Cosmic Reionization May Still Have Started Early and Ended Late: Confronting Early Onset with Cosmic Microwave Background Anisotropy and 21 cm Global Signals

and

Published 2021 June 14 © 2021. The American Astronomical Society. All rights reserved.
, , Citation Kyungjin Ahn and Paul R. Shapiro 2021 ApJ 914 44 DOI 10.3847/1538-4357/abf3bf

Download Article PDF
DownloadArticle ePub

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

0004-637X/914/1/44

Abstract

The global history of reionization was shaped by the relative amounts of starlight released by three halo mass groups: the first two groups are atomic-cooling halos (ACHs) with virial temperatures Tvir > 104 K, either (1) massive enough to form stars even after reionization (high-mass ACHs, ≳ 109 M) or (2) less massive (low-mass ACHs), subject to star formation suppression when overtaken by reionization, and the third group comprises (3) H2-cooling mini-halos (MHs) with Tvir < 104 K, whose star formation is predominantly suppressed by the H2-dissociating Lyman–Werner background. Our previous work showed that including MHs caused two-stage reionization—early rise to x ≲ 0.1, driven by MHs, followed by a rapid rise, late, to x ∼ 1, driven by ACHs—with a signature in cosmic microwave background (CMB) polarization anisotropy predicted to be detectable by the Planck satellite. Motivated by this prediction, we model global reionization semi-analytically for comparison with Planck CMB data and the Experiment to Detect the Global Epoch of Reionization (EDGES) global 21 cm absorption feature, for models with: (1) ACHs, no feedback; (2) ACHs, self-regulated; and (3) ACHs and MHs, self-regulated. Model (3) agrees well with Planck E-mode polarization data, even with a substantial tail of high-redshift ionization, beyond the limit proposed by the Planck Collaboration. No model reproduces the EDGES feature. For model (3), $\left|\delta {T}_{b}\right|\lesssim 60\,\mathrm{mK}$ across the EDGES trough, an order of magnitude too shallow, and absorption starts at higher z but is spectrally featureless. Early onset reionization by Population III stars in MHs is compatible with current constraints, but only if the EDGES interpretation is discounted, or else other processes we did not include account for it.

Export citation and abstract BibTeX RIS

1. Introduction

Cosmic reionization is commonly believed to have commenced with the birth of the first stars in the universe and ended when all of the intergalactic medium (IGM) became ionized due to the net production of ionizing photons surpassing the number of neutral hydrogen atoms. This whole epoch marks the epoch of reionization (EoR) in the history of the universe. There exist several observational constraints, including (1) the quasar Gunn–Peterson trough (Becker et al. 2001; Fan et al. 2002), in which reionization is likely to have ended at z ≳ 6, (2) the temporal evolution of the 21 cm background at redshift z ∼ 6 (Bowman & Rogers 2010), in which a sudden reionization scenario is ruled out, (3) the polarization anisotropy of the cosmic microwave background (CMB; Planck Collaboration et al. 2020), wherein the optical depth to the CMB photon, τes, is ${0.054}_{-0.0081}^{+0.0070}$(1σ), (4) the sudden change of the Lyα-emitter (LAE) population (Pentericci et al. 2011), in which the global ionized fraction, x, at z ≃ 7 is ≳0.1, and (5) the quasar proximity effect (Calverley et al. 2011), for which the metagalactic ionization rate Γ at z ≃ 6 is about 10−13 s−1. These constraints are nevertheless insufficient to constrain the full history of reionization.

Constraining the cosmic reionization process by observations requires theoretical modeling, of course. The simplest approach is to calculate the history of reionization using semi-analytical, one-zone models (see, e.g., Haardt & Madau 1996; Haiman & Loeb 1997; Haiman & Holder 2003; Furlanetto 2006). The more difficult approach is to simulate the process numerically (see, e.g., Iliev et al. 2007; Kohler et al. 2007; Trac & Cen 2007; Mesinger et al. 2011). These methods are mutually complementary. For example, the former allows a very fast exploration of the parameter space, while the latter enables simulating the three-dimensional (3D) structure of the process. If one is interested only in the averaged quantities, such as the global ionized fraction x and the kinetic temperature of the IGM Tk , the one-zone calculation can provide surprisingly reliable estimates very easily. Therefore the one-zone calculation has been used quite extensively. This method has become very useful in reionization-parameter estimation efforts, while one has to be wary about the fact that such an estimation is limited to specific reionization models that are assumed to be fully described by those parameters.

Some of the semi-analytical one-zone models used to have common "conventions" including the following: (1) only atomic-cooling halos (ACHs), with virial temperature Tvir ≳ 104 K, are considered as radiation sources, (2) the star formation rate (SFR) is proportional to the growth rate dfcoll/dt of the halo-collapsed fraction fcoll, and (3) feedback effects are neglected (Loeb & Barkana 2001; Furlanetto 2006; Pritchard & Furlanetto 2006). We categorize those models that follow these conventions as the vanilla model. First, the main justification for the first convention is from the fact that stars inside less-massive halos, or mini-halos (MHs; with Tvir ≲ 104 K and mass limited down to the Jeans mass of the IGM), are susceptible to the Lyman–Werner (LW) feedback effect and the Jeans-mass filtering. The LW feedback strongly regulates the amount of H2 inside MHs, the main cooling agent in the primordial environment, by photodissociation. The Jeans-mass filtering can happen if MHs are exposed to the hydrogen-ionizing radiation field, either from inside or from outside. Therefore, their contribution was believed to be negligible. Second, justification for SFR ∝ dfcoll/dt is not solid, and this assumption is equivalent to a null duty cycle of star formation (Section 2.1). Third, neglecting the feedback effect is a conventional simplification, while there may exist ACHs in the low-mass end that are affected by the Jeans-mass filtering if embedded in photoionized regions (Section 2.2.2).

Many numerical simulations inherited the convention of neglecting MHs, usually in those using a large (≳ a few tens of co-moving megaparsecs) simulation box. While the justification is the strong LW feedback effect, a practical reason is the numerical resolution limit that becomes worse as the size of the simulation box becomes larger. Therefore, large-scale simulations of reionization usually suffer from this limit. To overcome this limit and include the impact of MH stars, a sub-grid treatment of missing halos (Ahn et al. 2015a; Nasirudin et al. 2020) was included in a large simulation box by Ahn et al. (2012). They used an empirical, deterministic bias of MH population for any given density environment to successfully populate MHs over the domain of calculation. Using theoretical predictions on the formation of the first stars inside MHs, together with the impact of the LW feedback, they could then cover the full dynamic range of halos. They found that (1) MH stars cannot finish reionization but can ionize the universe up to ∼20% depending on the mass of Population III stars, (2) the resulting reionization history is composed of the early staggered stage with slow growth of $\left\langle x\right\rangle $ and the late rapid stage with rapid growth of $\left\langle x\right\rangle $, (3) the reionization is finished by photons from ACHs, and (4) there exists degeneracy in reionization histories that can result in the same τes and zreion.

There exists an interesting hint from the recent large-scale CMB polarization observation that is related to the result by Ahn et al. (2012). Constraining the full history of reionization has become available only recently, through observation of the large-scale CMB polarization anisotropy. Having such constraints used to be impossible with the given quality of CMB polarization data before, and thus only a two-parameter constraint on reionization with τes and the reionization redshift zreion had been available (e.g., Hinshaw et al. 2013; Planck Collaboration et al. 2014). The Planck 2015 data (Planck Collaboration et al. 2016) first allowed constraining the history of reionization beyond this two-parameter constraint, and it was claimed that the Planck data favored a type of reionization history composed of (1) the early, slow growth of x for a large range of redshift z ∼ 30–10 and (2) the late, rapid growth of x for a short range of redshift z ∼ 10–6 (Miranda et al. 2017; Heinrich & Hu 2018). While criticisms to this claim appeared due to the limitation of having only the Planck Low Frequency Instrument (LFI) data (Millea & Bouchet 2018), this "two-stage" reionization history had been indeed predicted by Ahn et al. (2012). They showed that the early stage could be dominated by strongly self-regulated formation of stars inside MHs, presumably Population III stars, and the late stage by stars inside ACHs, with weaker (inside low-mass ACHs, hereafter LMACHs) or even no modulation (inside high-mass ACHs, hereafter HMACHs) on star formation. The most refined, nonparametric constraint comes from the Planck 2018 data that also includes the High Frequency Instrument (HFI) data (Millea & Bouchet 2018; Planck Collaboration et al. 2020), and the constraint bears qualitative similarity with but some quantitative difference from the analyses by Miranda et al. (2017) and Heinrich & Hu (2018). First, the two-stage reionization is still favored at the ≲2σ level. Second, a "significant" amount of ionization at z ≳ 15 is disfavored, with a maximum of ∼10% ionized fraction allowed at z ∼ 15 at the 1σ level. Even though Planck Collaboration et al. (2020) and Millea & Bouchet (2018) stress the latter finding and even claim that the early stage of reionization dominated by self-regulated Population III stars is strongly disfavored, their constraint indeed allows for very extended ionization histories reaching to z ∼ 25 and z ∼ 30 for the 1σ and the 2σ constraints, respectively, in addition to showing a two-stage reionization feature. Therefore, we take the analysis on the Planck 2018 data (Millea & Bouchet 2018; Planck Collaboration et al. 2020) as a mild proof for the two-stage reionization, which will be tested quantitatively in this paper.

Meanwhile, observing the hydrogen 21 cm line background is believed to provide a direct probe of the dark ages and the EoR. The deepest (z = 14–26) observation so far, in terms of the sky-averaged global 21 cm background, has been delivered by the Experiment to Detect the Global Epoch of Reionization (EDGES), and the detection of ∼500 mK continuum absorption of the hydrogen 21-cm signal against the smooth background around z ∼ 17 was claimed (Bowman et al. 2018). It is hard to grasp such a large absorption in the standard picture, compared to the maximum ∼200 mK depth allowed in the Lambda cold dark matter (ΛCDM) universe. The overall shape of the signal over the redshift range is also incompatible with usual model predictions. This "conflict" stimulated many resolutions, which fall into roughly three categories: (1) the analysis of removing a smooth galactic foreground and obtaining another smooth absorption signal is unreliable (Hills et al. 2018), (2) nonstandard models beyond ΛCDM should be considered, such as large baryon–dark matter interaction (Tashiro et al. 2014; Barkana 2018) and an abnormally large expansion rate at high redshift (Hill & Baxter 2018), and (3) the contribution of yet unknown sources to the radio continuum (the "excess radio background") but still in the standard ΛCDM framework, may generate such a large absorption signal (e.g., Ewall-Wice et al. 2018; Feng & Holder 2018).

