to Visibilities physical models for the underlying emission processes come to and Back similar conclusions (Sathyanarayana Rao et al. 2017a, 2017b). In the preceding sections, we have described how radiation ( ) ( ) For example, Petrovic & Oh 2011 andBernardietal. 2015 from our universe (encoding information about fundamental fi nd that with reasonable assumptions regarding the energy physics as well as astrophysics) is transformed by an distribution of the electrons powering synchrotron radiation, it is interferometer into a set of visibilities. We have also discussed hard to produce spectra with large curvature in frequency. This how these visibilities are additionally corrupted by foreground conclusion is partly driven by the fact that the total observed contaminants and instrumental systematics. In the next few synchrotron spectrum is the convolution of the electron energy sections, our goal is to undo the aforementioned processing, in distribution and the spectrum of each electron in the distribution. order to extract the scientific insights that are encoded by our This convolution serves to further smear out spectral features. universe in the 21 cm line. Section 9 describes the problem of Similar conclusions can be made for the sea of unresolved calibrating one’s interferometer. Once calibration is complete, extragalactic point sources (Liu & Tegmark 2012).Tobefair, one may choose to make a map of the sky. This process is there do exist some sources of foregrounds that are not spectrally described in Section 10, although power spectrum estimation smooth. Radio recombination lines are a prime example of this. can occur with or without a preceding map-making step, as we However, both Oh & Mack (2003) and Petrovic & Oh (2011) discuss in Section 11. Section 12 examines the mitigation of find that they are unlikely to be bright enough to be of concern. foreground contamination. While many foreground suppres- We note that foregrounds, particularly the diffuse synchro- sion algorithms operate on the data prior to power spectrum tron emission from our own galaxy, can be highly polarized estimation, we choose to describe power spectrum estimation (Kogut et al. 2007) with the direction of the polarization related first because it often informs how one goes about removing to the local magnetic field. As this emission propagates through foregrounds. The process of going from power spectra to the magnetized ISM toward us, it is subject to Faraday rotation constraints on theoretical models is described in Section 13, which rotates the polarization vector ∝RM λ2, where RM is the and a discussion of statistics beyond the power spectrum rotation measure along the line of sight. Oppermann et al. follows in Section 14. (2012) find that typical values of RM can be ∼1 to 100 rad m−2 which corresponds to a full rotation of the polarization vector 9. Calibration ∼ within 10 MHz for frequencies around 300 MHz. This means As discussed in Section 6, in a real instrument one does not that there is tremendous spectral variation in the polarized necessarily measure ideal visibilities as described by (for intensity signal which has no counterpart in the total intensity example) Equation (12). Instead, the visibility V meas measured ( ) ij Wolleben et al. 2019 . In combination with polarization by a baseline consisting of the ith and jth antennas might be ( ) leakage within in the instrument see Section 6.3 this may given by irreversibly contaminate the observed total intensity with meas=+* true spectral structure from the polarized emission, limiting our VggVnij i j ij ij,71() ability to remove foregrounds. where gi and gj are complex-valued gain factors that are Another complication with using spectral smoothness to true unknown a priori, nij is the noise on this baseline, and Vij is sequester foregrounds is that observationally, one sees these the true visibility that we would measure with a perfectly ’ foregrounds through one s instrument. Thus, even if the calibrated noiseless interferometer. (For notational cleanliness, intrinsic properties of the foregrounds are such that they can for most of this section we avoid writing quantities as functions be limited to a few smooth spectral modes, instrumental effects of frequency and time, even though the dependence is always can cause mode proliferation, where foregrounds leak into implicitly there. We explicitly address calibration as a function modes where they might not be expected (Switzer & Liu 2014). ) true of frequency and time in Section 9.4. It is Vij that we This is of serious concern. In Section 6, we discussed the ultimately want, but to get to it, we need to somehow solve for systematics that may give rise to such leakage, and we will the gain/calibration factors gi and gj. Doing so is what we provide a few suggestions for how to deal with some of them in mean by calibration. later sections. It would be fair to say, however, that the In its most general form, we cannot solve Equation (71) for interplay of foreground contamination and instrumental what we want. With Nant antennas, there are Nant(Nant−1)/2 systematics is currently the chief obstacle in 21 cm cosmology, baselines from which we measure visibilities (excluding with the strict requirements on calibration, for example, driven baselines of length zero, i.e., those formed by forming the by the high dynamic range between foregrounds and the autocorrelation of an antennas with itself, as discussed in cosmological signal. This is a problem that has yet to be solved Section 3.1). But there are also Nant(Nant−1)/2 true visibilities definitively, and the near-term progress of the field will be that we would like to solve for along with Nant gain factors. We intimately tied to progress on this front. thus have more variables to solve for than we have

31 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw measurements with which to constrain them, which is of course more complicated than just a single point source, but are still mathematically impossible. simple enough for calibration. For example, one might model To proceed, we must back away from the most general form the sky as a sum of many point sources of varying brightness. of Equation (71) and make assumptions about the sky, our Calibration then amounts to minimizing the quantity instrument, or both. In the next few sections, we explore N N meas model 2 ant ant ∣∣VggVij - i * ij various options for making such assumptions. c2 = j ,75() åå 2 i==+11ji sij 9.1. Calibration by Fitting to a Sky Model s2 fi where ij is the noise variance on the baseline formed from the Consider rst the option of making assumptions about the model sky. Suppose, for example, that the sky consisted of just a ith and jth antennas, and Vij is a simulated visibility for this single bright point source with brightness I . If we work under baseline based on our sky model. A common way to minimize 0 χ2 the flat-sky approximation to simplify this toy example, then the shown here is to use an iterative approach. One begins ’ the true visibility V true (b) measured by a baseline b is by assuming various values for the free parameters of one s sky model. This then allows the simulation of ideal visibilities ⎛ b q ⎞ model ( ) true ⎜⎟· 0 Vij . With these visibilities in hand, Equation 71 can be VIi()b µ-0 exp 2p ,() 72 ⎝ l ⎠ used to solve for the complex gains (ignoring the noise term, which we cannot do anything about, except perhaps to average which we obtained by inserting a Dirac delta function centered it down in time; see Section 9.4). The complex gains can then on source position q into the I (q) term of Equation (12). 0 be used to solve for an improved set of ideal visibilities, again Using the fact that the baseline vector b connecting the ith and using Equation (71). The entire process can then be iterated, jth antennas (located at x and x , respectively) can be written as i j going back and forth between solving for calibration b =-xxwe have ij parameters and sky model visibilities until convergence. Note ⎡ ⎤ true ()·xxij- q0 that because this is a nonlinear minimization problem, there is VIµ-0 exp⎢ i 2p ⎥ ,() 73 ij ⎣ l ⎦ the potential for noise to bias the results, in the sense that time- averaged solutions need not necessarily converge to the correct ( ) and inserting this into Equation 71 , we obtain answer. Having said that, if the starting point of the iterations meas are relatively close to the truth, there is a reasonable VggInij =+i * 0 ij,74() j expectation that the bias will be small. These issues are where we have absorbed the phase factors into gi and gj. With discussed in Liu et al. (2010), albeit in the context of a different this restricted form of the equation, we have Nant(Nant−1) calibration strategy (the redundant calibration algorithm measurements (treating the real and imaginary parts of the described in Section 9.2), but the ideas are similar. visibilities as separate measurements) but only 2Nant+1 Naturally, the more realistic a sky model, the better one’s numbers to solve for (real and imaginary parts of our Nant calibration solution (Patil et al. 2016). In pipelines like those complex gains plus the amplitude of our single bright point described in Sullivan et al. (2012), Jacobs et al. (2016), Line source). Our calibration problem has now become possible to et al. (2017), and Noordam (2004), high-quality point source solve. catalogs (e.g., Carroll et al. 2016) are routinely used to model In reality, the sky does not consist of a single bright point the sky. A key challenge in this area is that catalogs will never source and nothing else. For this to even be a good be complete, for there will always exist many faint sources that approximation, one must observe a bright source using an are unaccounted for. This has been shown to result in instrument with a field of view that is narrow enough to calibration errors with artificial spectral features that make exclude other bright sources. Moreover, the angular resolution foreground mitigation difficult (Barry et al. 2016). (See of the instrument must be sufficiently low for the source to be a Sections 7 and 12 for extended discussions of the foreground point source, and not an extended source whose spatial problem.) Another challenge is the inclusion of model structure is resolved. While the angular resolution requirement components beyond point sources, such as diffuse Galactic might in some cases be satisfied given the low angular emission. One potential solution to this problem is to use resolution of compact interferometers that are used for 21 cm pulsar gating, as has been demonstrated in GMRT measure- cosmology, many of these instruments have fields of view that ments (Pen et al. 2009; Paciga et al. 2011). The key idea is to are too wide to enable an easy isolation of a single point source. difference two observations that are slightly separated in time, Despite its shortcomings, our toy single-point-source model one when the pulsar is “off” and one when the pulsar is “on.” illustrates the general principle that if the number of degrees of Assuming that all other sources of sky emission (and possibly freedom describing the sky are limited, then it is possible to RFI) are constant during the short separation in time, one can solve for calibration parameters. Generalizing our model then make the approximation that the sky flux is dominated by slightly, one can imagine that there are sky models that are a single source (the pulsar) in the differenced data. In principle,

32 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw this obviates the need for complicated sky modeling. In of degrees of freedom that one must solve for. For example, with practice, pulsar gating may not always be a viable technique, a rectangular nxy´ n array of regularly spaced antennas, there given that it not only requires the availability of a suitable are nxynnn() xy- 1 2 total baselines (and therefore measured meas pulsar, but also imposes particular requirements on one’s visibilities V ), but only 2nx ny− nx−ny unique visibilities. instruments, such as a reasonably narrow field of view (so that Adding to this are nxny gain parameters, yielding 3nxny− nx−ny other transient phenomena are not picked up) and good time that are unknown and must be solved for. This number, though, is resolution (to resolve the on/off behavior of the pulsar). in general smaller than the number of measured visibilities, In general, then, a good sky model beyond point sources is and redundant calibration solves for both the unique visibilities necessary. At the low radio frequencies of 21 cm cosmology, and the gain parameters at the same time by minimizing diffuse sky emission is often quite bright and non-negligible for Equation (75)(with our model for the visibilities simply being the required levels of calibration precision. Precisely how one the set of unique visibilities). This minimization can either be should parameterize and include diffuse models in calibration performed numerically in a nonlinear way using gradient descent is still very much of an open question. For now, many analyses algorithms (Marthi & Chengalur 2014; Grobler et al. 2018) or by attempt to circumvent this by performing calibration using only linearizing the expressions analytically in some way and solving long baselines. The concept here is that because diffuse sky the resulting linear equations (Liu et al. 2010). emission is—by definition—on spatially extended scales, its The advantage of redundant calibration is that although it influence should only dominate the visibilities of baselines that makes assumptions regarding the instrument (see Section 9.3 are short. In contrast, localized bright point sources contribute for some examples), it is relatively free of assumptions to the visibilities of all baselines. This means that the regarding sky emission. By solving directly for the unique approximation of a point source-dominated sky is better for visibilities, one need not decide ahead of time whether the sky long baselines than for short baselines. Thus, one should in is dominated by point sources or diffuse emission or (most principle be able to obtain better calibration solutions by likely) is some mixture of both. However, it is crucial to note performing a point source model calibration using only long that redundant calibration is not entirely free of modeling baselines, and then transferring the calibration solutions to all assumptions. This is because demanding self consistency in a baselines by using the fact that a particular antenna will in data set only allows one to solve for a relative calibration general be part of both short and long baselines. However, between antennas, not an absolute one. For example, redundant consistency checks following such procedures have often calibration is unable to fix an absolute scale for the calibration revealed suspicious discontinuities in data quality at the parameters and unique visibilities that it solves for. This is (arbitrarily defined) boundaries between “short” and “long” because one can always double all gain parameters while baselines, suggesting that further work is necessary in this area dividing all unique visibilities by four with no change to the χ2 (Gehlot et al. 2019). of the fit, since two gains are multiplied onto the true visibility in Equation (71). Similarly, the absolute phase cannot be obtained from redundant calibration, because the transforma- 9.2. Calibration Using Self Consistency ( j ) tion ggii exp( ij) where is some real number leaves An alternative to sky-based calibration is to calibrate an the equations unchanged. Finally, the transformations ( ψ ψ interferometer by requiring that a properly calibrated data set ggii exp(iix r, y x) and ggii exp(iiy r, y y) where x and y must be internally self consistent. Recall from Section 3.3 that are constants, and ri,x and ri,y are the x and y positions of the ith ) fi in order to boost sensitivity to the power spectrum, many antenna, respectively also leave the goodness of t unchanged. ( interferometers have their antennas arranged in a regular grid. This linear phase gradient can be intuitively thought of in the fi ) This results in multiple copies of the same baseline, whose narrow- eld limit as a rotation of the sky. Since redundant ’ visibilities—once calibrated—should be identical up to noise calibration relies only on the self-consistency of one s data, the fluctuations. Turning this consistency into a requirement allows system of equations that govern the calibration procedure ( ) one to work backwards and to calibrate one’s interferometer cannot be sensitive to details about the sky Liu et al. 2010 . ( using redundant calibration algorithms24 (Wieringa 1992; Liu We thus have four degrees of freedom global amplitude, ) et al. 2010). global phase, and two linear phase gradients that are model unconstrained by redundant calibration. In fact, there are more For an interferometer with redundant baselines, Vij depends such degenerate degrees of freedom if we consider more than only on the difference between ri and rj. This is another way to satisfy the requirement from Section 9.1 of reducing the number just the calibration of a single polarization, which is implicitly what we have assumed in our discussion of redundant 24 The phrase “redundant calibration” is a rather unfortunate one, for it seems calibration so far. With multiple polarizations, there are four to imply that our calibration is unnecessary! While a name like “identical degeneracies per polarization direction if data from each baseline calibration” might be more appropriate, the term “redundant calibration” is fairly widespread in the literature, and so we employ it in this polarization are treated independently. If assumptions about the paper. sky signals are made in order to relate the different polarization

33 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw directions (e.g., if one assumes that the sky does not contain a Taylor expansion, with the zeroth order term being the any circularly polarized sources), then the number of visibility that would be measured at the fiducial point, and the degenerate parameters can be reduced (Dillon et al. 2018).In first-order corrections being determined by gradients in the two any case, to fix the degenerate parameters, one must appeal to a directions of the uv plane and the “distance” between a minimal amount of sky modeling (Li et al. 2018). Thus, visibility and the fiducial point on the plane. Solving for the sky redundant calibration is never truly sky-independent, and many is then tantamount to solving for the fiducial visibility and the of the problems discussed in Section 9.1 with sky-based two gradients for each cluster. With an array layout that is calibration (such as sky model completeness) end up resurfa- sufficiently regular, there are still enough measurements and cing in the context of redundant calibration (Byrne et al. 2019). few enough degrees of freedom to enable well-defined solutions. The weakness of the aforementioned approach is that it can 9.3. Calibration Using Hybrid Techniques only be used to account for certain types of perturbations, and Given the importance of calibration to the precision moreover, is still agnostic to the possible form of the sky. A operation of an interferometer, it is worthwhile to combine better solution is the CorrCal25 algorithm (Sievers 2017).In some of the calibration techniques discussed above. At the CorrCal, the visibility data from the different baselines are simplest level, sky-based calibration is needed to fix the modeled as random variables drawn from a correlated higher- degenerate parameters of redundant calibration. Going a little dimensional Gaussian distribution.26 Let vtrue be a vector further, a rough sky-based calibration (even with an incomplete storing the true visibilities that would be measured if all gain ) ( true) sky model can be used to provide a starting point in parameter parameters were unity i.e., it is a vectorized version of Vij , space for gradient descent algorithms that attempt to minimize and similarly for vmeas. The two vectors are related by a the redundant calibration χ2 (Zheng et al. 2014; Li et al. diagonal gain matrix G, such that 2018). vGvnmeas=+ true ,76() More generally, it is possible to write down calibration algorithms that naturally combine sky model information and where n is a similarly vectorized version of the noise. Since we the self-consistency requirement of redundant calibration. are modeling the data as being Gaussian distributed, their Because redundant calibration does not depend on the sky probability distribution is given by p()vmeasµ-exp (c 2 2 ), (up to the fixing of degenerate parameters), it is essentially where marginalizing over all possible skies, including those that are c2=+vNGCGv meas()† - 1 meas,77 () unphysical or unlikely to be correct, given prior knowledge. In effect, one is attempting to solve a harder problem than is with C being the covariance of the true visibilities and necessary, because the fact that our knowledge of the sky is N ºánn† ñ being the noise covariance matrix. It is C that imperfect should not mean that all sky information is allows us to encode redundancy and sky information. For discarded! Similarly, the redundancy of two baselines should example, suppose we only have data from two baselines. If not be regarded as an all-or-nothing proposition. Two baselines they are perfectly redundant, then we might model C as of an array may not be perfectly redundant in reality, perhaps due to primary beam differences or positional uncertainties in C = a2 11,78() antenna placement (among many other possibilities), but their ()11 data will likely still be highly correlated. Treating the where α is a parameter controlling the amplitude of the sky visibilities measured by two such partially redundant baselines signal. The ones in this matrix ensure that this sky signal is as being completely independent will likely result in greater perfectly correlated (i.e., identical) between the redundant uncertainty in calibration solutions. On the other hand, simply baselines. Having written down a c2 for our measurement, our ignoring problems such as antenna position offsets leads to calibration procedure then amounts to performing gradient biases in calibration solutions (Joseph et al. 2018). descent in the parameter space spanned by elements of G. One approach to including partial redundancy is to do so Sievers (2017) provides a discussion of how this can be done in perturbatively within the framework of redundant calibration a computationally tractable way for large arrays. (Liu et al. 2010). An example of a type of non-redundancy The strategy of encoding all the details of our measurement where this can be done is positional errors in antenna in the covariance is one that reproduces redundant calibration placement. The effect of small positional errors in antenna as a special case. This can be proven by re-interpreting α as an placement is to shift nominally redundant baselines slightly on the uv plane. As a result, these baselines no longer lie perfectly 25 https://github.com/sievers/corrcal on top of each other on the uv plane, but instead have some 26 In reality, the statement that the data is Gaussian is certainly not rigorously fi true. However, Sievers (2017) argues that the forms of non-Gaussianity small spread about ducial points on the plane. The visibilities typically appearing in interferometric data are unlikely to cause bias in one’s measured by each baseline in the cluster can then be written as calibration solutions, and should only result in an increased variance.

34 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw adjustable parameter. Taking a ¥is equivalent to insisting While it is true that a smooth function of frequency and time that the redundancy of the array be perfectly respected, and is simply a special case of a flexible function that is free to vary with such a limit the χ2 becomes equivalent to one for a arbitrarily, there can be a considerable downside to the latter. A redundant calibration algorithm (Sievers 2017). But crucially, calibration solution that is derived using only data at a single CorrCal allows redundancy to be relaxed. For example, if timestamp and frequency channel may be quite noisy. two baselines are expected to be similar but not perfectly Applying such a calibration solution to the data would then redundant, one can include this information by slightly result in an incorrectly calibrated data set that is spectrally depressing the off-diagonal elements in Equation (78).In unsmooth, even if both the instrument response and the true general, the precise level of correlation between two baselines sky were in fact spectrally smooth. In other words, by can be modeled, and this information can be inserted into the overfitting the noise (or biases due to sky-model incomplete- covariance. An example of this would be the imperfect (but ness) during the calibration process, false spectral and temporal significant) correlation between two baselines that are nomin- signatures can be imprinted in the data (Barry et al. 2016). This ally redundant but fail to be because of primary beam can negatively impact one’s ability to measure the cosmolo- differences. Assuming that the sky corresponds to a Gaussian gical 21 cm signal, for as we will discuss in Section 12, most field (as CorrCal does), the correlation between two such foreground mitigation algorithms identify foreground contami- baselines can be readily computed. In this way, partial nants by their spectral smoothness, which is destroyed if one’s correlation can be accounted for in one’s calibration solution. calibration algorithm imprints unsmooth spectral structure on CorrCal also allows one to encode priors on sky the data. ’ information, because this information manifests itself as a To avoid unwarranted structure in one s calibration solu- particular correlation structure between different baselines. For tions, there are a number of strategies that can be employed. instance, as we discussed in Section 9.1, a sky dominated by a One strategy is to limit the exposure of the calibration solution single point source gives rise to visibilities with constant to pieces of data that are likely to, say, be chromatic. For ( ) ( ) complex magnitude across all baselines, while diffuse emission example, Ewall-Wice et al. 2017 and Orosz et al. 2019 have gives rise to visibilities that decrease in amplitude with found that false chromaticity can be mitigated by reducing the increasing baseline length. By including this correlation weight of data from long baselines when solving for a structure between baselines in the covariance matrix, one calibration solution. This is because long baselines are changes the χ2 function to allow it to favor fits to the data that intrinsically more chromatic, since u = b l, which means are consistent with one’s prior on the sky. Precisely what prior that long baselines sample a larger variety of u modes as a information one wishes to include can be done fairly flexibly. function of frequency. The result is that when data from a For example, when including information about known bright variety of baselines are brought together to solve for a “ ” point sources, one may choose to include strong priors on their calibration solution, spectrally unsmooth artifacts can leak from long baselines to short baselines. Excluding long positions but not their fluxes. This may be prudent because baselines excludes the source of unsmooth artifacts, improving positions are generally better known than fluxes, particularly the overall solution. Note however, that while this may work since existing radio surveys may not cover the same frequency well for mitigating spectral structure, one must be cognizant of bands as one’s current survey. CorrCal thus does not waste the fact that it is somewhat in conflict with the suggestion of prior information just because it is incomplete, and in our Section 9.1, where the exclusion of short baselines was example of bright source positions, this can substantially highlighted as a way to alleviate problems arising from sky improve the quality of one’s phase calibration. model incompleteness. Another strategy that can be employed to avoid excessive ʼ 9.4. Calibration Solutions as a Function of Frequency frequency structure is to extend CorrCal s system of and Time equations to include all frequency channels at once. In other words, rather than writing (and solving) a separate set of In our discussion of calibration so far, we have made no equations for each frequency, one can concatenate the vmeas mention of the frequency or time dependence of our calibration vectors from different channels into a longer vector and to solutions or our sky models; every algorithm that we have employ CorrCal as a single calibration step that calibrates all outlined can be applied on a frequency-by-frequency, time-by- frequencies at once. This allows one to encode covariance time basis. From a purely mathematical standpoint, this might information between not just the measurements of different even be considered necessary, as an instrument’s response baselines, but also measurements between different frequen- could in principle fluctuate greatly as a function of time and cies. Thus, just as one can impose priors on the sky’s spatial frequency. In practice, instruments are generally designed to be structures during calibration (as discussed in the previous as stable as possible, making it less important that a calibration section with the example of point sources), one can impose solution be able to respond to sudden changes. priors on its frequency structure. This enables the imposition of

35 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw smoothness requirements, allowing one to solve for a smooth power spectrum of true calibration variations. The identifica- bandpass calibration. tion of a two-component calibration “signal” plus noise model Of course, extending a system of calibration equations to then enables the construction of a Wiener filter (essentially a perform a simultaneous calibration across all frequencies is a “signal over signal plus noise” weighting in the Fourier dual computational demanding venture, since the relevant matrices space to time and frequency). Applying the resulting Wiener are now much bigger. In some cases, calibration routines have filter to the calibration solutions smooths out the solutions, in resorted to large-scale distributed computing. A notable principle providing the closest possible estimate to the true example of this is the SAGECal-CO algorithm. The original calibration variations as possible amidst noise. incarnation of this calibration, SAGECal27 (Yatawatta et al. Having described various techniques for ensuring smooth 2008; Kazemi et al. 2011), is essentially an iterative sky model- calibration parameters, we note that this pursuit of smoothness based scheme for performing direction-dependent calibration, is predicated on the assumption that the inherent instrumental using the space alternating generalized expectation (SAGE) response is spectrally smooth and stable in time. If, for maximization for which it is named. The -CO variant brings in example, one’s instrumental response is a complicated function the technique of consensus optimization (Yatawatta 2015, 2016, of frequency, the calibration should reflect this and inherit the 2018; Yatawatta et al. 2017). In the specific case of multi- same complication. One might think, then, that with current frequency radio interferometric calibration, one demands that hardware designs that often still show non-trivial fluctuations, the calibration solutions are well-fit by smooth functions of one should avoid imposing a smoothness prior. However, with frequencies (e.g., polynomials). The consensus optimization experiments currently still in the regime of setting upper limits, algorithm allows this to be achieved by a multi-stage iterative imposing smoothness on one’s calibration solutions is process. First, each distributed computational system calibrates generally a conservative choice. By doing so, one is essentially data from one specific frequency. The results from all the leaving residual spectral structure in the data, which generally different frequencies are then brought together and fit to the increases the level of the measured power spectrum. This gives fi chosen smooth parameterization. The best- t smoothed rise to a conservative result, as long as the increased power calibration solutions are then sent back to the distributed level is interpreted as an upper limit and not a detection of the systems for further iteration. Upon convergence, one has 21 cm signal. For a detection, it is likely that extensive end-to- effectively solved a constrained optimization that can be end simulation suites will be required to show that the practically scaled to large distributed computational systems. measured power spectra have not been biased by calibration As an alternative to computationally intensive calibration errors. Ideally, these simulations should also capture the schemes that explicitly optimize over frequency, one may covariance between calibration errors and the power spectrum ’ pursue the simple strategy of simply smoothing one s errors, such that one can marginalize over the former in a way calibration solutions after an initial calibration. In fact, this that self-consistently inflates the latter. may be the most computationally viable strategy when considering the fact that ultimately, calibration solutions should have some levels of smoothness not just in frequency, but also 10. Map-making in time. Smoothing a calibration solution in frequency and/or time can be done in an ad hoc way, or by averaging/filtering Once calibrated, we may regard the output data produced by the solution in a way that is informed by some prior a correlator to be a visibility sample for every pair of inputs, at expectations (or observed performance) of an instrument. every frequency channel as a function of time. As each One way to achieve the latter is to study one’s instrument in visibility is approximately a Fourier mode of some part of the detail. For example, Barry (2018) performed a thorough sky, they are not by themselves particularly easy to interpret. comparative study of calibration solutions as a function of Turning the raw data into something more useful is the different types of cables used within the MWA system, as well operation of map-making. as how these calibration solutions vary with measured However, we should emphasize that while map-making is an temperatures on the telescope site. Alternatively, one can try essential tool, it is not a prerequisite for further analysis of to model the behavior of the calibration solutions statistically. 21 cm data sets, particularly for areas like power spectrum Zheng et al. (2014), for example, study the MITEoR instrument estimation. In fact, analysis from maps can in some cases be by computing the power spectrum of their calibration solutions more problematic than proceeding directly from the underlying in frequency and in time. The resulting power spectra (of their visibilities. We will elaborate on this point later in the section calibration solutions) have the form of a white noise floor (i.e., and in Section 11. a flat power spectrum) plus an increase in power toward short The predominant method of map-making in radio astronomy timescales and over smooth frequency modes, indicative of a are variants on the CLEAN algorithm (Högbom 1974). However, certain properties of CLEAN make it problematic 27 http://sagecal.sourceforge.net/ for our purposes, so we will start our discussion by looking at

36 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw linear map-makers which are common in CMB cosmology where Pr(∣)vs is the probability distribution for v given s, and before returning to CLEAN in Section 10.2. C()xC, is a complex circularly symmetric Gaussian in x zero Generally, map-making does not couple frequencies, so for mean and covariance C.28 the rest of this section we will drop the frequency index and We will start with the easiest statistically meaningful map have it be implicit that the operations are performed on a that we can make. Starting with Equation (79), we apply the frequency by frequency basis. matrix N-12to pre-whiten the noise,29 giving To construct a linear map-maker we first need to define the ()()()Nv--12=+ NBsNn 12 - 12 .85 () forward model. In our case we need to first define a discretization of our visibility data, which is typically the data By pre-whitening we have transformed all the measured modes for every antenna-pair integrated into samples of fixed length in N-12v to have the same noise level. As all modes are now time. We will write this as a vector v. We also need to discretize equal in the eyes of statistics, to get back to something which the sky, usually into individual pixels in some scheme (e.g., looks like a map, we can simply project back with the HEALPix; Górski et al. 2005), which we denote as the vector s. transpose of the pre-whitened forward matrix, to give We describe the mapping between the two by --12† 12 sNBNvˆdirty = ()() ()86 vBsn=+ ()79 = BN† -1 v.87() which is the discretized version of Equation (51), where This quantity is called the dirty map, and in radio interferometry this operation (when performed in the uv plane) is also called gridding. We can interpret this operation as taking v()ijt= Vt ij () (80 ) the amount of signal observed in the data, and placing it back onto the sky with the (complex conjugate) of the transfer sr()pab= T ab(ˆ p)()81 function it was measured with. Though this the simplest way of transforming the data back into a map like quantity, in general a b -2piturij()·ˆ p ()Brr()()ijt pab = AtAtei (ˆp;;)(j * ˆp )() 82this does not yield a map which accurately reflects the sky. For

n()ijt=+nt ij ()discretization errors ( 83 ) an interferometer we typically think of the sky as having been convolved with the dirty beam, which is essentially the Fourier ( ) ( ) where the compound indices ijt and pab index over all the transform of the uv-plane sensitivity. fi feed pairs ij and time samples t, in the rst group, and all pixels Despite this, the dirty map is the basis for almost all map- ( ) fi p and polarization ab. As Equation 79 is discrete, for the xed making algorithms because it is a sufficient statistic for the true data to be equivalent to the true continuous case, we must sky s (meeting the Fisher–Neyman criterion). The dirty map is fi absorb a discretization error into the de nition of the discrete a biased estimator: by substituting Equation (79) into noise n. For a well chosen discretization scheme this Equation (86) and taking the expectation we find that contribution should not bias the output be and significantly † -1 smaller than the instrumental noise; in practice, achieving this áñ=sBNBsˆdirty ,88() this means choosing a pixelization with much higher resolution where not only can we expect the morphology to look different, than the smallest scales resolvable by the telescope. Provided but the map is not even dimensionally correct (having the scheme does not depend on the input data we would not dimensions of inverse temperature). expect it to bias the output maps. However, the form of Equation (88) goes give us a hint of how to make an unbiased estimator for the sky signal. We 10.1. Linear Map-makers could simply apply the inverse of the prefactor in Equation (88) Having discretized our problem we can proceed to build a to the sˆdirty. In the case where there are no unmeasured modes map-maker. Conceptually this is straightforward—we simply on the sky, this prefactor is in fact invertible, giving a new need to do a statistical inversion of Equation (79) to recover the estimate ( s) (v) sky represented by from our measurements . This sBNBBNvˆ = ()--11†† - 1.89 () inversion is statistical rather than exact, because we must ML account for the fact our data is noisy, and that in general the In actual fact, this is the maximum likelihood estimate for s, forward projection matrix B is not invertible. which we could have found more directly by maximisation of As this is a statistical inference, we need to account for the 28 We ignore any phase dependent couplings coming from the non-zero properties of the noise, which we assert to be Gaussian with ( ) † relation matrix on short baselines Myers et al. 2003 . zero mean and a known covariance N =ánn ñ. This allows us 29 By the negative square root we mean any factorization of the inverse noise to write down a likelihood function for the sky matrix such that N---11212= ()()NN†. This can be found either by Cholesky decomposition or, by eigendecomposition and replacing the eigenvalues with ()svsvBsN==-Pr (∣)C ( , ) ( 84 ) their square root.

