The following article is Open access

Lunar Orbit Measurement of the Cosmic Dawn's 21 cm Global Spectrum

, , , , , and

Published 2022 April 11 © 2022. The Author(s). Published by the American Astronomical Society.
, , Citation Yuan Shi et al 2022 ApJ 929 32 DOI 10.3847/1538-4357/ac5965

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/929/1/32

Abstract

A redshifted 21 cm line absorption signature is commonly expected from the cosmic dawn era, when the first stars and galaxies formed. The detailed traits of this signal can provide important insight on the cosmic history. However, high-precision measurement of this signal is hampered by ionosphere refraction and absorption, as well as radio frequency interference (RFI). Space observation can solve the problem of the ionosphere, and the Moon can shield the RFI from Earth. In this paper, we present simulations of the global spectrum measurement in the 30–120 MHz frequency band on the lunar orbit from the proposed Discovering the Sky at the Longest wavelength project. In particular, we consider how the measured signal varies as the satellite moves along the orbit and take into account the blockage of different parts of the sky by the Moon and the antenna response. We estimate the sensitivity for such a 21 cm global spectrum experiment. An rms noise level of ≤0.05 K is expected at 75 MHz after 10 orbits (∼1 day) observation, for a frequency channel width of 0.4 MHz. We also study the influence of a frequency-dependent beam, which may generate complex structures in the spectrum. Estimates of the uncertainties in the foreground and 21 cm model parameters are obtained.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Depending on whether the spin temperature of neutral hydrogen atoms is higher or lower than that of the cosmic background radiation, there is net emission or absorption of radiation of 21 cm wavelength. Such emission or absorption features from the early universe provide us with a useful tool to study the cosmic dark ages, the cosmic dawn (CD), and the epoch of reionization (EoR).

The 21 cm global spectrum experiments aim to measure the sky-averaged spectrum with high precision so as to probe the early epochs of the universe. There are a number of such ground-based experiments, including the Experiment to Detect the Global Epoch-of-Reionization Signature (EDGES; Bowman et al. 2008; Bowman & Rogers 2010; Monsalve et al. 2017), the Sonda Cosmológica de las Islas para la Detecciónde Hidrógeno Neutro (SCI-HI; Voytek et al. 2014), the Probing Radio Intensity at high-z from Marion (PRIzM; Philip et al. 2019), the Shaped Antenna measurement of the background RAdio Spectrum (SARAS; Patra et al. 2013; Singh et al. 2017, 2018), the Cosmic Twilight Polarimeter (CTP; Nhan et al. 2019), the Broadband Instrument for Global Hydrogen Reionization Signal (BIGHORNS; Sokolowski et al. 2015), the Large-Aperture Experiment to Detect the Dark Age (LEDA; Bernardi et al. 2016; Bernardi 2018; Price et al. 2018), and the Radio Experiment for the Analysis of Cosmic Hydrogen (REACH; de Lera Acedo 2019). Compared to the 21 cm tomography experiments, measurement of the global 21 cm emission has a higher raw sensitivity and requires smaller collecting area, so that it could be conducted even with a single antenna. The EDGES (Bowman et al. 2018) reported the detection of a strong absorption feature at ∼78 MHz, which has a cosmic dawn 21 cm spectrum interpretation, though it has an unexpectedly large amplitude (0.5 K) and an unusual flattened profile. If it originated from the cosmic 21 cm spectrum, this would suggest possibly new physics or astrophysics (Chen & Miralda-Escudé 2004, 2008; Creasey et al. 2011; Nelson et al. 2013; Barkana 2018; Fialkov et al. 2018; Fraser et al. 2018; Barkana et al. 2018; Slatyer & Wu 2018; Li et al. 2021; Houston et al. 2018; Hirano & Bromm 2018; Muñoz et al. 2018; Yang 2021), though this result was not confirmed by a recent measurement of the SARAS experiment (Singh et al. 2022). It is imperative to check this result and improve upon it with further and more precise observations.

In ground-based global spectrum experiments, a number of possible systematic effects, such as attenuation and differential refraction of the radio wave by the ionosphere (Vedantham et al. 2014; Shen et al. 2021), radio frequency interference (RFI), and the interactions of the antenna with nearby ground and underground features, can introduce distortions and errors, thus limiting the measurement precision. The magnitude of the ionospheric disruptions in the global spectrum varies in the range of ∼0.1 K–3 K depending on the ionospheric condition. These systematic effects can be avoided or mitigated by a space-based experiment. A number of space mission concepts have been proposed, and some are actively being studied at present, such as the Discovering the Sky at the Longest wavelength (DSL; also known as Hongmeng in Chinese, Chen et al. 2019, 2020), the Dark Ages Polarimetry PathfindER (DAPPER; Tauscher et al. 2018a) and its precursor the Dark Ages Radio Explorer (DARE, Burns et al. 2017), and the Farside Array for Radio Science Investigations of the Dark ages and Exoplanets (FARSIDE; Burns et al. 2019, 2021). By orbiting the Moon, or landing on the far side of the lunar surface, the RFI from Earth can be shielded, providing an ideal environment for such observations.