Motivated by the hint on the two-stage reionization from the Planck 2018 data and the compelling EDGES observation, we explore the possibility of the two-stage reionization leaving any characteristic signature in the CMB and the global 21 cm background. Toward this end, we use semi-analytical one-zone models to cover a wide range of reionization scenarios. We cast reionization models into three categories, from the vanilla model without feedback effect to progressively sophisticated ones, expanding the star-hosting halo species and considering feedback effects that regulate star formation in those halos. We then investigate how the history of reionization and the evolution of other relevant radiation fields shape the CMB the 21 cm background, with a special focus on finding whether or not the two-phase reionization models that carry the high-redshift ionization tail, constructed by the LW-regulated star formation in MHs, will leave any imprint on the CMB and the 21 cm background. While it is worthwhile to include the resolutions, alternative to the standard picture, for the deep absorption signal of EDGES in the parameter estimation effort (Mirocha et al. 2018; Mirocha & Furlanetto 2019; Mebane et al. 2020; Qin et al. 2020a, 2021), we limit our study to the standard ΛCDM framework without the excess radio background and other alternatives but instead explore any new possible observational predictions. We indeed find very unique and novel features of the two-phase ionization models both in the CMB and the global δ Tb, opening up exciting observational prospects.

The paper is organized as follows. In Section 2, we describe the details of the one-zone model, including the model categories and the numerical method. In Section 3, we present the estimates on the ionized fraction, the CMB polarization anisotropy, and the 21 cm background for various reionization models, with a special focus on the estimate from the strongly self-regulated models. We summarize and discuss the result in Section 4. Throughout this paper, we use "cMpc" to denote "co-moving megaparsec," or the co-moving length in units of megaparsecs.

2. Method

We take a simple one-zone model for studying the global evolution of the physical states of IGM and backgrounds including the 21 cm signal. Two crucial parameters are x and Tk , whose evolutions will be governed by physical properties, the evolution, and the feedback of radiation sources. Therefore, the main equations for reionization models are the rate equations for x and Tk .

The ionizing-photon production rate per baryon (PPR) is the source term that increases x, or the source term in the rate equation dx/dt where t is the cosmic time. PPR is the main parameter whose variance leads to the variance in the histories of reionization among different models. PPR during EoR is quite uncertain and will be parameterized by the species of host halos, spectral shape, and the star formation duty cycle. PPR is also affected by various feedback effects. These will be the main subjects to be described in Sections 2.12.2.

The recombination rate is the sink term in the ionization equation. The recombination rate does not have a model variance as large as that of PPR. Nevertheless, there exists some uncertainty in the recombination rate mainly due to the difficulty in quantifying the clumping factor. We conservatively take a combination of previous studies that quantified the clumping factor.

A few radiation background fields are crucial in determining Tk and the 21 cm spin temperature Ts . Variance in Tk is caused by variance in the heating efficiency, and variance in Ts by variance in Lyman-resonance-line backgrounds. Therefore, the model variance will be reflected also in the 21 cm background, which will be described in Sections 2.32.4. We defer the description of the used spectral energy distributions (SEDs) of the stellar radiation, which cause the variance in Ts even at a similar level of the ionization state, to Sections 3.1 and 3.3.

Throughout this paper, we use cosmological parameters reported by Planck Collaboration et al. (2020): h = 0.6732, Ωm,0 = 0.3144, Ωb,0 = 0.04939, ns = 0.966, and σ8 = 0.812, which are the present Hubble parameter in units of 100 km s−1 Mpc−1, the present matter density in units of the critical density, the present baryon density in units of the critical density, the power index of the primordial curvature perturbation, and the present variance of matter density at the filtering scale of 8h−1 Mpc. We use the mass function by Sheth & Tormen (1999) when calculating the halo-collapsed fraction.

2.1. Star Formation Rate Density and Duty Cycle

It is a usual approximation that PPR is proportional to the star formation rate density (SFRD: SFR per co-moving volume, in units of M s−1 cMpc−3). Rigorously, this becomes true only if t*, the lifetime of stars that we take as a constant for a given stellar species (e.g., Population II stars), is infinitesimal compared to the ionization time, i.e., x/(dx/dt). Even more rigorously, for PPR at time t to be the ionization rate at t, the photon travel time inside H ii regions should also be very small. In this paper, we take this usual assumption, PPR ∝ SFRD, which has been used extensively in semi-analytical modeling. The ionization rate ξ (in units of s−1) will then be given by

Equation (1)

where fesc, Nion, mb , and nb,0 are the ionizing-photon escape fraction out of halos, the total number of ionizing photons emitted per stellar baryon during the lifetime of a star, the baryon mass, and the average co-moving baryon number density, respectively.

We also modify the typical assumption of semi-analytical calculations to incorporate a nonzero duty cycle of star formation. Let us first briefly review the typical assumption in semi-analytical calculations: PPR and SFRD are assumed proportional to dfcoll/dt, the growth rate of the halo-collapsed fraction fcoll, such that

Equation (2)

where ξa denotes the ionization rate due to newly "accreted" matter, f* is the star formation efficiency in such an episode (e.g., Loeb & Barkana 2001; Furlanetto 2006), and fγ f* fesc Nion is the proportionality coefficient between ξ and dfcoll/dt. The second equality in Equation (2) is a rigorous form that considers only the newly accreted matter, whose maximal lookback time in the integration should be t*. Therefore, we can denote this assumption as the "mass-accretion dominated star formation" scenario (MADSF). In the limit of infinitesimal t* or slowly varying fcoll, the last approximation becomes valid. Because dfcoll in Equation (2) is the amount of matter that has been newly accreted into halos per time interval dt, considering ξa(t) only is identical to having a null duty cycle such that any gas that has once formed stars can never form stars again. Note also that dfcoll does not include halos whose masses have increased by a merger event at t, even though in nature "wet merger" events occur quite frequently. Because this null duty cycle assumption is somewhat extreme, we need to consider a more general case of nonzero duty cycle.

We quantify the nonzero duty cycle as the time fraction that stars emit radiation, fDCt*/(t* + tdorm), where tdorm is the duration that a stellar baryon remains dormant after the death of the host star until a new star formation episode occurs. The same account on tdorm has been addressed by Mirocha et al. (2018). Both t* and tdorm are average quantities, and consequently fDC will be a global parameter. Ergodicity is assumed such that this temporal duty cycle is the same as the spatially averaged fraction of gas that has once resided inside a star and now resides in a post-generation star. In this work, we further assume that fDC is constant over time for a given stellar species. With nonzero fDC, we obtain the ionization rate ξd due to "duty cycle":

Equation (3)

where only "old" gas at lookback time t* is allowed to regenerate stars, and the approximation holds when t* is treated as infinitesimal. The last equality absorbs fDC into the effective f* of halo gas, f*,d. Equation (3) obviously does not account for any newly accreted gas. Note also that Equation (3) holds due to ergodicity: f*,d is a spatially averaged quantity over many galaxies, which will be identical to the time average of star formation episodes on a single galaxy. We denote a scenario based on ξ = ξd as the "all-halo replenished star formation" scenario (AHRSF).

The most generic form for ξ should implement both contributions from old gas and newly accreted gas, or combining the episodes of MADSF and AHRSF, such that

Equation (4)

where f* of newly accreted gas is denoted by f*,a, to be distinguished from f*,d. The condition f*,a = 0, corresponds to the case where newly accreted gas waits longer than t* to start star formation activity, and practically there is no theoretical constraint on the relative strengths of f*,a over f*,d. With Equation (4), both cases of ξfcoll/t* and ξdfcoll/dt can be accommodated in terms of special cases of this generic form with {f*,d ≠ 0, f*,a = 0} and {f*,d = 0, f*,a ≠ 0}, respectively. Usually, the relation ξfcoll/t* has been used in numerical radiation transfer simulations (e.g., Iliev et al. 2007) and the relation ξdfcoll/dt in semi-analytical calculations (e.g., Loeb & Barkana 2001; Furlanetto 2006). In this paper, we only consider these two special cases, denoting the former by "F" and the latter by "dF" as the nomenclature for star formation scenarios. Because we use f*,a and f*,d mutually exclusively for these special cases, we drop the subscripts "a" (accretion) and "d" (duty cycle) from f* for simplicity.

As long as cosmology is fixed, the evolution of the volume ionized fraction will be uniquely determined for a given fγ in MADSF and for a given gγ f* fesc Nion/(t*/10 Myr) in AHRSF, because dfcoll/dt and fcoll are determined solely by cosmological parameters. Therefore, reionization models with MADSF and AHRSF will be parameterized by fγ (e.g., Furlanetto 2006) and gγ (e.g., Ahn et al. 2012), respectively. Of course, fγ and gγ need not be constant in time. fesc, f*, t*, and Nion can be time-varying individually and collectively (in terms of fγ and gγ ). Allowing fγ and gγ to change in time can yield even more variants of reionization scenarios, especially by reshaping dx/dz. For simplicity, we do not explore this possibility. Nevertheless, future CMB observations will become more accurate, and models inferred from the data could not be matched well by the constant values of fγ and gγ . In such a case, time-varying fγ and gγ may have to be considered in our models.

2.2. Model Variance

We consider three types of models: (1) the vanilla model, (2) the self-regulated model type I (SRI), and (3) the self-regulated model type II (SRII). From type (1) to type (3), these models become progressively sophisticated in the physical processes considered: the vanilla model does not consider any feedback, SRI considers the photo-heating feedback, and SRII considers both photo-heating and LW feedback effects. In each model, we allow both cases of ξdfcoll/dt and ξfcoll/t* that were described in Section 2.1. In all of these models, the source term is the ionization rate, and the sink term is the recombination rate, such that the change rate of the global (volume) ionization fraction x is

Equation (5)

where α is the hydrogen recombination coefficient (in units of cm3 s−1), C is the average clumping factor, and ne = nH + nHe is the (proper) number density of electrons inside H ii regions (e.g., Furlanetto 2006; Iliev et al. 2007). In this work, we use the case B recombination coefficient for α, or α = αB(T = 104 K) = 2 × 10−13 cm3 s−1, and adopt a fitting formula for C given by

Equation (6)

which is a conservative combination of work by Iliev et al. (2005), Pawlik et al. (2009), and So et al. (2014). While it is possible that the numerical resolution limit of previous numerical simulations may have led to underestimation of C (Mao et al. 2020), we do not consider this possibility in this paper. The model variance is mainly caused by the variance in ξ in Equation (5), which will be described in the following subsections.