37 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

--11† Equation (84). The inverse term (BN B ) attempts to extremes: if ε is zero, sˆW reduces to Equation (89), and one is deconvolve the dirty map into an accurate map of the sky. fully deconvolving the instrument’s response at the expense of Equation (89) is typically used for CMB map-making, although incorporating noise modes; if ε is large, the noisy modes are it has also been used in 21 cm cosmology (Masui et al. 2013). suppressed, but the resulting point spread functions relating the In the case that there are completely unmeasured modes of true sky and the obtained map are wider (Zheng et al. 2017b). the sky, the matrix BN-1 B† is singular and cannot be inverted. One of the major advantages of linear map-makers is that To avoid this we can write the maximum likelihood solution in their statistics are straightforward to calculate. For example for + terms of the Moore–Penrose pseudo-inverse (denoted with a ) the Wiener filter map-maker, the total covariance CW is defined ( )( sNBNvˆ = ()-+-12 12.90 ()in Equation 93 accurate when our model of the signal ML covariance is correct), and the noise covariance in map space is The distinction between this and Equation (89) is that singular † -1 modes are regularised by the Moore–Penrose pseudo-inverse to NCBNBCWW= W.95() have zero power. Though the maximum likelihood map is an unbiased and 10.2. CLEAN optimal estimate of the sky, in many ways it is not a practical The most commonly used map-making techniques within map-maker. As modes are essentially divided by their radio astronomy are variants of the CLEAN algorithm which sensitivity, in its pure form it will severely upweight noise- use a nonlinear process to perform the deconvolution of a dirty dominated modes in real data. This has dramatic effects when image. making a map and can add large amounts of noise below the CLEAN is an iterative algorithm with the (simplified) loop at resolution limit of the telescope, which may dominate any real iteration i being signal on larger scales. To resolve this we must find a way to regularize the inversion of these low sensitivity modes. 1. Form a new dirty image di from the set of visibilities at If we have some statistical knowledge about the sky in many the previous iteration vi-1 (if this is the first iteration use cases we can produce a much better map-maker by performing the raw visibilities v). That is a Bayesian inference of the sky. In the simplest case let us ii† --11 assume we do not know exactly what the sky looks like, but dBNv= .96() that we do have some knowledge of the two point statistics, 2. Select the brightest pixel in the dirty image which we can encode in a pixel-pixel covariance matrix S. This m = argmax di and add a fraction f of its flux into the allows us to place a Gaussian prior Pr()s on the sky maps. same pixel in the clean image si. Mathematically Bayes’ theorem tells us that ii-1 i Pr(∣)svµ Pr (∣) vs Pr () s , ( 91 ) ssaa=+dam, fdm.97() where Pr(∣)sv is the posterior distribution that we seek. Since 3. Update the residual visibilities by subtracting the both the prior and the likelihood (Equation (84)) are Gaussian, projection of the cleaned image to define the posterior is also Gaussian, which after some extensive ii manipulation we can write as vvBs=- .98() † -1 Pr(∣)sv=- ( s CW BN v , CW ) ( 92 ) This loop is processed until a sufficient number of components of have been added to the clean image. where the covariance CW is defined by Schematically, CLEAN works by assuming that the sky is a -1 --11† CSBNBW =+ .93() collection of point sources and constructing a model of them by Taking the expectation of this distribution (or equivalently the iteratively removing them from the dirty image. At radio maximum a posteriori point) gives us a new map-maker frequencies there are a significant number of point sources, and given that these have been a major area of research, a point- ˆ ----1111†† sSBNBBNvW =+[].94 ()source prior (combined with its algorithmic and computational This map-maker is a Wiener filter and has regularized the simplicity) has been very successful. However, this assumption inversion by only deconvolving modes where we expect the means that it is much more effective at producing maps of point signal amplitude to be larger than the noise (for these modes it sources than diffuse components like galactic synchrotron and is equivalent to the maximum-likelihood estimator). Modes 21 cm emission, though extensions to the algorithm can do where this is not true are suppressed by their signal to noise better (Schwab 1984; Cornwell 2008). This is in contrast to the ratio. We note that the Tikhonov regularisation (Eastwood et al. Wiener map-maker which struggles to deconvolve point 2018), where S-1 is set to eI, can be viewed as a special case sources (Berger et al. 2017). where we impose a white noise prior on the sky signal. The nonlinear and iterative nature of CLEAN (which comes Alternatively, it can also be viewed as a tradeoff between two from finding the maximum pixel in Step 2 above), means it is

38 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw particularly hard to analyze its statistical properties. Construct- is the telescope drift scanning)? In the latter two cases we are ing a covariance for the map output is not an easy process and forced to stitch together observations from different (but may be best tackled by Monte-Carlo simulation. overlapping) pointings into a single map, either by mosaicking (Ekers & Rots 1979), or more specialized techniques such as 10.3. Noise Covariance the m-mode formalism, which we consider in the next section. Finally, above we noted that CLEAN is particularly suited For maps to provide more than a simple visualization of our for imaging radio point sources whereas the Wiener map-maker data, we need to understand the uncertainties within them. This is more suited to diffuse imaging. In reality the sky is a mix of means that in addition to the map we need to be able to produce both components, and so map-making for 21 cm cosmology a covariance matrix. For linear map-makers, such as the Wiener may have to involve a mixture of CLEAN-like point source filter, the noise covariance is easy to calculate and does not peeling and linear diffuse imaging. depend on the map itself. In contrast the nonlinear process used by CLEAN makes it challenging to assess the noise covariance 10.5. Map-making with M-modes of the maps without large suites of Monte-Carlo simulations. Even where it is possible to calculate the noise covariance in In this section we describe the m-mode formalism, a recent the map basis it may be desirable more to produce power technique that allows wide-field polarimetric imaging to be spectra and other statistics from visibilities instead of from performed computationally efficiently. However, these gains maps. As the noise covariance is typically diagonal in visibility are at the expense of generality: the m-mode formalism works space between frequencies, times and baselines (or at the very only for drift scan telescopes, although these are predominant least has short ranged correlations between baselines) it is in upcoming 21 cm experiments (see Section 4). A full compactly represented. After we have transformed our data into treatment of the formalism can be found in Shaw et al. maps this is not the case and the noise will be correlated among (2015); we give an abbreviated version here, starting from pixels; typically, this coupling is between nearby pixels, though where we left off in Section 5. for some telescope configurations such as redundant arrays with large spacings this correlation can be non-local. 10.5.1. Introduction In Section 5 we described the sky emission in terms of the 10.4. Practicalities of Map-making coherency matrix Tab (rˆ); for the m-mode formalism it will be In the above discussion we have given short shrift to the useful to separate this explicitly into the individual fields for actual implementation of a realistic map-maker, focussing on each Stokes parameter. We can do that by writing the theoretical aspects of the various map-makers. In reality X TXab(rrˆ)(= å  ab ˆ)()99 these form only a small number of the factors that must be X considered. where XTQUVÎ [ ,,,] represents the different Stokes One major aspect is computational complexity. For mapping parameters in the sky. The polarization tensors  are simply large areas or working at high resolution the number of degrees ab the Pauli matrices of freedom we are solving for may be large as the number of fi pixels Npix, which is potentially in the billions. Super cially T 1 10 Q 1 10 2 ab ==, ab , any single projection such as Bs is an O()Npix operation and so 2 ()01 2 (01- ) this can be problematic both from a computational and storage 1 1 i U ==01, V 0 .() 100 perspective. This can be mitigated by taking advantage of the ab 2 ()10 ab 2 (-i 0 ) flat sky approximation to perform the matrix-vector products using FFTs (see Equation (69)). For the linear map-makers Using Equation (99) we can rewrite the visibility measure- there are still large matrix operations we must perform (i.e., ment Equation (51) as multiply by the matrix inverse for the deconvolution); if the full Vt()=W+ BX (rrˆ; tX)(ˆ)()() d nt 101 computation required for this is prohibitive, one option is to use ij å ò ij ij X the conjugate gradient algorithm to determine the inverse using fi X only the forward operations that can be applied via FFT (Myers where we de ne the beam transfer function Bij (rˆ; t) for each et al. 2003). In contrast CLEAN-like algorithms only require Stokes field as forward projections and so are likely to be computationally BtAtAteX (rrrˆ;)(= a ˆ ;)(b* ˆ ;)() X -2pitruˆ·()ij . 102 more straightforward. ij i j ab A second aspect is the observational strategy of the For drift scan telescopes, excepting noise for the moment, telescope. In particular, does it track a fixed location on the the time stream is periodic on the sidereal day and we can sky, tile multiple regions, or does it point at a fixed location in simply map time t into hour angle f. The presence of the local frame and let the sky rotate overhead (in other words, instrumental noise breaks the exact periodicity, but if we

39 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw average our time streams over sidereal days, the periodicity is 10.5.2. Application to Map-making approximately recovered. To use the m-mode formalism to construct a map-maker in To map this into the m-mode formalism we are going to the style of Section 10.1 we need to appropriately discretize make a series of transformations described in detail in Shaw and vectorize Equation (107). et al. (2014, 2015), which we can summarise as First, in the m-mode formalism the degrees of freedom are 1. Fourier transform the time stream with respect to hour naturally discrete as they are spherical harmonics of the sky. fi angle f. This transforms Vij()f  V ij; m, where m indexes Second, as the telescope has xed size, the sensitivity to large the different Fourier modes. values of l and m decreases exponentially beyond a certain 2. Take the spherical harmonic transform of the beam scale, and so we only need consider a finite range of each. We transfer functions and sky temperature and polarization. take l  2πDmax/λ and m  2πDEW/λ where Dmax is the X This maps Xa(rˆ)  lm, and Bij(rˆ; ff)()** B ij; lm and largest dimension of the array, and DEW is the total size in the transforms the sky integration to summation of l and m. east–west direction. See Shaw et al. (2015) for a more detailed 3. Apply the restriction that the pointing only changes by discussion. This means that for any given telescope we can Earth rotation. In terms of the Beam transfer matrices that easily represent all the quantities in Equation (107) as finite- imf is Bij;; lm()ff==Be ij lm (0 ) . Integrating over f now length vectors and matrices. generates a Kroenecker delta that we can sum over m. We can define our vectorisation as 4. Map the Q and U polarization into E and B mode []vB==VB ,, [ ] Y,d polarization. ()mij ij; m ()()mij¢ mYl ij; lm mm¢ []sn==anY , [ ] ( 108 ) After applying these transformations we can rewrite (101) as ()mYl lm ()mmaa; where the sky-like degrees of freedom have a compound index Y Y VBanij; m =+å ij; lm lm ij; m ()103 (mYl) comprised of m-mode, polarization, and spherical lY, harmonic l, and the visibility-like degrees of freedom are indexed as (mij±), which is comprised of m-mode, feed pair, where the polarization index Y Î [TEBV,,,]. This is an exact ( ) and positive or negative degree of freedom. Using this we can for transit telescopes mapping from the sky, described by ( ) Y write Equation 51 as multipoles alm, to our measurements, given by the Fourier coefficients of each visibilities time stream. The key thing to vBsn=+ ()109 note here is that the Fourier coefficient and the spherical from which we can apply all the results of Section 10.1. For harmonic order are the same, i.e., there is no summation instance we can use the Wiener map-maker, which we give over m. again here Our visibilities make complex valued measurements of the sSBNBBNvˆ =+[]----1111††. () 110 sky but the sky itself is real valued. This means that for any W m>0, while the coefficients Vij; m and Vij;- m are independent, Despite its complexity, there are significant advantages to the Y Y m-mode formalism. First, for drift scan telescopes it is the multipoles alm, and alm,- are not. To make this clear, we will group together the positive and negative m measurements essentially exact, making it naturally wide-field and fully together by defining polarized. While that was also true for the various linear map- makers in Section 10.1, as we noted in Section 10.4 the +- VVVVij; m ==ij; m ij;; m ij*- m ()104 extremely large matrices involved may be completely imprac- tical to invert or decompose. In contrast, for the m-mode formalism these operations are often completely tractable. BBBY,+-==-Y Y, ()1m BY* ( 105 ) ij; lm ij; lm ij; lm ij;, l- m To see why, let us note that the matrix B defined in Equation (108) is block diagonal, i.e., the interferometer itself +- nnnnij; m ==ij; m ij;; m ij*- m,() 106 does not couple m values. That means that if in Equation (110) the noise and sky covariances were also diagonal in m, the and now the m index only ranges for m 0. whole equation could be evaluated on an m-by-m basis. Let us Using this Equation (103) becomes look at these two contributions in turn:

 Y, Y 1. Noise—if the noise is stationary i.e., the correlations in VBana;m =+å a;lm lm a;m ()107 lY, the noise depend only on the time separation and not on the actual time itself for m 0. This is the m-mode measurement equation, á¢ñ=-¢ntnt()* ( ) N ( t t ), ( 111 ) describing how the polarised sky observed through our ij kl ()()ij kl telescope produces our measured visibilities. then the noise in m-mode space is also diagonal in m.

40 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

2. Sky—if the statistics of the sky do not depend explicitly on m, for instance if we model the sky as a Gaussian random field

Y Y¢¢YY áñ=aalm lm¢¢* Cl ddll¢¢ mm ()112 then the sky statistics are explicitly diagonal in m.In practice this is a good approximation for the 21 cm field, but may not be for the galactic foregrounds. If we can make both these approximations, the operations in Equation (110) can be applied m-by-m. This gives huge computational savings. For any matrix-vector multiplication we save a factor of O(mmax) in computation, and for any matrix- matrix operation such as eigendecomposition or inversion we 2  3 save a factor of O(mmax ). For many instruments mmax 10 and thus we can save 106 in computation. In practice, the approximation of noise stationarity is poor at Figure 9. A probability distribution function of the 21 cm brightness fi temperature image shown in Figure 2. The distribution is clearly non-Gaussian: low frequencies where the ampli er noise is small compared to = ( ( ( ) ) one sees not only a spike at Tb 0 due to patches of our universe that have the contribution from the sky making Equation 59 invalid . been reionized) but also a skewness in the rest of the histogram. These non- In this limit the noise is determined primarily by the brightness Gaussian features are not captured by measuring a power spectrum. of the sky which depends strongly on where the telescope is (A color version of this figure is available in the online journal.) pointing. However, for map-making, neglecting stationarity does not add any bias to the maps, it simply makes the fi estimation non-optimal. In many cases the signi cant computa- there is in principle some richly non-Gaussian information that tional savings may enable a more thorough analysis and can be extracted from data sets beyond the power spectrum. We ( ) outweigh the increased noise over the theoretical optimum of explore such possibilities in Section 14.2. the estimator. In this section, however, we focus on measurements of the power spectrum. Heuristically, Equation (16) suggests that the 11. Power Spectrum Estimation power spectrum can be estimated by casting the measured The ultimate richness of 21 cm cosmology is in its ability to 21 cm field in Fourier space and then squaring the result. While produce full, three-dimensional tomographic maps over a wide this is qualitatively correct, here we set up the problem of range of redshifts. This is in principle achievable using the power spectrum estimation in a more systematic fashion, which methods that we just discussed in Section 10. However, due to will enable a rigorous quantification of error properties. For sensitivity and systematic concerns, current experiments have concreteness, we illustrate how this works using the three- tended to focus on statistical characterizations of the cosmo- dimensional power spectrum P()k as an example (deferring a logical signal instead. A prime example of this is the power discussion of other types of power spectra to the examples in spectrum, defined by Equation (16). The power spectrum Appendix B).Wefirst discretize the continuous quantity P()k quantifies the variance (“power”) of spatial fluctuations in the into a series of piecewise constant bandpowers, which are then 21 cm brightness temperature field as a function of various grouped into a vector p, with each component signifying the length scales (“spectrum”). Measuring the power spectrum is a power at a different value of k. We then write a generic goal of all currently operating 21 cm instruments that are quadratic estimator pˆ of the power spectrum p: focused on spatial fluctuations. This emphasis is partly for pˆ º xE† a x,() 113 historical reasons, given the success of other cosmological a probes such as the CMB and galaxy surveys in extracting where x is the input data (discretized in some way and stored as information from their respective power spectra. However, it is a vector), Ea is a matrix whose form is chosen by the data important to stress that the power spectrum is an exhaustive analyst, and α indexes different components of p, i.e., P()k in summary of a data set’s information content only if the different bins/bands of predefined thickness in k-space. fluctuations follow Gaussian statistics. In the case of 21 cm The estimator that we have written is generic for any choice fluctuations, strong non-Gaussianities are present, particularly of basis for our input data. For example, one could imagine first during Cosmic Dawn and the EoR. This is illustrated in the preprocessing one’s data into a three-dimensional image cube clearly non-Gaussian probability distribution function of 21 cm of the 21 cm field. The vector x would then be a list of voxel brightness temperatures shown in Figure 9. Thus, while values in some predefined order, with each element of the measuring the power spectrum represents an excellent start, vector corresponding to one location in the cube. The matrix

41 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Ea might then take the (rough) form defined as

a iikraa··ji- kr ¶C Eij ~ ee ,() 114 Qa º ()118 ¶pa so that when inserted into Equation (113), the result is that the α two copies of x are each Fourier transformed before they are and is the response of the covariance matrix to the th power multiplied together to accomplish a squaring of the Fourier spectrum bandpower. That there is a linear relation between the coefficients—precisely what our heuristic recipe for power data covariance matrix C and the power spectrum is spectrum estimation suggested. As another example, imagine unsurprising, as the two quantities are just ensemble-averaged that the input data has already been Fourier transformed, such two-point statistics in different bases: the former in the native that different elements of x contain the Fourier coefficients for basis of the measurement and the latter in a discretized Fourier different k vectors. In such a basis, our matrix might take the basis. In fact, the quadratic estimator formalism that we are form developing here is not limited to the estimation of the rectilinear power spectrum P()k as defined by Equation (16). a Eij ~ ddijaa,() 115 Any two-point statistic can be estimated, including the angular power spectrum C (Equation (26)), the cross-angular power d i=j ℓ where ij is the Kronecker delta function, equal to 1 if and spectrum C (nn, ¢) (Equation (28)), and the spherical Fourier- zero otherwise. With this Ea matrix, one is essentially picking ℓ Bessel power spectrum Sℓ(k)(Equation (30)). The abstract out the αth Fourier coefficient to square. generality of the quadratic estimator formalism is what gives it In general, our data will not be written in bases that are as flexibility. However, this abstractness can make the rest of this easily relatable to the power spectrum as a three-dimensional section difficult to follow. Thus, we encourage readers who image cube or a set of Fourier coefficients. For a general prefer concrete examples to frequently consult Appendix B, quadratic estimator of the power spectrum, then, we require a where we explicitly work out some toy models of power formal link between the vector space of our input data and the fi spectrum estimation. Fourier space of our output. To do so, we de ne the covariance We can proceed here, however, with our quest for a general C x matrix of the data to be power spectrum estimator now that we have a precise relation Cxxºá† ñ,() 116 between the data covariance and the power spectrum band- powers. If we take the expectation value of Equation (113),we where (as with Equation (16)) the angle brackets á...ñ signify an obtain ensemble average.30 The covariance matrix (in whatever basis a * ab a ()0 it is written) can be related to the power spectrum via the áñ=pxxpˆa ååEEQECij áj i ñ=tr[]b + tr [ ]() . 119 ij b relation The second term is an additive bias term that must be ()0 a CC=+å pa Q,() 117 eliminated. If C0 represents an instrumental noise covariance, a for example, this bias term will not represent true sky power, which essentially makes the claim that the covariance matrix is but will instead be an instrumental noise bias. To understand its a linear function of the power spectrum. The matrix C()0 is a origin, recall that a power spectrum is a quantity that is constant term that does not depend on the power spectrum, quadratic in the data. Thus, even if the noise is equally likely to containing portions of the covariance that do not depend on the have positive or negative contributions in the data x, the sky (e.g., an instrumental noise covariance).31 The matrix Qa is squaring operation inherent in estimating a power spectrum makes the noise appear as a positive bias in the power spectrum 30 Of course, an ensemble average is not an operation that can be performed on estimate. To eliminate this bias, one must either model and real data. If the statistical properties of the data are well-understood, the fi covariance can be computed either analytically or via a large suite of subtract it off, or avoid incurring it in the rst place. The former simulations. Empirical estimations from the data are possible, for instance by approach requires having an exquisitely accurate model of C0, assuming stationary statistical properties in time and replacing the ensemble average with a time average. However, such approaches are not without risk, as and is therefore generally not favored. Instead, one often opts we describe in Section 12.5. In this section, though, it suffices to consider the for the latter approach, which can be accomplished by covariance an abstract theoretical quantity that is central to the framework of modifying Equation (113). Instead of having two copies of x, power spectrum estimation. a 31 In the way we have defined it, C()0 does not contain astrophysical one can instead compute an estimator of the form xE12 x, foregrounds. This means the estimated power spectrum will not just be the where x1 and x2 are two data sets with identical sky power spectrum of the cosmological signal. It becomes an effective power contributions but different realizations of the noise. For spectrum of total sky emission. Alternatively, one can choose to define C()0 such that it contains astrophysical foregrounds. The end result is the same, instance, if a telescope observes the same parts of the sky ( ) however, because then Equation 119 below acquires an additive foreground day after day, one might form x1 from data taken on odd- bias in addition to a noise bias. This foreground bias adds to the cosmological power spectrum, again giving an effective power spectrum of total sky numbered days, while x2 is formed from data taken on even- emission. numbered days. Assuming that the noise is uncorrelated across

42 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

† ( ) fi different days of observations, the cross-covariance áxx2 1 ñ accomplished by examining Equation 122 , which speci es between will not contain any noise terms, and the the relation between our power spectrum estimate pˆ and the Equation (119) will similarly not be additively biased. true power spectrum p. As written, it shows that each Assuming that the bias term has been eliminated, the bandpower estimate is a weighted sum of the true bandpowers; expectation value of our estimator can be written as for our estimate to be correctly normalized, the weights must be subject to some constraint. One reasonable choice would be to áñ=pWpˆ ,() 120 a å ab b impose the condition that the bandpowers are peak normalized, b so that Waa = 1. Another sensible choice would be to insist that where our weighted sum of true bandpowers be a weighted average. ab Mathematically, this would translate to the condition Wab º tr()EQ . () 121 åb Wab = 1. While the latter is more common in analyses of For notational convenience, we may write all of this as real data, we pick the former for simplicity in the derivation that follows. áñ=pWpˆ ()122 Aside from an overall normalization constraint, we would where we have grouped the different bandpowers of the true ideally also want a power spectrum estimator that has the power spectrum and the estimated power spectrum into the smallest possible error bars (“optimal”). To obtain error bars, vectors p and pˆ, respectively, and have grouped Wαβ values we compute the covariance of pˆa, which is given by into a window function matrix W. Each row of the window Vppppºáˆˆ ñ - á ˆ ñá ˆ ñ function matrix represents the window function for an estimate ab ab a b ( ) fi ab of a particular bandpower i.e., bin in k , which quanti es the =áñ-áñáñå EEij km ()()xxj i** xm xk xxj i * xm xk *. 123 linear combination of different k bins of the true power ijkm spectrum that are being probed. A sensible power spectrum estimator should be constructed to give bandpowers with To further simplify this expression, one must make assump- window functions that are relatively sharply peaked around tions about the underlying probability distribution function of x ( ) their target k bins. In the left column of Figure 10, we show the the data . First, we assume again for simplicity that the data window functions for a (completely artificial) survey of a one- consist of real-valued numbers. Essentially identical results can dimensional universe that is 200h-1 Mpc in extent. (This is be obtained with the more general assumption of complex- x essentially a one-dimensional version of the example that is valued data. Second, we assume that is drawn from a ) fi multivariate Gaussian field. Making this assumption allows one worked out in Appendix B.1. The different rows of the gure ’ / ’ ( show the window functions for different power spectrum to invoke Isserlis theorem Wick s theorem Isserlis 1918; ) x ( estimators (in other words, a different choices of Ea) that we Wick 1950 to express the four-point function of i.e., the expectation value containing four copies of x in will introduce below. For clarity, we show only four window ( )) functions, i.e., four rows of W, but in reality there is of course a Equation 123 as a sum of products of two-point functions. separate window function for every k bin that is probed. Phrased differently, all higher moments of a Gaussian Continuing with our development of the formalism, we distribution can be expressed in terms of its variance. For stress that the vector space inhabited by p , pˆ, and W is not the instance, one can show that same as the vector space inhabited by x and C. As illustrated in áñ=áñáñ+áñáñxxxijkm x xx ij x km x xx ik xx jm the examples above, the latter vector space is described by +áxx ñá xx ñ.() 124 whatever basis one chooses to describe the input data. The im jk former vector space, on the other hand, is the output Fourier α β a This allows the covariance between the th and th bandpower space of bandpowers. Linking the two spaces is Q , via to be written as Equation (117). As an example of this link,32 we note that if the ab data is written in position space and is free of practical Vab = 2tr()CE CE . () 125 observational considerations such as instrumental noise, then Equation (117) is simply a discretized version of the standard The covariances can be collected into a matrix V (which resides cosmological result that the correlation function is the Fourier in the output Fourier space of bandpowers) whose elements are transform of the power spectrum. We therefore expect that in a given by Vab. The error bars on our estimated power spectrum sensible power spectrum estimator, Ea should involve Qa. bandpowers correspond to the square root of the diagonal of In order to arrive at a specific form for Ea, one must impose this matrix, while the off-diagonal elements measure the extent several constraints on the problem. The first is that our power to which errors on different bandpowers are correlated. In the spectrum estimate be correctly normalized. This can be middle and right columns of Figure 10 we illustrate the diagonal and off-diagonal structures, respectively, for our toy 32 Again, for more concrete examples, see Appendix B. one-dimensional example. The diagonal errors are normalized

43 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Figure 10. Window functions, error bars, and error correlations for a minimum-variance estimator of the power spectrum (top row; Equation (130)), the MF= -1 estimator that has the slimmest possible window functions (second row), the estimator with uncorrelated errors on the bandpower where each row of the normalization matrix M is proportional to F-12(third row), and the Cholesky-based estimator discussed in Section 12.1.6 that is designed to minimize foreground contamination leakage from low k to high k (fourth row). The errors (middle column) are all plotted relative to the errors obtained from the minimum-variance estimator. In general, one sees that estimators with overly narrow window functions (second row) have the highest errors and neighboring bins that have negative error correlations (right column) between neighboring bins. The minimum variance error estimator achieves the smallest possible errors at the price of positive error correlations. The last two estimators strike a balance between the widths of their window functions and the sizes of their error bars, and are thus able to deliver power spectrum estimates with uncorrelated errors. (A color version of this figure is available in the online journal.)

to those of the minimum variance estimator that we are about to minimize Vaa subject to the constraint that Wαα=1. That is, derive. we minimize Continuing toward our goal of deriving an optimal estimator for the power spectrum, we use a Lagrange multiplier λ to  =-2tr()()CEaa CEl tr E aa Q . () 126

44 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Differentiating this with respect to Ea gives where the window functions are seen to be spikes at one k bin MF= -1 ¶ when . =-4CEaa Cl Q ,() 127 That the MF= -1 normalization gives an estimator with ¶Ea áppˆ ñ= is of theoretical significance. The result that áppˆ ñ= where we assumed that C, Ea, and Qa are all symmetric means that the estimator is one that has no multiplicative bias matrices. (Note that more generally, if x were complex the whatsoever. This is true regardless of whether one is referring relevant assumption here would be that the matrices are to a scalar multiplicative bias (where the overall scale of the Hermitian.) Setting the derivative equal to zero gives power spectrum is biased) or a matrix multiplicative bias (where the overall scale may be correct but different band- ECQCaaµ --11,() 128 powers of the true power spectrum may be mixed together in which in turns means that the estimator for a particular bandpower). Under the constraint of an unbiased estimator, the Cramer–Rao inequality guaran- pˆ µ xC† --11 Qa C x.() 129 a tees that the optimal estimator of the power spectrum has an Fixing the constant of proportionality can be done by imposing error covariance V that is given by the inverse of the Fisher our normalization constraint. This gives matrix, F-1. Such an estimator is optimal in the sense that no other estimator can have smaller error bars. Frustratingly, the 1 † --11a pˆa = xC Q C x,() 130 Cramer–Rao inequality does not provide a recipe for how to 2Faa construct this estimator. Fortunately, we have already inad- where Faa is the αth diagonal element of a more general object vertently written it down: our quadratic estimator with known as the Fisher matrix, which has elements MF= -1 is the optimal estimator of the power spectrum. To see this, note that in general, our quadratic estimator is defined 1 --11ab Fab = tr()CQCQ . () 131 by 2 a 1 --11a Equations (130) and (131) together define a power spectrum ECQCº å Mab ,() 136 2 estimator that is (by construction) a minimum variance b estimator. and combining this with Equation (125) yields Our minimum variance estimator is in fact just one example VMFM= t,() 137 of a broader class of optimal quadratic estimators. In deriving it, we assumed that the normalizing constant was a single scalar from which one can immediately see that VF= -1 if the number. More general, suppose we defined normalization matrix M is chosen to be F-1. -1 1 While the MF= estimator has theoretically attractive qˆ º xC† --11 Qa C x ()132 properties, in practice it has several shortcomings. For instance, a 2 although the Cramer–Rao inequality guarantees that our as unnormalized bandpowers. Grouping these into a vector estimator has the smallest possible error bars under the living in the output Fourier space (just as we did with p , pˆ, and constraint that áppˆ ñ= , it is possible to abandon this constraint W), we may normalize our bandpowers in a more general way in order to obtain an estimator with smaller errors. Indeed, by defining a normalization matrix M, such that Equation (130)(i.e., a quadratic estimator with a diagonal M) pMqˆˆ= .() 133 is precisely such an estimator, since it was derived by explicitly minimizing bandpower variances. The cost of this is that one Written in this way, our previous minimum variance estimator tends to obtain broader window functions, with W no longer is one where M was chosen to be diagonal. Another possible equal to I, as seen in Figure 10. Intuitively, one can make an choice for M would be to set MF= -1. Such an estimator has analogy to the binning of data. Narrow window functions (the the attractive property that áppˆ ñ= (i.e., W = I). This can be extreme example being W = I) essentially amount to narrow seen by observing that bins in k. Each bin contains very little information, and thus the errors are large. This is illustrated in the middle column of 1 --11ab áñºqpFpˆa ååtr()CQCQ b = ab b , () 134 Figure 10, where we show the expected errors relative to the 2 b b minimum-variance case. As expected, the narrower the window that is, áqFpˆñ= . As a result, one obtains ápMFpˆ ñ= , and functions, the larger the errors. Broad window functions are comparing this to Equation (122) gives equivalent to wide bins in k that average over many independent pieces of information, yielding small error bars WMF= .() 135 on the binned data. There is thus a trade-off between an One then sees that, indeed, picking MF= -1 gives W = I. estimator’s resolution in k and the error bars on each This is illustrated graphically in the second row of Figure 10, bandpower; in fact, this observation can be formalized into a

45 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw relation analogous to uncertainty principles found in quantum during the EoR, thanks to the complicated network of ionized mechanics (Tegmark 1995). bubbles. The presence of non-Gaussian fluctuations means that Another consideration in selecting an M matrix is the Equation (125) may not give strictly correct error information resulting correlation structure of errors. With its wide window (Mondal et al. 2015, 2016, 2017; Shaw et al. 2019). functions, the minimum-variance diagonal M estimator tends Additionally, our so-called optimal power spectrum estimator to give positive off-diagonal elements for V, and thus may not be strictly optimal with non-Gaussianities. However, neighboring bandpowers are positive correlated. This can be an inverse covariance weighting of the data is often still a good seen in the right column of Figure 10, where we plot the idea in practice. Moreover, it should be emphasized that even if correlation matrix of errors, where the ()ab, th element is our power spectrum estimator ceases to be an optimal one (and given by VabVV aa bb . One can see the light pink off-diagonal thus is not guaranteed to have the smallest possible error bars), bands, signifying a positive correlation between bins. On the it remains a perfectly valid estimator. Our estimator will other hand, MF= -1 generally gives negatively correlated continue to be correctly normalized, because a correctly bandpowers (notice the light blue off-diagonal bands).A normalized power spectrum is simply one where the window middle ground is to pick an M matrix that diagonalizes V. functions are correctly normalized. Being able to correctly From Equation (137), it is clear that one way to achieve this is normalize our estimator is then tantamount to being able to to pick an M matrix where each row is proportional to F-12 accurately calculate our window functions in the face of non- 33 (where the proportionality constants are chosen to satisfy the Gaussianity. Fortunately, Equation (121) satisfies this require- proper normalization of the window functions). With such an ment, since its derivation depended only on quadratic (i.e., M matrix, the V matrix of Equation (137) becomes diagonal, variance) statistics, and not higher-order statistics (like the four- and the estimated bandpowers have uncorrelated errors. As one point function of Equation (124)) that require extra assump- might expect, the window functions are not as narrow as they tions about Gaussianity in order for our algebraic simplifica- -1 were with MF= , nor are they as broad as they were with a tions to apply. diagonal M. Similarly, the error bars on the bandpowers are A second caveat to this is that our optimal estimator assumes intermediate in size. that the covariance matrix C is known. This is frequently not a In summary, we have derived a powerful framework for the good assumption, since C includes both instrumental noise and general problem of power spectrum estimation. With knowl- contributions from the sky. The former can be characterized to edge of the statistical properties of the data, via its covariance some extent using laboratory and in situ measurements, but the matrix C, the optimal quadratic estimator that we derived above latter is particularly uncertain, given that the low-frequency is guaranteed to deliver the smallest possible error bars. The radio sky has historically been poorly explored. (Although of -1 key is the application of the inverse covariance C to the data course this situation is changing with the advent of large scale ( ( )) e.g., in Equation 130 . An additional choice must be made 21 cm experiments!) One possible approach is to model the regarding the normalization matrix M. There is thus a whole covariance empirically from the data itself (Dillon et al. 2015a). family of optimal quadratic estimators, all of which weight the However, such empirical modeling must be performed with C-1 ( M) data by , but owing to different choices for , have extreme care, for it has the potential to cause signal loss (Cheng slightly different trade-offs between error bars, error covar- et al. 2018), i.e., a power spectrum estimate that is biased low. iances, and window functions. Note, however, that these We revisit this issue in much more depth in Section 12.5. quadratic estimators are all optimal so long as M is chosen to A third challenge is that of computational cost. A brute force be an invertible matrix (which is the case in all the examples application of Equation (129) results in rather large matrices. discussed above), since the multiplication of an invertible To be concrete, consider the length of the data vector x for a matrix does not change the information content of a vector. modern 21 cm interferometer. Such an interferometer might However, there are several important caveats to the approach consist of several hundred antennas. For concreteness, consider outlined here. First, certain portions of our derivation required a 300-element interferometer, which would then have visibility the assumption that x is drawn from a Gaussian distribution. In data from up to 300×299/2=44850 baselines. With each particular, while Equation (123) for the bandpower covariance baseline taking data over ()1000 frequency channels, x ends V does not assume Gaussianity, it does not simplify to up being a vector of length ∼50,000,000 per time integration! Equation (125) unless Gaussianity is assumed. In many 21 cm A typical time integration might be ∼1 s long, and to achieve applications, this assumption is violated. For example, the the required sensitivities, 21 cm experiments require a year to 21 cm brightness temperature field is highly non-Gaussian several years of total integration. One thus sees that naively, a 33 Note the plural, because each row may have a different proportionality brute force approach would result in x vectors that are rather constant. As such, it is incorrect to write MFµ -12(which would require unwieldy to store or manipulate. The problem becomes even every row to have the same proportionality constant). Unfortunately, this particular abuse of notation is fairly common in the literature, and for worse when one realizes that C has size equal to the length of x compactness we perpetuate this in Figure 10. squared, and needs to be inverted in Equation (129)!

46 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

To be fair, there are several generic ways in which the length while others are in configuration space (Trott et al. 2016). This of x can be reduced. First, even though an interferometer with last possibility is particularly popular for interferometric map- N antennas has N ()N - 12baselines, these baselines may not making over small patches of the sky (Dillon et al. 2015b), be unique. For example, instruments like CHIME, HIRAX, and since in the flat-sky approximation an interferometer is HERA have their antennas laid out on regular grids, which essentially sampling in Fourier space in the angular directions means that there are multiple baselines that in principle but configuration space in the frequency/radial direction. In all measure the same signals (up to some noise variations). of the above, map-making is lossless. Assuming that the data from these baselines have been Following map-making, data volumes are typically small correctly calibrated, the visibilities from identical copies of a enough that if one simply desires some non-optimal estimate given baseline type can be averaged together, dramatically of the power spectrum, the procedure is conceptually reducing the data volume. Another way in which the data straightforward. For example, if one had a three-dimensional volume can be reduced is to recognize the fact that typically, map in configuration space, one could simply implement ’ one does not analyze the full frequency bandwidth of one s Equation (19): Fourier transform all three directions, square the data set at once. This is because the full frequency range of a result, and normalize appropriately (with some binning during typical instrument maps to a large change in redshift, violating the process if desired). However, as we alluded to above, this is assumptions about stationary statistics that are inherent in a suboptimal process that does not produce the smallest power spectrum estimation.34 Finally, as the Earth rotates over possible errors. And importantly, the rigorous estimation of a full period of a day, the cosmological measurements repeat one’s error bars on the power spectrum is itself a challenge, themselves. Thus, some averaging can be done by folding perhaps even more so than the estimation of the power multiple sidereal days into one. Despite these reductions, power spectrum estimation con- spectrum itself. Said differently, implicitly accompanying the data x (now in tinues to be a computational challenge. Differing approaches to ) † this challenge have led to multiple styles of power spectrum the form of a map is its covariance Cxxºá ñ, which contains estimation that can be viewed as special cases of our general statistical error information that must be propagated through to fi quadratic estimator from Equation (113), if not our optimal the nal power spectrum estimate. In addition to covariances estimator of Equation (129). We now highlight the advantages that arise from the measurement process, the map-making and disadvantages of some common approaches. process itself may introduce extra correlations. Typically, even with the compressed size of a map compared to raw interferometric data, C is too large to be stored or manipulated. 11.1. Power Spectrum Estimation Via Map-making Thus, assumptions must be made regarding the structure of One way to deal with large data volumes is to compress the the C: raw time series into maps of the sky. This can be accomplished 1. Suppose the total covariance matrix C is decomposed using the methods outlined in Section 10. If done in this into the sum of a sky signal covariance matrix S and an manner, reducing a time series into a map is a lossless form of instrumental noise covariance matrix N. This is a ( ) information compression Tegmark 1997b , in the sense that reasonable decomposition as the instrumental noise is any error bars on parameter constraints derived from the maps generally uncorrelated with the sky signal. Had this not are provably just as good as those achievable from the raw time been the case, there would be cross-covariance terms series. Note that the resulting maps need not be images in between the signal and the noise. configuration space (i.e., with the axes of the maps being 2. The noise covariance is typically assumed to be diagonal comoving distances in three-dimensions). Maps can easily well in frequency. This is reasonable because the visibility be in harmonic space (e.g., spherical harmonic coefficients in data coming out of an interferometer is uncorrelated the angular directions and spherical Bessel function coefficients in the radial direction; Liu et al. 2016) or even some hybrid between different frequency channels, and the map- space where some directions are expressed in Fourier space making process usually proceeds independently for each frequency (unless the output of one’s map-making 34 In the definition of the power spectrum (Equation (16)), notice the presence pipeline is in harmonic space radially). It is important of the Dirac delta function, which says that different Fourier modes are to emphasize, however, that while the noise is indepen- independent. This is the result of assuming that the statistics of our universe are stationary, i.e., the statistical properties are the same in any location. To see dent, it is generally not white. Mathematically, this is why this is, consider how one would construct a feature—say, an anomalously equivalent to saying that the noise covariance is diagonal high-density spike—at some special location in our universe. To create such an outlier, a large number of the underlying Fourier modes would have to all peak but not proportional to the identity. This means that once at the chosen location. For such a conspiracy to occur, the Fourier modes must the data are Fourier transformed along frequency, the be correlated. We thus see that the assumption of independent Fourier modes (which is central to the definition of the power spectrum) is intimately tied to noise will cease to be uncorrelated, as the identity matrix the assumption of stationary statistics. is the only matrix that is diagonal in all bases.