Compared with lunar surface experiments, the lunar orbit experiment has the advantage that it is not affected by the transient ionosphere on the dayside of the lunar surface or the reflections of the lunar rocks. In terms of engineering, it is also simpler, as it does not require landing on the Moon, and it is easy to supply the required energy (solar power) and transmit the data back to Earth. In the DSL mission concept, an array of satellites will be launched together as an assembly into the lunar orbit by a single rocket and then released sequentially into a linear formation on the same circular orbit. During the mission, these satellites will make both interferometric imaging (Huang et al. 2018) and high-precision global spectrum measurement on the part of orbit behind the Moon, and the data will be transmitted to Earth on the near-side part of the orbit. Shi et al. (2022, hereafter referred as Paper I) have investigated the imaging quality and sensitivity of such an array in the presence of thermal noise for interferometric observation.

For the global spectrum experiment, the lunar orbit is also an ideal option, as the ionosphere refraction and ground reflection are largely avoided. Nevertheless, as the antenna beam may have frequency dependence, it is still a complicated problem. This effect has been investigated in a number of studies. In the context of the EDGES experiment, Mozdzen et al. (2016) simulated the chromatic beam effects for two dipole antennas used in observations and assessed the detectability of the global 21 cm signal. Further, Mahesh et al. (2021) validated the beam model used by EDGES by comparing different electromagnetic solvers and by comparing the simulated spectra with real observational data. More generally, in order to extract the 21 cm signal from the foregrounds convolved by a chromatic beam, one can construct training sets that include the foreground, the beam, and other instrument effects, which are then decomposed into basis vectors using singular value decomposition (SVD; Tauscher et al. 2018b; Rapetti et al. 2020; Tauscher et al. 2020, 2021; Bassett et al. 2021). There are also some models to deal with systematic distortions by a Bayesian nested sampling algorithm (Anstey et al. 2021).

In this paper, we explore the 21 cm global spectrum measurements on the lunar orbit. While our investigation is made in the context of the DSL mission, some of the results may also be applicable to other similar missions. This paper is structured as follows. Section 2 introduces some basic setups of the global signal experiment and the tools for generating the mock spectrum: the foreground map, the 21 cm signal, and the noise and beam model. In Section 3, we describe the simulation and the extraction of the foreground and the 21 cm signal from the mock data. We conclude in Section 4.

2. The Global Spectrum Measurements

The basic parameters of the DSL concept relevant to our end-to-end simulation of the global spectrum experiment are listed in Table 1. In the current DSL mission concept, a single dedicated microsatellite would measure the global spectrum in the 30–120 MHz frequency range, while a number of microsatellites would perform both intersatellite interferometry and single-antenna global spectrum measurements below 30 MHz. The high-frequency band (30–120 MHz) microsatellite uses a disk-cone antenna, while the low-frequency band (1–30 MHz) microsatellites use tripole antennas. To keep a stable flying formation, the microsatellites are tuned to have one face oriented to the Moon center and an axis pointed along the tangent direction of the orbit, so that in the rotating frame, it is kept at a stable configuration. Here, as a simple model for the circular orbit, we assume the axis of the antenna cone is pointing to the center of the Moon.

Table 1. Basic Parameters of the DSL Global Spectrum Experiment

Mission ParameterValues
Orbit height300 km
Orbit radius2037 km
Orbital plane inclination30°
Precession period1.3 years
Total observation time3–5 years
High-frequency band (HFB)30–120 MHz
Low-frequency band (LFB)1–30 MHz
Channel bandwidth of HFB0.4 MHz
HFB time resolution8–10 s

Download table as:  ASCIITypeset image

With the satellites moving on the lunar orbit, parts of the sky are blocked by the Moon at any time, and the blocked region changes with time. For any given instant, the visible part of the sky is determined by the position of the satellite. We treat the Moon as a solid sphere and take a 300 km circular orbit with an inclination angle of 30° with respect to the lunar equator as our fiducial orbit, ignoring the variation of orbital height induced by the gravity anomaly. The orbital plane precesses with respect to the inertial frame with a cycle period of about 1.3 yr, which we approximate by a uniform rotation. The simple simulation presented in Paper I shows that during about 31% of the total orbital time, Earth is shielded by the Moon, and about 10% of the time is "doubly good," during which both Earth and the Sun are shielded.

To simulate the spectrum measurements, we generate time-ordered mock data at different frequencies assuming a sky foreground map, a 21 cm signal model, and a noise model, taking into account the satellite orbit, the Moon blockage, and the antenna response. We then try to extract the 21 cm signal and assess the measurement errors on the 21 cm model parameters. The signal received by the antenna is modeled by the convolution of the sky signal with the antenna beam and the Moon blockage,

Equation (1)

where $\hat{{\boldsymbol{k}}}$ is the unit direction vector, $B(\nu ,\hat{{\boldsymbol{k}}},t)$ is the antenna primary beam response, and $S(\hat{{\boldsymbol{k}}},t)$, which we shall call the shade function, describes the blockage of sky by the Moon. Here we treat it as a simple geometric function, ignoring the frequency dependence induced by diffraction. We pixelize the sky with the HEALPix scheme (Górski et al. 2005), and the integral is then approximated by a sum over pixels:

Equation (2)

where k is the pixel index, Npix is the total number of pixels, and ΔΩ = 4π/Npix is the solid angle corresponding to one pixel. We have also checked the precision of integration by computing Equation (2) with Nside = 8, 32, 64, and 128 using the foreground model described below. The differences are smaller than 10−4 K for our sky map. Below we take HEALPIX Nside = 64, corresponding to a resolution of about 1°, which should be sufficient for the global spectrum evaluation.