Physical parameters governing ξ, such as fesc, f*, Nion, etc., would depend on halo properties. One of the crucial halo properties is the virial temperature Tvir. The natural borderline between MHs and ACHs is the temperature endpoint of the atomic line cooling (if dominated by hydrogen Lyα line cooling), ∼ 104 K. ACHs have Tvir > 104 K, and MHs have Tvir < 104 K. This classification can be cast into the mass criterion

Equation (7)

where we assume the mean molecular weight μ = 0.6 for fully ionized gas (Barkana & Loeb 2001). One important subtlety is that this redshift-dependent mass criterion is not commonly respected in numerical radiative transfer (RT) simulations, because the minimum mass of halos resolvable in accompanying N-body or uniform-grid simulations tends to be constant in time. 3 Therefore, many RT simulations take a constant-mass classification scheme, such that ACHs and MHs correspond to halos with M > Mth and M < Mth, respectively, with a constant-mass threshold Mth. When a simulation domain is increased to a few ∼10 cMpc, the numerical resolution limit quickly reaches ∼108 M, a common value of Mth for such classification. For example, in many large-volume simulations, MHs are ignored, and all halos that are numerically resolved, or, e.g., those with M > 108 M, are taken as ACHs. However, this scheme even loses track of those halos with Tvir > 104 K and M < 108 M at z ≳ 18. Therefore, one should be wary of the negligence when the high-redshift (z ≳ 20) astrophysics, for example, the early stage of cosmic reionization, is investigated with such a scheme. It may ignore not only MHs but also a substantial amount of ACHs, as long as we believe in the constant-temperature criterion as a natural distinction.

In order to have a fair comparison of the semi-analytical models to the previous numerical RT simulation results based on this constant-mass criterion for ACH/MH classification, we adopt a constant Mth (=108 M) in this work. Comparison will be made to numerical simulations that (1) further split ACHs into low-mass and high-mass species with a constant-mass classification (Section 2.2.2), and (2) cover the full dynamic range of halos, namely MHs, LMACHs, and HMACHs but again with a constant-mass classification (Section 2.2.3).

2.2.1. Vanilla Model

This model assumes a very simplified form for PPR: PPR is typically assumed to be proportional to dfcoll/dt, and no feedback effect on star formation is considered. The essential ingredient is not the relation ξdfcoll/dt but the lack of feedback, and thus we also allow the relation ξfcoll/t*. We therefore have

Equation (8)

for ξdfcoll/dt ("dF") and

Equation (9)

for ξfcoll/t* ("F"). These equations are generalized from Equations (2) and (3) with summation to accommodate different halo species i, with the superscript "(i)" denoting physical quantities of the halo species i. The simplicity of Equations (8) and (9) is the essence of vanilla models that ignore any feedback effects.

2.2.2. Self-regulated Model Type I

The work by Iliev et al. (2007) is among the first 3D RT simulations of self-regulated reionization based on multispecies halo stars, but with radiation sources restricted to ACHs. This simulation starts with classifying ACHs into two mass categories, namely (1) the HMACHs and (2) the LMACHs, with M ≳ 109 M and 108 MM ≲ 109 M, respectively. Again, even though it is more natural to classify halos in terms of Tvir, to make a direct comparison to Iliev et al. (2007), we adopt this convention in this work. While the accurate boundary does not exist, LMACHs defined this way roughly correspond to those halos that are subject to the "Jeans-mass filtering": if these halos are formed inside regions that have been already ionized, accretion of baryonic gas will not be efficient enough to form stars inside due to the high temperature (T ≳ 104 K) of the regions (Efstathiou 1992; Shapiro et al. 1994; Thoul & Weinberg 1996; Navarro & Steinmetz 1997; Gnedin & Hui 1998; Gnedin 2000; Dijkstra et al. 2004). The suppression of star formation is likely to be not as abrupt as assumed in Iliev et al. (2007) but rather gradual in halo mass (Efstathiou 1992; Navarro & Steinmetz 1997; Dijkstra et al. 2004). Nevertheless, in this work we simply adopt the self-regulation scheme of Iliev et al. (2007).

We denote such a type as SRI. Because HMACHs are unaffected by the feedback and LMACHs are suppressed inside H ii regions, the ionization rate will be given by

Equation (10)

for ξdfcoll/dt ("dF") and

Equation (11)

for ξfcoll/t* ("F"), where superscripts "H" and "L" denote HMACH and LMACH, respectively, and 0 < η < 1. The reason why such an amplified suppression term (1 − xη ) is used instead of (1 − x) is that LMACHs are clustered more strongly inside H ii regions than in neutral regions. The value η = 0.1 is the empirical one found in 3D numerical simulations by Iliev et al. (2007), such that xη roughly estimates the mass fraction of LMACHs occupied by H ii regions when the global ionization fraction is x. The effect of such a small value of η is to make star formation in LMACHs, and consequently the reionization history, more strongly regulated than the unrealistic case with η = 1 where LMACHs are uniformly distributed in space. The assumption for the Jeans-mass filtering in Equation (11) is that even those halos that have collapsed earlier cannot host new star formation episodes. Even though one can generalize ξ into a combined form of Equations (10) and (11), as we mentioned in Section 2.1, we restrict our models to these two categories of "dF" and "F."

2.2.3. Self-regulated Model Type II

The work by Ahn et al. (2012) is a unique 3D RT simulation of reionization in that (1) a full dynamic range of halos, from MHs to HMACHs, is treated as radiation sources, (2) both the Jeans-mass filtering and the LW feedback are considered, and (3) the simulation box is large (∼150 Mpc) enough to provide a reliable statistical significance. This simulation does not neglect MHs as most other large-box simulations do. MHs are subject both to the Jeans-mass filtering and the LW feedback. Jeans-mass filtering of MHs is obvious due to the smallness of the the virial temperature (T ≲ 104 K). The LW feedback occurs due to the fact that H2 is the main cooling agent in the primordial environment, and MHs are usually formed first in the primordial environment. Stars born in this environment will be Population III stars. Even inside MHs, after a few episodes of star formation, the chemical environment can gain metallicity beyond the critical value Z ≃ 10−3. However, dynamical feedback from supernova explosion inside MHs is believed to be very destructive (Greif et al. 2007; Yoshida et al. 2007), such that it may take longer than, e.g., the halo merger time for post-generation star formation episodes to occur in the same MH. It is possible that the supernova feedback in the massive MHs, which remained neutral even after photoionization from stars inside, could have been confined inside the halo and led to the next episode of star formation (Whalen et al. 2008). However, the dominant contribution to the number of MHs is from the least massive ones, and so it is appropriate to assume the destructive feedback. If one assumed the most destructive feedback effect of the first episode of star formation inside MHs, then it would be equivalent to assuming that only the newly forming MHs form stars. This assumption was taken in Ahn et al. (2012), which we implement in our modeling here as well.

We denote such a type as SRII. The ionization rate will be given by

Equation (12)

for ξdfcoll/dt ("dF") and

Equation (13)

for ξfcoll/t* ("F"), where the superscript M denotes MH, MIII is the mass of Population III stars per MH, nM is the co-moving number density of MHs, nb,0 ≃ 2 × 10−7 cm−3 is the co-moving number density of baryons, μ = 1.22 is the mean molecular weight of MH gas, and xLW is the LW intensity JLW normalized by the threshold intensity JLW,th given by

Equation (14)

to implement the suppression of star formation inside MHs (see Equations (12) and (13)) in a similar fashion with Ahn et al. (2012).

Because we only take newly forming MHs as sources, the last terms in Equations (12) and (13) are identical. Having $\xi \propto {{dn}}^{{\rm{M}}}/{dt}$ instead of $\xi \propto {{df}}_{\mathrm{coll}}^{{\rm{M}}}/{dt}$ for MHs is to accommodate the tendency found in numerical simulations of Population III star formation: Population III stars under the primordial environment will form mostly in isolation (Abel et al. 2000; Bromm et al. 2002) or as a few binary systems at most (Turk et al. 2009; Stacy et al. 2010), and the total mass of Population III stars in MHs is determined by the atomic physics (Hirano et al. 2014, 2015) and is not strongly dependent on the mass of MHs. The actual mass of Population III stars can vary substantially according to physical properties, such as the angular momentum, the local LW intensity, and the mass accretion rate of their host MHs (Hirano et al. 2014, 2015), and thus MIII should be taken as the average mass of Population III stars per MH.

We limit the MH mass to 105 MM ≤ 108 and use the Sheth–Tormen mass function (Sheth & Tormen 1999) based on the typical linear matter density perturbation obtained from the linear Boltzmann solver CAMB (Lewis et al. 2000) to calculate n M. Our choice of the minimum mass of MHs, ${M}_{\min }={10}^{5}\,{M}_{\odot }$, is also used in Ahn et al. (2012) and is indeed reasonable due to the following reasons. This value is about half of the Jeans mass of the neutral IGM when the baryon–dark matter streaming velocity is considered (Tseliakhovich et al. 2011). The actual minimum mass to host Population III stars could be somewhat larger than this Jeans mass and also redshift-dependent (see, e.g., Figure 1 of Glover 2013; Hirano et al. 2015). Even though some claim that ${M}_{\min }\gtrsim $ a few 106 M and thus the impact of MH stars on cosmic reionization is negligible (e.g., Kimm et al. 2017), other high-resolution numerical simulations (e.g., Hirano et al. 2015) find that massive Population III stars are hosted mostly by halos in the mass range ${M}_{\min }\simeq [{10}^{5},{10}^{6}]\,{M}_{\odot }$.

We also note that the ionization history x(t) would depend on ${M}_{\min }$ (or similarly on MIII) more weakly than JLW,th and fesc, and therefore determining an accurate value of ${M}_{\min }$ would not be too crucial. This is because star formation in MHs, during the time when MH stars are the dominant radiation sources, is regulated in a way to maintain JLWJLW,th (Ahn et al. 2012; see also Section 3.1). If ${M}_{\min }$ had been larger than our fiducial value and thus nM had been smaller, then MH stars would have produced less ionizing and LW radiation in the beginning and drive resulting suppression weaker (or (1 − xLW) larger). In this case, because of reduced suppression, MH star formation will soon be expedited until JLW reaches JLW,th and produce an x(t) evolution similar to the fiducial case thereafter. Similarly, any additional change in nM due to the baryon–dark matter streaming effect is likely to be unimportant in determining x(t).