47 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

3. In the angular directions, different Fourier modes are Vb (t) is thus given by frequently assumed to be uncorrelated. This approx- ~ -i2pnt imation is made particularly often when performing VdVeb()tnnfnº ò () (), ( 139 ) theoretical forecasts of an interferometer’s sensitivity to the power spectrum. The motivation is that an inter- where τ is the delay, and has dimensions of time. The function ferometer makes maps from visibilities, which are probes f(n) is an optional tapering function. If no taper is applied, this of a localized region in uv space. Rewriting Equation (69) is simply a tophat function that is zero outside the observed using the Convolution Theorem, we have frequency band. The delay-transformed visibility is a con- venient quantity to work with for two reason. First, the delay- ~ 2 VAIp=-()()uu¢¢¢ Tdu u , () 138 transformed visibility can be interpreted geometrically. If we ò consider a limiting case where the sky and the primary beam which suggests that if one’s bin size on the uv plane is are frequency independent, one sees that coarser than Ap, then each visibility provides an ⎡ ⎤ independent probe of a single uv pixel. We stress, ~ -+ic2pn(· t brˆ ) VdATdebp()tnfn=W (rrˆ)(ˆ)()⎣⎢ ⎦⎥ however, that this is a crude approximation, and that in òò practice one ought to use as fine a uv pixelization as is =WdAp (rrˆ)( Tˆ)(ft + br ·ˆ c). feasible, and then to quantify the (strong) covariances ò between nearby pixels. ()140 4. Some pipelines take advantage of the fact that the effective point-spread function of an interferometer (the If the observational band and the tapering function are  “synthesized beam” in interferometric parlance) is reasonably broad, f becomes narrow. Under such an approx- reasonably narrow once all baselines of an interferometer imation, f is zero unless t »-br· ˆ c, so each value of delay have been included in the analysis. Thus, there is limited receives its contribution only from parts of the sky that satisfy correlation between widely separated parts of the sky, and t »-br· ˆ c. This also motivates the term delay, because for one can split the sky into a small number of facets that are a given baseline b, the quantity b · rˆ c is the time delay assumed to be independent (Cornwell & Perley 1992; between the arrival of incoming radiation from direction rˆ at Tegmark & Zaldarriaga 2010; Dillon et al. 2015b). In this the first and the second antennas of the baseline. A particular manner, one is essentially making the approximation that delay mode thus receives radiation from a ring of constant one’s covariance matrices are block diagonal. Alterna- delay in the sky; taking a delay transform therefore enables tively, if a set of redundant overlapping facets are used, visibility data to be interpreted geometrically, essentially this is equivalent to assuming band diagonality of the allowing data from a single baseline to be “imaged” in 1D. covariance matrices. It is important to note, however, that Of course, this result is rigorously true only if the primary beam the assumption of independent facets can be dangerous and the sky emission are both frequency independent. Neither for interferometers with very regular antennas layouts. assumption holds in strict detail. However, it is frequently a This is because aliasing induced by strong grating lobes design goal of 21 cm experiments to produce a frequency of the synthesized beam may cause significant inter-facet independent (or at least a spectrally smooth) primary beam as correlations. much as possible. The sky emission is also a reasonably slowly varying function of frequency, since it is in practice dominated not by the redshifted 21 cm signal that we seek, but instead, by 11.2. Power Spectrum Estimation Via Delay Spectra foreground contaminants such as Galactic synchrotron radia- While a map-making approach uses visibility data from all tion (which we know from Section 7 are spectrally smooth). baselines of an array (minus any that are deemed to be The delay spectrum is therefore a useful tool for diagnosing and irrecoverably corrupted by systematics) to form a map before mitigating foregrounds. We discuss this further in Section 12. proceeding to a power spectrum, a delay spectrum approach In addition to being helpful for foreground mitigation, delay- does the opposite. Introduced by Parsons et al. (2012b) for transformed visibilities are natural stepping stones toward 21 cm power spectrum estimation, the delay spectrum approach power spectrum estimates from single baselines. Conceptually, is one where data from a single baseline (or possibly a group of the idea is again centered on the fact (discussed in Section 3.3) identical redundant baselines) is pushed through to a single that a single baseline’s visibility is a probe of a single k^ mode power spectrum estimate before combining data from multiple on the sky, with the mapping between the baseline vector and non-identical baselines. the k^ vector given by Equation (40). Since the different The delay spectrum is named after the delay transform, frequencies of the visibility map to different radial distances, a which is simply the Fourier transform of the frequency visibility is therefore a quantity that resides in a hybrid space spectrum of a single baseline. The delay transformed visibility where the angular directions are expressed in Fourier space

48 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

36 while the radial direction is in configuration space. Thus, a thus a sampling of T at ht» and u » n0 b c. Given this single Fourier transform of visibility data in frequency—a result, one expects that the absolute square of the delay delay transform—should be sufficient for recasting the data in transform should be proportional to the power spectrum (at the three-dimensional Fourier space where it can then be least in expectation). This is indeed the case: squared and binned to produce a power spectrum. ~ Importantly, however, one must recognize that this approach áñ=∣()∣Vdududdthhfthfth222  ( - )* ( - ) b ò 1 2 12 1 2 is only an approximate one. The reason for this is that ~~* Equation (40) is only approximate. A more exact expression for ´-AcAcp ()()nn0102bup bu - the mapping between the baseline vector b and k^ is given by  ´áTT()()uu1,,hh1 2 2 ñ 2 ~ 22 2pnb =--ò dudAhn∣p (0bu c )∣ ∣ fth ( )∣ P ( u, h ) k^ = ,() 141 cDc »=Pc()ubnht0 , = 2 ~ 22 which is identical to Equation (40) except the expression ´--ò dudAhn∣p (0bu c )∣ ∣ fth ( )∣ ν involves , the frequency of a visibility, rather than n0, the ⎛ ⎞ ⎜⎟n0b central frequency over some larger bandwidth. Our new ==P⎝u , ht =⎠ expression reveals that a single baseline in fact drifts through c 222 multiple k^ modes as a function of frequency. This means that ´ ò ddAqn∣p ( q )∣ ∣ fn ( )∣, ( 143 ) to properly perform a Fourier transform along the line of sight, one requires multiple baselines that have the same nb value where in the last equality we invoked Parseval’s theorem. In 35 across the radial extent of the transform. the preceding line, we made the approximation that the power The delay-spectrum approach to power spectrum estimation 2 2 spectrum varies slowly over the scales where ∣Ap∣ and ∣f∣ are avoids the complications of combining data from multiple non-zero. With this assumption, we were able to evaluate baselines. It assumes that for short baselines and relatively P()u, h at (ub,,hn)(= 0 c t) and to factor the term out of the short frequency ranges, nb does not vary substantially. In integral. This is mathematically reminiscent of the Feldman– ( ) effect, one is assuming that Equation 40 is a good Kaiser–Peacock (FKP) approximation that is commonly approximation. If this is the case, we can make the connection employed in galaxy surveys (Tegmark et al. 1998). between the delay spectrum and the power spectrum precise by With this derivation, one sees that an estimate of the power ( ) writing Equation 140 in terms of the three-dimensional spectrum can be obtained by squaring delay-transformed  ( Fourier transform T (u, h) of the sky emission, so that in the visibilities from a single baseline and dividing by the integral fl ) at-sky approximation on the right-hand side of Equation (143). While our derivation required an ensemble average, its omission in practice simply ~ 22-+icpn(·) t b q results in a power spectrum estimate that does not necessarily VddAebp()tqn= ()q fn () ò deliver precisely the true theoretical power spectrum, but ´ ò dude22hhi phn(·)+u q T()u, instead has some error bar associated with it (as it must). The delay spectrum approach provides a power spectrum 22 --i pn() t h ~ =-òòdudThhn()ubu, de fnn ()( Ap c ) estimation method that stays close to the primary data output of an interferometer: the visibilities from individual baselines. 2   ~ »--ò dudThhfthn()()(ubu,, Ap 0 c ) This has several advantages. First, it means that cuts to the data ( ()142 such as those due to malfunctioning baselines or radio frequency interference; see Section 12.4) can happen near the ’ where in the last line we made the approximation discussed end of one s analysis process. In addition, because power above, where nb is taken to be roughly constant and is spectra are estimated from individual baselines before they are averaged together, calibration requirements are in principle less evaluated at some representative frequency ν0. One sees from this expression that the delay-transformed visibility for a short stringent. For instance, suppose some calibration error caused ’ baseline is roughly the Fourier amplitude of a 3D Fourier each baseline s visibilities to be multiplied by a different complex phase. Such an error would have no effect on a delay- mode, because f and Ap are relatively compact. The integral is based power spectrum, since each baseline’s delay spectrum is 35 Of course, this problem only exists because we are using P(k) as an example in this section, which requires a Fourier transform along the line of 36 Conventionally, η is used to denote the true Fourier dual to frequency sight to estimate. It is not a problem for statistics such as Cℓ (n), for which only (constructed using data from multiple baselines), whereas τ is the Fourier dual single frequency channels are needed. Indeed, visibility-based estimators for to the frequency spectrum on a single baseline. The statement that η≈τ is Cℓ(ν) can be written down for this in a reasonably direct manner (Choudhuri therefore equivalent to saying that the delay approximation (i.e., that nnb » 0 b) et al. 2014). is a good one in the context of power spectrum estimation.

49 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw individually squared, causing complex phases to cancel out. On then average” algorithm, several methods have been proposed. the other hand, this type of calibration error (in essence, a For telescopes that track a particular field over time, Paul et al. relative calibration error between different baselines) would (2016) compute cross delay spectra of all possible pairs of time affect a power spectrum estimation method that is based on integrations over a long observation. For drift-scan telescopes, map-making, since map-making typically requires the combi- one can employ techniques such as fringe-rate filtering, which nation of data from multiple baselines into one map. Morales performs a weighted sliding average over a visibility time et al. (2019) provide a discussion of the different calibration series. If the weights of this average are carefully optimized, errors that arise in power spectrum pipelines that do not employ fringe-rate filtering can recover the full sensitivity of an the delay approximation versus those that do. “average then square” power spectrum estimation (Parsons The disadvantage of a delay spectrum approach is that it is et al. 2016). However, the result is a series of visibilities that often difficult to extract the full sensitivity of an array. Since have strong noise correlations between them, and care must be baselines are treated individually, any information about the taken to ensure that these correlations are properly propagated power spectrum that resides in the cross multiplication of two into one’s error budget. different baselines is lost. Baselines that are significantly In closing, we address some of the approximations made in different will have visibilities that are sensitive to significantly deriving the delay transform power spectrum estimator. In different Fourier modes of the sky; thus, their cross multi- deriving the normalizing scalar in Equation (143), we made use plication will not provide much extra sensitivity to the power of the flat-sky approximation. While many instruments spectrum. On the other hand, baselines that are similar in length designed for 21 cm cosmology have a large field of view and orientation (including those that become similar due to the (and therefore violate the strict regime of validity of the rotation synthesis effect discussed in Section 3.1) will be approximation), in practice one finds that the curved sky sensitive to the power spectrum. It is certainly possible to corrections to the derivation are negligible (Liu et al. 2016). generalize the delay spectrum estimates between two non- Another approximation is the FKP-like approximation of identical baselines—one essentially repeats the derivation that assuming that the power spectrum varies relatively slowly as a led to Equation (143), but with the cross multiplication áVV*ñ function of u and η. This assumption is typically violated if b1 b2 2 ’ between b1 and b2 in place of á∣∣Vb ñ. The result is simply a foreground contaminants are present in one s data, because slightly different normalization integral on the right-hand side foregrounds are orders of magnitude brighter than the of Equation (143) that must be divided out to obtain the power cosmological signal (see Section 7) and preferentially occupy spectrum (Zhang et al. 2018). However, by mixing information certain Fourier modes (see Section 12). As a result, when one from different baselines, the aforementioned calibration moves away from the foreground-dominated regions of Fourier advantages are lost. space, the power spectrum varies rapidly as u and/or η As presented here, the delay spectrum approach also does changes. Despite this issue, one can show that the normalizing not take full advantage of a long time integration to average scalar derived in Equation (143) still correctly normalizes the down noise. Implicit in the derivations here are that V ()n refers power spectrum. In other words, a power spectrum estimated to the visibility at a single instant in time. Typically, a power using the delay spectrum approach of this section remains a spectrum is estimated for each time instant, and the resulting perfectly valid estimator, provided its error properties are collection of power spectra are then averaged together. The properly accounted for. In particular, the FKP-like approx- delay spectrum approach is therefore a “square then average” imation in Equation (143) is tantamount to assuming that our method, whereas a map-making-based approach would be an power spectrum’s window functions are delta functions, since “average then square” method. The former represents an the square of the delay spectrum—once normalized—is equal incoherent averaging of the signal which does reduce noise, but to the power spectrum at a specific point in Fourier space, the latter is a coherent averaging that reduces noise more rather than being an integral over a finite region. Backing off quickly. More concretely, suppose that there are Nt number of from this approximation to the previous line, we see that the time samples in one’s data set. If Nt independent power spectra square of the delay spectrum is an integral over the power are formed, averaging them together should reduce error bars spectrum, and thus the window functions (with the FKP-like ) by ~1 Nt . On the other hand, if these Nt data samples are first approximation can be read off. In fact, because our estimator is combined coherently into a map of the 21 cm brightness a quadratic function of the data, the machinery of the quadratic temperature, it is the map’s error that goes down by ~1 Nt . estimator can be employed to provide a rigorous quantification Since power spectra are quadratic functions of the brightness of the statistics associated with the delay spectrum (Liu et al. temperature field, the resulting averaged power spectrum has 2014b). The quadratic estimator formalism can also be used to errors that scale roughly as ~1 Nt, which is a considerably quantify the effects of the delay approximation. This is faster scaling. To recover the lost sensitivity from a “square accomplished by constructing an estimator in the delay basis,

50 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw but avoiding the invocation of the delay approximation by Consider first the covariance matrix of the data in the m- computing window functions that show how a delay spectrum mode basis. This is given by estimate samples linear combinations of the true rectilinear i i¢ Fourier modes (Parsons et al. 2012b; Liu et al. 2014a). CAB º ⟨()()VVm nnm¢ ¢ *ñ i i¢* =+NAB å BBℓm ℓm¢¢ á aℓm()nn a ℓ¢¢ m (¢ )*ñ, ( 145 ) ℓℓ¢ 11.3. Power Spectrum Estimation Via m-mode Analysis where we have collapsed triplets of baseline, m mode, and A third way to achieve a computationally feasible power frequency indices into upper case Latin indices, and spectrum estimation pipeline is to use the m-mode formalism i i¢ * NAB ºánnm ()nnm¢ (¢ )ñ. One sees that elements of C with A that was presented in Section 10.5. Recall that the m-mode and B corresponding to mm¹ ¢ are not necessarily zero, formalism provides substantial computational savings by depending on the statistical properties of the spherical recognizing that for a drift scan telescope, if one Fourier harmonic coefficients on the right-hand side. The covariance transforms a set of visibilities in time, the mth Fourier matrix therefore mixes different m modes. This means that -1 coefficient Vm of the result depends only on spherical harmonic when enacting the C weighting of the x (as demanded by, mode coefficients aℓm with the same m indices. Suppressing e.g., Equation (129)), one loses the computational savings of polarization indices to focus our discussion on power spectrum working in an m mode basis, since all m modes must be estimation, we have simultaneously considered. To regain our computational savings, it is necessary to

i i i assume that the sky is statistically isotropic. With this VBanm ()nnnn=+å ℓm ()ℓm ()m (), ( 144 ) assumption, we can use the fact that ℓ áaaℓm()nn ℓ¢¢ m (¢ )*ñ= C ℓ ( nndd,¢ ) ℓℓ ¢ mm¢ , ( 146 ) where i is an index specifying the baseline, n i is an m where C (nn, ¢) is the cross power spectrum between Bi ℓ instrumental noise contribution, and ℓm is the beam transfer frequencies ν and n¢. With this simplification, the covariance matrix from Section 10.5. One sees that whereas there is a sum becomes over ℓ, there is no sum over m. Different m modes can therefore be solved for independently of one another. This is also a i i¢* CNAB=+ ABdnn mm¢ å BBℓm ℓm Cℓ (),¢ . () 147 conclusion that can be seen in the flat-sky derivation of ℓ Appendix A, where we bridge the intuition between the curved- The independence of different m modes is now manifest in the sky case and the uv plane. There, we see that only in the north– second term. The first term will exhibit the same dependence south direction is there a convolution of different Fourier provided the noise contributions are statistically stationary in modes together by the primary beam, whereas in the east–west time. Provided these assumptions are satisfied, we may proceed direction each Fourier mode remains independent. This is the by relating C (nn, ¢) to the power spectrum. In the flat sky flat-sky analog of different m modes being independent. ℓ approximation, for instance, one has The advantages of analyzing drift scan telescope data in the m-mode formalism in fact extend beyond map-making. This is 1 ¥ C ()nn, ¢ = dkcos ( kD D )() Pk , z ’ ℓ ò c important because as we remarked in Section 10.3, if one s end pDDc c¢ 0 ( ) goal is not a map but instead is, say, a power spectrum , then ⎡DDDk ⎤ ⎢ c  ⎥ there are advantages to dealing directly with visibilities in both » å cos()kDpD c a () 148 ⎣ pDD¢ ⎦ foreground subtraction and power spectrum estimation. That a c c the different m modes do not couple to one another makes where Dc and Dc¢ are the comoving line of sight distances many foreground subtraction algorithms presented in corresponding to ν and n¢, respectively, DDc is their difference, Section 12 easier to implement. and z is an effective (weighted mean) redshift between ν and n¢ The situation becomes more complicated in the context of (Datta et al. 2007; Shaw et al. 2014). A similar expression can power spectrum estimation, because further assumptions must be written down for a full curved-sky treatment (Datta et al. be made in order for different m modes to be independent. In 2007), and Eastwood et al. (2019) provides insight into when particular, if we look back at Equation (129), we see that we the flat-sky approximation is appropriate. For now, though, a need both C and Q to possess the same block-diagonal, m- continuing with our flat sky expression and inserting it into CAB decoupled structure. In Section 10.5.2, we in fact already yields an equation of the same form as Equation (117), with argued that C has the right structure, but we will do so again ⎡ ⎤ a i i¢* DDDkc  here in a more explicit way, in order to be able to compute Q = d ¢ BB ⎢ cos()kDD ⎥ . () 149 AB mm å ℓm ℓm ⎣ ¢  c ⎦ Qaaº¶C ¶p (Equation (118)). ℓ pDDc c

51 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

We thus see that the matrices needed for power spectrum exploring the space of possible power spectra and how well estimation (i.e., Qa and C) all satisfy the property that they do each one fits the observed data, a Bayesian approach in not couple different m modes. One can therefore use each m principle provides not just a best-fit answer to what the power mode to form its own estimate of the power spectrum, and then spectrum is, but also a full probability distribution of possible to average the different estimates together. In conclusion, the answers, thereby capturing a richer statistical description of computational savings of an m-mode analysis are preserved not one’s results than what a simple set of error bars provide. only for the map-making exercises of Section 10, but also for Mathematically, the key to pursuing the aforementioned power spectrum estimation. approach is Bayes’ theorem. Suppose we group a set of To arrive at our conclusion, it was necessary to assume a hypothetical power spectrum values into a vector q (with each statistically isotropic sky. This assumption is violated by element storing the value of the power spectrum in a different k foreground emission, and implies a coupling of different m bin, i.e., a bandpower). Computing the power spectrum is then modes. Foreground emission should be included in the tantamount to computing the probability distribution of q given modeling of C, for even if foreground subtraction has been some data d, which we denote Pr(∣)q d . To compute this performed prior to power spectrum estimation, the subtraction distribution, one invokes Bayes’ theorem, which states that will not have been perfect, and a residual covariance will be (∣d qq,Pr, ) (  ) present. To see why foreground covariances generally couple Pr(∣q d ,  )= ,() 150 different m modes, consider the fact that foreground emission Pr(∣d  ) in the Galactic plane is substantially stronger than that near the Galactic poles. To produce such a pattern of emission, the where  signifies the theoretical model, Pr(∣d ) is known as spherical harmonic modes that comprise the sky emission must the evidence (which we return to in Section 13.3), Pr()q is the constructively interfere at the Galactic plane, which cannot be prior on the parameters, and (∣)d q is the likelihood function. accomplished if the different m modes are uncorrelated as the The prior represents our previous belief (before folding in the different modes must collectively “know” to peak near the current data) in the distribution of the bandpowers. This is Galactic plane. Some advanced m-mode treatments allow updated by the likelihood function, which is where the forward limited relaxations of statistical isotropy (e.g., by including the modeling aspect of Bayesian inference enters. It is how details modeling of bright point sources; Berger et al. 2017), but have of the observation are incorporated into the analysis. As a toy yet to be shown to be free of artifacts that could compromise example, suppose we were “observing” a set of simulated the spectral smoothness of the data. In any case, it should be Fourier coefficients of a Gaussian random sky drawn from a noted that while the application of the above expressions for Qa power spectrum P()k . These Fourier coefficients are then and C to a statistically anisotropic sky is non-optimal (in the grouped into our data vector d = [()TTkk12, () ,...], and the sense of delivering the smallest possible error bars), it does not parameter vector that we wish to infer is a vector of result in biased power spectrum estimates. This can be shown bandpowers, i.e., q = [()PPkk12, ( ) ,...]. With no noise or by repeating the derivations following Equation (128), but with instrumental effects, the likelihood takes the form generic weighting matrices instead of C-1. This essentially 1 ⎛ 1 ⎞ mimics an incorrectly modeled covariance, and one can show ⎜⎟t -1 (∣)d q =-exp ⎝ dP[()]q d⎠, ( 151 ) that even in such a scenario, the resulting estimators satisfy ()2detp Nd [()]P q 2 áppˆ ñ= , i.e., they are unbiased.37 where Nd is the length of d, and P is a diagonal matrix with q 11.4. Bayesian Power Spectrum Estimation along the diagonal. This likelihood is simply a reflection of the fact that in this example our fluctuations are Gaussian, plus the As a final example of a power spectrum estimation method, fact that by definition the power spectrum is the variance of we discuss a Bayesian approach. Up until now, our discussions Fourier coefficients, as one can see from Equation (16). have centered on frequentist approaches to power spectrum If one thinks of the likelihood as a function of d with q fixed, estimation, where one follows a deterministic recipe for then it is a properly normalized probability distribution for the computing power spectra from the data. A Bayesian approach, observed data. However, if one interprets the likelihood as a on the other hand, begins with a procedure for forward function of q with d fixed at the observed values, then one does modeling some data given a possible underlying power not in general have a properly normalized probability spectrum. This forward-modeling process is then performed distribution, even if the constants of proportionality are many times for a wide variety of different possible power included in Equation (151). Fortunately, Bayes’ theorem is spectra (i.e., a wide variety of possible forms for P()k ), and the precisely what we need to obtain a probability distribution for resulting data are checked against the actual observed data. By q: one sees from Equation (150) that one just needs to multiply 37 Our claim is true provided the covariance modeling is not done empirically by the prior, and then to divide by the evidence Pr(∣d ). But (i.e., based on the data itself). See Section 12.5 for discussions of this subtlety. since the evidence does not depend on the parameters q,itis

52 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw simpler to think of this division as just a normalization of the be a set of visibilities that have been accumulated over time on posterior distribution. the uv plane. Sims et al. (2019) then take the approach of Conceptually, then, obtaining probability distributions for inferring a º [()TTkk12, () ,...] and p º [()PPkk12, ( ) ,...] the q is simple. The posterior is simply evaluated over a wide jointly.38 Using Bayes’ theorem gives variety of different q values using Bayes’ theorem. Computa- Pr(∣apd , , )µ (∣ dap , ,  )(∣) Pr ap ,  tionally, however, such a brute-force method quickly becomes infeasible, since the amount of parameter space volume that µ (∣da, ) Pr (∣ ap ,  ) Pr (∣ p  ) , ( 153 ) must be covered grows exponentially with the number of where in the second line we used the fact that Pr,()ap= parameters (i.e., with the length of the vector q). To deal with PrPr(∣)ap () p. We also dropped the p argument from the this, one generally employs a Markov Chain Monte Carlo likelihood because the data does not explicitly depend on p,since (MCMC) analysis. With an MCMC, one jumps around knowing a (the actual Fourier coefficients of our sky) completely parameter space with a predetermined set of rules, evaluating determines the data d, up to instrumental noise. Written in this the likelihood after each jump. The rules are chosen such that way, we have all the pieces that we need to evaluate the posterior: the resulting chain of parameter space locations are distributed Pr(∣p  )is simply the prior on our power spectrum, Pr,(∣ap  ) in a special way. In particular, forming a histogram of these is the probability distribution of the Fourier coefficients given the parameter space locations gives the posterior distribution power spectrum (which is precisely Equation (151),despitethe Pr(∣)q d . Sampling the distribution in this way mitigates the different notation),and(∣da,  )can be written as “curse of dimensionality,” where bad computational scalings ⎡ ⎤ with the number of dimensions make an explicit computation 1 † -1 (∣da,exp )µ--⎣⎢ ( dTaNdTa ) ( - )⎦⎥, ( 154 ) of the posterior prohibitive. 2 An MCMC analysis also makes it straightforward to assuming that the noise in the visibilities is Gaussian with marginalize over parameters that are not of interest. Algebrai- instrumental noise covariance matrix N. Here, T is a matrix θ cally, if one is not interested in, say, parameter 0, one simply that transforms Fourier coefficients a into visibilities. This (“ ”) averages over all plausible values of it marginalizing over it process can be written as the multiplication of a matrix because by integrating the posterior from -¥ to ¥ in θ0. In other going from {T()}ki to a configuration space map T ()r is a words, the quantity linear operation, as is Equation (12) for going from the map to ¥ visibilities. Pr(∣)(qq12 , ,...dd= Pr qqq 012 , , ,... ∣)()d q 0 152 ò-¥ With a full expression for Pr,(∣apd , ), one can sample from this distribution using MCMC techniques. Obtaining what is the correct marginalized probability distribution for the we are ultimately interested in—the final posterior Pr(∣)pd for remaining parameters. Taking this to the extreme, if one is the power spectrum p given the data d—then just requires ( interested in the distribution of a single parameter for instance, marginalizing over a. Sims et al. (2019) explore two different ) in order to quote a single error bar on the measurement , one ways to do this. One uses a Hamiltonian Monte Carlo method marginalizes over all parameters but the parameter of interest. to make the process computationally tractable. The other An MCMC analysis makes this trivial, as marginalization is performs an analytic marginalization over a. Either way, the equivalent to simply ignoring various parameters when end result is a full set of probability distributions for power constructing histograms. spectrum bandpowers. In fact, Sims & Pober (2020) goes ( Because MCMC analyses are ubiquitous in cosmology and further and incorporates a foreground model into the like- fi ) indeed, other elds as well , we will not review the details here, lihood, unifying power spectrum estimation with foreground ( ) instead referring readers to resources such as Mackay 2003 , mitigation. In doing so, one advantage of this Bayesian ( ) Hogg & Foreman-Mackey 2018 . We now return to the approach (over many of the approaches discussed in fi speci c problem of 21 cm power spectrum estimation and Section 12) is that it permits a statistically disciplined treatment fi provide a very brief and simpli ed introduction to the Bayesian of uncertainties in our understanding of foregrounds. methods pioneered by Sims et al. (2016, 2019), Sims & Pober (2020). 12. Foreground Mitigation While the likelihood function that we wrote down in Equation (151) is a good start, it assumes a set of perfectly Having established the general principles behind foreground measured Fourier coefficients {T()}k . If we had such a set of contamination and power spectrum estimation, we now give a i fi uncorrupted coefficients, power spectrum estimation would be sampling of various speci c mitigation methods that have been trivial. The challenge is to relate the power spectrum directly to proposed in the literature. While our implicit context will data coming off an interferometer. In other words, for our 38 fi This is in fact a simplified version of the Sims et al. (2019) approach, which Bayesian inference problem, we would like to de ne d to be includes a few extra parameters to account for spatial fluctuations that are on something closer to the raw data. For example, we could let d length scales beyond the extent of the survey volume being analyzed.

53 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw typically revolve around foreground mitigation for a power polynomials. The most slowly varying polynomials are then spectrum measurement, many of the methods we outline below discarded. These fits can be performed on intensity data that is are also applicable to other data products, such as images. We expressed as a function of frequency or in log–log space. The begin with an examination of astrophysical foreground latter has the advantage that log–log polynomials are power contamination, as those represent contaminants that are laws (with corrections that are small if the higher order terms unavoidable for any experiment are subdominant), which are a good fit to both empirical observations and simple physical models of foreground emission (see Section 7). However, in practice this can be 12.1. Astrophysical Contaminants difficult, because as we discussed in Section 3.1, interferom- ( ) 12.1.1. Parametrized Fits eters discard the zero mode i.e., the angular mean of the data. Thus, the spectra are not guaranteed to be positive, which Given that foreground sources are generally expected to be motivates avoiding log–log space. spectrally smooth, one may hope that foregrounds can be mitigated simply by fitting—and subtracting—smooth39 func- tions from the spectra. Some proposals involve first imaging 12.1.2. Non-parametric Fits ( the data e.g., by using some of the map-making techniques The aforementioned parametric fits are open to the criticism ) fi outlined in Section 10 , and then performing a pixel-by-pixel t that they require an arbitrary parameterization, i.e., an arbitrary ( of smooth functions to the spectra Di Matteo et al. 2002; choice of basis functions. On one hand, one requires only that ) Santos et al. 2005; Wang et al. 2006 . However, when taking the basis functions span the space of possible spectra. In other ( ) the resulting hopefully foreground-mitigated data through to a words, they simply need to form a proper basis. On the other fi power spectrum, one nds that this method performs rather hand, an unwise choice of basis could result in inefficient fi ( poorly on ne angular scales Bowman et al. 2009; Liu et al. foreground removal, where a large number of modes need to be ) fi 2009b . The reason is that ne angular scales are probed by the removed before the data are relatively foreground-free. This long baselines of an interferometer, and with many interfero- then runs the risk of removing the signal too, since in each fi metric array con gurations, this is precisely the regime where mode there (in principle) resides both foregrounds and baseline coverage becomes sparse. The consequences of this cosmological signal. In expanding one’s spectra in terms of a can be understood by thinking about the uv plane. Recall from set of pre-defined (but arbitrary) basis functions that are not ( ) Equation 11 that ∣u∣ ~ bcn . Now consider a single uv pixel connected to physical models of the foregrounds, one is simply on a part of the plane that is so sparse that it is measured by just hoping that two assumptions hold. First, that the foregrounds a single baseline, with no other baselines in the vicinity. While only occupy a small handful of the modes. Second, that they a baseline may probe this uv pixel for this frequency, at a completely dominate the cosmological signal in these modes. different frequency it probes a slightly different uv pixel. This Collectively, these assumptions—if true—minimize the prob- fi means that for a xed uv pixel, the baseline coverage is ability of accidentally projecting out the cosmological signal, intermittent as a function of frequency, and the alternation which is a subject that we will return to in Section 12.5. between zero and non-zero information content imprints Although it does not eradicate the aforementioned concerns, fi fi unsmooth spectral features that are dif cult to t out using an alternative approach to the problem is to use non-parametric smooth functions. To some extent, this can be mitigated by fitting techniques. Such techniques do not impose a priori fi performing the smooth spectral ts in a pixel-by-pixel manner functional forms to describe the foregrounds. Instead, they ( ) not in image space, but in uv space Liu et al. 2009a . This essentially provide a quantitative criterion by which to judge allows certain uv pixels to be given zero weight in the spectral the “smoothness” of the spectra, thus allowing the smooth fi ts at frequencies where there are no measurements. This portions of the data to be fit out. For example, Harker et al. partially alleviates the problem, but fundamentally, it is a (2009a) suggest a scheme known as Wp smoothing. In the Wp symptom of the intrinsic chromaticity of interferometry, which scheme, a curve is considered unsmooth (and therefore unlikely we alluded to in our discussion of delay spectra in Section 11.2. to contain foregrounds) if it has large changes in curvature over In the context of foregrounds, it makes spectrally smooth the frequency range of interest. The reason for selecting this foregrounds appear spectrally unsmooth when viewed through criterion rather than the actual curvature is that a function with the instrument. We discuss this in more detail in Section 12.1.5. persistent moderate levels of curvature integrated over a large fi In tting smooth functions to spectra, one must decide on frequency range would be considered an unsmooth function, fi what smooth functions to t. A popular choice is to decompose even though real foregrounds could easily exhibit such the spectra into a linear combination of orthogonal behavior. The Wp method identifies foregrounds by performing fi 39 least-squares ts to the data, but with a penalty for large We note that when we say “smooth,” we are not referring to the fi mathematician’sdefinition of “smooth,” i.e., infinitely differentiable. Rather, changes in curvature. The tting process is thus a balance we mean slowly varying. between two competing demands: the desire for a good fitto

54 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw the data (which would motivate the fit to pass through all data signal. With both the kernels defined, one can use the points even if this means a final spectrum with many wiggles) foreground kernel to write down the posterior distribution for and the desire to minimize changes in curvature (which drives new data points given the measured data, which are assumed to the fit to have fewer wiggles). The relative weights given to be distributed based on the sum of the two kernels. The result is these demands are adjusted to enable the best-possible a prediction—with quantified uncertainties—of the foreground identification of foregrounds in simulations. Once the smooth spectrum, which is then subtracted off the data to yield a (hopefully foreground-dominated) fit has been obtained, it is hopefully cleaned data set. subtracted from the original data to give (hopefully cosmolo- While GPR is a non-parametric approach to fitting in the gical signal-dominated) residuals. sense that it does not assume fixed functional forms for its basis As another example of a non-parameteric approach, we functions, one does have to make a choice about the form of the consider the proposal of Cho et al. (2012), which takes direct covariance functions. This is necessary to make the problem a advantage of the differing spectral coherence lengths of the well-conditioned one—it is simply impossible to fit data foregrounds versus the cosmological signal. Suppose one formed without making assumptions! However, to avoid making ( a stack of uv plane “images” at the different frequency channels assumptions that are too stringent which could come of observation. If these images are then averaged across dangerously close to declaring that we know a priori precisely ) frequency, the slowly varying foregrounds will accumulate what the foreground spectra are , the kernels used in most fi signal to noise more quickly than the rapidly varying (and GPRs possess tuneable parameters that are t for as part of the fi random) cosmological signal. Thus, the averaged map has a data tting exercise. The uncertainties in these parameters can fi suppressed cosmological signal, and can be used as an angular then be marginalized over in the nal predictions of the template for the foregrounds (either as an actual map or as a foreground spectra. While this does not change the fact that a statistical model in the form of an angular power spectrum).The form for the covariance had to be chosen, it does allow the data to drive the precise parameters of the covariance. For example, template can then be multiplied by an overall scaling factor at in our application we can allow the widths of the foreground every frequency and subtracted off the data, with the scaling and cosmological kernels to be learned from the data. In factors determined by minimizing the residuals of the fit. principle, one can (and should) even go one step further and A final example of a non-parametric approach is to use use statistical model selection techniques (such as the Bayesian Gaussian Process regression (GPR; see Rasmussen & Williams evidence; see Section 13.3) to confirm that the chosen 2006 for an introduction). Recently proposed in Mertens et al. covariance models are sensible choices for describing the data. (2018) for the 21 cm cosmology, the idea behind GPR foreground removal is to model the observed Nfreq data points of a spectrum as being drawn from an Nfreq-dimensional 12.1.3. Mode Projection × Gaussian distribution. The discrete Nfreq Nfreq covariance Mode projection is yet another way in which foregrounds matrix of the distribution is then assumed to be a discrete can be removed. By this, we mean expressing the data in some sampling of a continuous covariance K ()nn, ¢ ,whichtakesona basis, and then zeroing out the contribution from selected basis particular functional form that is chosen by the data analyst. With components that are believed to be dominated by foregrounds. knowledge of this continuous covariance, one has constraints on The polynomial subtractions of Section 12.1.1 are in fact an the correlations between the measured data points and a example of this, since fitting out low-order polynomials is ( ) hypothetical new data point anywhere on frequency axis. This equivalent to expressing the spectra in terms of orthogonal enables the prediction of a series of new data points as a function polynomials, and then projecting out the lowest order terms. of frequency, thus producing a fit to the data. Here, we examine proposed mode projection schemes that use For the application to 21 cm, the overall covariance function information beyond the simple observation that foregrounds is often modeled as the sum of covariance functions that tend to be spectrally smooth. describe each of the components. Thus, one might have One example of going beyond spectral smoothness is the KKK()nn, ¢ º+21 fg, where K21 and Kfg are covariance removal of bright point sources. Implicitly, the motivation for functions for the 21 cm signal and the foregrounds, respec- this is the (extremely reasonable) prior that anomalously bright tively. A popular model for these functions are the Matern pixels are almost certainly foreground sources. In CMB kernels. The Gaussian is a special case of a Matern kernel, and observations, bright point sources can be removed simply by qualitatively, the Matern kernels peak at n = n¢ and decay masking out the pixels where they reside (e.g., Planck away from this equality. The characteristic width of this decay Collaboration et al. 2018b). In the case of 21 cm interferom- is an adjustable parameter, and is typically adjusted so that Kfg eters, the synthesized beam of an instrument has considerable has a longer coherence length than K21,reflecting the fact that structure, which makes point source removal much more the foreground spectra are smoother than the cosmological complicated (Pindor et al. 2011). Thus, the projecting out of