2.1. Foreground Model

The mock observational data are generated with a physically motivated sky model. The input sky map can be obtained at a given frequency by extrapolation from observed data. A few programs are available for this purpose, e.g., the Global Sky Model (GSM; de Oliveira-Costa et al. 2008; Zheng et al. 2017), the Cosmology in the Radio Band (CORA; Shaw et al. 2014), and the Self-consistent whole-Sky foreground Model (SSM; Huang et al. 2019). However, interstellar medium (ISM) absorption becomes quite significant at low frequencies, which may affect the foreground 21 cm signal separation. In this work, we make use of the high-resolution Ultra-Long wavelength Sky model with Absorption (ULSA; Cong et al. 2021), which incorporated the free–free absorption effect of the ISM. 6

The sky maps at 30 MHz, 60 MHz, and 120 MHz are shown in Figure 1. The galactic free–free absorption lowers the temperature of the sky at lower frequencies toward H ii regions around massive stars or supernova remnants—an especially obvious one is the Gum Nebula (Woermann et al. 2001) at the lower-left corner (around l ∼ 150°) at 30 MHz. In comparison to the conventional absorption-free model, the inclusion of the free–free absorption effect results in decrements of 2.5% at 30 MHz and 0.5% at 120 MHz in sky brightness temperature.

Figure 1.

Figure 1. The whole-sky map from the ULSA model at 30 MHz, 60 MHz, and 120 MHz in the ecliptic coordinate system.

Standard image High-resolution image

The ULSA model considers both cases of a constant spectral index and a direction-dependent spectral index for the emission. In the case of a direction-dependent spectral index, the spectral index map is obtained by combining the Haslam 408 MHz map (Haslam et al. 1974, 1981, 1982; Remazeilles et al. 2015), the LWA maps (Dowell et al. 2017), and the Guzman 45 MHz map (Guzmán et al. 2011; through Equation (29) in Cong et al. 2021), after smoothing all maps with a Gaussian beam with FWHM = 5°. The whole-sky-averaged spectrum is shown in Figure 2. The solid lines are the absorption-included sky brightness, and the dashed lines are the absorption-free sky brightness. Here we show the result for both a constant and a direction-dependent spectral index of emission. We can see that there are significant differences at the lower frequencies due to absorption, which may affect the fitting of the spectrum for a given foreground model. It is therefore important to take into account the ISM absorption if one is to fit the spectrum at lower frequencies. Also, considering the importance of incorporating the spatial variation in the spectral index when accounting for the chromatic beam effect (Anstey et al. 2021), in this work, we shall use the model with absorption and direction-dependent spectral index to simulate the mock data.

Figure 2.

Figure 2. The input Galactic foreground global spectrum from the ULSA model. The solid lines account for the free–free absorption by the ISM, while the dashed lines are for absorption-free models. In each set of lines, the red line assumes a constant spectral index, and the black line adopts a direction-dependent spectral index.

Standard image High-resolution image

Then, in the data analysis, the foreground contribution to the global spectrum could be modeled as polynomials of frequency or its logarithm. In Table 2, we list some foreground models used in the literature along with short names (Pritchard & Loeb 2008, 2010; Harker et al. 2012; Bowman et al. 2018; Hills et al. 2018).

Table 2. Foreground Fitting Models, νc = 75 MHz

ModelFitting Function
(a) five-term Poly ${T}_{{\rm{F}}}(\nu )=\displaystyle \sum _{n\,=\,0}^{4}{a}_{n}{\left(\tfrac{\nu }{{\nu }_{{\rm{c}}}}\right)}^{n-2.5}$
(b) seven-term Poly ${T}_{{\rm{F}}}(\nu )=\displaystyle \sum _{n\,=\,0}^{6}{a}_{n}{\left(\tfrac{\nu }{{\nu }_{{\rm{c}}}}\right)}^{n-2.5}$
(c) three-term LogPoly ${T}_{F}(\nu )={\left(\tfrac{\nu }{{\nu }_{c}}\right)}^{-2.5}\displaystyle \sum _{n\,=\,0}^{2}{a}_{n}{\left(\mathrm{log}\tfrac{\nu }{{\nu }_{c}}\right)}^{n}$
(d) five-term LogPoly ${T}_{F}(\nu )={\left(\tfrac{\nu }{{\nu }_{c}}\right)}^{-2.5}\displaystyle \sum _{n\,=\,0}^{4}{a}_{n}{\left(\mathrm{log}\tfrac{\nu }{{\nu }_{c}}\right)}^{n}$
(e) three-term LogLogPoly $\mathrm{log}{T}_{{\rm{F}}}(\nu )=\displaystyle \sum _{n\,=\,0}^{2}{a}_{n}{\left(\mathrm{log}\tfrac{\nu }{{\nu }_{{\rm{c}}}}\right)}^{n}$
(f) five-term LogLogPoly $\mathrm{log}{T}_{{\rm{F}}}(\nu )=\displaystyle \sum _{n\,=\,0}^{4}{a}_{n}{\left(\mathrm{log}\tfrac{\nu }{{\nu }_{{\rm{c}}}}\right)}^{n}$