The way we implement the LW feedback as a multiplicative factor (1 − xLW) in Equations (12) and (13) roughly follows the work of Yoshida et al. (2003) and O'Shea & Norman (2008), where they find a gradual increase in ${M}_{\min }$, the minimum mass of halos that can form stars, as JLW increases if JLW < JLW,th. O'Shea & Norman (2008) also find that when JLWJLW,th ∼ 0.1 × 10−21 erg s−1 cm−2 Hz−1 sr−1, ${M}_{\min }$ and ${T}_{\mathrm{vir},\min }$ (the minimum virial temperature of halos that can form stars) jump to those of ACHs such that star formation in MHs are fully suppressed. Yoshida et al. (2003) find practically the same result, namely the full suppression of star formation inside MHs when JLWJLW,th ∼ 0.1 × 10−21 erg s−1 cm−2 Hz−1 sr−1 based on a combination of semi-analytical analysis and numerical simulation. Therefore, the exact functional form of the suppression is not important once JLW reaches JLW,th and the increasing number of MHs afterward try to produce more photons but fail to do so due to the self-regulation. As long as the condition that MHs become fully devoid of star formation when the condition JLWJLW,th is met, the formalism will correctly predict the self-regulation by LW feedback. This is exactly how the LW feedback is implemented in Equations (12)–(14). Other work (e.g., Mirocha et al. 2018; Qin et al. 2021) implementing the LW feedback on MH stars and forecasting the 21 cm background does not usually take this approach, which will be discussed in detail in Section 3.3.

The self-regulation of MH stars is predominantly governed by the LW feedback. In all of the SRII models we tested (see the detailed model parameters in Section 3.1), xLW > x, and thus the regulation factor $(1-\min \{\max ({x}_{\mathrm{LW}},x),1\})$ in Equations (12) and (13) is practically identical to $(1-\min \{{x}_{\mathrm{LW}},1\})$. We also note that we do not use biased suppression of star formation for MH stars, as quantified by η for LMACHs in Equations (10)–(13). This is because star formation inside MHs is likely to occur much more diffusively than that inside ACHs. This tendency is indeed observed in the numerical simulation by Ahn et al. (2012): ionized regions generated by MH stars are almost uniformly spread in space, in contrast to those by ACH stars (Figure 2 of Ahn et al. 2012). Therefore, we simply take an unbiased regulation factor $(1-\min \{{x}_{\mathrm{LW}},1\})$ for MH stars.

We take ${N}_{\mathrm{ion}}^{{\rm{M}}}=$ 50,000, which is a value suitable for very massive stars (M* ≳ 100 M). As was noted in Fialkov et al. (2013), quantifying suppression of SFR by LW intensity is not very straightforward when JLW < JLW,th, if, e.g., one considers the temporal evolution of LW intensity during halo formation. Our suppression scheme described by Equations (12)–(14) could instead perfectly mimic the full suppression by any threshold intensity JLW,th, and our ignorance of any other details is parameterized by the value of JLW,th. How JLW is evaluated is described in Section 2.3.2.

2.3. Background Radiation and Feedback

There are a few radiation backgrounds that determine the reionization history and the 21 cm background. Any unprocessed background intensity (erg s−1 cm−2 Hz−1 sr−1) at observing frequency ν and redshift z is given by

Equation (15)

where h is the Planck constant, ${{ \mathcal N }}_{\nu ^{\prime} }(z^{\prime} )$ is the photon-number luminosity density (s−1 Hz−1 cMpc−3) at source frequency $\nu ^{\prime} $ and redshift $z^{\prime} $, and τν is the optical depth from $z^{\prime} $ to z at ν. The soft-UV, LW, and X-ray backgrounds are all unprocessed types and thus given by Equation (15). The Lyα background is a processed type, and thus is not given by Equation (15); an appropriate description will be given instead in Section 2.3.4.

2.3.1. H-ionization by UV Background

Ionization of IGM by the UV background is believed to be very inhomogeneous, and the corresponding "patchy reionization" scenario is widely accepted. Unless a rather extreme scenario of X-ray-dominated reionization is assumed, patchy reionization will naturally occur in the universe. Well-defined H ii regions, almost fully ionized inside and connecting sharply with neutral IGM outside, will be created by UV sources in patchy reionization. In such scenarios, it is difficult to specify the background UV intensity by Equation (15) in our one-zone model.

The UV background is usually quantified in terms of the metagalactic H-ionizing rate per baryon, commonly denoted by Γ (in units of s−1), which is currently well constrained for the post-reionization epoch (e.g., Bolton & Haehnelt 2007; Calverley et al. 2011). Γ is a quantity that is determined after H-ionizing photons emitted from galaxies (and quasars) are filtered and reprocessed as they propagate through the IGM and dense gas clumps. Because we do not accurately model the clumping factor and we use practically a one-zone model, it is difficult to calculate Γ that is a processed quantity linked to IGM properties such as the photon mean free path λmfp. Instead, we can use a more transparent quantity, the UV-photon emissivity, which is blind to any physical properties of the IGM. The UV-photon emissivity is defined as the H-ionizing-photon production rate per co-moving volume (in units of s−1 cMpc−3), or equivalently ${\dot{N}}_{\mathrm{ion}}\equiv \xi {n}_{b,0}$, which can be linked to Γ by the relation ${\rm{\Gamma }}\propto {\lambda }_{\mathrm{mfp}}{\dot{N}}_{\mathrm{ion}}$ (e.g., Bolton & Haehnelt 2007 report ${\dot{N}}_{\mathrm{ion}}\simeq {10}^{50.5}\,{{\rm{s}}}^{-1}\,{\mathrm{cMpc}}^{-3}$ at z = 6). We simply check whether our models produce a reasonable value of ${\dot{N}}_{\mathrm{ion}}$ in Section 3.1.

2.3.2. H2-dissociation by Lyman–Werner Background

The frequency-averaged, global LW intensity is given by

Equation (16)

where ${\left\langle \right\rangle }_{\nu }$ and ${\left\langle \right\rangle }_{\nu ^{\prime} }$ are frequency-averages in the observed band at z and in the emitted band at $z^{\prime} $, respectively, $\nu ^{\prime} $ is limited below the Lyman limit (LL, $h\nu ^{\prime} \lt 13.6\,\mathrm{eV}$), and ${f}_{\mathrm{mod}}$ is the "picket-fence modulation factor" that accounts for the trimming of bands of radiation from a source at $z^{\prime} $ due to the redshifting of continuum into Lyman resonance lines (Ahn et al. 2009). As seen in Equation (16), ${f}_{\mathrm{mod}}$ replaces the attenuation factor ${e}^{-{\tau }_{\nu }}$ and is given approximately by

Equation (17)

where $\alpha \,\equiv \,{(h/0.7)}^{-1}{({{\rm{\Omega }}}_{m}/0.27)}^{-1/2}{[(1+z^{\prime} )/21]}^{-1/2}$ and rcMpc is the co-moving distance that light has traveled from $z^{\prime} $ to z, in units of cMpc. ${f}_{\mathrm{mod}}$ is useful when calculating the inhomogeneity of the LW background if inhomogeneous source distribution is given (Ahn et al. 2009), while in this work, it only serves as the relative weight that sources at $z^{\prime} $ contributes to JLW(z).

Because JLW is simply the frequency-averaged intensity, actual H2-dissociation rates by individual LW lines should be further implemented. This could be achieved by some multiplication factor weighted by line-wise dissociation rates, which would change the effective weight of source-contribution from a smooth form (${f}_{\mathrm{mod}}$) to a discrete form (fLW in Fialkov et al. 2013), or by interpreting JLW,th as the threshold intensity weighted by the same line-wise dissociation rates. We take the latter option in this work. We also note that in this one-zone model, Equation (16) is equivalent to the LW intensity that is averaged over the sawtooth-modulated spectrum (Haiman et al. 1997). Then, suppression of SFR by dissociation of H2 is implemented in the form of Equations (12)–(14).

2.3.3. Ionization and Heating by X-Ray Background

Global X-ray intensity determines the heating rate and the ionization rate of the IGM outside H ii regions (or "bulk IGM" as in Mirocha 2014). X-ray photon-number intensity is given by

Equation (18)

where we used Equation (15) and the fact that $\nu =\nu ^{\prime} (1+z)/(1+z^{\prime} )$. Once Nν (z) is known, we can calculate the photoionization rate

Equation (19)

the secondary ionization rate

Equation (20)

and the heating rate (erg s−1)

Equation (21)

where h νth = 13.6 eV is the hydrogen Lyman-limit energy, and σν,H I is the photoionization cross section of the hydrogen atom at frequency ν. The ionization rate equation for the bulk IGM is then given by

Equation (22)

where xb is used to denote the ionized fraction of the bulk IGM and to be distinguished from the volume ionized fraction x. Even though this will affect the volume ionization rate (Equation (5)) as well, we take the approximation that 1 − xb ≃ 1 until the reionization ends. This holds true for reionization scenarios we consider in this work. The energy rate equation is given by

Equation (23)

where kB is the Boltzmann constant, Tk is the kinetic temperature of gas, nb is the proper number density of baryons, epsiloncomp is the Compton heating (when TK < TCMB, and cooling when TK > TCMB) rate, and ${ \mathcal C }$ is the cooling rate. We only include the adiabatic cooling by cosmic expansion, which is dominant over the recombination cooling and the collisional excitation+ionization cooling.

2.3.4. Lyα Background

The hydrogen Lyα background is crucial in determining the 21 cm background by decoupling the spin temperature from the CMB temperature through the Lyα pumping process, or the Wouthuysen–Field mechanism (Wouthuysen 1952; Field 1958). The photon-number intensity of the Lyα background is given by

Equation (24)

where ${n}_{\max }$ (=23) is the effective maximum principal quantum number of Lyman resonances, frec(n) is the probability for an Lyn photon (Ly1 ≡ Lyα, Ly2 ≡ Lyβ, Ly3 ≡ Lyγ, etc.) to be converted to an Lyα photon, and $z{{\prime} }_{n}$ is the redshift satisfying (Pritchard & Furlanetto 2006)

Equation (25)

The Lyα background can also be generated by the collisional excitation of H atoms induced by energetic electrons generated by the X-ray background (e.g., Ahn et al. 2015b). However, we do not include this mechanism here because it is usually negligible when the X-ray efficiency is not extremely high. As seen in Section 2.4, we only impose a minimal level of X-ray background in this work.