55 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw point sources generally requires a forward-modeling effort, which are then used to form a set of modes that are ordered where point source catalogs are propagated through a by their signal-to-contaminant ratio (Shaw et al. 2014). In this simulation of instrument visibilities, which are then subtracted way, one has greater confidence that the first few modes (i.e., from the data (Bernardi et al. 2011; Sullivan et al. 2012). the ones that are projected out of the data) have negligible Another strategy is to use the fact that however the cosmological signal in them, minimizing the possibility of foregrounds behave, they are likely to be large in amplitude signal loss. However, there is still little guidance as to precisely compared to the cosmological signal. Thus, the dominant how many modes ought to be projected out, and as one modes of a data are likely to be dominated by foregrounds, increases the number of modes, one inevitably begins to incur making it in principle possible to extract information about the signal loss. Another difficulty is that in principle, the formation foregrounds from the data itself. For instance, one could of the KL modes requires good models for both the signal and estimate the frequency–frequency covariance matrix from the contaminants. This is easier in some contexts than in others. data, and then perform an eigenvalue decomposition to identify For instance, since the post-reionization 21 cm signal traces the the strongest modes in frequency. These modes can then be dark matter distribution, it can generally be modeled reasonably projected out. This is essentially a Principal Component well (perhaps with some uncertainties revolving around the Analysis (PCA; Liu & Tegmark 2012), which is often bias). This is more difficult for reionization experiments, since implemented in practice using a Singular Value Decomposition the astrophysics of the era is largely unconstrained by (Paciga et al. 2013). Compared to a simple projecting out of, observations. In practice, though, the KL projection algorithm say, low-order polynomials, there are both pros and cons to this seems to be relatively robust even if the covariance matrices are approach. One advantage is that the projected modes are simply slightly mismodelled, so long as the most general properties whatever modes are dominant in the data, whether they are (such as the spectral smoothness of foregrounds) are accounted spectrally smooth or not. This enables the method to capture for Shaw et al. (2014). the fact that even if the intrinsic foregrounds are smooth, they Further variants are possible for identifying dominant (and may no longer be smooth once they are viewed through the therefore hopefully foreground) modes in the data. In general, instrument. However, it is important to emphasize that such an the problem can be phrased as one where a series of sky maps approach is not assumption free. Consider, for example, how at different frequencies are arranged into an Nfreq by Npix array one might go about estimating a frequency–frequency D, where Nfreq is the number of frequency channels and Npix is covariance matrix. From a theoretical standpoint, the covar- the number of sky pixels. The goal is then to decompose this iance is given by data array into the product of two arrays C and M: Cxx()nn,¢ ºá ()() n* n¢ ñ , () 155 DCM» ,() 156 where x()n is a visibility spectrum. Implicit in this definition is where C is an Nfreq by Ncomp array and Ncomp by Npix, where the ability to take an ensemble average, signified by á...ñ.In Ncomp is the number of independent components used to model practice, of course, one is not able to take true ensemble the data. We thus have spectral modes encoded in C, and averages, and one must employ some proxy. One could, for spatial modes encoded in M. In general, there is no unique example, replace the ensemble average with a time average as solution to this problem, since one can define a new set of one’s telescope surveys different parts of the sky. However, in matrices C¢ and M¢ that are able to produce an identical fitto doing so one is essentially averaging over the sky, and thus the the data (i.e., CM¢¢= CM) provided CC¢ º Y and resulting covariance captures only the average frequency– MM¢ º Y-1 , where Y is an arbitrary invertible matrix (Zheng frequency correlations. Any differences in statistical behavior et al. 2017a). Moreover, if the data require a large number of between, say, the Galactic plane and the Galactic poles is not components Ncomp to be properly modeled, then the number of modeled by this procedure. parameters in the model (NfreqNNN comp+ comp pix) may exceed The scheme that we have just described is a blind, data- the number of measurements (NfreqNpix). Assumptions or priors driven scheme. This has the advantage that it can capture must therefore be made to regularize the problem. The PCA unknown characteristics in the instrument and/or the sky. treatment described earlier in this section is equivalent to However, this is also a disadvantage because without an populating columns of C with a small handful of eigenvectors underlying model for the data, it is difficult to guard against of DDt . One then solves (or rather, fits) Equation (156) for M. over subtracting and destroying part of the signal (Switzer et al. Zheng et al. (2017a) takes this one step further in an iterative 2015). Some of this can be slightly mitigated by more approach: once M has been solved for, it can be held constant advanced techniques like the Karhunen–Loève (KL) transform. in another fit of Equation (156) to solve for a refined C. The With the KL technique, rather than ordering modes by their process can then be repeated until convergence. Other contribution to the total (cosmological signal plus foreground possibilities for regularizing this general problem include plus noise) variance in the data, one models the signal and the Independent Component Analysis (ICA; Chapman et al. contaminant (foreground plus noise) covariance matrices, 2012; Wolz et al. 2014, 2017a) and Generalized Morphological

56 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Component Analysis (GMCA; Chapman et al. 2013). In ICA, mitigation becomes impossible. On the other hand, suppose the one maximizes the non-Gaussianity in the derived modes. The foregrounds have different statistical properties than the reasoning is that if an identified mode is implicitly a sum of cosmological signal. For instance, over small bandwidths, the multiple independent components, the central limit theorem cosmological signal possesses stationary statistics (i.e., its ensures that this sum will be closer to Gaussian-distributed than statistical properties are translation-invariant), whereas the the independent components. Maximizing non-Gaussianity is foregrounds behave differently at different frequencies, being therefore a way to constrain the underlying independent modes. typically brighter at low frequencies. The C-1 method takes In GMCA, the spatial structure of foregrounds are considered advantage of this by downweighting the data from low in a wavelet basis, and one performs a fit to Equation (156) frequencies when estimating a particular cosmological Fourier while imposing a sparsity prior. mode. This is possible because any particular Fourier mode receives contributions from every frequency, and thus its 12.1.4. Mode Weighting amplitude can be estimated even if certain frequencies are An alternative to the schemes outlined above is the idea of suppressed. downweighting modes rather than projecting them out Of course, for modes where foregrounds are strongly completely. The motivation here is that even if foregrounds dominant over the signal, the cosmological information content are dominant in certain modes, the cosmological signal is gain from recovering those modes will be negligible, and mode typically not precisely zero. There is thus useful information to projection algorithms are an excellent approximation to the be extracted from all modes, which is something that cannot be optimal treatment. For modes where the foregrounds and signal done if any modes have been projected out completely. are more comparable, mode weighting can in principle allow One way to implement a downweighting algorithm is to improved science constraints from power spectrum measure- simply employ the optimal quadratic estimator formalism. ments. However, a potential challenge with this approach is the Because C contains a covariance contribution from fore- modeling of foreground covariances. Good foreground covar- fi grounds, the C-1 step of Equation (132) has the effect of iance models are dif cult to construct in an a priori fashion. downweighting foregrounds (Liu & Tegmark 2011). Essen- One is therefore driven to semi-empirical constraints. Model- fi tially, one is treating the foregrounds as a very correlated form ling the low-frequency sky can be dif cult, however, given that of noise, and mitigating this noise by inverse covariance surveys of the relevant frequencies are substantially incomplete ( weighting. As a toy example, imagine that the non-foreground although the situation is rapidly improving with the advent of portion of the covariance is given by the identity matrix I, and 21 cm cosmology! See, e.g., Hardcastle et al. 2016; Su et al. ) that the foreground covariance has rank 1: 2017b; Eastwood et al. 2018 . An alternative to this is to use the data themselves to model the foregrounds, although as we discuss in Section 12.5, this carries with it the substantial risk † CI=+l uu,() 157 that part of the cosmological signal can be destroyed in the process. For initial detection-era level experiments, then, it may λ where is some parameter controlling the strength of the be advisable to opt for a clean projection of contaminated mode foregrounds, and u is a unit vector of the same length as the rather than a riskier downweighting of a possibly inaccurate ( ) data vector in whatever basis one has chosen and might be model. In general, the latter approach should only be chosen if some template for the foregrounds. The Woodbury identity fi ’ -1 it is deemed necessary to recover speci c modes for one s then shows that applying C is the same as scientific constraints.

luu† CI-1 »- .() 158 12.1.5. Avoiding Foregrounds in the Foreground Wedge 1 + l Because of modeling difficulties, an alternative to the One therefore sees that C-1 has the effect of subtracting off a explicit removal of foregrounds is to simply avoid them. If portion of u, but that this subtraction does not become a foregrounds can be shown to be sequestered to certain parts of complete projection unless l ¥. To gain more intuition for Fourier space, then one could simply choose to exclude those how this procedure is able to suppress foregrounds, note the contaminated regions when fitting one’s measured power following. If u took precisely the form of a cosmological spectra to theoretical models. Fourier mode, then applying C-1 would suppress this fore- In principle, the foreground avoidance paradigm is a very ground contribution in the quantity Cx-1 , but pushing through natural one to pursue, since the foregrounds are in fact naturally to the end of the quadratic estimator, one would find that the sequestered in Fourier space. Consider the Pk( ^, k) power normalization matrix M of Equation (133) undoes this spectrum defined in Equation (32), where k^ is the Fourier suppression. This is simply the statement that if the fore- wavenumber for modes perpendicular to the line of sight, and grounds look precisely like the cosmological signal, foreground k for modes parallel to the line of sight. Since the line of sight

57 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw direction corresponds to the frequency axis, spectrally smooth intensity in terms of its Fourier wavenumbers then gives foregrounds in principle map to the lowest k bins of a power spectrum measurement. Thus, the foregrounds are naturally ~ dk3 ik r  2 ( ) Vb()t » eT* ()krr ddAn ^^p ( ) compact in the k⊥, k Fourier space, and can be avoided òò()2p 3 simply by looking away from the lowest k modes. iikikr· --an2 pnbr· ^ + t ^^  ()cDc In practice, the subtleties of interferometry complicate the ´ e dk3 simple picture we have just painted, and the foregrounds = eTik r* ()kr de2 ikr^^· òò3 ^ proliferate beyond the lowest k bins. The problem is ()2p ⎛ ⎞ particularly acute for interferometers with a sparse set of br· ak ´++A ()r dtD ⎜ ^ ⎟,() 161 p ^ ⎝ ⎠ baselines. As an extreme example of a sparse array, consider cDc 2p the simplest possible case of a sparse interferometer—a single baseline with a baseline vector b. From Section 3.3, we have where we have made the approximation that the mapping that in the flat-sky limit this baseline measures between frequency and radial distance (Equation (36)) is approximately linear over the observational band. This enables us to write Dc »-r* an, where 22-icDpnbr· VdrTAebp()nnn= ò ^^ ()()rr, ^ ,^ c , () 159 11c ()+ z 2 a º ,() 162 n H Ez() where for convenience in this section, we have implicitly 21 0 ( )( applied Equation 37 and suppressed the resulting factors of with r* being a constant with dimensions of length, and the ) Dc to write this integral in terms of r^ rather than q. As argued cosmological factors having the same meanings as they did in before, this baseline probes a particular angular scale on the Section 3.3. This quantity α is approximated as frequency sky, and by Fourier transforming over frequency, one accesses independent, as is the the factor of Dc in the integral over ν. Fourier modes along the line of sight of the observation. With a single baseline, we may say without loss of generality However, our previous treatment neglected that a baseline of a that the baseline vector is aligned with the x direction, where given fixed physical length and orientation b probes multiple r^ º ()xy, . Evaluating the r^ integral then gives scales, since the fringe pattern of an interferometer is frequency dependent. Mathematically, the scale probed by a baseline b is 3 ~ dk ik r-+ ikxcDc ()apt k 2 kb~ 2pncD , which is proportional to ν. This means that if Vb()t » ee* 2pb T()k ^ c ò ()2p 3 we were to, say, estimate the power spectrum using the delay ⎡ cD ⎤ spectrum approach, then the required Fourier transform over iky y c ´+dye Ap ⎣⎢ ()()aptky 2 ,⎦⎥ , 163 frequency would not be independent of the Fourier transform ò 2pb given by Equation (159). We neglected this in Equation (142) ° of Section 11.2 by hiding the frequency dependence of the where in this step we implicitly assumed 180 rotational exponent in Equation (159) under the definition of a (presumed symmetry in the beam, such that App()rr^^=-A ( ), simply to ( frequency-independent) variable u º nb c. Treating this avoid notational clutter of extra minus signs. If one desires, the fi frequency dependence properly in the Fourier transform over same nal conclusions can be reached without assuming any ) frequency, we compute symmetries . Proceeding to a power spectrum by squaring Vb (t) and taking the ensemble average, we obtain

~ ~ -i2pnt PV()k µá ∣ ()∣t 2 ñ VdeVb()tnº ò b () n b 3 br· dk 2 -+i2pn^ t = ddrTnnnrr A e ()cDc = P()k ò ^^()(),,p ^ . ò ()2p 3 ()160 ⎡ ⎤ 2 ik y cDc ´+dyey Ap ⎢ ()aptky 2 ,⎥ . () 164 ò ⎣ 2pb ⎦ To proceed, we make the approximation that the primary beam Ap is frequency independent. In general, this is a better In the parlance of Section 11, the squared term inside the approximation than neglecting the frequency dependence of the integral correspond to the (unnormalized) window functions of fringe pattern because the primary beam often changes slowly the delay power spectrum estimator: they inform us of the with frequency, unlike the fringe pattern, whose phase often specific linear combination (i.e., the integral, in this continuous goes through multiple periods of 2π within one’s observational case) of the true power spectrum P()k that we probe when bandwidth. Making this approximation and expressing the sky forming delay-based power spectrum estimators. The window

58 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw functions therefore quantify leakage on the Fourier plane, and A simple way to interpret the wedge is that since each are key to investigating how foregrounds proliferate in Fourier foreground Fourier mode is observed by a decreasing range of space in particular. frequencies as we move to higher k^, this decreases the As a simple toy model for foreground emission, suppose that observed correlation length, and thus increases the spread in D 40 Pk()µ d ( k ). Such a power spectrum might represent a delay (or equivalently k). We illustrate this in Figure 12. population of dim unresolved point sources that are unclustered That the foreground wedge has such a characteristic shape is (hence the k^-independent form that would make a picture of an opportunity, because it limits how badly foregrounds can the sky look like white noise), but with flat (i.e., frequency- proliferate. To see this, note that the primary beams of most independent) spectra. Inserting this into our expression and instruments fall off from their pointing centers, reaching zero at simplifying the result yields an angle q0 away (which, for many antenna designs, may not be until the horizon). Thus, the proliferated foreground power ⎡ ⎤ ⎡cD t ⎤ k c ()1 + z 2 ( ) PdyA()k µº22⎢ c , yA⎥ ⎢ ⎥,() 165 goes to zero when the argument of Ap in Equation 165 is ò p ⎣ ⎦ p ⎣ ⎦ b k^ H0 Ez() equal to Dc q0 (recall our unusual convention in this section where we express angles on the sky in terms of the 2 fi ) where Ap is the squared primary beam pro le integrated over perpendicular comoving distances that they subtend . This the direction perpendicular (“y,” in our notation) to the baseline means that the foregrounds do not extend above the line vector, and we have inserted the fact that a baseline of length b mostly probes k » 2pnbDc, and the delay mode τ mostly HDEz00c ()q ^ c kk = ^ ,() 166 probes k » 2pt a (refer back to Equations (141) and (162), cz()1 + plus the delay approximation th» of Section 11.2). What this shows is that even with an input foreground sky where all which is a conclusion that turns out to be almost exact (up to slight ambiguities about how to define q ) even in the presence emission resides in k = 0 modes, the resulting power spectrum 0 of full curved-sky effects that our flat-sky derivation ignored Pˆ ()k estimate contains power in a whole variety of k ¹ 0  (Liu et al. 2016). Equation (166) enables a strategy of modes. This has been termed mode mixing in the literature foreground avoidance, where one simply makes measurements (Morales et al. 2012). Because of the inherently chromatic outside the wedge in what is sometimes termed the Epoch of response of an interferometer, foregrounds have leaked from Reionization window. k = 0 to other modes. This can be seen by plotting the window  A strategy of avoidance is conceptually attractive in that one functions of our power spectrum estimator (Figure 11). One does not need to worry about an exquisite modeling and sees that when measuring power spectra high k , low k power ^  subtraction of the relevant foregrounds. However, it is not a (i.e., foreground power) is inadvertently mixed in. panacea for the formidable foreground problem of 21cm Mode mixing is both a problem and an opportunity. The cosmology. One limitation is that while the observational problem is that the foreground modes have proliferated evidence suggests that foregrounds do tend to be mostly (Switzer & Liu 2014), contaminating more modes than one sequestered to the wedge (Pober et al. 2013a), there is usually would naively expect. However, from Equation (165), we see some low-level leakage beyond the horizon line of that this proliferation has a clear signature in Fourier space— Equation (166). Because the foregrounds are orders of the contamination decreases as one moves toward higher k ,  magnitude brighter than the cosmological signal, even a little with the falloff taking the form of the squared primary beam leakage can make cosmological measurements prohibitive. profile. Lines of constant contamination on the k –k plane are ^  This leakage can occur for a variety of reasons. First, in our given by k µ k . This gives rise to the characteristic shape ^  derivation we assumed that foregrounds were perfectly flat in known as the foreground wedge (Datta et al. 2010; Morales frequency. In practice, the foreground spectra (while smooth) et al. 2012; Parsons et al. 2012b; Trott et al. 2012; Vedantham will not be perfectly flat. This means that their intrinsic spectra et al. 2012; Hazelton et al. 2013; Pober et al. 2013a; extend beyond the k = 0 mode, and the resulting wedge Thyagarajan et al. 2013; Liu et al. 2014a, 2014b), which is  moves up to higher k , reducing the available Fourier space for illustrated in Figure 5. With detailed simulations going beyond  cosmological measurements. Second, our derivation assumed a our toy model (including realistic instrument beams and systematics-free instrument. In practice, calibration errors can realistic foregrounds), the detailed profile of the wedge cause leakage beyond the wedge, as can instrumental effects becomes more complicated (such as the pitchfork effect for such as cable reflections (Ewall-Wice et al. 2016b; Kern et al. some instrument and foreground combinations; see Thyagar- 2019, 2020). As a result of these concerns, it may be advisable ajan et al. 2015a, 2015b for details). However, the overall to pursue hybrid strategies where foreground subtraction is shape remains a robust prediction. used in conjunction with foreground avoidance (Kerrigan et al. ) 40 For a similar calculation that uses a more sophisticated foreground model, 2018 . Precisely how successful such a hybrid approach might see Murray et al. (2017). be is still an open question. In forecasting papers in the

59 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Figure 11. An example set of window functions for a power spectrum estimator for a typical radio interferometer. Each window function is centered on a different point on the (kk^, ) plane. When one constructs a power spectrum estimator for a particular point, the measurement is in fact a linear combination of Fourier modes near that point. This linear combination is given by the relevant window function. One sees that the window functions are generally quite compact, but toward high k⊥, they develop long tails toward low k. The result is a leakage of foregrounds toward higher k⊥, which gives rise to the foreground wedge feature seen in Figure 5. (A color version of this figure is available in the online journal.) literature, one often sees the success of foreground subtraction have so many short baselines. Such short baselines primarily parametrized in terms of q0. One where foreground subtraction probe the small k⊥ parts of Fourier space, where the wedge does not allow one to push down to lower k than the line does not extend to such high k (and therefore k), preserving the indicated in Equation (166) with q0 ~ ()1 (representative of opportunity to measure high sensitivity low-k modes. Even so, the horizon) is often termed a “horizon wedge” scenario. A the mere act of cutting out the wedge modes results in a loss of more optimistic scenario of a “primary beam wedge” is often sensitivity. Additionally, it is important to note that even assumed, corresponding to a situation where k modes can be without the excision of the wedge, sensitivity can be reduced ( cleaned down to the line where θ0 is about the primary by error correlations between different Fourier modes Liu beamwidth. Even if foreground subtraction is not this et al. 2014a). There is thus a reduction in the number of successful (i.e., it does not allow a recovery of modes inside independent measurements, and this occurs throughout Fourier the horizon wedge), it may still be useful. This is because it space because the physics of mode mixing applies to all modes may still reduce the foreground amplitude in contaminated (e.g., Equation (164) shows that modes proliferate for all k). Its modes, such that low-level leakages of these reduced modes effect is just more apparent at low k because of the large outside the horizon end up being below the amplitude of the dynamic range between the foregrounds and the cosmological cosmological signal. signal. Another downside to a strategy of avoidance is that one In limiting one’s measurements to modes above the wedge, gives up on measuring certain Fourier modes. This results in one is taking advantage of the statistical isotropy of our lost sensitivity (Chapman et al. 2016), particularly since one is universe to probe a given k mode by measuring low k⊥ modes 2212 ( ) preferentially losing modes at low k, which is where signal to where k =+()kk^  » k Morales & Hewitt 2004 . How- noise is expected to be highest. This is because the instrumental ever, recall from Section 3.2 that even if the intrinsic power noise tends to be fairly flat as a function of k, which is often spectrum is isotropic, peculiar velocities induce redshift-space approximately k, since k is much larger than k⊥ for many distortions that make the measured power spectrum anisotropic. instruments, especially those probing reionization redshifts. In This anisotropy contains cosmological information (see contrast, the signal drops by orders of magnitude in P(k) as one Equation (178)), and so measuring it is of considerable interest. moves to higher k. It is for this reason that instruments such as The signature of redshift space distortions is a dependence of HERA (which was designed with explicit consideration of the the power spectrum on the orientation of one’s k mode, which foreground wedge; Dillon & Parsons 2016; DeBoer et al. 2017) is often parametrized using the parameter m º kk so that one

60 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

the right instrument and the right analysis, the wedge can be undone. However, as a practical matter, it is typically difficult to entirely rid oneself of the effects of the wedge. Consider first some proposed analysis strategies mitigating the wedge. One possible method would be to take advantage of the fact that picking MF= -1 in Equation (133) gives delta function window functions (see Figure 10 for a one-dimen- sional toy demonstration), where the power spectrum estimated for a particular Fourier mode is sourced only by power that intrinsically resides there. This can be used in the choice MFº ¢12P F- 1, where F-1 first attempts to sequester fore- ground power to their native low-k modes (essentially undoing the integral in Equation (164)), followed by a projection matrix P that zeros out such modes before F¢12performs what is effectively a smoothing operation on the Fourier plane to avoid the anti-correlation and high-error problems that were high- Figure 12. Each baseline (shaded as different colors in the above plot), lighted in Section 11 with a pure F-1 treatment. (We write integrates over a distinct and linearly increasing set of k⊥ modes at each F 12rather than F12to denote the minor adjustment that must frequency. The foregrounds are very highly correlated in frequency, but the ¢ apparent correlation length for each baseline can only be as long as Δν, the be made to account for the fact that in projecting out low-k distance in frequency that a single foreground Fourier mode is observed for modes, a small amount of cosmological signal will be lost, (i.e., the height of each shaded region at fixed k⊥). This distance Δν is which biases the final answer if no adjustments are made.) inversely proportional to the baseline length b. When performing the delay In simulations, wedge decorrelation using the M matrix has transform we observe a spread in delay of length D~tn1 Dµb because of ( ) the finite correlation in frequency. This spread is what we call the foreground had mixed success Liu et al. 2014b . On one hand, the wedge. Noting that k ∝ τ and k⊥ ∝ b we find the standard prediction that Δk technique does appear to reduce foreground contamination at ∝ k⊥, and if we carefully trace through the coefficients in this argument, we high k. However, numerical stability is an issue, since the reproduce Equation (166) exactly. decorrelation relies on small differences in the long tails of (A color version of this figure is available in the online journal.) different window functions on the k^–k plane. In addition, the notion that F-1 should be able to undo window function effects is constraining P(k, μ) rather than P(k). Excluding wedge is strictly true only if one assumes that the measured fi modes therefore reduces the range of μ values that are available temperature eld has stationary statistics, i.e., it has transla- for constraining the functional form of P(k, μ)(or more tionally invariant statistical properties and is therefore describ- typically, for fitting parametrized forms for P(k, μ)). This can able by a power spectrum. This is violated by foregrounds, reduce signal to noise in redshift space distortion investigations which typically have a higher amplitude at lower frequencies. (Pober 2015), as well as introduce biases in the spherically One variant of M matrix decorrelation is to use a judicious averaged power spectrum P(k)(Jensen et al. 2016). Addition- choice of M not to completely undo the effects of the window ally, it can make it difficult to employ reconstruction functions, but to instead limit leakage beyond the wedge. techniques (Seo & Hirata 2016) for undoing nonlinear Suppose that one were to order the elements of pˆ such that the evolution that have been successfully used to sharpen BAO bandpowers in the vector go from modes that are suspected to signatures in galaxy redshift surveys. be heavily contaminated (i.e., those in the wedge) to those that are less contaminated (typically those outside the wedge). Recalling that WM= F, if we select an M matrix such that W 12.1.6. Decorrelating the Foreground Wedge becomes upper triangular, then the window functions will The prevalence of interferometric approaches to 21 cm possess long tails away the wedge region, but cut off abruptly cosmology means that the wedge occupies a central role in toward the wedge. In other words, predominantly cosmological the foreground mitigation literature. In the past, there has been power is permitted to leak from outside the wedge to inside the considerable confusion as to whether the phenomenon of the wedge, but predominantly foreground power cannot leak from wedge is fundamental or not. More precisely, one can ask if it inside the wedge to outside the wedge. This can be is possible to undo the effects of the wedge, putting accomplished if we first perform a Cholesky decomposition † foregrounds back into the low-k Fourier modes that they on F, such that F º LL , where L is a lower triangular matrix. intrinsically occupy. Whether the foreground wedge can be Then, one sees that if MLµ -1 (with a row-by-row normal- undone is a question whose answer depends on both ization to make sure that rows of W sum to unity), then instrumental design and analysis strategy. In principle, with WLµ †, giving the upper-triangular, wedge-isolating form that

61 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw we desire. This is demonstrated for a toy one-dimensional foreground bias, they can still suffer from inflated variance survey in the bottom row of Figure 10. In one-dimension, the compared to a hypothetical foreground-free survey. This is wedge-contaminated modes are simply the low-k modes, and because the variance of the measured power takes the form one sees that the window functions permit leakage of power from high k values to low k values, but not the other way 22 Var()Pcross ~áxy ñ around. ~áxy22 ñ+á xy 22 ñ+á xy 22 ñ+á xy 22 ñ, An alternative to picking M wisely is to attempt to undo the s s s f f s f f effects of the wedge earlier in the pipeline, prior to the squaring ()168 step of forming power spectra. Conceptually, the wedge arises because interferometers are inherently chromatic instruments, and thus we see that the foregrounds do contribute to the with chromatic point spread functions. In principle, one could variance (and therefore the errors). Performing foreground attempt to undo these point spread functions in an intermediate subtraction on each data set prior to cross correlation does map-making step of one’s pipeline. In Equation (89) from reduce this variance, but the point remains that the effect of Section 10, for example, the inverse term undoes the point- foregrounds cannot be ignored simply because one is spread function of an instrument. If this inverse is computable, performing cross correlations. it will therefore have the effect of undoing the wedge. Our argument here is exactly analogous to ones that we However, the inverse does not always exist. For sparse made earlier regarding noise. In Section 3.1 we argued that interferometers with poor uv coverage, the inverse generally because interferometric visibilities are formed by cross-multi- does not exist. To undo the effects of the wedge, a necessary plying two independent antennas, they do not possess a noise condition is to have an instrument with sufficiently dense bias. However, they certainly do possess a noise variance baseline coverage that the true Fourier transform along the line (otherwise, interferometric measurements would have no of sight can be taken, using multiple baselines with equal nb uncertainties associated with them!) Similarly, in Section 11, (see Figure 12 and Section 11.2). Thus, removing or we argued that a noise bias does not arise when forming a suppressing the wedge requires not just an appropriate analysis cross-correlation power spectrum between two data sets taken pipeline, but also a suitable instrument such as a single dish at different times. But again, the power spectrum still ends up telescope (which can be thought of as an interferometer with a with a noise error bar. fully sampled uv plane up to the size of the aperture) or an Despite these caveats, cross correlations have been used to appropriately optimized interferometer (Murray & Trott 2018). great effect in the past. For example, to date the only detections of cosmological 21cm fluctuations have come from cross 12.1.7. Cross Correlations correlating GBT and Parkes data with galaxy survey data (Chang et al. 2010; Masui et al. 2013; Anderson et al. 2018a). Another way to combat foregrounds is to avoid measuring For higher redshifts there are fewer candidates for cross the 21cm auto power spectrum, but instead to measure the correlating with 21cm measurements, but a promising avenue cross power spectrum with some other tracer of cosmic might be [C II] intensity mapping surveys (Gong et al. 2012; structure. The key idea is that while each tracer may have its Chang et al. 2015; Kovetz et al. 2017, 2019a; Beane & share of foreground contaminants, the sources of these Lidz 2018; Dumitru et al. 2019). Although the aforementioned contaminants is different among the different tracers. Thus, challenges would have to be dealt with in the interpretation of a the foregrounds will be uncorrelated with one another, and if cross-correlation measurement, a consistent [C II]-H I measure- xxº+ xis the measurement from the first tracer and sf ment would be highly complementary to an H I auto power yyº+ yis the measurement from the second tracer, sf spectrum, boosting one’s confidence in an initial detection. heuristically we have 12.2. Interloper Lines P ~áxy ñ~á xy ñ+á xy ñ~á xy ñ.() 167 cross s s f f s s In our discussions so far, we have focused on mitigating the effect of broadband foregrounds, i.e., those that emit across a One sees that only the signal portion remains. While appealing, wide range of frequencies. In principle, foreground contamina- this result should be interpreted with caution. First, note that it tion can also occur because of discrete line emission from other only applies in expectation. In practice one can use the transition lines. More concretely, two different lines (say, with common cosmological practice of replacing the ensemble rest rest) rest frequencies n1 and n2 will appear at the same observed average with a spatial average. However, the quality of frequencies in our data if they are emitted at two different foreground mitigation is now tied to parameters such as one’s redshifts (denoted z1 and z2) such that survey area, since spatial averages are prone to unlucky fl nnrest rest uctuations when the spatial area is small. Moreover, while 1 = 2 .() 169 cross correlations can in principle yield results that are free of 11+ zz1 + 2

62 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

This particular form of the foreground problem is often called apparent structure of astronomical sources can also change (i.e., the problem of interloper lines. the effective point-spread function of a telescope can broaden In the case of H I 21 cm mapping, interloper lines are and acquire extra structure; seeing in the optical parlance). generally not a problem—in a fortunate accident, there are Finally, the ionized nature of the ionosphere causes the essentially no strong lines with the right wavelengths emitting polarization direction of radio waves to be Faraday rotated. at the right redshifts to contaminate the 21 cm signal. However, This can be troublesome for foreground subtraction because it this is not the case for other lines that are used in intensity can cause spectrally smooth foreground emission to appear mapping. For example, an experiment targeting emission from spectrally unsmooth. To see this, recall from Section 6.3 that in the 157.7 μm [C II] line at z=7 might easily confuse the [C II] general, the visibility measured by a baseline of a radio emission with the ladder of CO rotational lines that emit at interferometer exhibits polarization leakage, where linearly frequencies J×115 GHz as CO transitions from its (J + 1) th polarized emission can leak into what one interprets as rotational state to its Jth state. Estimates suggest that these unpolarized Stokes I emission. Now, Faraday rotation is a rotational lines are sufficiently strong at z∼0.9 (for the 43 frequency-dependent phenomenon, where the polarization transition),atz∼1.4 (for 5  4),atz∼1.9 (for 65 ), and direction of linearly polarized emission is rotated by an angle at z∼2.4 (for 7  6)(Lidz & Taylor 2016). proportional to λ2. Since polarization angles must be between 0 Several solutions to the problem of interloper lines have been and π, the result of this is that after Faraday rotation, a proposed. One is to mask strong point sources, which can be measurement that is projected along one linear polarization axis substantial contributors to the interloper problem (Breysse et al. can acquire a strongly oscillating frequency dependence. When 2015; Silva et al. 2015; Yue et al. 2015). Another possibility is this leaks into one’s Stokes I measurement, the result is a to match putative detections of a line with companion lines that spectrally unsmooth foreground that may be difficult to are also expected to be present at the same redshift, given distinguish from the cosmological signal (Moore et al. 2013; reasonable assumptions about the physics of galaxies (Kogut Kohn et al. 2016). Together, the various effects of the et al. 2015). Taking a statistical approach, one might also cross ionosphere can result in non-negligible biases in measured correlate with external data sets (Visbal & Loeb 2010; Gong 21 cm power spectra, particularly when the ionosphere is et al. 2012, 2014; Switzer 2017; Switzer et al. 2019). Another exhibiting strong activity (Trott et al. 2018). attractive option is to use the fact that a misidentified spectral Fortunately, each of the aforementioned ionospheric effects line will be misplaced when converting to a cosmological can be mitigated, at least in principle. Ionospheric absorption is coordinate system, but only in the radial direction and not in expected to be relatively spectrally smooth (Vedantham et al. the angular directions. The presence of interloper lines will 2014), making it a relatively benign influence on 21 cm therefore induce statistical anisotropies with predictable measurements. (Provided one is observing above the iono- signatures that can be fit for in data analyses (Cheng et al. sphere’s plasma frequency of ∼1 to few MHz, below which the 2016; Lidz & Taylor 2016; Liu et al. 2016). ionosphere is opaque). The situation is a little different for the issue of refractive shifts in position. This is typically viewed as a problem for direction-dependent calibration. For example, 12.3. Ionosphere one can use known positions of cataloged bright point sources Closer to Earth than astrophysical emission sources is the to simultaneously solve for antenna gains and ionospheric ionosphere. The ionosphere is a turbulent layer of our phase shifts (Mitchell et al. 2008; Jordan et al. 2017; Gehlot atmosphere that is ionized by incoming solar radiation (Kintner et al. 2018). In some cases, Global Positioning Satellite & Seyler 1985). The existence of the ionosphere impacts low- measurements can be used to provide supplemental information frequency radio observations in many ways. In fact, many to enhance solutions obtained under such a scheme (Arora et al. interferometers that were designed for cosmology have turned 2015). Note that even if one is able to obtain reasonable out to be excellent instruments for ionospheric studies (e.g., Loi direction-dependent calibration solutions, residual ionospheric et al. 2015a, 2015b, 2016; Mevius et al. 2016)! effects will remain, partly due to the imperfections in the In the context of cosmology, the ionosphere affects calibration process and partly because the ionosphere has observations in several ways. First, there is some level of variations on timescales shorter than the typical temporal absorption (Vedantham et al. 2014). More troublesome than the cadence at which one calibrates. These residual effects manifest absorption are the scintillation effects. As an ionized plasma, themselves in the data as an extra ionospherically sourced the ionosphere acts as a screen for radio waves, imprinting scintillation noise (Vedantham & Koopmans 2015). The effect extra time- and position-dependent phases. This can cause of this scintillation noise depends on the array configuration, refractive shifts in the apparent locations of radio sources on because of the interplay between baseline lengths and the the sky (known as the tip-tilt effect in optical astronomy). It can spatial and temporal coherence lengthscales of the ionosphere. also cause amplitude scintillations (twinkling in the optical Widely separated antennas, for example, are receiving incom- parlance) by focussing and defocussing the wavefronts. The ing radiation that has been scattered by different patches of the