Download table as:  ASCIITypeset image

Figure 3 shows the residuals of fitting the global spectrum using the models listed in Table 2. Here we show both a fit in the frequency range of 30–120 MHz (solid lines) and a fit in the frequency range of 60–120 MHz (dash lines) for each model. The rms of residuals of models (a)–(f) are 78.37 mK (0.06 mK), 7.58 mK (0.01 mK), 172.74 mK (33.69 mK), 8.14 mK (0.17 mK), 117.60 mK (1.21 mK), and 16.08 mK (0.01 mK) for the solid lines (dashed–dotted lines), respectively.

Figure 3.

Figure 3. The residuals after fitting the six models listed in Table 2, in the range of 30–120 MHz (solid lines) and 60–120 MHz (dashed–dotted lines). The lines are shifted vertically for clarity. (a) Five-term Poly (+0.6 K), (b) seven-term Poly (+0.4 K), (c) three-term LogPoly (+0.2 K), (d) five-term LogPoly, (e) three-term LogLogPoly (−0.2 K), and (f) five-term LogLogPoly (−0.4 K).

Standard image High-resolution image

As shown by the figure and the residues, the primary source of foreground, the galactic synchrotron radiation, has a nearly power-law form for its spectrum, which can be easily fitted with a few terms in a polynomial expansion. However, the ISM free–free absorption at low frequency breaks the power law. So when the fitting is expanded to the 30–120 MHz frequency range, the five-term Poly and three-term LogLogPoly models, which worked well in the 60–120 MHz, range become inadequate. The five-term LogPoly, five-term LogLogPoly, and seven-term Poly models can fit well in all cases. We take the five-term LogPoly model as our fiducial model for the foreground below.

2.2. Models of the 21 cm Signal

We consider here two models of the 21 cm global signal: one is the form adopted by the EDGES experiment in Bowman et al. (2018), another is a Gaussian fit for the 21 cm through with adjustable amplitude and width (Hills et al. 2018).

EDGES 21 cm model: This is motivated by the EDGES experiment observation. We assume the signal is

Equation (3)

where

Equation (4)

Gaussian model: as demonstrated in Cohen et al. (2018), different astrophysical parameters can produce a variety of global 21 cm spectra with absorption trough depths in the range of −200 mK < Tb < −25 mK. Xu et al. (2021) also took into account the inevitable nonlinear density fluctuations in the intergalactic medium, shock heating, and Compton heating and found that these effects can reduce the maximum absorption signal by 15% at redshift 17. Here we use a simpler three-parameter Gaussian form given in Equation (5) as our global signal model:

Equation (5)

Then we consider four cases for the global 21 cm signal profile:

Case 1: No detectable 21 cm global signal.

Case 2: The EDGES 21 cm signal given by Equation (3), with A = 0.5 K, ν0 = 78.0 MHz, w = 19.0 MHz, τ = 7.

Case 3: A Gaussian signal given by Equation (5), with A = 0.20 K, ν0 = 75.0 MHz, w = 6.0 MHz.

Case 4: A Gaussian signal given by Equation (5), with A = 0.155 K, ν0 = 75.0 MHz, w = 10.0 MHz.

Figure 4 illustrates the 21 cm global signal for these four scenarios.

Figure 4.

Figure 4. The 21 cm signal for different scenarios at 30–120 MHz. Case 1 (no 21 cm signal) is shown by the gray dashed line, Case 2 (EDGES signal) is shown by the blue solid line, and Case 3 and Case 4 (two Gaussian forms) are shown by the yellow and green solid lines, respectively.

Standard image High-resolution image

2.3. Thermal Noise

The thermal noise on the observed signal is

Equation (6)

where N is the number of independent measurements, Δν is the channel bandwidth, which we set to 0.4 MHz, and tint is the integration time. The system temperature is given by

Equation (7)

where the average sky temperature is given by

Equation (8)

Note that while it is important to account for the ISM absorption when fitting the low-frequency spectrum, which would affect the extracted 21 cm signal several orders of magnitude lower than the foregrounds, the absorption effect makes little difference when estimating the thermal noise. Therefore, a simple power-law model is assumed here. Trcv gives the equivalent antenna temperature induced by the noise in the receiver. As the antenna and receiver impedances are mismatched, the actual power that goes into the receiver is (1 − ∣Γ∣2)Pant, where Γ is the reflection coefficient between the antenna and receiver, and Pant is the power of the antenna output. So,

Equation (9)

where ${T}_{\mathrm{rcv}}^{0}$ is the noise temperature of the receiver. For our current receiver design, we have measured ${T}_{\mathrm{rcv}}^{0}\approx 200\,{\rm{K}}$ in the given band. Based on antenna simulation, we find that

Equation (10)

We plot Trcv in Figure 5. It is about 200 K near 80 MHz.