2.4. 21 cm Background

The spin temperature Ts is a parameter representing the ratio of up to down states of the hyperfine structure (Field 1958):

Equation (26)

where n1 and n0 are the number of hydrogen atoms in the up (triplet) state and the down (singlet) state, respectively, and T* = 0.0628 K is the energy difference of the two states in terms of temperature. In the absence of Lyman resonance photons, Ts is driven to the CMB temperature TCMB radiatively. Otherwise, the absorption and re-emission of Lyman resonance photons can drive Ts to the color temperature of Lyman lines. The dominant radiative coupling is by the Lyα photons, and the repeated scattering of Lyα photons against thermalized gas brings the Lyα color temperature Tα into Tk . The mechanical pumping by collision, separately, drives Ts into Tk . Then the spin temperature becomes

Equation (27)

where xc is the collisional coupling coefficient and xα is the Lyα pumping coefficient. xc is given by

Equation (28)

where A10 = 2.85 × 10−15 s−1 is the spontaneous emission coefficient, and κ(1 − 0) is the collisional deexcitation coefficient defined and tabulated as a function of Tk in Zygelman (2005). xα is given by Pritchard & Furlanetto (2006)

Equation (29)

where me is the electron mass and Sα is a correction factor of the order of unity that accounts for the distortion of the line profile by thermalized atoms and peculiar motion (Chen & Miralda-Escudé 2004; Chuzhoy & Shapiro 2006; Hirata 2006). In this paper, we adopt the functional form of Sα suitable for co-moving gas without peculiar motion (Chuzhoy & Shapiro 2006):

Equation (30)

3. Result

3.1. Reionization History

We cover a limited but representative set of parameters in calculating reionization histories. For the vanilla model, we use parameters that are sampled similarly to Furlanetto (2006) and Bernardi et al. (2015). For SRI, we use parameters including those of Iliev et al. (2007), which first suggested the self-regulation scheme of the model. For SRII, we use parameters including those of Ahn et al. (2012), which provide the physical basis of the model. Parameters and some characteristics of reionization are listed in Tables 13. For the vanilla model and SRI, we accommodate both dF and F star formation scenarios (Section 2.1). The resulting x(z)'s are plotted in Figures 25, and overlaid on the 68% and 95% constraints from the Planck Legacy Data ("PLD"; Planck Collaboration et al. 2020). Note that in these figures we show the ionized volume fraction in terms of the electron fraction ${x}_{{\rm{e}}}\equiv \left\langle {n}_{{\rm{e}}}\right\rangle /\left\langle {n}_{{\rm{H}}}\right\rangle $, with $\left\langle \,\,\right\rangle $ being the volume average, such that ${x}_{{\rm{e}}}=(1+\left\langle {n}_{\mathrm{He}}\right\rangle /\left\langle {n}_{{\rm{H}}}\right\rangle )x\,=\,1.079\,x$ if helium atoms are assumed singly ionized in H ii regions. The PLD constraints are in fact on xe, and they are shown as shaded regions in Figures 25.

Table 1. Model Parameters and Resulting Characteristics of the Vanilla and the SRI

Model fγ gγ ${f}_{\gamma }^{{\rm{H}}}$ ${f}_{\gamma }^{{\rm{L}}}$ ${g}_{\gamma }^{{\rm{H}}}$ ${g}_{\gamma }^{{\rm{L}}}$ zbegin zend Δz3−97 τes τes(15–30) χ2/ν
V-L_dF (P)19.6215.875.508.180.059983.97E-40.81
V-M1_dF23.5516.326.217.980.066385.11E-40.86
V-M2_dF34.8817.017.277.660.076477.57E-41.19
V-H_dF56.6917.858.527.330.088991.23E-32.31
V-L_F0.407712.215.495.060.047942.73E-50.86
V-M1_F (P)0.704213.006.255.100.055554.71E-50.81
V-M2_F1.47214.067.325.110.066519.85E-50.87
V-H_F3.14015.138.435.100.078522.10E-41.32
SRI-L0_dF (P)26.02012.825.815.560.054362.01E-50.82
SRI-LL_dF26.0291.9316.046.057.960.063804.14E-40.83
SRI-LH_dF26.02919.319.116.8210.410.093112.77E-32.93
SRI-HL_dF91.9391.9316.228.256.150.080574.60E-41.46
SRI-HH_dF91.93919.319.128.788.540.100832.81E-34.54
SRI-0H_dF0919.319.110.076592.75E-31.80
SRI-L0_F0.8673010.815.783.840.047561.53E-60.87
SRI-LL_F (P)0.86738.67314.416.246.490.061401.16E-40.81
SRI-LH_F0.867386.7317.096.938.340.088168.63E-42.22
SRI-HL_F6.9388.67314.598.274.820.075621.25E-41.16
SRI-HH_F6.93886.7317.108.796.740.093108.71E-42.95
SRI-0H_F086.7317.094.1010.120.085978.62E-41.98

Note. SRI-0H_dF does not reach the end of reionization, not even x = 0.97, until z = 0. Denoted by "(P)" are models that best fit the observed ${C}_{l}^{\mathrm{EE}}$ of the PLD.

Download table as:  ASCIITypeset image

All models are described by simple ordinary differential equations (ODEs), and thus can be easily integrated with ODE solvers. We start numerical integration from z = 40, when the contribution of any type of halo to reionization and heating is believed to be negligible. The initial value of x is set to an arbitrarily small value, because the volume occupied by H ii regions at z = 40 must be negligible. Ionization rate equations to solve are not stiff, and we use a fourth-order, adaptive Runge–Kutta integrator with both the relative tolerance and the absolute tolerance of x set to 10−5. For SRII, we need an extra effort to calculate JLW(t) at any time t, because JLW(t) regulates SFR inside MHs at t via xLW and impacts PPR (Equations (12) and (13)). Therefore, we calculate x(t) and J(t) at each incrementally increasing time step, by integrating Equation (5) with Equation (13) (or Equation (12) if dF assumed) and using Equations (14) and (16).

The vanilla model shows a smooth and monotonic evolution of x (Figure 1). The monotonic behavior of x(t) is easily explained by the fact that ξ is proportional to the monotonically increasing fcoll (F scenario) or dfcoll/dt (dF scenario). This characteristic makes τes and zend tightly correlated, and thus lacks the "leverage" to accommodate reionization scenarios that are different in x(z) but degenerate in τes and zend, as long as one type of star formation scenario is chosen from F or dF. One can of course make a somewhat more sophisticated variant of this model by, e.g., allowing multiple species of halos with different fγ 's (a mixture of Population II and Population III stars; Furlanetto 2006). Nevertheless, due to the lack of any self-regulation, the resulting reionization histories of such variants would still remain similar to the original vanilla model. The duration of reionization is in general more extended in the dF scenario than the F scenario. This is due to the fact that dfcoll/dt grows more slowly than fcoll. Therefore, for given zend, the dF scenario produces larger τes than the F scenario. This tendency is clearly presented in Figure 1 and Table 1.

Figure 1.

Figure 1. Evolution of the global ionized fraction in vanilla ("V") reionization models with varying fγ . Model specifications are shown in legends, and are listed in Table 1. Left panel: vanilla models when the "dF" scheme is used are shown. Right panel: vanilla models when the "F" scheme is used are shown. Throughout Figures 15, the ionized fraction is in terms of xene/nH = 1.079 x, which reaches the maximum value of 1.079, based on assuming that all helium atoms inside H ii regions are singly ionized before the epoch of helium reionization occurring at z ≃ 3.5.

Standard image High-resolution image

The SRI model adds a little more complexity to the characteristics compared to the vanilla model (Figure 2). Due to the existence of self-regulation, the SRI is expected to have more extended reionization histories than the vanilla model. This has indeed been shown to be the case for a consistent halo-selection criterion for HMACHs and LMACHs (Iliev et al. 2007). However, one subtlety in our modeling scheme complicates such an expectation. Because we use a constant-mass criterion in SRI, while a constant-temperature criterion in the vanilla model, we find that in some cases, the duration of SRI models can be shorter than that of vanilla models. Had we used the same halo-selection criterion, SRI models would have larger Δz than vanilla models, which we actually tested and confirmed. Aside from this complication, which is not essential, the general trend is that (1) the larger the value of ${g}_{\gamma }^{{\rm{L}}}/{g}_{\gamma }^{{\rm{H}}}$ is, the larger the duration of reionization becomes, and (2) the addition of LMACHs to HMACH-only scenarios extends the duration of reionization. Also, cases with very large ${g}_{\gamma }^{{\rm{L}}}/{g}_{\gamma }^{{\rm{H}}}$ (e.g., the SRI-0H_F case in Figure 2) slow down reionization significantly at the end of reionization, producing histories as symmetric as the tangent-hyperbolic model that has been used extensively in the analysis of the CMB data. It is easy to understand this behavior: ${g}_{\gamma }^{{\rm{L}}}/{g}_{\gamma }^{{\rm{H}}}$ is a rough measure of the relative contribution of LMACHs to reionization to that of HMACHs, and the self-regulation becomes stronger as x becomes larger. One very extreme case is SRI-0H_dF, which never finishes reionization due to a strong self-regulation. The trend that Δz is larger in the dF scenario than in the F scenario is the same as in the vanilla model. The general trend of the vanilla and SRI models can also be seen in Figure 3 in terms of ξ.

Figure 2.

Figure 2. Evolution of the global ionized fraction in the model category "self-regulated I." The model variance is specified by the set of $\left\{{g}_{\gamma }^{{\rm{H}}},{g}_{\gamma }^{{\rm{L}}}\right\}$, which is specified in legends, with the nomenclature "SI-${g}_{\gamma }^{{\rm{H}}}$ ${g}_{\gamma }^{{\rm{L}}}$" but gγ 's with letters "0" for null, "L" for low, and "H" for high. These models are listed in Table 1.

Standard image High-resolution image
Figure 3.