63 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw ionosphere. Vedantham & Koopmans (2016) find that for operated on the ground. In many cases, these RFI sources need minimally redundant array configurations, scintillation noise not be local. For example, ionized meteor trails are known to enters at approximately the level as thermal noise, and is reflect RFI from faraway locations. At the other extreme, self- confined to the same wedge-like region in Fourier space (see generated RFI can be produced by insufficiently shielded Section 12.1.5) as foregrounds are in a power spectrum electronics in the telescopes themselves. In general, all these analysis. Thus, one might reasonably expect that foreground sources of RFI can be orders of magnitude brighter than avoidance techniques that take advantage of the wedge will sources of astrophysical origin (including foregrounds), and successfully mitigate scintillation noise. For compact, redun- therefore must be mitigated. dant array configurations with a high filling factor, scintillation The most basic strategy for RFI mitigation is to avoid the noise may be larger. However, the noise tends to be spatially RFI in the first place by building telescopes in remote locations. and spectrally coherent (Vedantham & Koopmans 2016), again Some sites are legally protected as radio-quiet zones. Examples enabling foreground mitigation techniques to do double duty include the Dominion Radio Astrophysical Observatory in 41 and suppress scintillation noise along with foregrounds. British Columbia, Canada, the Green Bank Observatory in The issue of polarization leakage can also be dealt with by West Virginia, USA, and the Square Kilometre Array sites in performing exquisite polarized calibration. Recall again from the South African Karoo desert and the Western Australian Section 6.3 that whereas polarization leakage cannot be desert (Bowman & Rogers 2010a; Sokolowski et al. 2016).Of properly corrected for using information from a single baseline, course, these sites are typically only protected against RFI in — — it can at least in principle be corrected for if one is able to certain frequency ranges, so a site that is appropriate for post- fi combine visibilities from a suf ciently large collection of reionization 21 cm measurements may not be appropriate for different baselines. In other words, in the limit where an array reionization measurements and vice versa. Other popular fi is con gured in the right way to image the sky, one has access locations are not necessarily legally protected, but are simply to information as a function of angular position, and so one can isolated or have natural geographical features (such as correct for polarization leakage, which is also a function of favourable combinations of mountains and valleys). Examples angular position. This is not a luxury that all telescopes have. of sites like this include the Owens Valley Radio Observatory For example, arrays that are configured to have a sparse set of (in California, USA; Eastwood et al. 2018), Marion island regular, repeated baselines cannot pursue such a strategy. (located halfway between South Africa and Antarctica; Philip However, one mitigating factor is that if observations are et al. 2019), the Forks (in Maine, USA; Zheng et al. 2014), the repeated night after night, the astrophysical signal should Timbaktu Collective (in southern India; Singh et al. 2018b), the remain constant while the specific realization of the random Ladakh site (in the Himalayas; Singh et al. 2018a), and the ionospheric fluctuations is different every night. Thus, folding Castrillon quarry (a defunct gold mine located in northern multiple days of data into an averaged data set can suppress ) polarized leakage systematics from the Faraday rotation caused Uruguay; Battye et al. 2016 . Looking ahead, there even exist by the ionosphere (Moore et al. 2017; Martinot et al. 2018). futuristic proposals to place radio arrays on the far side of the ( Ironically, then, the randomness of the ionosphere helps to Moon, avoiding both RFI and the ionosphere Lazio et al. ) partially solve the problems it has created! 2009 . Beyond site selection, RFI must also be removed during data analysis. A common strategy is to identify RFI based on its 12.4. Radio Frequency Interference characteristics as a function of frequency, time, amplitude, position, polarization, and any other axes over which In addition to contaminating signals arising from astro- observations are taken (Ekers & Bell 2002). Mobile radio physical sources and the ionosphere, low-frequency radio transmitters, for example, tend to be bright but might be used instruments must also contend with terrestrial contaminants. only sporadically. This leads to a signal that is localized in These contaminants are collectively known as radio frequency time, enabling an RFI mitigation strategy that involves flagging interference (RFI). There are a variety of ways that RFI can (and subsequently ignoring) time integrations that are extreme occur. Radio stations, for instance, broadcast strongly in many outliers in amplitude. Radio stations, on the other hand, tend to frequency ranges of interest to 21 cm cosmology, a prime be persistent in time, but only transmit at specific narrow example being the FM broadcast band from 88 to 108 MHz ranges in frequencies. One can thus flag specific outlier (corresponding to z∼12 to 15). Satellite or aircraft commu- frequency channels, an example of which can be seen in nications also contribute to RFI, as do radio transmitters being Figure 13. Generally, this is reasonably effective, but care must 41 Note that many of these results were derived in the weak-scattering regime, be taken to ensure that the brightness of the RFI does not push where the ionospheric phase fluctuations are assumed to be small over the first one’s electronics into a nonlinear regime, which can cause Fresnel zone. Observationally, the weak-scattering approximation seems to be ( a reasonable one except during short periods of high solar activity (H. narrowband RFI to leak into other frequencies Morales & Vedantham 2019, private communication). Wyithe 2010; Zheng et al. 2014).

64 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Figure 13. A qualitative representation of RFI in typical radio interferometric data. Because RFI (in red) tends to show up as outliers in data, one way to visualize it is to fit a model to the data (and one’s calibration solution) assuming no RFI, and to look for parts of the data set where the χ2 per degree of freedom is too high. Here we show this quantity as a function of time and frequency using data from the MITEoR telescope. One sees clear indications of broadband RFI localized to specific times, persistent RFI localized to specific frequencies, and intermittent RFI sources. For more details regarding the observations, please see Zheng et al. (2014). (A color version of this figure is available in the online journal.)

There are also more sophisticated algorithms than simple invariant in the sense that any expansions or contractions threshold flagging to combat RFI. As a first extension, we can (whether as a function of frequency or time) in some RFI note that very often RFI is not perfectly isolated to single pattern results in an identical expansion or contraction in the times/frequencies. Thus, rather than simply threshold flagging pattern of flags. The key idea here is that RFI often appears on a channel-by-channel or time-by-time basis, one might also with similar shapes at a variety of frequency and timescales consider additional flagging if neighboring times/frequencies (i.e., it is often scale invariant) depending on the details of the both exceed a lower threshold. This is the basis behind emissions sources and the instrument in question. The algorithms such as SumThreshold (Offringa et al. 2010).As algorithm for detecting such RFI should therefore respect this an extension of these algorithms, there also exist post- symmetry. fl / processing methods where agged regions shapes on the Another powerful method for identifying RFI is Principal – frequency time plane are expanded to include neighboring Component Analysis (PCA). Suppose we considered a set of times and frequencies (e.g., as applied in Winkel et al. 2007). raw visibilities coming from Nbl baselines, each having Such methods are sometimes referred to as morphological recorded data over Nfreq frequency channels for Nt time algorithms, since they expand flagged shapes in time and samples. Forming an Nt ´ NNbl freq array X with the data, one frequency. Taking advantage of morphology can be preferable can estimate a covariance matrix by computing XX†. In doing to lowering one’s thresholds as a strategy for catching so, one is producing a covariance estimate by averaging over additional RFI. This is because overly aggressive (i.e., low) time to yield an N NNN´ matrix. Performing an amplitude thresholds can result in a large number of false bl freq bl freq positives and additionally skew the statistics of the true signal. eigenmode decomposition of this covariance, the most For example, suppose our data consisted of Gaussian- dominant eigenmodes will tend to be sourced by persistent distributed signal plus some extremely bright RFI spikes. With sources of RFI, which survive the time-averaging operation. a high threshold, one removes only the RFI and preserves the The eigenvectors paired with the largest few eigenvalues can signal. As the threshold is lowered, however, one begins to then serve as templates for how RFI appears in the data. Armed mistakenly flag samples that reside in the extreme tails of the with such templates, one possibility would be to project out Gaussian-distributed data. By eliminating these data, one these RFI modes. Alternatively, note that a single eigenvector introduces an artificial skew in the data distribution. Thus, is a vector of length Nbl Nfreq, and thus can be thought of as a morphological methods may be particularly well suited for visibility data set in its own right (albeit without the time axis). detecting low-level RFI. An advanced example of a morpho- This set of visibilities can be imaged, providing a near-field logical algorithm is the one implemented in AOFlagger, image of the positions of RFI sources. This provides a literal which expands RFI shapes using a scale-invariant rank (SIR) real-world map that can used to physically locate RFI sources, operator (Offringa et al. 2012). The SIR operator is scale which can then be dismantled before further data is taken.

65 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

(Provided one has permission to do so!) Such a strategy has large numbers to obtain a small number, i.e., the cosmological been successfully deployed in Paciga et al. (2011).42 signal. One of the key challenges to this is that one may not be Recently, machine learning techniques have also been able to perform foreground mitigation well enough, leaving brought to bear on the problem of RFI flagging. Using residuals that are small but still dominate the cosmological techniques borrowed from the computer vision community, one signal. However, equally problematic (and perhaps more can treat a waterfall plot (i.e., data on the time-frequency plane, pernicious) is the problem of oversubtraction, where the such as what is shown in Figure 13) as an image with RFI cosmological signal itself is suppressed in the foreground features to be identified. Neural networks can then be trained subtraction process. Often termed signal loss in the literature, with example data (either from simulations where one knows this can lead to overly aggressive—and therefore incorrect— precisely where RFI was inserted, or from real data that have upper limits on the strength of the 21 cm signal. been flagged with a trusted algorithm) until they correctly Signal loss is particularly harmful when it is undiagnosed. identify RFI. By using waterfall plots of both visibility Already, this has led to multiple revisions of upper limits on the amplitude and phase as inputs to the neural network, Kerrigan 21 cm signal that have been published in the literature. As two ( ) fi et al. 2019 nd that they are able to train neural networks that particularly instructive case studies, consider Paciga et al. fl are generally competitive with state-of-the-art agging algo- (2011) versus Paciga et al. (2013) and Ali et al. (2015) versus rithms. An important exception is that their neural networks Ali et al. (2018). In the former case, Paciga et al. (2011) “ ” often miss RFI blip events that are localized to single pixels attempted to remove foregrounds from GMRT data by first in time-frequency space. However, machine learning RFI modeling the foregrounds using what amounts to a moving fl aggers are still in their infancy, and thus many improvements boxcar averaging of the data in the spectral direction. This are likely in the next few years. Moreover, they tend to be extracts the spectrally smooth component of the data, which is orders of magnitude quicker than traditional algorithms in then subtracted from the data. Upon further analysis, however, ( ) computational speed once the neural networks are trained , Paciga et al. (2013) found that the subtracted model and can therefore yield substantial computational savings for substantially overlapped with the Fourier modes where limits large next-generation telescopes that have enormous data on the cosmological signal were being placed. In other words, volumes. the cosmological signal was being subtracted along with the In closing, we point out that while RFI is in general a foregrounds. In Ali et al. (2015), foreground suppression was nuisance to be avoided, there are some limited situations where performed by the inverse covariance weighting method it can be helpful. For example, one would generally consider described in Section 12.1.4. However, without a priori the ORBCOMM satellite constellation (which broadcasts at covariance models, an empirical model for the covariance 137 MHz) to be an irritating source of RFI for 21 cm matrix was constructed by replacing the ensemble average in measurements targeting reionization. While this is true, the fi Cxxºá† ñ, with a time known broadcast frequency of these satellites enables them to the de nition of the covariance matrix, fi serve as calibrator sources that can be used to provide empirical average. If an in nite series of time samples had been available, mappings of the primary beam patterns of an antenna. With 30 this should have converged to the true covariance matrix. fi such satellites in low-Earth orbits that precess over time, these However, with a nite number of time samples, there can be ( sources transit through an antenna’s field of view quickly and errors in the empirical covariance estimate Dodelson & ) frequently, enabling a reasonably densely sampled mapping of Schneider 2013; Taylor & Joachimi 2014 . In the case of Ali ( ) the primary beam (Neben et al. 2015; Line et al. 2018). Similar et al. 2015 , this issue was exacerbated by the fringe rate fi techniques are being developed using drones that are outfitted ltering described in Section 11.2, where visibilities were pre- with radio transmitters (Jacobs et al. 2017). For drift-scan averaged together in order to increase signal to noise, at the telescopes, known astrophysical point sources can also be used. expense of decreasing the number of independent data samples However, with the motion of the sources governed entirely by for covariance estimation. To make worse, having few the Earth’s rotation relative to the celestial sphere, one cannot independent samples not only results in an inaccurate in general solve for a full beam profile without making covariance estimate, but also one where the estimate retains additional symmetry assumptions (Pober et al. 2012). “memory” of the random fluctuations in the actual data. With the covariance matrices now a strong function of the specific 12.5. Signal Loss realization of x itself (as opposed to just the statistical properties of x), the so-called quadratic estimate of the power With our need to subtract, suppress, or otherwise avoid spectrum in Equation (129) ceases to be a quadratic form. As a strong contaminants in our foregrounds, one is essentially in result, the usual normalization of the power spectrum in the business (whether directly or indirectly) of subtracting two Equation (130) does not apply, as it was derived assuming a 42 In that paper, the PCA is implemented using a Singular Value quadratic form for the power spectrum estimator. If the usual Decomposition, but the principle is the same. normalization is used anyway, the resulting power spectrum

66 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw estimate can be shown to be typically biased low (Cheng et al. our final power spectrum given our (considerable) uncertainties 2018), which is equivalent to saying that signal loss has occurred. in the behavior of the foregrounds. With these larger errors, the Signal loss is in principle not a problem if it is understood power spectrum is in principle unbiased, because the posterior and corrected for. Since signal loss is tantamount to employing distributions fully capture the range of possible outcomes a biased estimator for the power spectrum, an accurate within the foreground-signal degeneracy, although care must quantification of the bias enables corrections. Often this be taken to ensure that any summary statistics are not defined in quantification takes the form of simulations, where one is a way that gives the mistaken impression of bias in the essentially computing a transfer function that accounts for the presence of significantly asymmetric posteriors. signal loss due to foreground subtraction (or in principle, any In closing, we note that foreground mitigation is not the only other form of loss). One subtlety with such simulations is that culprit in the problem of signal loss. Calibration errors, for in many foreground subtraction schemes, the cleaned data is example, can very easily result in signal loss. Continuing the not a linear function of the input data. (This is equivalent to the discussion from Section 9.4, imagine a calibration scheme that observation we made above, where the resulting power allows the gains of an instrument to be adjusted frequency ’ spectrum estimator—which involves a squaring of the data— channel by frequency channel. Allowing such freedom in one s fl is no longer a quadratic function of the initial data.) Thus, it is data analysis means that the spectrally uctuating cosmological fi insufficient to simulate only the cosmological signal portion of signal can be perfectly t out and eliminated by misestimated the data through the foreground cleaning process, even though calibration parameters, particularly in the case of direction- ( ) it is ultimately the attenuation of this signal that we are dependent calibration Mouri Sardarabadi & Koopmans 2019 . interested in. Doing so misses the impact of correlations The spectral smoothing of calibration solutions described in between the cosmological signal and any foreground models Section 9.4 can in principle mitigate this problem, but to ensure estimated from the data, which include the cosmological signal that this is indeed the case, one should ideally employ end-to- ’ (Paciga et al. 2013; Switzer et al. 2015; Ali et al. 2018; Cheng end simulations that capture as much of one s analysis pipeline as possible (see Sullivan et al. 2012; Lanman & Pober 2019; et al. 2018). Another complication with quantifying transfer Lanman et al. 2019 for examples of progress toward this). functions through simulations is that at low radio frequencies, Another valuable approach along this theme is to test multiple even state-of-the-art foreground models (and for that matter, analysis pipelines against one another (Jacobs et al. 2016).In state-of-the-art instrument models) are unlikely to be particu- general, it is clear that whatever one’s approach for calibration, larly accurate, making it hard to know what to simulate. One foreground suppression, map-making, and power spectrum approach that has been employed to get around this is to estimation, a careful validation of the chosen analysis approach perform cosmological signal injection on real data, effectively will be crucial for precision 21 cm science. using the real data as a template for the foregrounds. However, interpreting such simulations can be difficult (Cheng et al. 2018), and thus in general it is preferable to employ algorithms 13. Power Spectra to Parameters/Science fi that are less susceptible to signal loss in the rst place. Measuring the power spectrum is not an end in itself but ( ) Alternatively, it is at least in principle possible to sidestep rather a means to an end. The ultimate goal is to use some of these issues via the Bayesian approach for joint measurements to place constraints on (or better yet, rule out) ( foreground and power spectrum estimation Sims & theoretical models. Power spectra should therefore be thought ) fl Pober 2020 that we brie y discussed in Section 11.4. of as a compressed version of the data that can be used to Fundamentally, signal loss can occur during foreground distinguish between different theoretical scenarios. If the mitigation because at least some part of the cosmological fluctuations that are being measured are Gaussian-distributed, signal can be traded for foregrounds in one’s model without then this compression is lossless. If not, then the compression is sacrificing the goodness of fit to the data. This degeneracy is lossy, and the power spectrum contains less information than fully captured in a Bayesian approach, which produces full the original data. Regardless of whether this is the case, the joint probability distributions of all parameters in the model exercise of quantitatively comparing theoretical and observa- (whether they are foreground or signal parameters). When tional power spectra is generally still a fruitful one that enables marginalizing over foreground parameters as nuisance para- the measurement of key parameters that govern physical meters, the error bars on the power spectrum bandpowers will models. increase to reflect degeneracies between foregrounds and The process of extracting parameters from power spectrum signal. This fully captures what we have been calling signal measurements is generally tackled by Bayesian inference. loss, although in this probabilistic framework “signal loss” is a Again, we may invoke Bayes’ theorem (Equation (150)),aswe slight misnomer as nothing has really been “lost”—we have did in Section 11.4 to constrain the power spectrum. However, simply produced a more realistic estimate of the uncertainties in whereas before the data vector d consisted of interferometric

67 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw visibilities and the parameters of interest q were the power then the covariance matrix is no longer diagonal, and the power spectrum bandpowers, here the data d are the power spectrum spectrum alone does not capture everything. bandpowers and the parameters q are those in the physical There are several solutions to this problem. One is to simply models. With this setup, if the errors on the power spectrum are capture the relevant off-diagonal elements. This only needs to assumed to be Gaussian,43 the likelihood takes the form be done along the frequency direction, since it is redshift evolution that is responsible for destroying translation 1 t -1 invariance. Thus one possibility is to compute a quantity such (∣)dddVddqqqµ--exp [M ()] [ -M ()], ( 170 ) {}2 as the cross-angular power spectrum from every pair of frequency channel in one’s data, i.e., Cℓ (nn, ¢) (Santos et al. where dM ()q is the prediction of our theoretical model for the 2005; Datta et al. 2007; Bharadwaj et al. 2019). This quantity power spectrum bandpowers, and V is the covariance matrix of records full correlation information in the frequency direction, fi ( ) the observed bandpowers, as de ned in Equation 137 . and in the angular direction assumes statistical isotropy to Sampling from this distribution using an MCMC algorithm justify the use of angular power spectra. Such a formalism is and multiplying by a prior then produces a sampling of the particularly fruitful for low-redshift 21 cm cosmology, where posterior distribution Pr,(∣q d  )of the parameters given the (at least on linear scales) the evolutionary effects can be data and the model (see Section 11.4). This is a fairly standard relatively straightforwardly modeled by a growth factor that way to constrain model parameters, but in the rest of this independently multiplies each Fourier mode by some scalar section, we highlight several subtleties that arise in 21 cm data factor. This makes it simple to relate the measured quantity, Cℓ analysis that may not be important in other cosmological (ν, ν′), to theoretical parameters, in a way that is not true for analyses. experiments targeting reionization and Cosmic Dawn redshifts. Cℓ 13.1. Light Cone Effects One potential weakness of computing quantities such as (ν, ν′) is that it may be too general of a way to capture the One of the key features of 21 cm surveys is their ability to statistics of non-translation invariant fields. By capturing the probe large cosmological volumes. While this of course results cross-correlation between any possible pair of frequencies, one in statistically powerful scientific constraints, it can result in can encode arbitrary violations of translation invariance. This subtleties in analysis that arise because of cosmological gives up on a powerful constraint on the problem: the fact that evolution efforts. In particular, the survey volumes have the for ν and ν′ that are close to one another, the field should be property that the lowest frequency edge of the volume may be approximately translation invariant. One way to incorporate at a significantly higher redshift than the highest frequency this information is to try to search for an alternative to the edge. In other words, cosmological evolution is typically non- Fourier basis, essentially searching for a basis where the two- negligible in intensity mapping surveys, and one is surveying a point covariance is diagonal. This is a difficult problem to solve past light cone of our universe rather than a three-dimensional because doing so in rigorous generality requires modeling the fi volume at xed time. exact nature of the cosmological evolution. However, in the fi With signi cant evolutionary effects, our surveyed temper- spirit of this approach, wavelet transforms have been shown to fi ature elds are no longer statistically translation invariant. This be a promising possible way forward (Trott 2016). implies that the power spectrum no longer captures all the Of course, one can simply choose to ignore the light cone fi information contained in the elds. To see this, consider effect. For small enough frequency ranges, this is likely a fair ( ) fi Equation 16 , which de nes the power spectrum. If one approximation. In any case, even if light cone effects are imagines a discrete version of the equation where there are a significant, the power spectrum is still a reasonable statistic to finite and discrete set of k values, then the left-hand side is compute—it is simply not one that captures all the information. essentially a covariance matrix between every “pixel” in Additionally, the light cone effect is arguably a higher order Fourier space and every other pixel in Fourier space. The right- effect that is to be tackled only after high signal to noise hand side says that if the field is statistically translation measurements of the power spectrum are available. Indeed, this invariant, then this covariance matrix is diagonal, with the has been the stance taken with recent upper limits that have diagonal elements given by the power spectrum. Thus, with been placed on the 21 cm power spectrum. One simply splits up assumptions of translation invariance (and Gaussianity; see the entire observational bandwidth into smaller sub bands, Section 14.2), the power spectrum contains all the information measuring an evolution-averaged power spectrum in each of about the field. If translation invariance is broken, however, many redshift bins. However, it should be noted that in 43 In general, the errors on the power spectrum are not expected to be principle, light cone effects do have the potential to affect the Gaussian, but Gaussianity may be a reasonable approximation if the errors are interpretation of even upper limits, since they can potentially small. Alternatively, if the Bayesian methods for power spectrum estimation from Section 11.4 are used, one can incorporate information from the non- depress 21 cm power on a some scales during reionization Gaussian power spectrum posterior into the likelihood here. (Datta et al. 2012; La Plante et al. 2014).

68 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

13.2. Computationally Expensive Model Predictions ratio of the H I power spectrum to the matter power spectrum) becomes nonlinear at small scales (Wang et al. 2019), and only A key assumption in the use of MCMCs to constrain model recently have robust theoretical frameworks been developed to parameters is that one has a way to generate model predictions model this behavior (Villaescusa-Navarro et al. 2018). While quickly once the model parameters are provided. In other these frameworks were strongly motivated by detailed hydro- words, MCMCs allow an efficient exploration of parameter dynamical simulations (and therefore agree with them reason- space provided the evaluation of d ()q at any given parameter M ably well), it would be fair to say that revisions and refinements space location q is quick. In many cases, this is not the case. At may be necessary as 21 cm measurements mature. post-reionization redshifts, complicated models are in principle With the underlying theoretical frameworks uncertain, one required to model the total H I content and distribution relative option is to build models that are general and flexible, and to dark matter, although some of these complications can be some semi-analytic codes are headed in this direction (Park omitted if the focus is restricted to linear scales or robust et al. 2019). An alternative is to perform model selection prior signatures like BAOs. At reionization redshifts and above, to model fitting. One way to do this is to simply examine which astrophysical effects on various scales have a complex models are able to fit the data well. However, this tends to interplay with cosmological fields, necessitating either radia- reward models that have an excessive number of parameters, tion-hydrodynamical simulations or semi-analytic treatments some of which may not be physically well-motivated and exist for any theory predictions. These can be slow to run, with even simply to improve the fit. A possible alternative to this would semi-analytic codes potentially taking several hours per run to be to compute the Bayesian evidence. Recall from Section 13 obtain predictions at the wide variety of redshifts needed to that the evidence is given by Pr(∣d ), i.e., the probability of compare to predictions. the observed data d given the theoretical model . This can be One solution to this problem is to forgo the exact evaluation computed by noticing that because Equation (150) must be a d ()q in favor of a surrogate or emulated model, dem (q), that M M properly normalized posterior probability distribution, margin- is quick to evaluate. Such an emulator can be constructed by alizing over all the parameters in the numerator gives us the evaluating the exact model at various carefully chosen grid fi em denominator, and thus points in parameter space, and tting to them to create dM (q) as a higher dimensional interpolation function. This approach Pr(∣dd )= (∣qq , ) Pr ( , )dnq , ( 171 ) has been successfully employed in the past in the CMB and ò ( galaxy survey communities Heitmann et al. 2006, 2009, 2014; where n here is the number of parameters. In model selection Fendt & Wandelt 2007; Habib et al. 2007; Schneider et al. we are ultimately interested in a closely related quantity: the ) 2011; Aslanyan et al. 2015 , and has recently been demon- probability of the theoretical model given the observed data, strated in forecasts of 21 cm measurements at reionization and Pr(∣) d . This is a quantity that we can obtain using Bayes’ ( Cosmic Dawn redshifts Kern et al. 2017; Schmit & theorem, which gives Pritchard 2018). An example of this from Kern et al. (2017) Pr(∣d  ) Pr ( ) is shown in Figure 14 for a hypothetical HERA power Pr(∣) d = ,() 172 spectrum measurement in the range 5<z<25. With Pr()d ( emulators, it is possible to perform MCMC forecasts and where p() is the prior on the model. If we have two eventually data analyses) over parameter spaces that include competing theoretical models, 1 and 2, we may compare not just astrophysical parameters but also cosmological them by computing the ratio of their probabilities. This is given parameters. With the high sensitivities of current and upcoming by experiments, this has been shown to be necessary for accurate ( ) Pr(∣d  ) Pr ( ) constraints Liu & Parsons 2016 . 11.() 173 Pr(∣d 22 ) Pr ( ) 13.3. Uncertainties in the Underlying Theoretical Models We thus see that in selecting between different models, one In the discussion above, we have implicitly assumed that computes the ratio of the evidences, adjusting for one’s priors there is an underlying theoretical model  that is known to be on the plausibility of each model. Which model is ultimately correct, and that only their adjustable parameters are unknown favored depends on whether the ratio is greater or less than and need to be constrained. Of course, this is in general not unity; precisely how far the ratio needs to be from unity for one true. For example, at z>6 the astrophysics of reionization and model to be favored over the other is in some ways a matter of Cosmic Dawn is extremely uncertain, to the extent that it is not taste, although a rough guide that is frequently used is the one just model parameters that are unconstrained, but also the given by Kass & Raftery (1995). underlying frameworks. At z<6 the situation is slightly better, The Bayesian evidence has so far been employed in since the H I gas density is expected to be a biased tracer of forecasting exercises for potential power spectrum measure- matter. However, the H I bias (defined as the square root of the ments, suggesting that upcoming instruments should have the

69 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Figure 14. Forecasted HERA constraints on theoretical model parameters from a hypothetical z=5toz=25 power spectrum measurement. Forecasts (and eventually, fits to the real data) like these that involve both cosmological and astrophysical parameters are made possible by the emulators discussed in Section 13.2. −1 The cosmological parameters here are the root-mean-square amplitude of density fluctuations on 8 h Mpc scales, σ8; the dimensionless Hubble parameter, h; the normalized baryon density, Ωb; the normalized cold dark matter density, Ωc; and the spectral index of the primordial scalar fluctuations, ns. The astrophysical min fi ζ parameters are the minimum virial temperature of ionizing halos, Tvir ; the ionizing ef ciency, ; the mean free path of ionizing photons in ionized regions, Rmfp; the luminosity of X-ray sources, fX; the spectral index of X-ray emission, αX; and the minimum photon energy for escape into the intergalactic medium, νmin. Note that there is considerable freedom as to how one chooses to parameterize reionization; this is simply one possible parameterization, and others are possible (e.g., see Park et al. 2019). For details on the forecasting methodology and detailed instrumental assumptions, please see Kern et al. (2017). (A color version of this figure is available in the online journal.)

70 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw sensitivity to distinguish between different classes of reioniza- as the Spectro-Photometer for the History of the universe, tion models (Binnie & Pritchard 2019). It has also been Epoch of Reionization, and Ices Explorer (SPHEREx; Doré proposed as a way to select between different parameterizations et al. 2018), and possibly by futuristic proposed missions such of the foregrounds (Harker 2015; Sims et al. 2020, in as the Cosmic Dawn Intensity Mapper (Cooray et al. 2019) and preparation), and has been applied to real global signal data the Origins (Battersby et al. 2018; Meixner (described in Section 15) to evaluate different models for the et al. 2018). cosmological signal, foregrounds, and systematics. For a cross-correlation measurement to have high signal to noise, it is necessary but not sufficient for the two participating 14. Power Spectrum Alternatives and Variants surveys to overlap in configuration space (i.e., for the surveys For the bulk of this paper, we have focused on measurements to look at the same part of the sky and the same redshifts);itis of the 21 cm power spectrum. The primary reason for this is also necessary for the surveys to overlap in Fourier space. In that the power spectrum is what most of the current 21 cm other words, the surveys must be sensitive to the same length experiments are attempting to detect and characterize (with the scales. This can be achieved either by having telescopes that exception of a small number of global signal experiments have comparable angular and spectral resolutions, or by described in Section 15). There is no deep, fundamental reason ensuring that the finer data set also be sensitive to large-scale this. The power spectrum is simply a convenient summary fluctuations that the coarser data set is exclusively sensitive to. statistic for describing the spatial fluctuations being measured. This is not simply a matter of hardware design; care must also Moreover, power spectra have a long history in cosmology, and be taken to make sure that one’s data analysis does not destroy have the advantage of being well-defined quantities that are Fourier modes that are important for cross correlation. As an relatively straightforward to compute from theory and/or example of this, note that any foreground subtraction algorithm simulations. that relies on suppressing smooth modes along the line of sight However, the 21 cm power spectrum is ultimately limited as will inevitably produce maps that destroy power at low k∼0 a summary statistic, for both theoretical and practical reasons. modes. Because of this, a cross correlation of such maps with Theoretically, the power spectrum may not contain all the any integral field (i.e., one that is projected along the line of fi ( information contained in the 21 cm eld or the underlying sight, such as any CMB map) will have very low signal to ) quantities that determine the 21 cm brightness temperature . noise. One way around this would be to simply skip foreground Practically, it is prudent to include other summary statistics subtraction, and hope that the foregrounds are uncorrelated when reducing 21 cm data, since these alternatives may have between the two surveys and thus will not contribute. However, complementary systematics. In this section, we highlight a few as we discussed in Section 12.1.7, this is only the case in possible power spectrum alternatives and variants. expectation, and residual foreground variance may remain. It is therefore likely that foreground mitigation will play a role even 14.1. Cross Correlations for cross-correlation studies, depressing the final signal-to- As a first variation on the 21 cm power spectrum, we noise ratio in the final measurement. consider cross power spectra between the 21 cm line and some other tracer of structure in our universe. The cross power spectrum is defined in much the same way as the 21 cm power 14.2. Higher Order Statistics spectrum, i.e., as a two-point covariance function in Fourier To increase the signal to noise on a cross correlation between space as defined in Equation (16), except with one copy of the an integral field and a foreground-suppressed 21 cm map, one 21 cm brightness temperature field and one copy of the other possibility is to go beyond a two-point cross-correlation tracer rather than two copies of the 21 cm field. There are a fi large number of tracers that are suitable candidates for such a between two elds and to instead use higher order n-point cross correlation, especially at z<6. For instance, the first statistics. If the signals in question were Gaussian, higher order detections of the cosmological 21 cm line were made in cross statistics would not yield any new information: n-point correlation with traditional galaxy surveys at z∼0.8. At functions would vanish for odd n by symmetry and reduce to slightly higher redshifts, CO rotational lines and [C II] powers of the two-point functions for even n. Moreover, hyperfine transitions have been proposed for cross-correlations. statistical translation invariance guarantees that different Four- It is possible that CO may be insufficiently abundant at z  6 ier modes are not correlated with one another, giving rise to a due to photodissociations (Lidz et al. 2011), but [C II] intensity two-point function that is non-zero only if the two “points” mapping observations have been planned for up to z∼9 coincide in Fourier space, which we recall from Equation (16) (Stacey et al. 2018). At even higher redshifts, there is the is what we denote the power spectrum. If non-Gaussianities are potential of cross-correlating with Lyα,Hα, [O II], and [O III] present, different Fourier modes couple to one another and intensity maps that will be provided by satellite missions such higher order statistics contain new information. For example,