Figure 5.

Figure 5. The equivalent receiver temperature.

Standard image High-resolution image

With such a system temperature, for a single antenna (N = 1), after about 24 hr (10 orbits) of integration time, the rms noise level is ≤15 mK at 75 MHz, assuming a channel bandwidth of 0.4 MHz. Even if we assume that only the data obtained during the "doubly good time," when both Earth and Sun are shielded, can be used, which is about 10% of the total orbit time, it takes only 10 days for the thermal noise to be suppressed down to this level according to Equation (6).

A detailed study of the global spectrum in the 1–30 MHz band will be presented elsewhere, but here we note in passing that at 25 MHz, the sky temperature is about 3.6 × 104 K, according to Equation (8). After a full three years of integration, the thermal noise of a single antenna could be suppressed to a level of about 10 mK for 0.4 MHz channel bandwidth. There are five to eight satellites in the DSL mission, and each is equipped with three pairs of antennas of different polarizations. So the DSL mission has the potential to measure the 21 cm global spectrum from the dark ages and cosmic dawn. However, we note that this estimate accounts only for the thermal noise; a number of other factors or systematic errors, such as the self-generated RFI, stability, dynamic range, and spectral response of the system will probably limit the precision of the measurement. It is necessary to design the system such that these systematic errors will be smaller than the expected 21 cm signal.

2.4. Beam

The global spectrum in the 30–120 MHz band will be measured by an antenna on a single satellite in the DSL mission. The primary requirement of the design is to have low chromaticity, i.e., its beam pattern should be as frequency independent as possible, to avoid convolving the spatial structure of the sky into the frequency domain. This can be achieved by using an antenna that is electrically small, i.e., with a physical size smaller than the half-wavelength, so that the resonance frequency is above the observational band. The small size of the antenna is also consistent with the requirement on the payload of the mission.

In the present design, we consider a disk-cone antenna model, which is composed of a radiating cone and a reflective disk, as shown in Figure 6. The size parameters of the cone and disk are labeled on the figure. The design of this antenna will be presented elsewhere; here, we take a particular design as our fiducial model. The beam of this antenna is simulated using the software FEKO. 7 The cross section of the axial-symmetric beam profiles, cut at an arbitrary ϕ with varying θ, is shown in the top panel of Figure 7; the actual beam in the 3D space can be generated by rotating the cross-section figure around the central axis (θ = 0°). The beam pattern of these electrically small disk-cone antennas is quite similar to that of a short dipole; there is a single lobe of the beam (it peaks at θ = 90° and 270° on the cross-section plot). To see the effect of frequency variation of the beam more clearly, we also show the logarithmic frequency gradient of the beam $\partial B(\nu ,\theta )/\partial \mathrm{log}\nu =\nu \partial B(\nu ,\theta )/\partial \nu $ in the bottom panel of Figure 7. The gradient for this disk-cone antenna is fairly small. In the next section, we shall mainly use this antenna beam model, but to explore how the chromatic beam affects the global spectrum measurement result, we shall also consider beam models that are more frequency dependent.

Figure 6.

Figure 6. The disk-cone antenna. The size parameters are D1 = 40 cm, D2 = 4 cm, H1 = 30.5 cm, and H2 = 4.5 cm. The diameter of the optimal metallic reflector D3 is 55 cm in our fiducial model.

Standard image High-resolution image
Figure 7.

Figure 7. The cross section of the beam profiles at a constant ϕ for the fiducial model at three frequencies (top) and the corresponding frequency gradients at various directions and frequencies (bottom).

Standard image High-resolution image

3. Simulation and Results

With the setup described above, we simulate the observed global spectrum. First we investigate how the measured global spectrum varies as a function of time, as the satellite moves to different parts of the orbit. Next we estimate the measurement error after some time of observation. Then we consider how the systematic errors, such as the chromatic beam effect, and the nonflat spectral response of the system would affect the 21 cm signal measurement.

3.1. Variation of Observing Position

As there are large-scale variations in the foreground, and the Moon blocks different parts of the sky for the satellite orbiting to the different positions, we first check how much the variation is for different locations. For this purpose, we select four points, marked as A, B, C, and D, which are located 90° apart on the lunar orbit. These are essentially the cross-over points of the orbit with the ecliptic plane and the highest and lowest points away from the elliptic plane along the orbit. The observed global spectrum at these points are the averaged sky temperature as weighted by S( n )B( n ) at these locations. In Figure 8, we show the product of the antenna beam and the Moon blockage function, S( n )B( n ), at the four locations. As expected from the symmetry, the pair A and C have exactly opposite patterns, and so does the pair B and D.

Figure 8.

Figure 8. The beam seen from the four positions A, B, C, and D, which are located 90° apart, around one orbit. The value is normalized to fall in the range [0, 1].

Standard image High-resolution image

We then calculate the foreground spectra as would be observed at these points with our fiducial disk-cone antenna model described above. In Figure 9, we plot the global spectra in the 30–120 MHz range. There are obvious differences in the amplitude for the four locations, but the shapes of the spectra are still quite similar to each other, and the spectra from locations A and C almost coincide.