Figure 3. Top panel: ionization rate and ionizing-photon emissivity (ξ in Equation (5) and ${\dot{N}}_{\mathrm{ion}}\equiv \xi {n}_{b,0}$) of PLD-favored vanilla and SRI models, compared to the z = 6.02 and z = 6.42 constraints by Calverley et al. (2011). Models are specified in the subtitle. For SRI cases, contributions by LMACHs (red, dashed) and HMACHs (blue, dotted–dashed) are shown separately. The net quantities (orange, solid) are also plotted. Note that the constraint by Calverley et al. (2011) comprises values of observed Γ (Section 2.3.1) translated into ${\dot{N}}_{\mathrm{ion}}$ based on a specific set of λmfp and the spectral hardness of H-ionizing photons: λmfp = 10 cMpc and αs = αb = 2 in Equation (21) of Bolton & Haehnelt (2007).

Standard image High-resolution image

The SRII model has features richer than the vanilla and the SRI models (Figures 4 and 5). The most notable feature is the existence of the early, extended, and slowly increasing phase in x. This is due to the self-regulation of star formation, even stronger than that in the SRI model, which takes place inside MHs. The star formation inside MHs is mainly regulated by the LW background JLW, which quickly builds up to reach JLW,th. Continuum photons below the Lyman limit and emitted at redshift z travel a cosmological distance (∼100{(1 + z)/21}−0.5 cMpc), in contrast to the hydrogen-ionizing UV photons that travel up to the ionization front and are then absorbed. Therefore, any newly forming MHs will be under the influence of the LW background long before being exposed to the ionizing photons, and any pre-ionized region would have been under the overcritical LW intensity (JLW > JLW,th). We find that this is indeed the case: when tested with $\max ({x}_{\mathrm{LW}},x)$ replaced by xLW in Equation (13), the resulting x(z) was not affected.

Figure 4.

Figure 4. Evolution of the global ionized fraction in the model category "self-regulated II," with MIII = 100 M per MH. Each panel corresponds to varying sets of $\{{g}_{\gamma }^{{\rm{H}}},{g}_{\gamma }^{{\rm{L}}},{M}_{\mathrm{III}}\}$, denoted in the format "SRII-${g}_{\gamma }^{{\rm{H}}}{g}_{\gamma }^{{\rm{L}}}$-MIII/M." The letters "0," "L," and "H" denote "null," "low," and "high" values of gγ , respectively, in a relative sense (listed in Table 1). In all panels, the set of ${f}_{\mathrm{esc}}^{{\rm{M}}}$ and JLW,th (in units of 10−21 erg s−1 cm−2 Hz−1 sr−1) is specified as legends: {0.1, 0.05}, black dotted–long-dashed; {0.1, 0.10}, red solid; {0.5, 0.05}, blue dotted; {0.5, 0.10}, magenta dotted–short-dashed; {1.0, 0.05}, cyan dashed; and {1.0, 0.10}, brown dotted–dotted–dashed.

Standard image High-resolution image
Figure 5.

Figure 5. Same as Figure 4 but with MIII = 300 M. Each panel corresponds to varying ${f}_{\gamma }^{{\rm{H}}}$ and MIII as specified in the title. In all panels, the set of ${f}_{\mathrm{esc}}^{{\rm{M}}}$ and JLW,th (in units of 10−21 erg s−1 cm−2 Hz−1 sr−1) are specified as legends using the same line convention as Figure 4.

Standard image High-resolution image

Another notable feature of SRII is that, in some cases where the contribution of ionizing photons by MHs is as significant as to drive x beyond ≳10%, there exists a phase where x decreases in time. 4 This is mainly due to the fact that (1) the LW feedback renders JLWJLW th when MHs dominate as the main radiation sources (Figure 11), and (2) the large difference in the number of soft-UV (h ν = ∼ 11 − 13.6 eV) photons per ionizing photon of Population II and Population III stars (see the detail in Section 2.4), and (3) the drop of fesc from MH values (fesc ∼ 1) to ACHs (fesc ∼ 0.2). Then, LMACHs (assumed to host Population III stars) and HMACHs (assumed to host Population II stars) start to generate soft-UV photons to make up JLW near JLW th. This is achieved at the expense of ACHs putting out a much smaller amount of ionizing photons than MHs (assumed to host Population III stars). Consequently, the IGM gains a chance to recombine faster than ionization (Figure 6: see dips in ξ). In practice, however, this recombination is slight and not as dramatic as the "double reionization" that has been suggested by Cen (2003). Given the same set of $\left\{{g}_{\gamma }^{{\rm{H}}},{g}_{\gamma }^{{\rm{L}}}\right\}$ as in SRI, the SRII model has zend that is practically identical to that of SRI while it has τes and Δz that are both boosted from those of SRI (Tables 13). This is simply due to the existence of additional photon sources, or MH stars, that ionize the IGM only to a limited extent (x ≲ 15% at most under our parameter range, but may be increased if MIII and JLW,th are pushed to higher values) such that zend is not much affected but can increase τes and extend the duration of reionization substantially by strongly regulated ionization history.

Figure 6.

Figure 6. Top panels: same as Figure 3. Bottom panels: the LW band intensity in terms of xLWJLW/JLW,th (Equation (14)), for two selected cases among PLD-favored SRII models. Line convention is identical to Figure 3, and contributions by MHs (black, dotted) are shown in addition.

Standard image High-resolution image
Figure 7.

Figure 7.  χ2/ν of the E-mode power spectrum ${D}_{l}^{\mathrm{EE}}$ of each model with respect to the 2018 PLD. In each panel, plotted points are grouped in subcategories specified by legends, and arrows point to the minimum-χ2/ν cases in each panel. Models marked by arrows correspond to those denoted by "(P)" in Tables 13.

Standard image High-resolution image

3.2. Comparison with CMB Observations

The 2018 PLD provides, among improvements from the 2015 data, so far the most precise measurement of the large-scale CMB polarization anisotropy. The large-scale E-mode autocorrelation angular power spectrum, ${C}_{l}^{\mathrm{EE}}$, is strongly affected by the history of reionization. The quadrupole moment of the CMB anisotropy generates linear polarization after Thomson scattering from the viewpoint of an electron, and the polarization signal is observed after being modulated by the relevant wave-modes, resulting in affecting ${C}_{l}^{\mathrm{EE}}$ mostly in the low-l (≲20) regime (Hu & White 1997; Haiman & Knox 1999; Dodelson 2003).

The extended high-redshift (z ≳ 15) ionization tail predicted by Ahn et al. (2012) and SRII models here has been advocated by Miranda et al. (2017) and Heinrich & Hu (2018), based on their principal component analysis (PCA) of the Planck 2015 ${C}_{l}^{\mathrm{EE}}$ data observed through the LFI. They claimed that a specific SRII model with a substantial high-redshift tail, corresponding roughly to the L2M1J2 case of Ahn et al. (2012), was favored over the vanilla model at ∼ 1σ level. Later, Millea & Bouchet (2018) used both the LFI and the HFI (proprietary at the time) data, with a well-handled physicality (x > 0) prior, to claim against too much contribution from the z ≳ 15 epoch. They constrained the optical depth from 15 ≤ z ≤ 30, or τ(15, 30), to τ(15, 30) < 0.015 at the 2σ level. The Planck 2018 analysis based only on the low-l E-mode polarization further reduces this value to τ(15, 30) < 0.007 at the 2σ level and ${\tau }_{\mathrm{es}}={0.0504}_{-0.0079}^{+0.0050}$ at the 1σ level (Planck Collaboration et al. 2020).

In light of the constraints described above, we compare our model x(z)'s from Section 3.1 to PLD. The main purpose of this task is to (1) understand whether any class of our models are preferred by observation and (2) whether the degeneracy of models in τes and zov can be broken. For example, Ahn et al. (2012) showed that SRII models with an extended tail in x(z) could be distinguished from the vanilla- or SRI-type models, even when the models have the same τes (=0.085) and zov (=6.8). The pictorial comparison of x(z) to the Planck constraint (1σ and 2σ constraints shown in the shaded regions) is shown in Figures 15. We test the relative goodness of several selected models by calculating the reduced chi square,

Equation (31)

where ${D}_{l}^{\mathrm{EE}}\equiv l(l+1){C}_{l}^{\mathrm{EE}}/(2\pi )$ and ${\tilde{D}}_{l}^{\mathrm{EE}}\equiv l(l+1){\tilde{C}}_{l}^{\mathrm{EE}}/(2\pi )$ are the E-mode power spectra corresponding to a model x(z) and PLD, respectively, and σl is the standard deviation of ${\tilde{D}}_{l}^{\mathrm{EE}}$ due to the cosmic variance and the noise. 5 ${C}_{l}^{\mathrm{EE}}\equiv \left\langle {\left|{a}_{{lm}}^{{\rm{E}}}\right|}^{2}\right\rangle $ for given l averaged over m = [−l, l], with the spherical-harmonics decomposition of the E-mode anisotropy $E(\theta ,\phi )={\sum }_{{lm}}{a}_{{lm}}^{{\rm{E}}}{Y}_{{lm}}(\theta ,\phi )$. In calculating ${D}_{l}^{\mathrm{EE}}$, we use a version of the Boltzmann solver CAMB that was modified to allow for a generic shape of x(z) (Mortonson & Hu 2008; downloadable from http://background.uchicago.edu/camb_rpc/). For the base cosmology, we use the best-fit parameter set of PLD. While this is not a full likelihood analysis including other data products such as the temperature anisotropy, the value of χ2/ν from Equation (31) can indicate the relative goodness of models because the impact of reionization histories is the strongest in the E-mode (see, e.g., an identical approach by Qin et al. 2020b). E-mode power spectra of selected models against the PLD are plotted in Figure 8.

Figure 8.

Figure 8. Low-multipole E-mode autocorrelation power spectra predicted by models (lines with model specification on the top-left panel) with smallest χ2/ν and that of PLD (points with 1σ error bar). We select only two models with smallest χ2/ν from SRII models, to avoid crowdedness of lines. CMB lensing is considered in all cases, even though the net effect in this range of l is weak. Different plotting schemes (top-right: logarithmic; bottom-left: linear; bottom-right: linear-zoomed) are used. SRII models with the high-z ionization tail achieve larger values of ${D}_{l}^{\mathrm{EE}}$ for 14 ≲ l ≲ 24 than the vanilla and SRI models, even though all of the models plotted here share about a single value of τes (≃0.06).

Standard image High-resolution image

It is interesting to note that the constraint on reionization by the Planck observation provides a good match to the observed Γ (Section 2.3.1). If the observed Γ (e.g., Calverley et al. 2011) is translated into ${\dot{N}}_{\mathrm{ion}}$ with a reasonable set of physical parameters (λmfp = 10 cMpc: the IGM mean free path to H-ionizing photons at z ≃ 6; and αs = αb = 2: the spectral hardness of H-ionizing photons in Equation (21) of Bolton & Haehnelt 2007), PLD-favored models have a good agreement with the observed Γ at z ≃ 6 (Figures 3 and 6). This consistency between the two independent observations, even though uncertainties are large, is encouraging.