71 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw the bispectrum B()kk123,, k can be defined as couplings can bias statistical estimators of lensing, which are designed to detect lensing by looking for correlations between 3 D BTTT()()()()()()kk123,, kº++áñ 2pd k123 k k k 1 k 2  k 3 , modes. It is possible, however, to write down bias hardened ()174 estimators that are protected against such contamination (Foreman et al. 2018). where one sees that as long as kk123++= k0 (i.e., the three In addition to the effects described above (which affect all vectors form a closed triangle), different Fourier modes can be redshifts), there are also strong non-Gaussian signatures that correlated with one another. are intrinsic to the 21 cm field during reionization (Iliev et al. One way in which mode correlations can occur is through 2006; Mellema et al. 2006). This arises because (among other nonlinear gravitational evolution, and it is through such a factors) the 21 cm brightness temperature involves the product mechanism that it may be possible to cross-correlate a 21 cm of the neutral fraction and the density. Even if we restrict map with an integral field. Nonlinear evolution results in a ourselves to scales where the density field is Gaussian, the coupling between large-scale tidal fields and small-scale neutral fraction field will necessarily be non-Gaussian given fi density elds. To access the information contained in this that it is restricted to take on values between 0 and 1. Aside coupling, one can either try to directly measure the higher-point from this line of purely mathematical reasoning, there are also statistics or attempt tidal reconstruction. With the latter physical reasons to expect non-Gaussianity. In particular, approach, small-scale measurements can in principle be used galaxies form at the peaks of the density field (i.e., they are ( ) to reconstruct large-scale modes Zhu et al. 2016 , recovering biased tracers of density), and thus the ionization structures cosmological information that was lost in data analysis. Such they produce around them are caused by the high-end tails of tidal reconstruction schemes can in principle be applied to a the mass function (Wyithe & Loeb 2007), resulting in ionized variety of different cross-correlations between 21 cm data and structures that do not depend linearly on overdensity (Wyithe ( other measurements. Examples include the CMB via Inte- & Morales 2007). This leads to strong non-Gaussianities (Lidz grated Sachs–Wolfe or kinetic Sunyaev–Zel’dovich contribu- et al. 2007). Similar effects are present at even higher redshifts tions) or photometric galaxy surveys (which behave in a similar (prior to reionization) due to inhomogeneities in X-ray heating way to integrated fields since large redshift uncertainties mean (Watkinson & Pritchard 2015; Watkinson et al. 2019). that often only angular correlation information is trustworthy) To access the non-Gaussian information contained in the (Li et al. 2019; Zhu et al. 2018). While tidal reconstruction high redshift 21 cm field, one possibility is to simply measure forecasts vary in their levels of optimism regarding the the bispectrum (Watkinson et al. 2017). Recent studies have approach (see, e.g., Goksel Karacayli & Padmanabhan 2019 pointed out that sensitivity to the bispectrum scales rather for a more pessimistic evaluation), its enormous scientific − / − favourably with time (scaling as t 3 2 rather than t 1 for the potential clearly justifies further investigation. power spectrum, as discussed in Section 3.5), and that Another mechanism for inducing mode correlations—and kk k therefore another probe of the underlying physics—is gravita- foregrounds may not be prohibitive for certain ()123,, fi ( ) tional lensing by large scale structure. Statistical characteriza- triangle con gurations Trott et al. 2019b . Theoretical tions of lensing have been used to great effect in the CMB predictions have suggested that the bispectrum may break community, for instance in placing neutrino mass constraints certain parameter degeneracies that arise in analyses utilizing ( ) (Planck Collaboration et al. 2018a) or as one possible way to only the power spectrum Shimabukuro et al. 2017 . The break the various geometric degeneracies in earlier CMB bispectrum also contains distinctive signatures that mark the experiments, such as that between the normalized curvature progression of reionization in the form of sign changes (Majumdar et al. 2018). Such sign changes occur because the parameter ΩK and H0 (Efstathiou & Bond 1999; Howlett et al. 2012). In 21 cm cosmology, similar lensing measurements are 21 cm bispectrum can be written as the sum of all possible in principle possible, although there are some crucial bispectrum combinations between density and ionization (e.g., differences. For one, in CMB lensing there is just a single three-point functions of density or three-point functions source plane at high redshifts, whereas in 21 cm cosmology consisting of two factors of density and one factor of neutral there are a series of source planes. Having multiple source fraction, and so on). Some of these combinations may probe planes can increase the signal to noise of a lensing detection, fields that are correlated to one another, while others probe an but care must be taken to ensure that correlations between the anti-correlated set of fields. This results in sign changes in the source planes are taken into account, since the different source total bispectrum as different combinations dominate the total planes are part of a larger correlated three-dimensional field bispectrum as reionization progresses. Additional interesting (Zahn & Zaldarriaga 2006). An additional complication with signatures can be obtained by forming the bispectrum not from 21 cm lensing is that the source planes themselves probe matter the three-point function of the Fourier modes themselves, but fluctuations that contain mode couplings due to the aforemen- their phases, where each copy of T in Equation (174) is tioned gravitational evolution effects. These intrinsic mode replaced by TT∣∣. Such a bispectrum of phases has been

72 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw shown in theoretical studies to be a probe of the characteristic functions. For instance, beyond the bispectrum there is the length scales of ionized regions during reionization (Gorce & trispectrum (or equivalently, the four-point correlation func- Pritchard 2019). tion). One alternative to this is to simply compute the Although theoretically promising, the experimental feasi- probability distribution function (PDF) of voxels in a 21 cm bility of bispectrum measurements is still an open question. map (Wyithe & Morales 2007; Harker et al. 2009b; Watkinson While current and planned instruments may have the sensitivity & Pritchard 2014; Shimabukuro et al. 2015), which then to potentially detect the bispectrum (Yoshiura et al. 2015), most captures the full extent of non-Gaussianity in one statistic. The studies assume that foregrounds can be removed to extremely weakness of this approach is that it does not capture correlation high precision. Moreover, calibration errors are typically not information between pixels, since the value of each pixel is included in theoretical studies. One way to (in principle) simply binned into a histogram (i.e., a PDF) without regard for sidestep the need for high precision calibration is to measure other pixels. Some correlation information can be recovered by the phase of the bispectrum (in contrast to the bispectrum of forming two-point PDFs that give the joint probability of phases described above). In particular consider the phase of the measuring various brightness temperature values in two pixels. angular bispectrum, where the bispectrum is computed using Closely related to this is the idea of difference PDFs, where one interferometric data from a single frequency channel. In the flat quantifies the probability distribution function of differences sky approximation, each baseline of an interferometer probes a between the values of two voxels separated by a given distance different angular Fourier mode (see Equation (40) and (Barkana & Loeb 2008; Gluscevic & Barkana 2010; Pan & Section 3.2), which means that the bispectrum is simply the Barkana 2012). three-point correlation function of three interferometric visibi- Many of these PDFs have shown promise in capturing lities whose baseline vectors form a closed triangle. The important features of the underlying physics. However, they triangular requirement is imposed by the Dirac delta function of have yet to be shown to be practical in the face of foreground Equation (174), and can be satisfied by using the visibilities contamination. The foreground challenge is greater for PDFs from baselines that connect three antennas. Computing the than for statistics such as the power spectrum. This is because phase of the angular bispectrum is then equivalent to the foregrounds do not enter in an additive manner in the PDFs; meas meas meas meas computing the phase of V12 VV23 31 where Vij is the instead, the PDF of foregrounds is convolved with the PDF of measured visibility from a baseline formed between the ith and the cosmological signal, since the PDF of a sum of two random jth antennas. This has the attractive property of being immune variables is given by the convolution of the constituent PDFs. to per-antenna calibration effects considered in Section 9, since Typically, the foreground brightness temperature will vary with we can insert Equation (71) to see that a larger amplitude from pixel to pixel (up to 100s of K in variation, depending on the frequency) than the cosmological meas meas meas f = arg()VVV12 23 31 signal does (which might have ∼mK-level variations). This ***true true true means that the foreground contribution makes the measured = arg()gg1 2 V12 gg2 3 V23 gg3 1 V31 (i.e., foreground contaminated) PDF very broad, and the = arg(∣∣∣gggVVV2 ∣∣2 ∣2 true true true ) 1 2 3 12 23 31 foregrounds must be known to exquisite precision in order for true true true = arg()VVV12 23 31 , () 175 their PDF to be deconvolved out to leave the narrow cosmological signal PDF. Having such precision in our (L) fi where arg signi es the phase of a complex number, and in knowledge of foregrounds is difficult at the low radio the last equality we used the fact that multiplying a complex frequencies that are relevant. However, recent work has number by a real number does not affect its phase. We see that suggested that with multiple tracers (e.g., galaxy surveys in f ( our bispectrum phase which is also known as the closure conjunction with 21 cm maps), it may be possible sidestep this phase in more traditional radio astronomy contexts; Cornwell requirement in the foreground PDF deconvolution (Breysse ) & Fomalont 1999; Thompson et al. 2017; Carilli et al. 2018 is et al. 2019). independent of per-antenna gain calibration factors, making it To increase signal to noise, one might also consider particularly attractive a quantity to measure with real data. Of measuring the higher order moments (e.g., the skewness and/ course, aside from calibration errors one must also contend or the kurtosis) of PDFs. Condensing non-Gaussianity with foreground contamination. Simulations have suggested signatures into a few moments means that there are fewer though, that subsequently computing the delay spectrum of the numbers to measure, thereby increasing sensitivity. Studies f may potentially allow foregrounds to be separated from the have suggested that even though the skewness and the kurtosis ( ) cosmological signal Thyagarajan et al. 2018 . do not capture the full richness of a PDF, they are able to capture some interesting non-Gaussian effects of reionization 14.3. Probability Distribution Functions (Wyithe & Morales 2007; Kittiwisit 2019). These statistics are In principle, to capture the full effects of non-Gaussianity, potentially measurable even in maps that have been prepro- one must compute an infinite series of higher-order correlation cessed to filter out the foreground wedge region of Fourier

73 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw space described in Section 12.1.5 (Kittiwisit 2019). The Importantly, such a scheme works reasonably well even if one filtering, however, does affect the moments (beyond just assumes that foreground mitigation causes certain Fourier removing the foregrounds), and so a correct interpretation of modes to be irretrievably lost within the wedge-shaped region the results will likely require simulation-based forward models. described in Section 12.1.5. Another way to use images is to use them to identify extremes. For reionization studies, this can be helpful both at 14.4. Imaging the beginning and the end of reionization. Toward the end of Capturing the full richness of spatial fluctuations in the reionization, recent studies have suggested that reasonably 21 cm line—including its cosmological evolution, probability large neutral islands can persist even when our universe is distribution and higher-order correlations—is an image of the almost completely ionized (Becker et al. 2015; Malloy & 21 cm field. However, the information-completeness of images Lidz 2015; Keating et al. 2020; Kulkarni et al. 2019). comes at a cost. First, images will generally have lower signal Conversely, luminous quasars are able to generate large to noise than averaged/binned statistics such as the power ionized bubbles even if our universe is mostly neutral (Wyithe spectrum, necessitating larger and more sensitive arrays. & Loeb 2004). These large features are potentially detectable Second, an image is not a statistical property of the underlying by current-generation arrays, particularly using analysis field; instead, it is a specific realization of the underlying techniques such as matched filtering (Malloy & Lidz 2013). statistics. However, further forecasts need to be made to definitively Having said this, the fact that an image is a realization rather quantify the effect of foregrounds and instrumental systematics than a statistical property can be an advantage. Consider, for on such measurements. example, the possibility of comparing 21 cm images with the Matched filtering is an example of an analysis scheme where locations of high-redshift galaxies during reionization. The a specific feature of the 21 cm maps is searched for, with the 21 cm images can in principle provide context for the detected feature selected by the analyst ahead of time because of its high-redshift galaxies. A galaxy’s location within an ionized potential to constrain the underlying physics. An alternative to bubble, for instance, might indicate whether the galaxy resides this is to use machine learning, where the features are not in a newly ionized region of our universe, or one where selected by hand, but instead are the result of an optimization reionization happened long ago. In practice, there is unfortu- process (“learning”) that searches for features that are best able nately a mismatch in angular scales: by design, intensity to constrain the ultimate quantities of interest. Suppose that one mapping surveys such as 21 cm surveys do not seek to resolve is interested in extracting a parameter vector q from an image individual galaxies, unlike with high-redshift galaxy observa- whose voxel values are stored in a data vector d. With a tions, which in turn tend to have fairly narrow fields of view. machine learning approach, we seek a function f that provides As an example of this severe mismatch of scales, we note that good estimates qˆ of our parameter vector, i.e., we seek a the entire Great Observatories Origins Deep Survey South field function where (GOODS South field; covered by the , the , and the Chandra X-ray Observa- qq»ºˆ f ()wd, , () 176 tory) is approximately the same size as the resolution of a with w being a set of adjustable parameters. Starting with a single pixel in maps produced by HERA! Future widefield general form for f, the adjustable parameters are determined by surveys covering sky areas comparable to what is envisioned requiring that q » qˆ over a series of training examples of d for the Wide Field Infrared Survey Telescope would be better 44 matched for the scales probed by 21 cm instruments. However, where the true parameters are known. Precisely what it means ˆ fi the importance of line of sight fluctuations in 21 cm maps for q » q is quanti ed by a cost function. If q were to be a set ( means that one ideally requires spectroscopic galaxy redshifts of parameters governing some theoretical model like we saw ) to avoid the washing out of radial fluctuations that results from in Section 13, for example , then one possible choice might be photometric redshift estimates. Additionally, the low signal-to- the mean squared error between the recovered parameters and noise of a high redshift 21 cm map would likely necessitate a the true parameters. The training process is then tantamount to fi statistical cross correlation rather than an analysis of the maps nding w such that themselves. 2 å∣()∣q - f wd,i () 177 Perhaps more promising for the direct analysis of 21 cm i maps is to use them to provide probabilistic models of the context in which high-redshift galaxies reside. In particular, is minimized, where the index i runs over the different training fi although resolution effects make it difficult to definitively data sets. Once the best w vector is found, it is xed and f can determine whether a high-redshift galaxy resides in an ionized 44 In this paper, we will only discuss supervised learning algorithms where the bubble or not, it is possible to use 21 cm maps to predict the correct answers are known in the training data, as opposed to unsupervised probability that a pixel is ionized or not (Beardsley et al. 2015). learning algorithms, where this is not the case.

74 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw be applied to either test data sets (for validation) or to real data 15.1. Science Applications of the Global 21 cm Signal to provide parameter estimates. 15.1.1. The Low Redshift (z  6) Global Signal In machine learning, one typically selects forms for f that are extremely flexible. A particularly popular choice is to have f At different redshifts, the global signal enters analyses in take on the form of a Convolutional Neural Network (CNN; different ways. For low-z intensity mapping, the global signal is see, e.g., Goodfellow et al. 2016). In a CNN, the input data often regarded as a nuisance factor. It essentially sets the (i.e., the input image) is passed through a series of convolu- overall normalization of the power spectrum. Of course, this tions, downsamplings, and linear transformations. After each of normalization is not known a priori, which can cause these steps the data is additionally passed through a nonlinear degeneracies that prohibit constraints on parameters of interest. function that is predetermined by the data analyst. The w vector For example, suppose one is interested in using redshift-space encapsulates the details of the various stages of the CNN. For distortions to constrain the growth of structure. In such studies, instance, some of its elements might capture the shapes of the one often attempts to constrain the quantity f σ8, where σ8 is the convolution kernels that the data is put through. With a CNN, root mean square amplitude of matter fluctuations smoothed on − one can in principle mimic any reasonably well-behaved 8h 1 Mpc scales, fdDdaº ln ln is the dimensionless higher-dimensional function (Cybenko 1989). It is for this growth rate, D is the growth function of perturbations in linear reason that they are powerful, although care must be taken to perturbation theory, and a is the scale factor. With redshift ensure that they are not too powerful, in the sense that one must space distortions, one measures (to linear order in perturbation) avoid overfitting. When a CNN is overfit to the training data, it PTbfPk()k =+2 (ssm22 ) () s2 , ( 178 ) is essentially memorizing the training data set and its b H8I 8 m 8,fid corresponding parameter values. Fortunately, there are well- where we have omitted all noise terms for simplicity, m º kk , established techniques such as dropout (Srivastava et al. 2014) 2 fi Tb is the global signal, bH I is the H I bias, s8,fid is a ducial to prevent these problems. value for σ8. One sees that there is a perfect degeneracy In several proof of concept studies, CNNs have shown some σ 2 between f 8 and Tb , since they enter as a multiplicative promise as ways to analyze 21 cm data. CNNs have been combination and can therefore be perfectly traded off for one shown in simulations to be able to take in 21 cm images and another. Fortunately, recent theoretical work has demonstrated successfully recover theoretical parameters (such as the that by going to the mildly nonlinear regime, the next-order fi fi ionizing ef ciency of the rst galaxies during reionization; terms go as higher powers of fσ , while T 2 remains an overall ) ( 8 b Gillet et al. 2019 or phenomenological parameters such as the normalization. This allows the degeneracy to be broken, and duration of reionization; La Plante & Ntampaka 2019) directly forecasts suggest that competitive constraints for fσ8 can be from 21 cm images. CNNs can also potentially perform model obtained (Castorina & White 2019). Another possibility is to selection, determining whether a 21 cm image favors a use multi-tracer techniques, where information from futuristic galaxies-driven reionization scenario or an active galactic galaxy surveys can be combined with 21 cm surveys to break nucleus-driven reionization scenario even when power spec- the degeneracy (Fialkov et al. 2020). trum measurements are unable to distinguish between the two (Hassan et al. 2019). Despite its immense promise, the 15.1.2. The High Redshift (z  6) Global Signal application of CNNs to 21 cm cosmology is still in its infancy, and further studies (particularly regarding the presence of At high-z, the global signal is seen as a particularly foregrounds and systematics in one’s images) are necessary interesting signal in its own right. As Figure 15 illustrates, before CNNs can be robustly applied to real data. the global signal is an incisive probe of the various phenomena during Cosmic Dawn. During the reionization epoch, the evolution of the global signal is essentially a measure of the 15. Global Signal Measurements ionization history, with the signal going slowly to zero with the For most of this paper, we have focused on the mapping of neutral fraction of our universe. At higher redshifts, the global α the spatial fluctuations of the cosmological signal. However, signal is driven by X-ray heating and Ly coupling, and serves there exists a complementary signal—the global 21 cm signal as a probe of energy injection into the IGM. —which is also of interest. The global signal refers to the mean 15.2. Measuring the Global 21 cm Signal in Practice brightness temperature of the 21 cm line as a function of redshift, averaged over all angular directions of the sky. In the Because of differences in their target signals, experiments language of spherical harmonics, the global signal is the pursuing the spatial fluctuations of the 21 cm line tend to be (ℓ, m)=(0, 0) monopole mode, which contains independent quite different from those attempting to measure the global information. An example theory prediction for what the global signal. The latter, by definition, does not need to resolve spatial 21 cm signal might look like can be seen in the middle panel of features on the sky. Thus, global signal experiments are most Figure 1. commonly conducted using single antennas that have wide

75 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Cosmic Dawn might have a spectral resolution of Δν∼50 kHz, whereas a global signal experiment designed for the same redshifts might have Δν∼1 MHz. With the ability to integrate over larger bandwidths, noise levels quickly beat down in less than a day to levels much lower than the expected amplitude of the cosmological signal. Of course, allowances must be made for non-ideal observing conditions, but even so, far less time is needed to reach acceptably low noise levels for global signal experiments than for power spectrum experiments. Despite the obvious appeal of their conceptual simplicity, global 21 cm experiments are still challenging to perform because the systematics are formidable. Just as with the spatial fluctuation experiments, global signal experiments need to contend with foregrounds (Shaver et al. 1999; Bernardi et al. 2015). Most experiments pursue a similar strategy to the fluctuation experiments, in that they base their foreground separation strategy on the spectral smoothness of the fore- grounds. This may be more difficult for global experiments, because the signal itself can be quite smooth. For example, consider an experiment targeting the reionization epoch, when the amplitude of the global signal is expected to transition from non-zero to zero, tracing the evolution of the IGM from neutral to ionized. If reionization is a relatively extended process, then the expected global signal is one that smoothly and monotonically decreases as one goes from low frequencies (high redshifts) to high frequencies (low redshifts). This is Figure 15. Example global 21 cm signal curves. The top panel shows the effect essentially the same behavior as one sees for astrophysical vir fi of changing Tmin , the minimum virial temperature of ionizing halos. The foregrounds, making foreground mitigation extremely dif cult. bottom panel shows the effect of changing the proportionality constant fX It is for this reason that global 21 cm signal experiments between X-ray luminosity and star formation rate. These example curves were observing the reionization epoch may be confined to ruling out ( ) generated using the Accelerated Reionization Era Simulations ARES code very rapid reionization. This is also why many recent efforts (Mirocha et al. 2015, 2017). have instead targeted the absorption feature during Cosmic (A color version of this figure is available in the online journal.) Dawn, taking advantage of its non-monotonicity to separate it from the foregrounds. The crux of the global signal foreground problem is that beams, although creative approaches using interferometry have global signal measurements are, in a sense, too simple. By been proposed and attempted (Mahesh et al. 2014; Presley et al. reducing the entire measurement to a single spectrum, there are 2015; Singh et al. 2015; Vedantham et al. 2015; Venumadhav far fewer degrees of freedom in the measurement, limiting the ) et al. 2016; McKinley et al. 2018 . For the single-element number of ways in which the cosmological signal can differ experiments, their wide beams integrate instantaneously over a from the foregrounds. Because of this, there have been substantial fraction of the sky, which provides an excellent proposals for using more than just spectral information in the approximation to the monopole, since the typical correlation measurement of the monopole. For instance, if one surveys the lengths of the anisotropies are on much smaller angular scales sky with reasonably fine angular resolution, the parts of the sky than the beam size. These experiments do not have the extreme with the brightest foregrounds can be selectively avoided, sensitivity requirements of those seeking to measure spatial rather than having the entire sky of foregrounds averaged into fluctuations. This is because global signal experiments seek our global spectrum. This is equivalent to saying that because only to capture the relatively gradual cosmic evolution (and we seek to measure the monopole of the cosmological signal, therefore frequency evolution) of the signal, and not to resolve anything appearing in the higher multipoles must be a the fine spectral features caused by the spatial fluctuations foreground, providing us with another mechanism for fore- along the line of sight. Global signal experiments can therefore ground rejection beyond the spectral information. Note, employ relatively coarse frequency channels. As a comparison, however, that the foregrounds will still have a monopole a typical experiment targeting spatial fluctuations during component, and thus spectral methods are still required. By

76 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw avoiding the brightest parts of the sky, one is simply reducing (Vedantham et al. 2014; Datta et al. 2016). Current experiments the amplitude of the foregrounds from being the average over deal with this by careful data excision and calibration (often the whole sky to being its minimum value. To go beyond this using similar techniques to those described in Sections 12.3 and and to take full advantage of the spatial information to actively 12.4), but there have also been proposed experiments such as subtract foregrounds requires a detailed knowledge of its the Dark Ages Radio Explorer (DARE; Burns et al. 2012, correlations and heteroscedasticity between different parts of 2017) and the Dark Ages Polarimetry PathfindER (DAPPER; the sky; if this information is available, foregrounds can in Tauscher et al. 2018a) that propose going to the far side of the principle be distinguished from the signal even if their spectral Moon to avoid these problems entirely. signatures are extremely similar (Liu et al. 2013). Finally, global signal experiments require exquisite control An alternate way to take advantage of spatial foreground of systematics (Rogers & Bowman 2012), such as those due to information is to use polarization information. Consider a linear antenna beam chromaticity (Vedantham et al. 2014; Bernardi polarization-sensitive telescope that is pointed at a celestial et al. 2015; Mozdzen et al. 2016; Singh et al. 2018a), cable pole. Anisotropic foregrounds appear in the two linear reflections (Monsalve et al. 2017) and environmental condi- polarizations at different strengths due to projection effects. tions (Bradley et al. 2019). In many ways, the calibration As the Earth rotates, this induces polarization signals that are requirements for global signal experiments are more extreme modulated in time (Nhan et al. 2017). This provides a way to than those for interferometric experiments, since single-element distinguish anisotropic foregrounds from the isotropic fore- experiments possess a noise bias (as discussed in Section 3.1) grounds and the cosmological signal without requiring large that must be well-characterized and subtracted. To achieve this telescopes that resolve the characteristic scales of the level of calibration, experimental groups tend to rely on a anisotropies. Initial progress has been made in implementing combination of laboratory measurements (Rogers & Bowman ( such an approach using the Cosmic Twilight Polarimeter Nhan 2012; Patra et al. 2013; Monsalve et al. 2016, 2017), in situ ) et al. 2019 . measurements (Patra et al. 2017), and principal component- Yet another proposal for distinguishing foregrounds from the based modeling from the actual data (Tauscher et al. 2018b). ( global 21 cm signal is to use velocity information Slosar 2017; Typically, the instrumental effects are calibrated out of the ) ’ Deshpande 2018 . In particular, the solar system s motion measurement by introducing nuisance calibration parameters. relative to the cosmic rest frame induces a dipole pattern in However, given the limited number of degrees of freedom in a what would otherwise be a pure monopole signal. It can be global signal experiment, one must take care to design helpful to consider the dipole as being the sum of two terms. experiments in a way that avoids the need for a large The first is a boosting of the signal intensity, analogous to the proliferation of calibration parameters (Switzer & Liu 2014). CMB dipole. The second is a frequency dependent effect. Consider motion toward the CMB, which results in the blueshifting of photons received by one’s telescope. If the 15.3. The Current Status of Global Signal Experiments global signal were independent of frequency, this would have no effect, since for a given frequency channel there would be Most global signal experiments have so far focused on just as many photons blueshifting into the channel from lower Cosmic Dawn and reionization, given the richer phenomenol- frequencies as there are photons blueshifting out of the channel ogy accessible in those epochs compared to the post- into higher frequencies. The size of the effect is therefore reionization era. proportional to the slope of the global signal curve, and because Early global experiments provided measurements of the there are some reasonably large gradients in typical predictions spectral index of the diffuse radio emission at low frequencies ( α ( ) for the global signal (see Figure 15), it is this second frequency- i.e., the syn parameter of Equation 70 ; Rogers & Bowman dependent effect that dominates the dipole signature. In 2008; Patra et al. 2015; Mozdzen et al. 2017, 2019). More absolute terms the dipole 21 cm signal is still small compared recently, the timing of reionization has been constrained by the to the monopole, but it is—crucially—different from the dipole Shaped Antenna measured of the background RAdio Spectrum caused by the modulation of the foreground monopole, since (SARAS; Singh et al. 2018a) experiment and the Experiment to our velocity vector relative to the Milky Way is not aligned Detect the Global Epoch of reionization Signature (EDGES; with our velocity vector relative to the cosmic rest frame. This Bowman et al. 2008). SARAS has ruled out rapid reionization can potentially provide a clean signature for further separating in the interval 6<z<10 where the rate of change dTdzb of foregrounds from the cosmological signal, although other the global signal Tb with redshift is less than 114 mK (Singh techniques would still have to be employed in order to isolate et al. 2017, 2018b). Similar results have come out of EDGES, extragalactic foregrounds. but in terms of a phenomenological model where reionization is Aside from astrophysical foregrounds, global signal experi- parameterized in terms of a midpoint redshift zmid,a ments must also contend with RFI and the ionosphere characteristic duration Δz, and the globally averaged neutral

77 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

( ) fraction xH I given by instrumental systematics Bradley et al. 2019 and subtleties in the analysis methods (Bowman et al. 2018b; Hills et al. ⎡ ⎛ ⎞ ⎤ 1 ⎜ zz- mid ⎟ 2018; Jana et al. 2019; Singh & Subrahmanyan 2019). xzH I ()= ⎢tanh+ 1⎥ . ( 179 ) 2 ⎣ ⎝ Dz ⎠ ⎦ Fortunately, other experiments such as the Probing Radio Intensity at high-z from Marion experiment (PRIZM; Philip Early EDGES data ruled out the possibility that reionization et al. 2019), the Large-aperture Experiment to Detect the Dark was more abrupt than Δz∼0.06 (Bowman & Rogers 2010b). Age (LEDA; Price et al. 2018), and the Radio Experiment for Additional data sharpened that constraint, ruling out Δz  1if the Analysis of Cosmic Hydrogen (REACH; de Lera the midpoint of reionization occurred at zmid∼8.5, and Δz  Acedo 2019) are collecting data and may soon confirm or 0.4 if it occurred anywhere in the range 6.5<zmid<14.8. refute the EDGES claims. With the large impact of a first Combining some of this data with EDGES observations detection of a 21 cm signal from Cosmic Dawn, these centered on Cosmic Dawn redshifts as well as constraints verifications will clearly be well worth the effort. from the CMB and high redshift quasars has favored a 4.5 minimum virial temperature Tvir in the range 10 Me to 5.7 16. Conclusion 10 Me (Monsalve et al. 2018), a CMB optical depth τe of less than 0.063, and a mean free path Rmfp of ionizing photons Now is an exciting time for 21 cm cosmology. The field is that is greater than 27.5 Mpc (Monsalve et al. 2019). rapidly becoming one of experimental reality rather than Of considerable recent interest are constraints on Cosmic theoretical promise. A large number of experiments have been Dawn. SARAS has ruled out a variety of models with an built, many of which have sufficient sensitivity for a detection extremely cold IGM that would result in large absorption and characterization of the cosmological 21 cm signal. troughs (Singh et al. 2017, 2018b). However, EDGES has Although sensitivity is not necessarily the limiting factor in recently claimed a detection of just such a signature at 78 MHz these observations, having greater signal to noise also allows (Bowman et al. 2018a). If confirmed, the purported EDGES for a more incisive diagnosis and mitigation of systematics. In signal would have dramatic implications on our understanding the last decade, 21 cm cosmology has gone from a data-starved of Cosmic Dawn. For example, the timing and narrowness of field to one where there is plenty of data with which to test a the EDGES signal suggests a more rapidly evolving star plethora of analysis ideas. formation rate than expected (Mirocha & Furlanetto 2019). The primary challenges of 21 cm cosmology remain the Additionally, the amplitude of the absorption is unexpectedly extreme sensitivity (Section 3), foreground mitigation large, signifying a large temperature contrast between the (Sections 7 and 12), and systematic control requirements background radiation temperature and the neutral hydrogen (Sections 6 and 9). Importantly, these challenges do not exist in spin temperature (see Equation (2), which shows that large isolation, and much of the problem comes from interactions 21 cm absorption signals occur when Ts is much cooler than between these effects. For example, in Section 12.1.5 we Tγ). Achieving such a large contrast would require at least one discussed the way in which the already-formidable foreground of the following revisions to our theoretical models. One problem is made worse by the way in which the observed possibility is the existence of a population of previously foregrounds seen through an interferometer are less spectrally undetected high-redshift radio sources. This would increase the smooth than the true foreground emission on the sky. This temperature contrast by boosting the background radiation results in a larger number of Fourier modes (“the foreground temperature, essentially replacing Tγ in Equation (5) with some wedge”) that are likely to be irretrievably contaminated by higher temperature (Ewall-Wice et al. 2018, 2019; Feng & foregrounds than one would predict from intrinsic foreground Holder 2018; Sharma 2018; Fialkov & Barkana 2019; Jana properties. Thus, one sees that the data analysis techniques et al. 2019). Alternatively, there may be exotic physics at play discussed in this paper for calibration (Section 9.1), map- that allows the spin temperature to cool below the temperature making (Section 10), power spectrum estimation (Section 11), expected for a gas that cools adiabatically with our universe’s foreground mitigation (Section 12), and parameter estimation expansion (Barkana 2018; Berlin et al. 2018; Clark et al. 2018; (Section 14) are all inextricably linked to one another as well as Costa et al. 2018; Falkowski & Petraki 2018; Hektor et al. one’s hardware. This is true not only for experiments targeting 2018; Hirano & Bromm 2018; Kovetz et al. 2018, 2019b; spatial fluctuations in the 21 cm line (which were emphasized Mitridate & Podo 2018; Moroi et al. 2018; Muñoz & for most of this paper) but also the global signal measurements Loeb 2018; Muñoz et al. 2018; Safarzadeh et al. 2018; discussed in Section 15. In general, 21 cm experiments truly are Schneider 2018; Slatyer & Wu 2018; Yoshiura et al. 2018; software telescopes, where careful hardware-aware analysis Cheung et al. 2019; Chianese et al. 2019; Jia 2019; Lawson & pipelines are required for success. Zhitnitsky 2019). It would be fair to say that data analysis in 21 cm cosmology The EDGES claim is not without controversy. Discussions remains an unsolved problem, in that there is still no consensus and debates are ongoing regarding various possible in the community regarding the optimal way to go from raw

78 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Figure 16. For a tracking telescope the part of the uv-plane that we measure as the Earth rotates follows an ellipse, probing different Fourier modes of the sky (left panel or Figure 3). Through this we fill in the gaps in the uv-plane, building up a larger set of measured modes. In contrast, a drift-scan telescope measures the same set of Fourier modes (right panel), but its observing strategy means that it can directly distinguish Fourier modes in the u direction within the footprint of the primary beam. (A color version of this figure is available in the online journal.) radio telescope data to the science that is in principle enabled the Berkeley Center for Cosmological Physics and the Berkeley by measurements of the redshifted 21 cm line. In many ways, Radio Astronomy Lab for their hospitality when this work was however, this is beneficial, given that multiple highly sensitive started. A.L. acknowledges support from National Science experiments now exist as testbeds for the vast landscape of Foundation National Science Foundation grant No. 1836019, analysis ideas in the literature. One may thus reasonably look the New Frontiers in Research Fund Exploration grant forward to accelerating progress in the field, with confirmation program, a Natural Sciences and Engineering Research Council or refutation of the tentative detections of the global 21 cm of Canada (NSERC) Discovery Grant and a Discovery Launch signal, a continued downward trend in upper limits on Supplement, as well as the Canadian Institute for Advanced fluctuations in the 21 cm signal at all redshifts, and eventually Research (CIFAR) Azrieli Global Scholars program. a set of robust detections and high-significance measurements. These will unlock the great potential of 21 cm cosmology, Appendix A enabling a new generation of cosmological measurements that Drift Scan Observations explore our universe in both space and time to unprecedented precision and accuracy. Many upcoming 21 cm experiments are designed to perform drift scan observations, where they simply point to a fixed The authors are delighted to acknowledge helpful discus- location relative to the ground and use the Earth’s rotation to sions and/or help with figures from Marcelo Alvarez, Joëlle- build up sky coverage. This is in contrast to more traditional Marie Begin Miolan, Gianni Bernardi, Judd Bowman, Chris tracking observations where the telescope has steerable Carilli, Phil Bull, Tzu-Ching Chang, Eloy de Lera Acedo, antennas which are pointed to a fixed location on the sky Avinash Deshpande, Josh Dillon, Gary Hinshaw, Danny during the observation. In this section we analyze the behavior Jacobs, Nick Kern, Piyanat Kittiwisit, Matt Kolopanis, Henri of two telescopes differing only in whether they track, or drift Lamarre, Zac Martinot, Andrei Mesinger, Jordan Mirocha, scan to show how these two different modes of observation are Steven Murray, Michael Pagano, Nima Razavi-Ghods, Peter conceptually quite distinct. A visual summary of this section Sims, Harish Vedantham, and Jeff Zheng. We also gratefully can be found in Figure 16. We will also connect the m-mode thank Jeff Mangum for his editorial patience, and the formalism outlined in Section 10.5 with the more traditional anonymous referee for an astoundingly speedy yet thorough uv-plane formalism for radio interferometry. report. A.L. would additionally like to thank Rose Bagot, To make this comparison we will use the flat sky Khanh Huy Bui, Daryl Haggard, Corinne Maurice, and Kieran approximation; however, rather than modeling the sky as a O’Donnell for writing group support. R.S. would like to thank plane around some fixed point, we will approximate it as a