Figure 9.

Figure 9. Observed foreground spectra at positions A, B, C, and D.

Standard image High-resolution image

Just as shown by the spectra at these four points, there is a large variation in the total intensity. In Figure 10, we plot the simulated intensity variation at 60 MHz, as the satellite takes an observation over 24 hr. In this plot, the red solid line marks the part of data obtained when Earth is shielded, while the black dashed line are for the rest of the time. The figure shows that the signal received at a fixed frequency has a regular periodic variation, with a period of 8246 s corresponding to one cycle of the orbit, and an amplitude of ∼10% level.

Figure 10.

Figure 10. Signal received at 60 MHz over 24 hr of observation. The black dashed line is the spectra for 1 day continuous observation, and the red solid line accounts for the one-third effective observation time, during which the satellite is shielded by the Moon from Earth.

Standard image High-resolution image

Actually the global spectrum is measured over a period of time as the satellite orbits the Moon. We can compute the sky average for a segment of the orbit. Despite large variations in the overall amplitude of the antenna temperature at different parts of the orbit, as shown in Figure 10, the variation in the shape of the spectrum is not large.

We fit each of the integrated spectrum for the AB, BC, CD, and DA segments of the orbit separately with the five-term LogPoly foreground model and plot the residuals in Figure 11. Here the input sky model includes only the foreground, while the mock observation is performed with simulated noise, Moon blockage, and the fiducial model of antenna response. For clarity, the results are vertically shifted by different amounts in the figure. We find that the beam-weighted signals can be reconstructed with a high accuracy. The four segments yield similar residuals, and these residuals (0.012 K, 0.009 K, 0.011 K, 0.008 K, respectively) are all noticeably smaller in magnitude than the expected global 21 cm signal in typical cosmic dawn models, showing that despite the variation of the foreground with sky direction, this would not seriously affect the 21 cm signal extraction.

Figure 11.

Figure 11. Residuals of the observed foreground spectra for the four segments of orbit, after fitting the five-term LogPoly model.

Standard image High-resolution image

3.2. Projected Error

In Figure 12, we plot the averaged spectrum at 30–120 MHz after a total of 30 orbits, but take observations only when Earth is shielded, which is about one-third of the effective time. In the same figure we also plot the fitting residual of the foreground model for both the noise-free case and with noise given at a level corresponding to tobs = 10 orbits. The data are assumed to be taken with our fiducial antenna model.

Figure 12.

Figure 12. Top: the average spectrum for 10 orbits of effective observation (total 30 orbits but one-third are shielded by the Moon from Earth). Bottom: the residuals of foreground fitting for the noise-included (black solid line) and noise-free (red dashed–dotted line) cases.

Standard image High-resolution image

We find that the LogPoly model is able to account for the global spectral structure by expanding around the expected power-law spectral index of −2.5. Here a0 is an overall foreground scale factor, $\tfrac{{a}_{n}}{{a}_{0}}\ll 1\ (n\geqslant 1)$ is the correction to the typical spectral index of the foreground by capturing anisotropies of the spectral index and other higher-order spectral corrections, such as the ISM absorption. The five-term LogPoly model can provide a good fit in this foreground-only sky model, with the foreground model parameters taking reasonable values as shown in Figure 13. There is also no serious degeneracy in these parameters. Note that with a microsatellite on the lunar orbit, our fitting implicitly assumes that the parameter dependencies encode only the foreground properties and does not account for ionospheric effects.

Figure 13.

Figure 13. Foreground model parameter fitting results for 10 orbit observations. We plot the contour maps with 1σ, 2σ, and 3σ for the LogPoly foreground coefficients a0, a1, a2, a3, and a4. The solid lines are the 1D probability distribution functions of the parameters.

Standard image High-resolution image

Next we consider the 21 cm signal extraction from the mock data simulated with the foreground, the 21 cm signal, and the noise. We plot in Figure 14 the five-term LogPoly model0fitting results for the EDGES 21 cm model (left panel) and the two Gaussian 21 cm models (middle and right panels). We see that in all cases, the 21 cm model signals can be well recovered by fitting the mock observation data.

Figure 14.

Figure 14. Fitting results for three 21 cm models after an effective 10 orbit observation. The thermal noise is σT ∼ 0.05 K. The blue solid lines denote residuals after fitting and removing only the foreground, while the orange lines denote residuals after fitting and removing the combined foreground and 21 cm signal. The black dashed lines denote the recovered 21 cm signals with noise.

Standard image High-resolution image

We also list the fitted model parameters for the EDGES 21 cm model in Table 3 and for the two Gaussian 21 cm models in Table 4 (the four rows correspond to the satellite completing 1, 2, 5, and 10 orbits, respectively). The thermal noise level drops with observation time as ${t}_{\mathrm{obs}}^{-1/2}$. For the frequency resolution of 0.4 MHz, the residual rms induced by the thermal noise drops to ∼0.05 K after 10 orbits, allowing the extraction of the 21 cm signal with good precision. The uncertainties of the parameters are estimated by Markov Chain Monte Carlo simulations. If the experiment is properly designed so that the error is dominated by thermal noise on this timescale, one should be able to discriminate between the EDGES model and the Gaussian models, and also between the two Gaussian models with different parameters introduced earlier. We also checked the results generated at different starting positions of precession for a 10 effective-orbit observation; the recovered 21 cm signals generally agree with each other within 0.01 K.