We now claim that some SRII models with substantial high-redshift tails are still among those highly favored by PLD. Because χ2/ν is a measure of the goodness of fit, we can use the value of χ2/ν to find the PLD-favored models. We indeed find many models can explain the PLD low-l ${C}_{l}^{\mathrm{EE}}$ fairly well even though the variance in τes of such models is substantial. Models that fit the PLD ${C}_{l\leqslant 30}^{\mathrm{EE}}$ best are marked by arrows in Figure 7 and denoted by "(P)" in model names in Tables 13: all of these models have almost the same likelihood with χ2/ν = 0.80–0.82, but with a substantial spread on τes with τes ≃ [0.0544, 0.0643]. If allowance is extended to models with χ2/ν ≤ 1, then the allowed optical depth becomes τes ≃ [0.044, 0.072]. This indicates that some of our SRII models are still well within the PLD constraint and are favored as much as those models without high-redshift ionization tails. As seen in Tables 13, the most favored models with the smallest values of χ2/ν, χ2/ν = 0.80, are indeed SRII-L0-100-e1.0-J0.1 and SRII-L0-300-e1.0-J0.1, which have a substantial high-redshift tail that reaches maximum x = 0.12 at z = 13.6 and x = 0.15 at z = 13.6, respectively. Such tails contribute to τes(15–30) substantially: τes(15–30) = 0.0063 and 0.0078 for SRII-L0-100-e1.0-J0.1 and SRII-L0-300-e1.0-J0.1, respectively. We also note that these maximum-likelihood models are clustered around τes = 0.06, substantially different from the inferred value τes = 0.0504 by PLD-${C}_{l\leqslant 30}^{\mathrm{EE}}$. The reason why such a difference occurs is unclear; this is nevertheless a very important issue and a further investigation is warranted.

The model with the strongest ionization tail of all, SRII-L0-300-e1.0-J0.1, is worthy of close attention. Compared to the 2σ constraint of PLD, τes(15–30) < 0.007, SRII-L0-300-e1.0-J0.1 actually violates this constraint with τes(15–30) = 0.0078 but is still the best fit (to PLD ${C}_{l\leqslant 30}^{\mathrm{EE}}$) of all of the models we tested. This model also has τes = 0.0596, which is about 2σ away from the E-mode-only best-fit estimate by PLD, τes = 0.0504. If we do not consider other CMB observables and assume a flat prior, we can conclude that this model is as good as or just slightly better than other tailless models with χ2/ν = 0.81–0.82. It is interesting to see that there exists a weak tension between τes's estimated by the low-l E-mode polarization and the CMB lensing: the low-l E-mode data of PLD prefers such a low ${\tau }_{\mathrm{es}}(={0.0504}_{-0.0079}^{+0.0050})$, while the CMB lensing of PLD prefers higher τes at around τes ≃ 0.08 (Planck Collaboration et al. 2020). This tendency of the CMB lensing favoring large values of τes, even though the uncertainty is large, is on par with favoring two-stage reionization models with substantial ionization tails. Our findings are in slight disagreement with the PLD constraint that was constructed using nonparametric Bayesian inference. Based on our forward modeling and goodness-of-fit approach, we argue that a family of models with a substantial high-redshift ionization tail reaching ${x}_{\max }\sim 0.12$ is still just as strong a contender at the moment as the tailless models.

Will there be a chance to probe a high-redshift ionization tail in the future? The high-redshift tail tends to boost ${C}_{l}^{\mathrm{EE}}$ at 14 ≲ l ≲ 24 (e.g., Ahn et al. 2012; Miranda et al. 2017). SRII models in Figure 8 produce ${C}_{14\leqslant l\leqslant 24}^{\mathrm{EE}}$ larger than that of the rest of models, and in particular, SRII-L0-300-e1.0-J0.1 has the strongest ${C}_{l}^{\mathrm{EE}}$ at 8 ≤ l ≤ 24. In principle, models degenerate in τes can have different ${C}_{l\lt 30}^{\mathrm{EE}}$'s due to the variance in x(z). Ahn et al. (2012), using a PCA, had indeed predicted that high-precision CMB observations could break the degeneracy in τes and probe (or disprove) the existence of the high-redshift ionization tail. We stress this point again through Figure 9, showing two models from Ahn et al. (2012) that are degenerate both in τes and zend but are clearly different in x(z), especially in the existence of the high-z ionization tail, and in the resulting ${C}_{l}^{\mathrm{EE}}$. From Figures 8 and 9, we observe that the boost of ${C}_{14\leqslant l\leqslant 24}^{\mathrm{EE}}$ in the two-stage reionization models with the high-z ionization tail against the tailless models is a universal effect. As seen in Figure 7, the relation between τes and χ2/ν is not exactly monotonic but instead there exists some scatter in χ2/ν for the same τes and vice versa. Such a scatter increases as τes increases, which is due to the increased freedom in constructing x(z) for given τes. However, because the PLD E-mode power spectrum prefers such a low τes, as of now the leverage of having a pronounced tail has somewhat diminished from that prediction. Nevertheless, it is possible that observation by a more accurate apparatus might find a preference for a higher τes than Planck that is still hampered by the large noise in measuring the polarization anisotropy. Therefore, we need a better apparatus than Planck to (1) see whether or not τes could get larger than the estimate by PLD to allow for more pronounced two-stage reionization models and (2) break the model degeneracy in τes better than Planck to probe the ionization tail even when the tail is weak.

Figure 9.

Figure 9. Breaking degeneracy in τes and zend by E-mode polarization observation. Left panel: x(z)'s of two selected reionization models from Ahn et al. (2012), sharing the same τes (=0.086) and zend (=6.77) but are grossly different in x(z) with (L2M1J1: black, solid) and without a substantial high-z ionization tail (g2.609C_165.2: red, dotted), are plotted against the PLD constraint (gray shades). Right panel: E-mode polarization power spectra of the two models, against the PLD data (data points with error bar), showing a difference at l ≲ 24.

Standard image High-resolution image

We also briefly describe another type of constraint from CMB observations. The kinetic Sunyaev–Zel'dovich effect can arise from the peculiar motion of H ii bubbles during EoR and can affect the small-scale (l ∼ a few thousand) temperature anisotropy power spectrum ${C}_{l}^{\mathrm{TT}}$. Measurement of ${C}_{l}^{\mathrm{TT}}$ by the South Pole Telescope, especially ${C}_{l=3000}^{\mathrm{TT}}$ (Reichardt et al. 2012), was used by Zahn et al. (2012) to constrain the duration of reionization Δzz(x = 0.25) − z(x = 0.99) to Δz < 4 − 7 at the 2σ level (depending on the assumed correlation between the thermal Sunyaev–Zel'dovich effect and the cosmic infrared background; see also the similar assessment by Mesinger et al. 2012 and Battaglia et al. 2013). Without MH stars, reionization occurs always in a patchy way and thus any addition of electrons, or equivalently extension of Δz, increases ${C}_{l=3000}^{\mathrm{TT}}$ monotonically, as was assumed in Zahn et al. (2012), Mesinger et al. (2012), and Battaglia et al. (2013). However, Park et al. (2013) readdressed this issue with a variety of reionization scenarios including the SRII-type, and found that the added duration of reionization beyond this limit could still be accommodated by the measured ${C}_{l=3000}^{\mathrm{TT}}$. As claimed in Park et al. (2013), H ii regions by MHs are distributed almost uniformly (Ahn et al. 2012), and thus the increase in Δz in SRII models does not guarantee an increase in ${C}_{l=3000}^{\mathrm{TT}}$. Therefore, the largeness of Δz3−97z(0.03) − z(x = 0.97) of many SRII models (see, e.g., those denoted by "(P)" in Tables 2 and 3) should not be considered as a violation of such a constraint. Instead, constraining Δz using the small-scale ${C}_{l}^{\mathrm{TT}}$ should be restricted to only a limited set of models without MHs.

Table 2. Model Parameters and Resulting Characteristics of SRII, with MIII = 100 M