79 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw cylinder around some fixed decl. In this limit, the sky is flat in where q = ()qqxy, and umn =D()mn2,p . Substituting the sense there is no curvature we need to worry about, but it is these into Equation (180) we find that periodic in R.A. This approximation is reasonable provided the VBT()bubu,fp=D 2˜() - ; , f˜() () 184 primary beam of our telescope is sensitive to a limited range of å mn mn mn declinations such that we measure the sky on a thin strip. We will also assume that the primary beam of the telescope is and thus to calculate the visibilities we simply need to be able limited in R.A., such that we can use small angle to evaluate the Fourier transform of B()q;,b f ˆ approximations. We define a group fixed coordinate system with be (points fi δ ˆ In the equations below we will de ne decl. as the central east) and bn (points north) tangent to Earth’s surface at the decl. of the strip we are observing, which has a width Δ. θx and telescope location. In terms of the usual cartesian unit vectors θy are the cartesian coordinates within the observed strip in the these are east–west and north–south directions, with the origin of θx bxyˆe =-sinffˆˆ + cos() 185 coinciding with the origin of the R.A. axis, and the origin of θy being at the center of the strip, i.e., decl. δ . The telescope is decl. bxyzˆn =-sindflat cosˆˆ - sin df lat sin + cos d lat ˆ () 186 located at a latitude of δlat. We write the local sidereal time (as an angle) as f, such that at f=0 the meridian is θx=0. The and the unit vector pointing to a general position on the sky ( θ θ ) drift-scan telescope is permanently pointed at the meridian, paramterized by x and y is such that it is centered at R.A. α=f. The tracking telescope is rxˆ()qqxy,coscos=+ ( ddecl. q y ) q xˆ pointed at a fixed R.A. which we choose to be the origin of our ++cos()dqqdecl.yx sinyzˆ ++ sin() dq decl. yˆ . cylindrical coordinate system θx=0 (to make this a general point we just need to apply an offet to f). ()187 We will use the vector u as the Fourier conjugate to the To investigate the behavior of the telescope we need to angular position in our strip. To avoid confusion we will write construct the product b · rˆ in these coordinates. We can do that the baseline vector in the ground frame as b, which (as is by breaking the baseline vector b into east–west be and north– convention) will be expressed a physical length (and not in south b components as b =+bbbbˆˆ. Having done that we wavelengths). n ee nn can calculate the two components To proceed let us write the standard equation for an unpolarized visibility as an integral over our cylindrical surface brˆeyx· ˆ =+-cos()()()dqqfdecl. sin , 188

--2pfibrr()·[ˆˆ0] l brˆ · ˆ =-sinddqqf cos()() + cos - VAeTd()br,;ff=Wò p (ˆ )( rˆ) nyxlat decl. 2p D 2 ++cosddlat sin() decl. qy . () 189 » ddAqqqqfxyp¢ ()xy,; òò0 -D 2 For both drift-scan and tracking telescopes we have made the

--2,pfibr()·[ˆ() qqxy rˆ0] l approximation that the field of view is limited such that we do ´ eT()qqxy, . () 180 not have contributions to the integral in Equation (180) far from fi For convenience we have de ned Ap¢ ()qqfxy,; º the direction we are pointing and thus we can make small angle cos()(dqdecl. + ypA rˆ ; f), which incorporates the decl. depen- approximations of the above terms. In both cases we can treat θ dent term from the measure dΩ. We have also included a y as a small angle, i.e., we see only a narrow strip in decl. In θ −f phase-centring term that references the complex phase to center the drift scan case x is a small angle as we do not see far θ of the field we are pointing at which we call rˆ0. For the drift- from the meridian, whereas for the tracking case x itself is a scan telescope this will be at the center of the observed strip, on small angle. Using this we can expand out expressions for the fi the meridian (i.e., θx=f, θy=0), for the tracking telescope it phase to rst order. will be at the origin (θx=θy=0). Let us proceed with the drift-scan telescope case. Here the Let us define the transfer function expansion of the phase factor is fairly simple, and we find that

--2,pfibr()·(ˆ() qqxy rˆ0) l br·[ˆˆ-= r0]()bexcosdqf decl. sin - BAe()()qqxy, ;b , f= p¢ qqfxy , ; , () 181 +-bnycos()ddqdecl. lat sin , () 190 and write B and T in terms of their Fourier series: and we also note that for a drift-scan telescope the behavior ˜ 2piumn·q TTe()q = å ()umn ()182 under Earth rotation of the primary beam is simple for mn Ap¢ ()(qqfxy,;= Ap¢ qxy- fq ,). As both the baseline phase component and the beam simply translate in the azimuthal

˜ 2piumn·q direction as the Earth rotates we can take advantage of the BBe()q;,bubff= å ()mn ;, () 183 mn Fourier shift theorem when evaluating the Fourier transform of

80 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw the transfer function location (dlat and ddecl.). However, we only convolve over the north–south direction (the n-mode), meaning that the observed BeB˜()ub;,ff==-imf ˜()() ub ;, 0. 191 mn mn mʼs map directly to the mʼs on the sky. Using this we only need to evaluate the transform at f=0 The result for the tracking telescope is more complex. The which can be done as phase factor after expansion to first order in small parameters is

1 2p D 2 BA()ubmn;,f == 0 òò p¢ ()q 2pD 0 -D 2 br·[ˆˆ- r0] --+2pqlli[·brˆ () br ·ˆ u ·] q ´ edd0 mn qqxy ()192 =-()bbencosfd cosdecl. sin d lat cos d decl. sin f sin q x +-( bbensindfdecl. sin + sin ddf lat sin decl. cos 1 2p D 2 » òò Ap¢ ()qqxy, + bnycosddlat cos decl.) sin q . 2pD 0 -D 2 -++-+D2pdlpqddlqib[( cos m 2 ) ( b cos ( ) n ) ] ()198 ´ eddexnydecl. decl. lat qqxy ()193 Similar to the drift-scan telescope let us define an effective ˜ »+Apmn(())()uubdrift,dd decl. , lat 194 vector

⎛ ⎞ bbencosfd cosdecl.- sin d lat cos d decl. sin f ubtracking(),,fd decl. , d lat = ⎜ ⎟ ()199 ⎝-+bbensindfdecl. sin sin ddf lat sin decl. cos + b n cos dd lat cos decl.⎠ where to get to the second line we needed to expand small such that angles in sin qq» and in the final line we implicitly introduced the Fourier transform of the beam function br·[ˆˆ-= r0]( u tracking b,fd , decl. , d lat )·()q 200 A˜pmn()u and have defined fi ⎛ b cos d ⎞ at rst order in small angles. Note that unlike the drift-scan ub(),,dd= ⎜ e decl. ⎟.() 195 ( f) drift decl. lat ⎝b cos()dd- ⎠ case, this vector is a function of time i.e., , and also that n decl. lat ˆ ˆ udrift(bub,,dd decl. lat)(== tracking , f 0,, dd decl. lat).

Remembering that A˜p is a discrete quantity, the final line is VA()bubuu,,,,.ffdd=-˜pmnmn((tracking decl. lat ))() T only exact when 2pdu cos decl. and Dv cos(dddecl.- lat) are å mn integers.45 ()201 To account for the time direction, let us take the Fourier f transform of the visibilities V (b, f) in the direction. Using This does not have any simple reduction across the time ( ) ( ) ( ) fi Equations 184 , 191 , and 194 we come to our nal result domain in the same way that the drift-scan telescope does, as for the drift-scan telescope the shift vector is a function of time. We can interpret the observations as sampling part of the uv-plane where the 1 -imf VVedm ()bb= (,ff ) ( 196 ) position moves as an ellipse through the plane as function of 2p ò the Earth’s rotation. For more details (including the full three- dimensional treatment) see Thompson et al. (2017). =-AT˜ ((ub,dd , ) u )() u . () 197 å pmnmndrift decl. lat One important conclusion from this is that for a drift-scan n telescope, the coverage in the uv-plane must be instantaneously This is the flat sky version of the m-mode formalism outlined in sufficient, observing for extended periods of time never Section 10.5. For the drift-scan telescope we can easily see that changes the overall area of Fourier modes that we have access. the modes probed by the telescope are fixed, i.e., the primary However observing the changing of the visibilites in time gives beam footprint A˜p is shifted by an amount which depends only us access to modes within the footprint of the primary beam, on the baseline separation (b) and the telescope and field and if we observe for a full sidereal day down to each individual azimuthal m-mode. This is stark contrast to the 45 This limitation can be removed in the azimuthal direction by performing the transform without making the small angle approximation. This yields a behavior of a tracking telescope where Earth rotation synthesis convolution with a Bessel function Jum (2cospddecl.), which for large m large can dramatically improve the overall footprint of modes we can approximates the delta function shift generated by the small angle approx- measure, but we have no access to higher resolution Fourier imation. Unfortunately this does not work in the polar direction as it is not strictly periodic. modes than the primary beam.

81 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Appendix B the power spectrum P()k via Equation (16). Inserting this into Explicit Quadratic Estimator Examples our integral above and defining the integral containing * 0 In this appendix, we construct some explicit, concrete áTTother ()kkabother ()ñ to be Cij then gives the right-hand side of ( ) examples of the powerful and general—but abstract—quadratic Equation 204 . estimator formalism presented in Section 11. Many of our Having obtained the covariance matrix, our next step is to a results will be approximate, for the pedagogical sake of differentiate with respect to the bandpower p . This requires analytical tractability. discretizing our power spectrum. The most straightforward way to do this is to take our power spectrum to be piecewise B.1. Measuring P(k) from a Three-dimensional constant, such that within the αth k-space voxel, the value of Configuration-space Map the power spectrum is pa. Under this assumption, we have dk3 ~~ Suppose our input data consists of a three-dimensional map CC=+0 pa yy*()kk (), ( 206 ) ij ij å ò 3 i j in configuration space, which is something that one might a Vka ()2p obtain from one of the map-making algorithms described in where the bandpower has come out of the integral because of Section 10. To be precise, let the ith component of our data our piecewise constant assumption, and V signifies the region vector x be given by ka of Fourier space covered by the αth k-space voxel. The 3 3 dk ~*  derivative is now straightforward, yielding xdrTii==y ()rr () yi () kkT () (202 ) òò3 3 ()2p ¶Cij dk ~~ Q a º = yy*()kk (). ( 207 ) ij a ò 3 i j where T ()r is the (continuous) temperature field and yi ()r is ¶p Vka ()2p the footprint of the ith voxel of our map. In the second equality, We now have all the pieces that are required to construct the we made use of Parseval’s theorem to rewrite the integral in optimal quadratic estimator of the power spectrum. Recall from Fourier space. As a concrete example, we might imagine a set Equation (129) that an optimal estimator for the αth bandpower of uniform cuboid voxels, where is given by pˆaaµ xC† --11 Q C x. Inserting our expression for ⎧ Qa into this and simplifying the result gives ⎨1, ifrrr is withinD 2 of i yi ()r = (203 ) ⎩0, otherwise, ⎡ 3 ⎤ a --1 dk ~~* 1 pˆ µ å()Cxi⎢ò yyi () kjj ()( k⎥ Cx ) ⎣ Vk 3 ⎦ with Dr ºDDD( xyz,,) denoting the extent of the voxels. ij a ()2p Recall from Section 11 that the key to establishing a link 3 2 dk ~ -1 between the data and the power spectrum that we wish to = ò å yii()(kCx ). ( 208 ) † Vka 3 estimate is the covariance matrix Cxxºá ñand its response ()2p i with respect to changes to a generic bin (say, the αth bin) of the To gain some intuition for this, consider the voxelization of our power spectrum, Qaaº¶C ¶p . We can obtain the covariance data that we defined in Equation (203). Taking the Fourier matrix by inserting Equation (202) into the definition of the transform to get y ()k gives covariance. The result is that the (i, j)th element of the i ⎛ ⎞ covariance is given by ~ ⎛ kxD ⎞ kyy D -ikr· i ⎜⎟x yi ()k =DDD (xyze ) sinc sinc⎜ ⎟ 3 3 ⎝ 2 ⎠ ⎝ 2 ⎠ dkab dk ~~* * Cij =áñò yyi ()kkajb ()TT () k a () k b ⎛ ⎞ pp3 3 kzz D ()()22 ´ sinc⎜⎟»DDD()xyze-ikr· i, () 209 3 ⎝ 2 ⎠ 0 dk ~~* =+Cijò yy i ()kkkj ()P (), ( 204 ) ()2p 3 where the last approximation can be made if the resolution of fi where in the second equality we decomposed T into a sum of the survey is very ne compared to the lengthscales of interest, kr· D  1 contributions from the cosmological signal T and other such that and the sinc terms are approximately sig unity. With this approximation, we have contaminants or noise, Tother. This allowed us to write 3 2 áTT()kkab* ()ñ as dk pˆa µ e--ikr· i ()Cx1 . () 210 ò 3 å i **    * Vka ()2p i áñ+áñ+áñTTsig ()kkabsig () TTsig () k aother () k b Tother () kk ab Tsig () In words, this says that to optimally estimate the power +áTTother ()kkab* () ñ, ( 205 ) other spectrum one should first apply an inverse covariance 3 D * ( which reduces to (2pd)(kkkaba-+áñ )()PTTother ()() k aother k b weighting to the data this is what makes the optimal quadratic because the cosmological signal is on average uncorrelated estimator an optimal estimator, as shown in Section 11). with the “other” signals and the signal–signal term is related to Following that, notice that the sum in our expression is

82 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw precisely a discrete Fourier transform of the data. Once the where k is (kkij+ ) 2, and we have assumed that the survey Fourier transform is taken, the results are squared before they volume is large, so that f()r has a large footprint while f is are simply summed (integrated) over the Fourier space volume a sharply peaked. With this assumption, Qij is approximately

Vka of the Fourier space voxel of interest. The normalization of zero unless kkij» . If we assume that Vka is relatively small this power spectrum estimate can be computed using the (i.e., we have discretized our Fourier space reasonably finely) expressions given in Section 11. One sees that the optimal and denote the location of the Fourier space volume by ka,we quadratic estimator formalism essentially yields the same thing a come to the conclusion that kkka »»ijfor Qij to be nonzero. as our guess for how to compute the power spectrum in In other words, we have Q a » dd , which means our ( ) ij ij ia Equation 19 , with two differences: the data is inverse quadratic estimator is covariance weighted in order to achieve the best possible error aa† --11 - 12 bars in the final result, and the quadratic estimator formalism pˆ µ=xC Q C x∣( C x )a ∣. ( 212 ) allows one to generalize to more complicated situations where In words, this says that if our data is already in Fourier space, the pixelization might be complicated. power spectrum estimation is straightforward: simply inverse covariance weight the data, and square the component B.2. Measuring P(k⊥, k) or P(k) corresponding to the wavevector that we are interested in. As discussed in Section 3.2, it is often appropriate to report a While this may seem an obvious result, note that it was the binned version of P()k . For diagnosing systematic effects, it is result of assuming a large survey volume. In practice, this helpful to bin our Fourier space into modes parallel and approximation will never be exact. Fortunately, the full matrix fi perpendicular to the line of sight, given the differences in how formulation of the quadratic estimator formalism allows nite parallel and perpendicular information is obtained in 21 cm survey volumes to be treated properly. cosmology. For a final cosmological result, all directions B.4. Measuring Cℓ(ν) from an Angular Sky Map should be statistically equivalent (ignoring redshift-space distortions), so binning k down into a scalar k is appropriate. The quadratic estimator formalism can be used to estimate Within the quadratic estimator formalism, obtaining any quadratic statistic of the data. Here, we demonstrate that by Pk( ^, k) or P(k) is trivial. Take P(k) as an example. Forming using it to estimate the angular power spectrum Cℓ that was such a power spectrum requires binning together all k with the defined in Equation (26). same magnitude k (i.e., those that fall on a particular spherical For convenience, suppose that our data comes in the form of ) fi ν shell in Fourier space . If we simply de ne Vka to be the Fourier a sky map at a particular frequency , expressed in angular space region falling on the αth shell, all our expressions from coordinates. Our pixelization is then given by the previous section follow. The same is true for Pk( , k) with ^  i* appropriately adapted definitions of the bandpower bins. xdii=Wò y (nnˆ)( Tˆ)() =å waℓm ℓm, 213 ℓm B.3. Measuring P(k) from a Fourier Space Map where in the second equality we expressed the sky in terms of its spherical harmonic expansion aℓm ºWòdYℓm* (nnˆ)( Tˆ),where As we remarked in Section 11, the quadratic estimator Y (nˆ) is the spherical harmonic function with quantum numbers formalism allows the input data to be expressed in any basis. ℓm ℓ and m. We similarly defined wi ºWdY* (nnˆ)(y ˆ) and To illustrate this, suppose our data is in Fourier space. This ℓm ò ℓm i assumed that yi (nˆ) is real. An appropriate choice for yi might be can be accommodated in our formalism by setting yi ()r = f()rkrexp (-i · ), where f()r is 1 inside our survey and 0 i ⎧1, ifnnˆˆ is withinDW of outside. With this choice of y , Equation (202) implies that our y (nˆ) = ⎨ i ()214 i i ⎩0, otherwise, data vector x is in Fourier space. Taking the Fourier transform gives y =-f()kk. In turn, ii where DW is the angular area of a single pixel. Following an this gives analogous set of steps as we did for our rectilinear power spectrum, we obtain 3 a dk ~~* 0 i j Qij = ò yyi ()kkj () CCij =+ij wwCℓm* ℓm ℓ ()215 Vka 3 å ()2p ℓm dk3 =--ff*()()kk kk ò 3 ij and Vka ()2p ⎧  2 ¶C ⎨∣(f kkij- )∣ if k is in V ka ℓ ij i j µ ()211 Qij º = å wwℓm* ℓm.() 216 ~ ⎩ 0otherwise, ¶Cℓ m

83 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Note the unfortunate near-clash of notations, where the symbol by writing the ith frequency channel of the visibility as C is used both for the covariance matrix and the angular power ⎛ ⎞ spectrum. The former will always be written in boldface or with 2 ⎜⎟2pn VddAiip»-ò nqgn()()()qq,,exp n T n ib · q two indices; the latter contains just a single index ℓ. We also ⎝ c ⎠ note that Equation (216) represents a very specific choice for ⎡ ⎤ 3 HEz021ngn()i ( ) -ic2pnb·q the binning of our bandpowers, with each spherical harmonic ℓ = dr⎢ Aep ()q,,n ⎥ T ()r ò ⎣ 2 2 ⎦ mode being its own bin. While we made this choice here for the czDz()()1 + c sake of pedagogy, there are often very good reasons to use ()218 thicker bins that encompass multiple ℓ modes. For example, fi incomplete sky coverage will cause different ℓ modes to be where gi ()n describes the pro le of the ith frequency channel. coupled to one another, making it reasonable to average them In the first equality we made the narrow-field, flat-sky together (Alonso et al. 2019). Alternatively, such couplings approximation, and in the second equality we rewrote our can be computed using the window function and error visibility in terms of comoving cosmological coordinates, with covariance formulae presented in Section 11, which can then the tacit understanding that q, ν, and z are all implicit functions be incorporated into one’s comparisons between data and of r. Storing the visibilities from our different frequency theory. channels in one data vector x, we have xViiº , and comparing Continuing with constructing our power spectrum estimator, this expression to Equation (202), we see that the measurement we have equation of an interferometer can be viewed as a generalized pixelization of the sky. The term within the square brackets is then simply a complicated yi ()r function. Taking this and 2 inserting it into all the subsequent expressions like ˆ † --11ℓ i - 1 Cwℓ µ=xC QC xååℓm () C xi Equations (206) and (207) yields a quadratic estimator of the mi power spectrum that operates directly on the visibilities. If 2 the frequency dependence of everything but g ()n is neglected, the =Wå òdYℓm* (nnˆ)( dˆ)(), 217 i m result is the delay-spectrum estimator discussed in Section 11.2. Including the frequency dependence in the complex exponential where we have eschewed the symbol pˆ in favor of the more term incorporates the phenomenology of the foreground wedge a ( ) explicit Cˆ and have defined d (nnCxˆ)(º y ˆ)(-1 ) . Given the Section 12.1.5 into the estimator. This is worked out in detail in ℓ åi i i ( ) form of Equation (214), this is essentially an inverse Liu et al. 2014a . covariance-weighted map of the sky that is pixelized but remapped into a continuous function (or as close to continuous ORCID iDs as possible given the discrete nature of real data). After this, our Adrian Liu https://orcid.org/0000-0001-6876-0928 expression for Cˆℓ instructs us to take the spherical harmonic transform, square, and sum over m. With a proper normal- References ization, this treatment agrees with Equation (27) up to the Ali, Z. S., Parsons, A. R., Zheng, H., et al. 2015, ApJ, 809, 61 inverse covariance weighting. Ali, Z. S., Parsons, A. R., Zheng, H., et al. 2018, ApJ, 863, 201 Also of interest in 21 cm cosmology (and intensity mapping Alonso, D., & Ferreira, P. G. 2015, PhRvD, 92, 063525 in general) is the cross power spectrum between two Alonso, D., Sanchez, J., Slosar, A. & LSST Dark Energy Science Collaboration fi ( ) 2019, MNRAS, 484, 4127 frequencies, Cℓ (nn, ¢),asde ned in Equation 28 . Formally, Anderson, C. J., Luciw, N. J., Li, Y. C., et al. 2018a, MNRAS, 476, 3382 one can go through the same process as we did above, but with Anderson, M. M., & Hallinan, G. 2017, in Radio Exploration of Planetary ( ) all the data from different frequencies stacked into one long Habitability AASTCS5 , 49, 401.02 Anderson, M. M., Hallinan, G., Eastwood, M. W., et al. 2018b, ApJ, 864, 22 data vector. Working through the algebra then gives a Arora, B. S., Morgan, J., Ord, S. M., et al. 2015, PASA, 32, e029 Aslanyan, G., Easther, R., & Price, L. 2015, JCAP, 9, 5 similar expression for Cˆℓ, except with a cross multiplication fi Bandura, K., Addison, G. E., Amiri, M., et al. 2014, Proc. SPIE, 9145, 914522 between spherical harmonic coef cients from different fre- Bandura, K., Bender, A. N., Cliche, J. F., et al. 2016a, JAI, 5, 1641005 quencies rather than the squaring of coefficients from a single Bandura, K., Cliche, J. F., Dobbs, M. A., et al. 2016b, JAI, 5, 1641004 frequency. Barkana, R. 2018, Natur, 555, 71 Barkana, R., & Loeb, A. 2008, MNRAS, 384, 1069 Barnes, D. G., Staveley-Smith, L., de Blok, W. J. G., et al. 2001, MNRAS, 322, 486 B.5. Measuring a Power Spectrum from Visibilities Barry, N. 2018, PhD thesis, Univ. Washington Barry, N., Hazelton, B., Sullivan, I., Morales, M. F., & Pober, J. C. 2016, As a final example in this appendix, we consider how one MNRAS, 461, 3135 Barry, N., Wilensky, M., Trott, C. M., et al. 2019, ApJ, 884, 1 might estimate a power spectrum directly from the visibility Battersby, C., Armus, L., Bergin, E., et al. 2018, NatAs, 2, 596 measured by a single baseline of an interferometer. We begin Battye, R., Browne, I., Chen, T., et al. 2016, arXiv:1610.06826

84 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Battye, R. A., Browne, I. W. A., Dickinson, C., et al. 2013, MNRAS, Ciardi, B., Inoue, S., Mack, K., Xu, Y., & Bernardi, G. 2015b, Advancing 434, 1239 Astrophysics with the Square Kilometre Array (AASKA14),6 Battye, R. A., Davies, R. D., & Weller, J. 2004, MNRAS, 355, 1339 Ciardi, B., Labropoulos, P., Maselli, A., et al. 2013, MNRAS, 428, 1755 Beane, A., & Lidz, A. 2018, ApJ, 867, 26 Clark, S. J., Dutta, B., Gao, Y., Ma, Y.-Z., & Strigari, L. E. 2018, PhRvD, 98, Beardsley, A. P., Hazelton, B. J., Sullivan, I. S., et al. 2016, ApJ, 833, 102 043006 Beardsley, A. P., Johnston-Hollitt, M., Trott, C. M., et al. 2019, PASP, Cooray, A., Chang, T.-C., Unwin, S., et al. 2019, arXiv:1903.03144 36, e050 Cornwell, T., & Fomalont, E. B. 1999, in ASP Conf. Ser. 180, Synthesis Beardsley, A. P., Morales, M. F., Lidz, A., Malloy, M., & Sutter, P. M. 2015, Imaging in Radio Astronomy II, ed. G. B. Taylor, C. L. Carilli, & ApJ, 800, 128 R. A. Perley (San Francisco, CA: ASP), 187 Beardsley, A. P., Thyagarajan, N., Bowman, J. D., & Morales, M. F. 2017, Cornwell, T. J. 2008, ISTSP, 2, 793 MNRAS, 470, 4720 Cornwell, T. J., & Perley, R. A. 1992, A&A, 261, 353 Becker, G. D., Bolton, J. S., & Lidz, A. 2015, PASA, 32, e045 Costa, A. A., Landim, R. C. G., Wang, B., & Abdalla, E. 2018, EPJC, 78, 746 Berger, P., Oppermann, N., Pen, U.-L., & Shaw, J. R. 2017, MNRAS, Crighton, N. H. M., Murphy, M. T., Prochaska, J. X., et al. 2015, MNRAS, 472, 4928 452, 217 Berlin, A., Hooper, D., Krnjaic, G., & McDermott, S. D. 2018, PhRvL, 121, Cybenko, G. 1989, Math. Control Signal Systems, 2, 303 011102 Dalal, N., Pen, U.-L., & Seljak, U. 2010, JCAP, 2010, 007 Bernardi, G., McQuinn, M., & Greenhill, L. J. 2015, ApJ, 799, 90 Datta, A., Bowman, J. D., & Carilli, C. L. 2010, ApJ, 724, 526 Bernardi, G., Mitchell, D. A., Ord, S. M., et al. 2011, MNRAS, 413, 411 Datta, A., Bradley, R., Burns, J. O., et al. 2016, ApJ, 831, 6 Bharadwaj, S., Pal, S., Choudhuri, S., & Dutta, P. 2019, MNRAS, 483, 5694 Datta, K. K., Choudhury, T. R., & Bharadwaj, S. 2007, MNRAS, 378, 119 Binnie, T., & Pritchard, J. R. 2019, MNRAS, 487, 1160 Datta, K. K., Mellema, G., Mao, Y., et al. 2012, MNRAS, 424, 1877 Book, L., Kamionkowski, M., & Schmidt, F. 2012, PhRvL, 108, 211301 Davis, G. 2012, PhD thesis, Univ. British Columbia, https://open.library. Bowman, J., & Rogers, A. E. E. 2010a, RFI Mitigation Workshop, 30 ubc.ca/collections/ubctheses/24/items/1.0072797 Bowman, J. D., Cairns, I., Kaplan, D. L., et al. 2013, PASA, 30, e031 de Lera Acedo, E. 2019, Int. Conf. on Electromagnetic for Advanced Bowman, J. D., Morales, M. F., & Hewitt, J. N. 2009, ApJ, 695, 183 Applications 2019, ICEAA 2019 Bowman, J. D., & Rogers, A. E. E. 2010b, Natur, 468, 796 de Oliveira-Costa, A., Tegmark, M., Gaensler, B. M., et al. 2008, MNRAS, Bowman, J. D., Rogers, A. E. E., & Hewitt, J. N. 2008, ApJ, 676, 1 388, 247 Bowman, J. D., Rogers, A. E. E., Monsalve, R. A., Mozdzen, T. J., & DeBoer, D. R., Parsons, A. R., Aguirre, J. E., et al. 2017, PASP, 129, 045001 Mahesh, N. 2018a, Natur, 555, 67 Denman, N., Amiri, M., Bandura, K., et al. 2015, 2015 IEEE 26th Int. Conf. on Bowman, J. D., Rogers, A. E. E., Monsalve, R. A., Mozdzen, T. J., & Application-specific Systems, Architectures and Processors (ASAP) Mahesh, N. 2018b, Natur, 564, E35 Deshpande, A. A. 2018, ApJL, 866, L7 Bradley, R. F., Tauscher, K., Rapetti, D., & Burns, J. O. 2019, ApJ, 874, 153 Di Matteo, T., Perna, R., Abel, T., & Rees, M. J. 2002, ApJ, 564, 576 Breysse, P. C., Anderson, C. J., & Berger, P. 2019, PhRvL, 123, 231105 Di Valentino, E., Melchiorri, A., & Mena, O. 2017, PhRvD, 96, 043503 Breysse, P. C., Kovetz, E. D., & Kamionkowski, M. 2015, MNRAS, 452, 3408 Di Valentino, E., Melchiorri, A., & Silk, J. 2016, PhLB, 761, 242 Bull, P., Ferreira, P. G., Patel, P., & Santos, M. G. 2015, ApJ, 803, 21 Dillon, J. S., Kohn, S. A., Parsons, A. R., et al. 2018, MNRAS, 477, 5670 Burns, J. O., Bradley, R., Tauscher, K., et al. 2017, ApJ, 844, 33 Dillon, J. S., Liu, A., Williams, C. L., et al. 2014, PhRvD, 89, 023002 Burns, J. O., Lazio, J., Bale, S., et al. 2012, AdSpR, 49, 433 Dillon, J. S., Neben, A. R., Hewitt, J. N., et al. 2015a, PhRvD, 91, 123011 Byrne, R., Morales, M. F., Hazelton, B., et al. 2019, ApJ, 875, 70 Dillon, J. S., & Parsons, A. R. 2016, ApJ, 826, 181 Callister, T. A., Anderson, M. M., Hallinan, G., et al. 2019, ApJL, 877, L39 Dillon, J. S., Tegmark, M., Liu, A., et al. 2015b, PhRvD, 91, 023002 Carilli, C. L., Gnedin, N. Y., & Owen, F. 2002, ApJ, 577, 22 Dinh-v-Trung 2009, MNRAS, 396, 2319 Carilli, C. L., Nikolic, B., Thyagarayan, N., & Gale-Sides, K. 2018, RaSc, Dodelson, S., & Schneider, M. D. 2013, PhRvD, 88, 063537 53, 845 Doré, O., Werner, M. W., Ashby, M. L. N., et al. 2018, arXiv:1805.05489 Carroll, P. A., Line, J., Morales, M. F., et al. 2016, MNRAS, 461, 4151 Dumitru, S., Kulkarni, G., Lagache, G., & Haehnelt, M. G. 2019, MNRAS, Castorina, E., & White, M. 2019, JCAP, 6, 025 485, 3486 Challinor, A., & Lewis, A. 2011, PhRvD, 84, 043516 Eastwood, M. W., Anderson, M. M., Monroe, R. M., et al. 2018, AJ, 156, 32 Chang, T. C., Gong, Y., Santos, M., et al. 2015, Advancing Astrophysics with Eastwood, M. W., Anderson, M. M., Monroe, R. M., et al. 2019, AJ, 158, 84 the Square Kilometre Array (AASKA14),4 Efstathiou, G., & Bond, J. R. 1999, MNRAS, 304, 75 Chang, T.-C., Pen, U.-L., Bandura, K., & Peterson, J. B. 2010, Natur, 466, 463 Eisenstein, D. J., Hu, W., & Tegmark, M. 1998, ApJL, 504, L57 Chang, T.-C., Pen, U.-L., Peterson, J. B., & McDonald, P. 2008, PhRvL, 100, Ekers, R. D., & Bell, J. F. 2002, in IAU Symp. 199, The Universe at Low 091303 Radio Frequencies, ed. A. Pramesh Rao, G. Swarup, & Gopal-Krishna, 498 Chapman, E., Abdalla, F. B., Bobin, J., et al. 2013, MNRAS, 429, 165 Ekers, R. D., & Rots, A. H. 1979, in IAU Coll. 49: Image Formation from Chapman, E., Abdalla, F. B., Harker, G., et al. 2012, MNRAS, 423, 2518 Coherence Functions in Astronomy, ed. C. van Schooneveld (Berlin: Chapman, E., Zaroubi, S., Abdalla, F. B., et al. 2016, MNRAS, 458, 2928 Springer), 61 Chen, X., Meerburg, P. D., & Münchmeyer, M. 2016, JCAP, 9, 023 Evans, N. J. I., Hills, R. E., Rydbeck, O. E., & Kollberg, E. 1972, PhRvA, Cheng, C., Parsons, A. R., Kolopanis, M., et al. 2018, ApJ, 868, 26 6, 1643 Cheng, Y.-T., Chang, T.-C., Bock, J., Bradford, C. M., & Cooray, A. 2016, Ewall-Wice, A., Bradley, R., Deboer, D., et al. 2016a, ApJ, 831, 196 ApJ, 832, 165 Ewall-Wice, A., Chang, T. C., Lazio, J., et al. 2018, ApJ, 868, 63 Cheung, K., Kuo, J.-L., Ng, K.-W., & Tsai, Y.-L. S. 2019, PhLB, 789, 137 Ewall-Wice, A., Chang, T.-C., & Lazio, T. J. W. 2019, MNRAS, 492, 6086 Chianese, M., Di Bari, P., Farrag, K., & Samanta, R. 2019, PhLB, 790, 64 Ewall-Wice, A., Dillon, J. S., Hewitt, J. N., et al. 2016b, MNRAS, 460, 4320 CHIME/FRB Collaboration, Amiri, M., Bandura, K., et al. 2018, ApJ, 863, 48 Ewall-Wice, A., Dillon, J. S., Liu, A., & Hewitt, J. 2017, MNRAS, 470, CHIME/FRB Collaboration, Amiri, M., Bandura, K., et al. 2019a, Natur, 1849 566, 235 Ewall-Wice, A., Dillon, J. S., Mesinger, A., & Hewitt, J. 2014, MNRAS, CHIME/FRB Collaboration, Amiri, M., Bandura, K., et al. 2019b, Natur, 441, 2476 566, 230 Ewall-Wice, A., Hewitt, J., Mesinger, A., et al. 2016c, MNRAS, 458, 2710 Chluba, J. 2018, arXiv:1806.02915 Falkowski, A., & Petraki, K. 2018, arXiv:1803.10096 Cho, J., Lazarian, A., & Timbie, P. T. 2012, ApJ, 749, 164 Fender, R., Stewart, A., Macquart, J. P., et al. 2015, Advancing Astrophysics Choudhuri, S., Bharadwaj, S., Ghosh, A., & Ali, S. S. 2014, MNRAS, with the Square Kilometre Array (AASKA14),51 445, 4351 Fendt, W., & Wandelt, B. 2007, ApJ, 654, 2 Christian, P., & Loeb, A. 2013, JCAP, 9, 014 Feng, C., & Holder, G. 2018, ApJL, 858, L17 Cianciara, A. J., Anderson, C. J., Chen, X., et al. 2017, JAI, 6, 1750003 Fialkov, A., & Barkana, R. 2019, MNRAS, 486, 1763 Ciardi, B., Inoue, S., Abdalla, F. B., et al. 2015a, MNRAS, 453, 101 Fialkov, A., Barkana, R., & Jarvis, M. 2020, MNRAS, 491, 3108