Table 3. The Best-fit Parameters of the 21 cm Global Signal and rms Values for the EDGES Model

21 cm model tobs [orbit]1 rms [K] A τ ν0 ω
Input0.507.0078.0019.00
Recovered10.133 ${0.501}_{-0.013}^{0.013}$ ${7.2}_{-1.1}^{1.5}$ ${77.96}_{-0.14}^{0.13}$ ${19.00}_{-0.31}^{0.31}$
 20.077 ${0.5017}_{-0.0093}^{0.0092}$ ${6.84}_{-0.82}^{0.94}$ ${77.89}_{-0.10}^{0.10}$ ${19.00}_{-0.23}^{0.22}$
 50.061 ${0.5001}_{-0.0067}^{0.0061}$ ${6.92}_{-0.54}^{0.67}$ ${77.960}_{-0.065}^{0.065}$ ${19.02}_{-0.15}^{0.14}$
 100.041 ${0.5004}_{-0.0044}^{0.0043}$ ${7.02}_{-0.40}^{0.46}$ ${77.979}_{-0.045}^{0.046}$ ${19.02}_{-0.10}^{0.10}$

Note. Here, as a simple model for the circular orbit, the satellite will orbit the Moon with a period of 8248.7 s (∼2.3 hr).

Download table as:  ASCIITypeset image

Table 4. The Best-fit Parameters of the 21 cm Global Signal and the rms Values for the Two Gaussian Models

21 cm model tobs [orbit]rms [K] A ν0 ω
Input0.2075.006.00
Recovered11.149 ${0.201}_{-0.015}^{0.015}$ ${74.85}_{-0.52}^{0.49}$ ${6.11}_{-0.55}^{0.61}$
 20.095 ${0.200}_{-0.010}^{0.010}$ ${75.04}_{-0.36}^{0.34}$ ${5.93}_{-0.39}^{0.38}$
 50.074 ${0.2000}_{-0.0067}^{0.0067}$ ${74.90}_{-0.23}^{0.22}$ ${5.90}_{-0.24}^{0.26}$
 100.041 ${0.2020}_{-0.0049}^{0.0048}$ ${74.91}_{-0.17}^{0.16}$ ${6.04}_{-0.18}^{0.20}$
Input0.15575.0010.00
Recovered10.117 ${0.170}_{-0.026}^{0.032}$ ${75.3}_{-1.4}^{1.2}$ ${10.2}_{-1.5}^{1.6}$
 20.092 ${0.160}_{-0.020}^{0.017}$ ${75.27}_{0.85}^{-0.94}$ ${9.6}_{-1.0}^{1.1}$
 50.060 ${0.162}_{-0.012}^{0.014}$ ${75.11}_{-0.67}^{0.62}$ ${10.22}_{-0.70}^{0.76}$
 100.047 ${0.1582}_{-0.0082}^{0.0089}$ ${75.07}_{-0.44}^{0.42}$ ${10.06}_{-0.48}^{0.50}$

Download table as:  ASCIITypeset image

3.3. Chromatic Beam Effect

To explore how the chromatic beam affects the global spectrum measurement, we now consider some nonideal cases where the beam is more frequency-dependent. To be concrete, we still consider disk-cone antennas but with different disk diameters (D3 in Figure 6). However, we find that as long as the size of the antenna is smaller than the half-wavelength, the beam is pretty similar to that of the short dipole and the frequency dependence is weak. This is good news for the design of the antenna for the space experiment. However, for this illustration, we decide to employ two "villain" models, which are intentionally designed to have a larger frequency variation of the beam. These are realized by keeping the cone part unchanged, while increasing the disk to a size comparable with the half-wavelength. In the first case, we increase the disk to 200 cm from the 55 cm in the fiducial model while keeping the cone unchanged. In the second case, the disk size is increased to 300 cm. One can obviously make antenna less frequency dependent, but we use them as illustrative examples because they are derived from electromagnetic simulations of an antenna configuration.

Figure 15 shows the beam profile of these two villain models in the top row and the corresponding logarithmic frequency gradient of the beam in the bottom row. Comparing with Figure 7, in these two cases, the beams vary with frequency much more significantly, and the logarithmic frequency gradient is also much larger. When convoluted with the spatial variations of the foreground, this may generate fake features in the global spectrum. Furthermore, in the model with a 300 cm disk (right panels of Figure 15), an additional sidelobe appears. The presence of a sidelobe could induce even larger features in the global spectrum from the sky variation, as we will find out below.

Figure 15.

Figure 15. The cross sections of beam profiles at a constant ϕ for the two villain models at three frequencies (top), and the corresponding frequency gradients at various θ and frequencies (bottom). The left panels (a), (c) are for a disk-cone antenna with a 200 cm disk, and the right panels (b), (d) are for a disk-cone antenna with a 300 cm disk.