85 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Fialkov, A., Barkana, R., Tseliakhovich, D., & Hirata, C. M. 2012, MNRAS, Jones, M. G., Haynes, M. P., Giovanelli, R., & Moorman, C. 2018, MNRAS, 424, 1335 477, 2 Fialkov, A., Barkana, R., Visbal, E., Tseliakhovich, D., & Hirata, C. M. 2013, Jordan, C. H., Murray, S., Trott, C. M., et al. 2017, MNRAS, 471, 3974 MNRAS, 432, 2909 Joseph, R. C., Trott, C. M., & Wayth, R. B. 2018, AJ, 156, 285 Fialkov, A., Cohen, A., Barkana, R., & Silk, J. 2017, MNRAS, 464, 3498 Kapahi, V. K., & Ananthakrishnan, S. 1995, BASI, 23, 265 Foreman, S., Meerburg, P. D., van Engelen, A., & Meyers, J. 2018, JCAP, Kass, R. E., & Raftery, A. E. 1995, J. Am. Stat. Assoc., 90, 773 2018, 046 Kazemi, S., Yatawatta, S., Zaroubi, S., et al. 2011, MNRAS, 414, 1656 Furlanetto, S., Bowman, J. D., Mirocha, J., et al. 2019, arXiv:1903.06212 Keating, L. C., Weinberger, L. H., Kulkarni, G., et al. 2020, MNRAS, Furlanetto, S. R., & Loeb, A. 2002, ApJ, 579, 1 491, 1736 Furlanetto, S. R., & Oh, S. P. 2016, MNRAS, 457, 1813 Keeley, R. E., Joudaki, S., Kaplinghat, M., & Kirkby, D. 2019, JCAP, 12, 035 Furlanetto, S. R., Oh, S. P., & Briggs, F. H. 2006, PhR, 433, 181 Kent, J., Dowell, J., Beardsley, A., et al. 2019, MNRAS, 486, 5052 Gehlot, B. K., Koopmans, L. V. E., de Bruyn, A. G., et al. 2018, MNRAS, Kern, N. S., Liu, A., Parsons, A. R., Mesinger, A., & Greig, B. 2017, ApJ, 478, 1484 848, 23 Gehlot, B. K., Mertens, F. G., Koopmans, L. V. E., et al. 2019, MNRAS, Kern, N. S., Parsons, A. R., Dillon, J. S., et al. 2019, ApJ, 884, 105 488, 4271 Kern, N. S., Parsons, A. R., Dillon, J. S., et al. 2020, ApJ, 888, 70 Gillet, N., Mesinger, A., Greig, B., Liu, A., & Ucci, G. 2019, MNRAS, Kerrigan, J., La Plante, P., Kohn, S., et al. 2019, MNRAS, 488, 2605 484, 282 Kerrigan, J. R., Pober, J. C., Ali, Z. S., et al. 2018, ApJ, 864, 131 Gluscevic, V., & Barkana, R. 2010, MNRAS, 408, 2373 Kintner, P. M., & Seyler, C. E. 1985, SSRv, 41, 91 Goksel Karacayli, N., & Padmanabhan, N. 2019, MNRAS, 486, 3864 Kittiwisit, P. 2019, PhD thesis, Arizona State Univ., https://search.proquest. Gong, Y., Cooray, A., Silva, M., et al. 2012, ApJ, 745, 49 com/docview/2235423537 Gong, Y., Silva, M., Cooray, A., & Santos, M. G. 2014, ApJ, 785, 72 Knox, L., & Millea, M. 2020, PhRvD, 101, 043533 Goodfellow, I., Bengio, Y., & Courville, A. 2016, Deep Learning (Cambridge, Kocz, J., Greenhill, L. J., Barsdell, B. R., et al. 2014, JAI, 3, 1450002 MA: MIT Press) Kogut, A., Dunkley, J., Bennett, C. L., et al. 2007, ApJ, 665, 355 Gorce, A., & Pritchard, J. R. 2019, MNRAS, 489, 1321 Kogut, A., Dwek, E., & Moseley, S. H. 2015, ApJ, 806, 234 Górski, K. M., Hivon, E., Banday, A. J., et al. 2005, ApJ, 622, 759 Kohn, S. A., Aguirre, J. E., La Plante, P., et al. 2019, ApJ, 882, 58 Greig, B., & Mesinger, A. 2017, MNRAS, 472, 2651 Kohn, S. A., Aguirre, J. E., Nunhokee, C. D., et al. 2016, ApJ, 823, 88 Grobler, T. L., Bernardi, G., Kenyon, J. S., Parsons, A. R., & Smirnov, O. M. Koopmans, L., Pritchard, J., Mellema, G., et al. 2015, Advancing Astrophysics 2018, MNRAS, 476, 2410 with the Square Kilometre Array (AASKA14),1 Habib, S., Heitmann, K., Higdon, D., Nakhleh, C., & Williams, B. 2007, Kovetz, E., Breysse, P. C., Lidz, A., et al. 2019a, BAAS, 51, 101 PhRvD, 76, 83503 Kovetz, E. D., Cholis, I., & Kaplan, D. E. 2019b, PhRvD, 99, 123511 Hamaker, J. P., Bregman, J. D., & Sault, R. J. 1996, A&AS, 117, 137 Kovetz, E. D., Poulin, V., Gluscevic, V., et al. 2018, PhRvD, 98, 103529 Hardcastle, M. J., Gürkan, G., van Weeren, R. J., et al. 2016, MNRAS, Kovetz, E. D., Viero, M. P., Lidz, A., et al. 2017, arXiv:1709.09066 462, 1910 Kramer, M., & Stappers, B. 2015, Advancing Astrophysics with the Square Harker, G., Zaroubi, S., Bernardi, G., et al. 2009a, MNRAS, 397, 1138 Kilometre Array (AASKA14),36 Harker, G. J. A. 2015, MNRAS, 449, L21 Kulkarni, G., Keating, L. C., Haehnelt, M. G., et al. 2019, MNRAS, 485, L24 Harker, G. J. A., Zaroubi, S., Thomas, R. M., et al. 2009b, MNRAS, 393, La Plante, P., Battaglia, N., Natarajan, A., et al. 2014, ApJ, 789, 31 1449 La Plante, P., & Ntampaka, M. 2019, ApJ, 810, 110 Haslam, C. G. T., Klein, U., Salter, C. J., et al. 1981, A&A, 100, 209 Lanman, A., Hazelton, B., Jacobs, D., et al. 2019, JOSS, 4, 1234 Haslam, C. G. T., Salter, C. J., Stoffel, H., & Wilson, W. E. 1982, A&AS, 47, 1 Lanman, A. E., & Pober, J. C. 2019, MNRAS, 487, 5840 Hassan, S., Liu, A., Kohn, S., & La Plante, P. 2019, MNRAS, 483, 2524 Lawson, K., & Zhitnitsky, A. R. 2019, PDU, 24, 100295 Hazelton, B. J., Morales, M. F., & Sullivan, I. S. 2013, ApJ, 770, 156 Lazio, J., Carilli, C., Hewitt, J., Furlanetto, S., & Burns, J. 2009, Proc. SPIE, Heitmann, K., Higdon, D., Nakhleh, C., & Habib, S. 2006, ApJL, 646, L1 7436, 74360I Heitmann, K., Higdon, D., White, M., et al. 2009, ApJ, 705, 156 Lewis, A., & Challinor, A. 2007, PhRvD, 76, 083005 Heitmann, K., Lawrence, E., Kwan, J., Habib, S., & Higdon, D. 2014, ApJ, Li, D., Zhu, H.-M., & Pen, U.-L. 2019, PhRvD, 100, 023517 780, 111 Li, W., Pober, J. C., Hazelton, B. J., et al. 2018, ApJ, 863, 170 Hektor, A., Hütsi, G., Marzola, L., et al. 2018, PhRvD, 98, 023503 Li, Y.-C., & Ma, Y.-Z. 2017, PhRvD, 96, 063525 Hickish, J., Abdurashidova, Z., Ali, Z., et al. 2016, JAI, 5, 1641001 Lidz, A., Furlanetto, S. R., Oh, S. P., et al. 2011, ApJ, 741, 70 Hills, R., Kulkarni, G., Meerburg, P. D., & Puchwein, E. 2018, Natur, 564, E32 Lidz, A., & Taylor, J. 2016, ApJ, 825, 143 Hincks, A. D., Shaw, J. R. & Chime Collaboration 2015, in ASP Conf. Ser. Lidz, A., Zahn, O., McQuinn, M., et al. 2007, ApJ, 659, 865 495, Astronomical Data Analysis Software an Systems XXIV (ADASS Lidz, A., Zahn, O., McQuinn, M., Zaldarriaga, M., & Hernquist, L. 2008, ApJ, XXIV), ed. A. R. Taylor & E. Rosolowsky (San Francisco, CA: ASP), 523 680, 962 Hirano, S., & Bromm, V. 2018, MNRAS, 480, L85 Line, J. L. B., McKinley, B., Rasti, J., et al. 2018, PASA, 35, 45 Hoare, M., Perez, L., Bourke, T. L., et al. 2015, Advancing Astrophysics with Line, J. L. B., Webster, R. L., Pindor, B., Mitchell, D. A., & Trott, C. M. 2017, the Square Kilometre Array (AASKA14), 115 PASA, 34, e003 Högbom, J. A. 1974, A&AS, 15, 417 Liu, A., & Parsons, A. R. 2016, MNRAS, 457, 1864 Hogg, D. W. 1999, arXiv:astro-ph/9905116 Liu, A., Parsons, A. R., & Trott, C. M. 2014a, PhRvD, 90, 023018 Hogg, D. W., & Foreman-Mackey, D. 2018, ApJS, 236, 11 Liu, A., Parsons, A. R., & Trott, C. M. 2014b, PhRvD, 90, 023019 Howlett, C., Lewis, A., Hall, A., & Challinor, A. 2012, JCAP, 2012, 027 Liu, A., Pritchard, J. R., Tegmark, M., & Loeb, A. 2013, PhRvD, 87, 043002 Hurley-Walker, N., Callingham, J. R., Hancock, P. J., et al. 2017, MNRAS, Liu, A., & Tegmark, M. 2011, PhRvD, 83, 103006 464, 1146 Liu, A., & Tegmark, M. 2012, MNRAS, 419, 3491 Iliev, I. T., Mellema, G., Pen, U.-L., et al. 2006, MNRAS, 369, 1625 Liu, A., Tegmark, M., Bowman, J., Hewitt, J., & Zaldarriaga, M. 2009a, Isserlis, L. 1918, Biometrika, 12, 134 MNRAS, 398, 401 Jacobs, D. C., Burba, J., Bowman, J. D., et al. 2017, PASP, 129, 035002 Liu, A., Tegmark, M., Morrison, S., Lutomirski, A., & Zaldarriaga, M. 2010, Jacobs, D. C., Hazelton, B. J., Trott, C. M., et al. 2016, ApJ, 825, 114 MNRAS, 408, 1029 Jacobs, D. C., Pober, J. C., Parsons, A. R., et al. 2015, ApJ, 801, 51 Liu, A., Tegmark, M., & Zaldarriaga, M. 2009b, MNRAS, 394, 1575 Jana, R., Nath, B. B., & Biermann, P. L. 2019, MNRAS, 483, 5329 Liu, A., Zhang, Y., & Parsons, A. R. 2016, ApJ, 833, 242 Jensen, H., Majumdar, S., Mellema, G., et al. 2016, MNRAS, 456, 66 Loeb, A., & Furlanetto, S. R. 2013, The First Galaxies in the Universe Jia, L.-B. 2019, EPJC, 79, 80 (Princeton, NJ: Princeton Univ. Press) Johnston-Hollitt, M., Govoni, F., Beck, R., et al. 2015, Advancing Loeb, A., & Zaldarriaga, M. 2004, PhRvL, 92, 211301 Astrophysics with the Square Kilometre Array (AASKA14),92 Loi, S. T., Murphy, T., Cairns, I. H., et al. 2015a, GeoRL, 42, 3707

86 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Loi, S. T., Murphy, T., Cairns, I. H., et al. 2016, RaSc, 51, 659 Muñoz, J. B. 2019b, PhRvL, 123, 131301 Loi, S. T., Trott, C. M., Murphy, T., et al. 2015b, RaSc, 50, 574 Muñoz, J. B., Ali-Haïmoud, Y., & Kamionkowski, M. 2015, PhRvD, 92, Lutomirski, A., Tegmark, M., Sanchez, N. J., et al. 2011, MNRAS, 410, 2075 083508 Maartens, R., Abdalla, F. B., Jarvis, M., Santos, M. G. & SKA Cosmology Muñoz, J. B., Dvorkin, C., & Loeb, A. 2018, PhRvL, 121, 121301 SWG, f. t. 2015, PoS, 215 Muñoz, J. B., & Loeb, A. 2018, Natur, 557, 684 Mack, K. J., & Wyithe, J. S. B. 2012, MNRAS, 425, 2988 Murray, S. G., & Trott, C. M. 2018, ApJ, 869, 25 Mackay, D. J. C. 2003, Information Theory, Inference and Learning Murray, S. G., Trott, C. M., & Jordan, C. H. 2017, ApJ, 845, 7 Algorithms (Cambridge: Cambridge Univ. Press) Myers, S. T., Contaldi, C. R., Bond, J. R., et al. 2003, ApJ, 591, 575 Madau, P., & Haardt, F. 2015, ApJL, 813, L8 Neben, A. R., Bradley, R. F., Hewitt, J. N., et al. 2015, RaSc, 50, 614 Mahesh, N., Subrahmanyan, R., Udaya Shankar, N., & Raghunathan, A. 2014, Neben, A. R., Bradley, R. F., Hewitt, J. N., et al. 2016, ApJ, 826, 199 arXiv:1406.2585 Newburgh, L. B., Addison, G. E., Amiri, M., et al. 2014, Proc. SPIE, 9145, Majumdar, S., Pritchard, J. R., Mondal, R., et al. 2018, MNRAS, 476, 4007 91454V Malloy, M., & Lidz, A. 2013, ApJ, 767, 68 Newburgh, L. B., Bandura, K., Bucher, M. A., et al. 2016, Proc. SPIE, 9906, Malloy, M., & Lidz, A. 2015, ApJ, 799, 179 99065X Mao, Y., Tegmark, M., McQuinn, M., Zaldarriaga, M., & Zahn, O. 2008, Nhan, B. D., Bordenave, D. D., Bradley, R. F., et al. 2019, ApJ, 883, 126 PhRvD, 78, 023529 Nhan, B. D., Bradley, R. F., & Burns, J. O. 2017, ApJ, 836, 90 Marthi, V. R., & Chengalur, J. 2014, MNRAS, 437, 524 Nityananda, R. 2003, Low Frequency Radio Astronomy, ed. J. N. Chengalur, Martinot, Z. E., Aguirre, J. E., Kohn, S. A., & Washington, I. Q. 2018, ApJ, Y. Gupta, & K. S. Dwarakanath (Pune: National Centre for 869, 79 Radio Astrophysics), http://www.ncra.tifr.res.in/ncra/gmrt/gmrt-users/ Masui, K. W., & Pen, U.-L. 2010, PhRvL, 105, 161302 low-frequency-radio-astronomy Masui, K. W., Switzer, E. R., Banavar, N., et al. 2013, ApJL, 763, L20 Noordam, J. E. 2004, Proc. SPIE, 5489, 817 McKinley, B., Bernardi, G., Trott, C. M., et al. 2018, MNRAS, 481, 5034 Obuljen, A., Castorina, E., Villaescusa-Navarro, F., & Viel, M. 2018, JCAP, McQuinn, M., Lidz, A., Zahn, O., et al. 2007, MNRAS, 377, 1043 2018, 004 McQuinn, M., & O’Leary, R. M. 2012, ApJ, 760, 3 Offringa, A. R., de Bruyn, A. G., Biehl, M., et al. 2010, MNRAS, 405, 155 McQuinn, M., Zahn, O., Zaldarriaga, M., Hernquist, L., & Furlanetto, S. R. Offringa, A. R., van de Gronde, J. J., & Roerdink, J. B. T. M. 2012, A&A, 2006, ApJ, 653, 815 539, A95 Meixner, M., Cooray, A., Leisawitz, D., et al. 2018, in The Cosmic Wheel and Oh, S. P., & Mack, K. J. 2003, MNRAS, 346, 871 the Legacy of the Archive: From Galaxies and Stars to Planets and Oppermann, N., Junklewitz, H., Robbers, G., et al. 2012, A&A, 542, A93 Life, 59 Orosz, N., Dillon, J. S., Ewall-Wice, A., Parsons, A. R., & Thyagarajan, N. Mellema, G., Iliev, I. T., Pen, U.-L., & Shapiro, P. R. 2006, MNRAS, 372, 679 2019, MNRAS, 487, 537 Mena-Parra, J., Bandura, K., Dobbs, M. A., Shaw, J. R., & Siegel, S. 2018, Oyama, Y., Kohri, K., & Hazumi, M. 2016, JCAP, 2016, 008 JAI, 7, 1850008 Paciga, G., Albert, J. G., Bandura, K., et al. 2013, MNRAS, 433, 639 Mertens, F. G., Ghosh, A., & Koopmans, L. V. E. 2018, MNRAS, 478, 3640 Paciga, G., Chang, T.-C., Gupta, Y., et al. 2011, MNRAS, 413, 1174 Mesinger, A., Ewall-Wice, A., & Hewitt, J. 2014, MNRAS, 439, 3262 Pan, T., & Barkana, R. 2012, arXiv:1209.5751 Mesinger, A., Furlanetto, S., & Cen, R. 2011, MNRAS, 411, 955 Park, J., Mesinger, A., Greig, B., & Gillet, N. 2019, MNRAS, 484, 933 Mesinger, A., Greig, B., & Sobacchi, E. 2016, MNRAS, 459, 2342 Parsons, A., Pober, J., McQuinn, M., Jacobs, D., & Aguirre, J. 2012a, ApJ, Mevius, M., van der Tol, S., Pandey, V. N., et al. 2016, RaSc, 51, 927 753, 81 Mirocha, J., & Furlanetto, S. R. 2019, MNRAS, 483, 1980 Parsons, A. R., Backer, D. C., Foster, G. S., et al. 2010, AJ, 139, 1468 Mirocha, J., Furlanetto, S. R., & Sun, G. 2017, MNRAS, 464, 1365 Parsons, A. R., Liu, A., Aguirre, J. E., et al. 2014, ApJ, 788, 106 Mirocha, J., Harker, G. J. A., & Burns, J. O. 2015, ApJ, 813, 11 Parsons, A. R., Liu, A., Ali, Z. S., & Cheng, C. 2016, ApJ, 820, 51 Mitchell, D. A., Greenhill, L. J., Wayth, R. B., et al. 2008, ISTSP, 2, 707 Parsons, A. R., Pober, J. C., Aguirre, J. E., et al. 2012b, ApJ, 756, 165 Mitridate, A., & Podo, A. 2018, JCAP, 2018, 069 Patil, A. H., Yatawatta, S., Koopmans, L. V. E., et al. 2017, ApJ, 838, 65 Mondal, R., Bharadwaj, S., & Majumdar, S. 2016, MNRAS, 456, 1936 Patil, A. H., Yatawatta, S., Zaroubi, S., et al. 2016, MNRAS, 463, 4317 Mondal, R., Bharadwaj, S., & Majumdar, S. 2017, MNRAS, 464, 2992 Patra, N., Bray, J. D., Roberts, P., & Ekers, R. D. 2017, ExA, 43, 119 Mondal, R., Bharadwaj, S., Majumdar, S., Bera, A., & Acharyya, A. 2015, Patra, N., Parsons, A. R., DeBoer, D. R., et al. 2018, ExA, 45, 177 MNRAS, 449, L41 Patra, N., Subrahmanyan, R., Raghunathan, A., & Udaya Shankar, N. 2013, Monsalve, R. A., Fialkov, A., Bowman, J. D., et al. 2019, ApJ, 875, 67 ExA, 36, 319 Monsalve, R. A., Greig, B., Bowman, J. D., et al. 2018, ApJ, 863, 11 Patra, N., Subrahmanyan, R., Sethi, S., Udaya Shankar, N., & Raghunathan, A. Monsalve, R. A., Rogers, A. E. E., Bowman, J. D., & Mozdzen, T. J. 2017, 2015, ApJ, 801, 138 ApJ, 835, 49 Paul, S., Sethi, S. K., Morales, M. F., et al. 2016, ApJ, 833, 213 Monsalve, R. A., Rogers, A. E. E., Mozdzen, T. J., & Bowman, J. D. 2016, Pen, U.-L., Chang, T.-C., Hirata, C. M., et al. 2009, MNRAS, 399, 181 ITMTT, 64, 2631 Peterson, J. B., Bandura, K., & Zen, U. L. 2006, arXiv:astro-ph/0606104 Moore, D. F., Aguirre, J. E., Kohn, S. A., et al. 2017, ApJ, 836, 154 Petrovic, N., & Oh, S. P. 2011, MNRAS, 413, 2103 Moore, D. F., Aguirre, J. E., Parsons, A. R., Jacobs, D. C., & Pober, J. C. 2013, Philip, L., Abdurashidova, Z., Chiang, H. C., et al. 2019, JAI, 8, 1950004 ApJ, 769, 154 Pindor, B., Wyithe, J. S. B., Mitchell, D. A., et al. 2011, PASA, 28, 46 Morales, M. F. 2005, ApJ, 619, 678 Planck Collaboration, Adam, R., Ade, P. A. R., et al. 2016a, A&A, 594, Morales, M. F. 2011, PASP, 123, 1265 A10 Morales, M. F., Beardsley, A., Pober, J., et al. 2019, MNRAS, 483, 2207 Planck Collaboration, Ade, P. A. R., Aghanim, N., et al. 2016b, A&A, Morales, M. F., Hazelton, B., Sullivan, I., & Beardsley, A. 2012, ApJ, 752, 137 594, A13 Morales, M. F., & Hewitt, J. 2004, ApJ, 615, 7 Planck Collaboration, Aghanim, N., Akrami, Y., et al. 2018a, arXiv:1807. Morales, M. F., & Wyithe, J. S. B. 2010, ARA&A, 48, 127 06209 Moroi, T., Nakayama, K., & Tang, Y. 2018, PhLB, 783, 301 Planck Collaboration, Akrami, Y., Ashdown, M., et al. 2018b, A&A, in press Mouri Sardarabadi, A., & Koopmans, L. V. E. 2019, MNRAS, 483, 5480 Pober, J. C. 2015, MNRAS, 447, 1705 Mozdzen, T. J., Bowman, J. D., Monsalve, R. A., & Rogers, A. E. E. 2016, Pober, J. C., Liu, A., Dillon, J. S., et al. 2014, ApJ, 782, 66 MNRAS, 455, 3890 Pober, J. C., Parsons, A. R., Aguirre, J. E., et al. 2013a, ApJL, 768, L36 Mozdzen, T. J., Bowman, J. D., Monsalve, R. A., & Rogers, A. E. E. 2017, Pober, J. C., Parsons, A. R., DeBoer, D. R., et al. 2013b, AJ, 145, 65 MNRAS, 464, 4995 Pober, J. C., Parsons, A. R., Jacobs, D. C., et al. 2012, AJ, 143, 53 Mozdzen, T. J., Mahesh, N., Monsalve, R. A., Rogers, A. E. E., & Prandoni, I., & Seymour, N. 2015, Advancing Astrophysics with the Square Bowman, J. D. 2019, MNRAS, 483, 4411 Kilometre Array (AASKA14),67 Muñoz, J. B. 2019a, PhRvD, 100, 063538 Presley, M. E., Liu, A., & Parsons, A. R. 2015, ApJ, 809, 18

87 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Prestage, R. M., Constantikes, K. T., Hunter, T. R., et al. 2009, IEEEP, Su, H., Hurley-Walker, N., Jackson, C. A., et al. 2017b, MNRAS, 465, 3163 97, 1382 Subrahmanya, C. R., Manoharan, P. K., & Chengalur, J. N. 2017, JApA, 38, 10 Price, D. C. 2016, arXiv:1607.03579 Sullivan, I. S., Morales, M. F., Hazelton, B. J., et al. 2012, ApJ, 759, 17 Price, D. C., Greenhill, L. J., Fialkov, A., et al. 2018, MNRAS, 478, 4193 Switzer, E. R. 2017, ApJ, 838, 82 Pritchard, J. R., & Loeb, A. 2012, RPPh, 75, 086901 Switzer, E. R., Anderson, C. J., Pullen, A. R., & Yang, S. 2019, ApJ, 872, 82 Pritchard, J. R., Loeb, A., & Wyithe, J. S. B. 2010, MNRAS, 408, 57 Switzer, E. R., Chang, T.-C., Masui, K. W., Pen, U.-L., & Voytek, T. C. 2015, Rasmussen, C. E., & Williams, C. K. I. 2006, Gaussian Processes for Machine ApJ, 815, 51 Learning (Cambridge, MA: MIT Press) Switzer, E. R., & Liu, A. 2014, ApJ, 793, 102 Remazeilles, M., Dickinson, C., Banday, A. J., Bigot-Sazy, M.-A., & Switzer, E. R., Masui, K. W., Bandura, K., et al. 2013, MNRAS, 434, L46 Ghosh, T. 2015, MNRAS, 451, 4311 Tauscher, K., Rapetti, D., & Burns, J. O. 2018a, JCAP, 2018, 015 Robertson, B. E., Ellis, R. S., Dunlop, J. S., McLure, R. J., & Stark, D. P. 2010, Tauscher, K., Rapetti, D., Burns, J. O., & Switzer, E. 2018b, ApJ, 853, 187 Natur, 468, 49 Taylor, A., & Joachimi, B. 2014, MNRAS, 442, 2728 Rogers, A. E. E., & Bowman, J. D. 2008, AJ, 136, 641 Tegmark, M. 1995, ApJ, 455, 429 Rogers, A. E. E., & Bowman, J. D. 2012, RaSc, 47, RS0K06 Tegmark, M. 1997a, PhRvD, 56, 4514 Safarzadeh, M., Scannapieco, E., & Babul, A. 2018, ApJ, 859, L18 Tegmark, M. 1997b, ApJL, 480, L87 Santos, M. G., Cooray, A., & Knox, L. 2005, ApJ, 625, 575 Tegmark, M., Eisenstein, D. J., Hu, W., & de Oliveira-Costa, A. 2000, ApJ, Sathyanarayana Rao, M., Subrahmanyan, R., Udaya Shankar, N., & Chluba, J. 530, 133 2017a, AJ, 153, 26 Tegmark, M., Hamilton, A. J. S., Strauss, M. A., Vogeley, M. S., & Sathyanarayana Rao, M., Subrahmanyan, R., Udaya Shankar, N., & Chluba, J. Szalay, A. S. 1998, ApJ, 499, 555 2017b, ApJ, 840, 33 Tegmark, M., & Zaldarriaga, M. 2009, PhRvD, 79, 083530 Schmit, C. J., & Pritchard, J. R. 2018, MNRAS, 475, 1213 Tegmark, M., & Zaldarriaga, M. 2010, PhRvD, 82, 103501 Schneider, A. 2018, PhRvD, 98, 063021 Thompson, A. R., Moran, J. M., & Swenson, G. W., Jr. 2017, Interferometry Schneider, M., Holm, Ó., & Knox, L. 2011, ApJ, 728, 137 and Synthesis in Radio Astronomy (3rd ed.; Berlin: Springer) Schwab, F. R. 1984, AJ, 89, 1076 Thyagarajan, N., Beardsley, A. P., Bowman, J. D., & Morales, M. F. 2017, Scott, D., & Rees, M. J. 1990, MNRAS, 247, 510 MNRAS, 467, 715 Semelin, B. 2016, MNRAS, 455, 962 Thyagarajan, N., Carilli, C. L., & Nikolic, B. 2018, PhRvL, 120, 251301 Seo, H.-J., & Eisenstein, D. J. 2003, ApJ, 598, 720 Thyagarajan, N., Jacobs, D. C., Bowman, J. D., et al. 2015a, ApJL, 807, L28 Seo, H.-J., & Hirata, C. M. 2016, MNRAS, 456, 3142 Thyagarajan, N., Jacobs, D. C., Bowman, J. D., et al. 2015b, ApJ, 804, 14 Sharma, P. 2018, MNRAS, 481, L6 Thyagarajan, N., Udaya Shankar, N., Subrahmanyan, R., et al. 2013, ApJ, Shaver, P. A., Windhorst, R. A., Madau, P., & de Bruyn, A. G. 1999, A&A, 776, 6 345, 380 Tingay, S. J., Goeke, R., Bowman, J. D., et al. 2013, PASA, 30, e007 Shaw, A. K., Bharadwaj, S., & Mondal, R. 2019, MNRAS, 487, 4951 Trott, C. M. 2014, PASA, 31, e026 Shaw, J. R., Sigurdson, K., Pen, U.-L., Stebbins, A., & Sitwell, M. 2014, ApJ, Trott, C. M. 2016, MNRAS, 461, 126 781, 57 Trott, C. M., Fu, S. C., Murray, S. G., et al. 2019a, MNRAS, 486, 5766 Shaw, J. R., Sigurdson, K., Sitwell, M., Stebbins, A., & Pen, U.-L. 2015, Trott, C. M., Jordan, C. H., Murray, S. G., et al. 2018, ApJ, 867, 15 PhRvD, 91, 083514 Trott, C. M., Pindor, B., Procopio, P., et al. 2016, ApJ, 818, 139 Shimabukuro, H., Yoshiura, S., Takahashi, K., Yokoyama, S., & Ichiki, K. Trott, C. M., Watkinson, C. A., Jordan, C. H., et al. 2019b, PASA, 36, e023 2015, MNRAS, 451, 467 Trott, C. M., Wayth, R. B., & Tingay, S. J. 2012, ApJ, 757, 101 Shimabukuro, H., Yoshiura, S., Takahashi, K., Yokoyama, S., & Ichiki, K. van Haarlem, M. P., Wise, M. W., Gunst, A. W., et al. 2013, A&A, 556, A2 2017, MNRAS, 468, 1542 Vedantham, H., Udaya Shankar, N., & Subrahmanyan, R. 2012, ApJ, 745, 176 Shiraishi, M., Muñoz, J. B., Kamionkowski, M., & Raccanelli, A. 2016, Vedantham, H. K., & Koopmans, L. V. E. 2015, MNRAS, 453, 925 PhRvD, 93, 103506 Vedantham, H. K., & Koopmans, L. V. E. 2016, MNRAS, 458, 3099 Sievers, J. L. 2017, arXiv:1701.01860 Vedantham, H. K., Koopmans, L. V. E., de Bruyn, A. G., et al. 2014, MNRAS, Silva, M., Santos, M. G., Cooray, A., & Gong, Y. 2015, ApJ, 806, 209 437, 1056 Sims, P. H., Lentati, L., Alexander, P., & Carilli, C. L. 2016, MNRAS, Vedantham, H. K., Koopmans, L. V. E., de Bruyn, A. G., et al. 2015, MNRAS, 462, 3069 450, 2291 Sims, P. H., Lentati, L., Pober, J. C., et al. 2019, MNRAS, 484, 4152 Venumadhav, T., Chang, T.-C., Doré, O., & Hirata, C. M. 2016, ApJ, 826, 116 Sims, P. H., & Pober, J. C. 2020, MNRAS, 492, 22 Villaescusa-Navarro, F., Genel, S., Castorina, E., et al. 2018, ApJ, 866, 135 Singh, S., & Subrahmanyan, R. 2019, arXiv:1903.04540 Visbal, E., & Loeb, A. 2010, JCAP, 11, 016 Singh, S., Subrahmanyan, R., Shankar, N. U., et al. 2018a, ExA, 45, 269 Wang, X., Tegmark, M., Santos, M. G., & Knox, L. 2006, ApJ, 650, 529 Singh, S., Subrahmanyan, R., Udaya Shankar, N., et al. 2017, ApJL, 845, Wang, Z., Chen, Y., Mao, Y., et al. 2019, arXiv:1901.02724 L12 Watkinson, C. A., Giri, S. K., Ross, H. E., et al. 2019, MNRAS, 482, 2653 Singh, S., Subrahmanyan, R., Udaya Shankar, N., et al. 2018b, ApJ, 858, 54 Watkinson, C. A., Majumdar, S., Pritchard, J. R., & Mondal, R. 2017, Singh, S., Subrahmanyan, R., Udaya Shankar, N., & Raghunathan, A. 2015, MNRAS, 472, 2436 ApJ, 815, 88 Watkinson, C. A., Mesinger, A., Pritchard, J. R., & Sobacchi, E. 2015, Slatyer, T. R., & Wu, C.-L. 2018, PhRvD, 98, 023013 MNRAS, 449, 3202 Slosar, A. 2017, PhRvL, 118, 151301 Watkinson, C. A., & Pritchard, J. R. 2014, MNRAS, 443, 3090 Smits, J. M., Stappers, B. W., Macquart, J. P., Ramachandran, R., & Watkinson, C. A., & Pritchard, J. R. 2015, MNRAS, 454, 1416 Kuijpers, J. 2003, A&A, 405, 795 Wayth, R. B., Lenc, E., Bell, M. E., et al. 2015, PASA, 32, e025 Sobacchi, E., & Mesinger, A. 2014, MNRAS, 440, 1662 Wayth, R. B., Tingay, S. J., Trott, C. M., et al. 2018, PASA, 35, 33 Sokolowski, M., Wayth, R. B., & Lewis, M. 2016, 2015 IEEE Global Wick, G. C. 1950, PhRv, 80, 268 Electromagnetic Compatibility Conference (GEMCCON) (Piscataway, Wieringa, M. H. 1992, ExA, 2, 203 NJ: IEEE) Winkel, B., Kerp, J., & Stanko, S. 2007, AN, 328, 68 Srivastava, N., Hinton, G., Krizhevsky, A., Sutskever, I., & Salakhutdinov, R. Wolleben, M., Landecker, T. L., Carretti, E., et al. 2019, AJ, 158, 44 2014, J. Mach. Learn. Res., 15, 1929 Wolz, L., Abdalla, F. B., Blake, C., et al. 2014, MNRAS, 441, 3271 Stacey, G. J., Aravena, M., Basu, K., et al. 2018, Proc. SPIE, 10700, 107001M Wolz, L., Blake, C., Abdalla, F. B., et al. 2017a, MNRAS, 464, 4938 Staveley-Smith, L., & Oosterloo, T. 2015, Advancing Astrophysics with the Wolz, L., Blake, C., & Wyithe, J. S. B. 2017b, MNRAS, 470, 3220 Square Kilometre Array (AASKA14), 167 Wuensche, C. A. & the BINGO Collaboration 2019, J. Phys.: Conf. Ser., 1269, Staveley-Smith, L., Wilson, W. E., Bird, T. S., et al. 1996, PASA, 13, 243 012002 Su, H., Hurley-Walker, N., Jackson, C. A., et al. 2017a, MNRAS, 472, 828 Wyithe, J. S. B., & Loeb, A. 2004, ApJ, 610, 117

88 Publications of the Astronomical Society of the Pacific, 132:062001 (89pp), 2020 June Liu & Shaw

Wyithe, J. S. B., & Loeb, A. 2007, MNRAS, 375, 1034 Yoshiura, S., Shimabukuro, H., Takahashi, K., et al. 2015, MNRAS, 451, Wyithe, J. S. B., & Loeb, A. 2008, MNRAS, 383, 606 266 Wyithe, J. S. B., Loeb, A., & Geil, P. M. 2008, MNRAS, 383, 1195 Yoshiura, S., Takahashi, K., & Takahashi, T. 2018, PhRvD, 98, 063529 Wyithe, J. S. B., & Morales, M. F. 2007, MNRAS, 379, 1647 Yue, B., Ferrara, A., Pallottini, A., Gallerani, S., & Vallini, L. 2015, MNRAS, Xu, Y., Chen, X., Fan, Z., Trac, H., & Cen, R. 2009, ApJ, 704, 1396 450, 3829 Xu, Y., Wang, X., & Chen, X. 2015, ApJ, 798, 40 Zahn, O., & Zaldarriaga, M. 2006, ApJ, 653, 922 Yatawatta, S. 2015, MNRAS, 449, 4506 Zhang, Y. G., Liu, A., & Parsons, A. R. 2018, ApJ, 852, 110 Yatawatta, S. 2016, arXiv:1605.09219 Zheng, H., Tegmark, M., Buza, V., et al. 2014, MNRAS, 445, 1084 Yatawatta, S. 2018, arXiv:1805.00265 Zheng, H., Tegmark, M., Dillon, J. S., et al. 2017a, MNRAS, 464, 3486 Yatawatta, S., Diblen, F., & Spreeuw, H. 2017, arXiv:1710.05656 Zheng, H., Tegmark, M., Dillon, J. S., et al. 2017b, MNRAS, 465, 2901 Yatawatta, S., Zaroubi, S., de Bruyn, G., Koopmans, L., & Noordam, J. 2008, Zhu, H.-M., Pen, U.-L., Yu, Y., & Chen, X. 2018, PhRvD, 98, 043511 arXiv:0810.5751 Zhu, H.-M., Pen, U.-L., Yu, Y., Er, X., & Chen, X. 2016, PhRvD, 93, 103504

89