Standard image High-resolution image

To examine how this affects the measurement of the global spectrum, we simulate the observation of a total of 10 orbits with the foreground-only model without noise. Here for simplification we have used the data from the full orbit but ignored any possible RFI.

In Figure 16, we plot the residuals after fitting and subtracting a five-term LogPoly model for our fiducial model, as well as for the two villain models. The residuals for the two villain models are about two orders of magnitude higher than our fiducial model, so we multiply a factor of 100 to the fiducial model residue to show them on the same plot. The rms of the residuals of our fiducial model without adding the noise is about 10 mK.

Figure 16.

Figure 16. Residuals after fitting and subtracting a five-term LogPoly model in frequency from the beam-weighted signal. The solid, dashed, dotted lines correspond to results for D = 55 cm, 200 cm, and 300 cm antennas, respectively. The residuals for the D = 55 cm case is multiplied by a factor of 100 so that they can be plotted in the same figure.

Standard image High-resolution image

As was shown in Figure 7, νB/∂ν ∼ 10−2 for our fiducial antenna beam model, and as shown in Figure 10 the amplitude of the global spectrum varies at 10% level. We may then expect the chromatic beam may introduce a deviation at the 10−3 level over the observation band, or ∼1 K for the 103 ∼ 104 K sky temperature of this band. However, the variation in the spectral shape or spectral index induced by the varying position/orientation is much smaller. As a result, this deviation varies smoothly over the frequencies, and so when we fit with the five-term LogPoly model, the bulk of it is removed. In the end, we obtain a residual with an rms of only 10 mK. This is still a very substantial and important source of error for the global spectrum measurement.

The two villain models are both exaggerated bad cases, but these illustrate how the beam may vary with frequency over the observational band. Even for better designed cases, the frequency gradient of the beam may have a pattern similar to them, though with smaller gradient values. In particular, we note that the additional sidelobe may greatly increase the error. Thus, as shown in Figure 16, the disk-cone antenna with a 300 cm disk, which has an additional sidelobe in its beam, has much larger rms residuals than the one with a 200 cm disk.

It is likely that the real antenna has a beam more like that of our fiducial model than that of the two villain models shown above. Nevertheless, there may be frequency-dependent deviations due to imperfect manufacturing or additional components necessary for a satellite, which will modify the antenna beam pattern. Such deviations could adversely affect the performance of the global spectrum measurement. It is critical to minimize such effects for the experiment to succeed.

4. Conclusion and Discussions

In this work, we have presented the end-to-end simulation results for a global 21 cm spectrum experiment on a lunar orbit operating in the 30–120 MHz frequency band. By adopting the specific parameters of the DSL mission concept, we assessed the feasibility of obtaining the 21 cm global spectrum from a strong foreground. A variety of practical problems, such as the Moon blockage and antenna beam (including the chromatic beam effect), have been taken into account.

We first generate an input sky map with ISM free–free absorption effect and make a mock observation of the global spectrum. We find the absorption is significant at lower frequencies, which affect the fitting of the foreground model. We simulate the observation at different positions on the lunar orbit. As the Moon blocks different parts of the sky at different positions and the antenna orientation also changes with the position, there is variation in the measured sky signal. However, we find that although the intensity is changing in the received signal, the change in the spectrum shape (or spectral index) is much smaller. Adopting a five-term LogPoly model to fit the foreground, we find that the residual error drops to a level of ∼0.05 K within a short span of 10 orbit observations. The statistical noise would soon drop to a negligible level.

We then consider systematic errors, such as the effect of a chromatic beam. For our fiducial model with a disk-cone antenna with D = 55 cm for the disk, the beam chromaticity is actually very small. The variation of the gain is at the 10−4 to 10−3 level within the 30–120 MHz range and does not induce a significant effect on the recovered 21 cm signal. We also introduced two "villain models," which illustrated the effect of the chromatic beam with a poorly designed antenna, having a disk size of D = 200 cm or D = 300 cm, exceeding the half-wavelength. In this case, the beam chromaticity can introduce a lot of complexity into the result, especially if there is an additional sidelobe in the beam profile. In the real world, we expect the beam to be close to our fiducial case, perhaps with a slightly distorted profile and a little higher frequency gradient due to the imperfections, but not large enough to affect the final result. In this study, we have treated the Moon as a simple opaque sphere and ignored the thermal emission of the Moon and its reflection and diffraction of radio waves, which can be complicated due to its rugged terrain and varying composition. These problems will be further investigated in future studies.

Now it is feasible to conduct a new experiment toward the detection of the global 21 cm signal from cosmic dawn. The lunar orbit provides an unparalleled opportunity by allowing high-precision global spectrum measurements. Our simulations show that the 21 cm global signal could be recovered very well on a lunar orbit.

This work is supported by National SKA Program of China No. 2020SKA0110401, the Chinese Academy of Sciences (CAS) Strategic Priority Research Program XDA15020200, the CAS Key Instrument grant ZDKYYQ20200008, the National Natural Science Foundation of China (NSFC) grants 11973047 and 11633004, and the CAS Frontier Science Key Project QYZDJ-SSW-SLH017.

Footnotes

Please wait… references are loading.
10.3847/1538-4357/ac5965