Model ${g}_{\gamma }^{{\rm{H}}}$ ${g}_{\gamma }^{{\rm{L}}}$ MIII ${f}_{\mathrm{esc}}^{{\rm{M}}}$ JLW,th zbegin zend Δz3−97 τes τes(15–30) χ2/ν
SRII-L0-100-e0.1-J0.050.867301000.10.0510.865.783.860.048013.45E-40.86
SRII-L0-100-e0.1-J0.100.867301000.10.1011.035.783.880.048506.18E-40.86
SRII-L0-100-e0.5-J0.050.867301000.50.0511.185.783.900.049831.72E-30.85
SRII-L0-100-e0.5-J0.100.867301000.50.1022.155.794.070.052263.08E-30.83
SRII-L0-100-e1.0-J0.050.867301001.00.0522.915.793.960.052103.44E-30.83
SRII-L0-100-e1.0-J0.10 (P)0.867301001.00.1025.385.794.510.056966.17E-30.80
SRII-LL-100-e0.1-J0.05 (P)0.86738.6731000.10.0514.486.246.500.061683.75E-40.81
SRII-LL-100-e0.1-J0.10 (P)0.86738.6731000.10.1014.906.246.530.061996.37E-40.81
SRII-LL-100-e0.5-J0.05 (P)0.86738.6731000.50.0515.266.246.540.062841.44E-30.82
SRII-LL-100-e0.5-J0.100.86738.6731000.50.1022.146.246.760.064462.78E-30.83
SRII-LL-100-e1.0-J0.05 (P)0.86738.6731001.00.0522.906.246.590.064322.79E-30.82
SRII-LL-100-e1.0-J0.100.86738.6731001.00.1025.386.2413.930.067625.50E-30.86
SRII-LH-100-e0.1-J0.050.867386.731000.10.0517.116.938.350.088321.01E-32.23
SRII-LH-100-e0.1-J0.100.867386.731000.10.1017.196.938.350.088471.15E-32.24
SRII-LH-100-e0.5-J0.050.867386.731000.50.0517.226.938.360.088951.60E-32.29
SRII-LH-100-e0.5-J0.100.867386.731000.50.1022.056.938.410.089732.37E-32.37
SRII-LH-100-e1.0-J0.050.867386.731001.00.0517.466.938.370.089762.35E-32.36
SRII-LH-100-e1.0-J0.100.867386.731001.00.1025.386.938.500.091353.91E-32.55
SRII-HL-100-e0.1-J0.056.9388.6731000.10.0514.628.274.820.075863.48E-41.17
SRII-HL-100-e0.1-J0.106.9388.6731000.10.1014.718.274.830.076105.70E-41.18
SRII-HL-100-e0.5-J0.056.9388.6731000.50.0514.798.274.840.076841.26E-31.20
SRII-HL-100-e0.5-J0.106.9388.6731000.50.1022.138.274.910.078072.40E-31.26
SRII-HL-100-e1.0-J0.056.9388.6731001.00.0515.358.274.860.078072.41E-31.25
SRII-HL-100-e1.0-J0.106.9388.6731001.00.1025.388.275.040.080574.70E-31.38
SRII-HH-100-e0.1-J0.056.93886.731000.10.0517.128.796.740.093251.01E-32.96
SRII-HH-100-e0.1-J0.106.93886.731000.10.1017.198.796.750.093391.16E-32.98
SRII-HH-100-e0.5-J0.056.93886.731000.50.0517.228.796.750.093871.60E-33.04
SRII-HH-100-e0.5-J0.106.93886.731000.50.1022.058.796.790.094642.35E-33.15
SRII-HH-100-e1.0-J0.056.93886.731001.00.0517.438.796.760.094672.34E-33.14
SRII-HH-100-e1.0-J0.106.93886.731001.00.1025.388.796.880.096223.87E-33.38
SRII-0H-100-e0.1-J0.05086.731000.10.0517.114.1010.120.086121.01E-31.99
SRII-0H-100-e0.1-J0.10086.731000.10.1017.194.1010.130.086271.15E-32.00
SRII-0H-100-e0.5-J0.05086.731000.50.0517.224.1010.130.086761.60E-32.04
SRII-0H-100-e0.5-J0.10086.731000.50.1022.054.1010.180.087542.37E-32.12
SRII-0H-100-e1.0-J0.05086.731001.00.0517.464.1010.150.087562.35E-32.11
SRII-0H-100-e1.0-J0.10086.731001.00.1025.384.1010.270.089163.92E-32.27

Note. In the table, MIII and JLW,th are in units of M and 10−21 erg s−1 cm−2 Hz−1 sr−1, respectively. Denoted by "(P)" are models that fit the observed ${C}_{l}^{\mathrm{EE}}$ of the PLD best.

Download table as:  ASCIITypeset image

Table 3. Model Parameters and Resulting Characteristics of SRII, with MIII = 300 M

Model ${g}_{\gamma }^{{\rm{H}}}$ ${g}_{\gamma }^{{\rm{L}}}$ MIII ${f}_{\mathrm{esc}}^{{\rm{M}}}$ JLW,th zbegin zend Δz3−97 τes τes(15–30) χ2/ν
SRII-L0-300-e0.1-J0.050.867303000.10.0510.895.783.860.048174.50E-40.86
SRII-L0-300-e0.1-J0.100.867303000.10.1011.105.783.890.048757.75E-40.86
SRII-L0-300-e0.5-J0.050.867303000.50.0511.455.783.920.050632.29E-30.84
SRII-L0-300-e0.5-J0.10 (P)0.867303000.50.1023.265.794.140.053503.88E-30.82
SRII-L0-300-e1.0-J0.05 (P)0.867303001.00.0523.505.794.020.053654.50E-30.82
SRII-L0-300-e1.0-J0.10 (P)0.867303001.00.1027.155.7915.280.059607.79E-30.80
SRII-LL-300-e0.1-J0.05 (P)0.86738.6733000.10.0514.516.246.500.061774.49E-40.81
SRII-LL-300-e0.1-J0.10 (P)0.86738.6733000.10.1015.266.246.540.062157.73E-40.82
SRII-LL-300-e0.5-J0.05 (P)0.86738.6733000.50.0520.006.246.550.063301.83E-30.82
SRII-LL-300-e0.5-J0.100.86738.6733000.50.1023.266.246.860.065283.45E-30.83
SRII-LL-300-e1.0-J0.050.86738.6733001.00.0523.506.246.650.065273.57E-30.83
SRII-LL-300-e1.0-J0.100.86738.6733001.00.1027.156.2414.770.069446.96E-30.88
SRII-LH-300-e0.1-J0.050.867386.733000.10.0517.126.938.350.088361.04E-32.23
SRII-LH-300-e0.1-J0.100.867386.733000.10.1017.216.938.360.088561.23E-32.25
SRII-LH-300-e0.5-J0.050.867386.733000.50.0517.266.938.360.089171.77E-32.30
SRII-LH-300-e0.5-J0.100.867386.733000.50.1023.226.938.420.090222.78E-32.41
SRII-LH-300-e1.0-J0.050.867386.733001.00.0517.696.938.380.090242.74E-32.40
SRII-LH-300-e1.0-J0.100.867386.733001.00.1027.156.938.540.092354.76E-32.64
SRII-HL-300-e0.1-J0.056.9388.6733000.10.0514.638.274.820.075944.10E-41.17
SRII-HL-300-e0.1-J0.106.9388.6733000.10.1014.758.274.830.076236.85E-41.18
SRII-HL-300-e0.5-J0.056.9388.6733000.50.0514.918.274.850.077221.58E-31.22
SRII-HL-300-e0.5-J0.106.9388.6733000.50.1023.268.274.930.078752.97E-31.28
SRII-HL-300-e1.0-J0.056.9388.6733001.00.0523.508.274.880.078803.01E-31.28
SRII-HL-300-e1.0-J0.106.9388.6733001.00.1027.158.275.120.081985.90E-31.45
SRII-HH-300-e0.1-J0.056.93886.733000.10.0517.128.796.740.093291.05E-32.97
SRII-HH-300-e0.1-J0.106.93886.733000.10.1017.208.796.750.093491.24E-32.99
SRII-HH-300-e0.5-J0.056.93886.733000.50.0517.278.796.750.094111.78E-33.06
SRII-HH-300-e0.5-J0.106.93886.733000.50.1023.228.796.810.095132.77E-33.21
SRII-HH-300-e1.0-J0.056.93886.733001.00.0517.588.796.770.095122.69E-33.19
SRII-HH-300-e1.0-J0.106.93886.733001.00.1027.158.796.910.097214.70E-33.50
SRII-0H-300-e0.1-J0.05086.733000.10.0517.124.1010.120.086171.04E-31.99
SRII-0H-300-e0.1-J0.10086.733000.10.1017.214.1010.130.086371.24E-32.01
SRII-0H-300-e0.5-J0.05086.733000.50.0517.274.1010.130.086981.78E-32.06
SRII-0H-300-e0.5-J0.10086.733000.50.1023.224.1010.190.088032.79E-32.16
SRII-0H-300-e1.0-J0.05086.733001.00.0517.694.1010.150.088022.71E-32.14
SRII-0H-300-e1.0-J0.10086.733001.00.1027.154.1010.310.090164.77E-32.36

Note. The unit convention is the same as that in Table 2. Denoted by "(P)" are models that fit the observed ${C}_{l}^{\mathrm{EE}}$ of the PLD best.

Download table as:  ASCIITypeset image

3.3. 21 cm Background and Comparison with EDGES Observation

The main variants determining δ Tb are the X-ray heating efficiency and the Lyα intensity, which determine TK and xα , respectively. The X-ray efficiency is not a direct product of the stellar radiation and is thus the main cause of the uncertainty in δ Tb . The Lyα intensity, on the other hand, is almost solely determined by the stellar radiation and is closely related to the ionizing PPR and the SED. We do not consider the creation of Lyα photons due to the excitation of H atoms by the X-ray-induced electrons, which is a good approximation unless the X-ray efficiency is extremely high (fx ≫ 1 with fx in Equation (32)). For the X-ray efficiency, we use the common parameter fX (Furlanetto 2006; Mirocha 2014), defined as the fudge parameter connecting the co-moving X-ray luminosity density ${{ \mathcal L }}_{\nu }$ ($=h\nu {{ \mathcal N }}_{\nu };$ in erg s−1 Hz−1 cMpc−3) to SFRD (Section 2.1):

Equation (32)

where the additional proportionality coefficient cX is fixed to cX = 3.4 × 1040 erg s−1(M yr−1)−1, an extrapolation of the 2–10 keV relation between ${{ \mathcal L }}_{\nu }$ and SFRD (or equivalently between Lν and SFR on average galaxies) by Grimm et al. (2003) to h ν ≥ 0.2 keV. Here, we limit ${{ \mathcal L }}_{\nu }$ to the energy range h ν = [0.2, 30] keV and a power-law SED ${{ \mathcal L }}_{\nu }\propto {\nu }^{-1.5}$. For the Lyα intensity and the LW intensity, we take a simple distinction between Population II and Population III stars. Population III stars are assumed to have Nion = 50,000, Nα L = 4800, and Nβ L = 2130, where Nα L and Nβ L are the number of photons emitted by a stellar baryon during the stellar lifetime in the energy range from Lyα to LL and from Lyβ to LL, respectively. Population II stars are assumed to have Nion = 6000, Nα L = 9690, and Nβ L = 3170. This makes Nα L /Nion and Nβ L /Nion of Population III stars about an order of magnitude smaller than those of Population II stars, respectively. Nα L /Nion and Nβ L /Nion strongly affect JLW (Equation (16)) and Nα (Equation (24)) for a given PPR. We use the following SED conventions for each category of models:

  • 1.  
    Vanilla model: Population II SED.
  • 2.  
    SRI model: Population III SED for LMACH; Population II SED for HMACH.
  • 3.  
    SRII model: Population III SED for LMACH and MH; Population II SED for HMACH.

The claimed detection of a ∼500 mK absorption dip around ν ≃ 78 MHz by EDGES has been a matter of debate, mainly due to the fact that it is impossible to explain such a large amplitude in the standard ΛCDM framework, if the background after a successful foreground removal is composed only of the CMB and the 21 cm background. In the ΛCDM universe, the kinetic temperature of the IGM is limited to the adiabatically cooled value (TK ∼ 10.2 K[(1 + z)/21]2), and even at the maximum Lyα coupling, $| \delta {T}_{b}| $ is limited to $\sim 200\,{\rm{mK}}$. Another difficulty faced by the EDGES result is the existence of a peculiar spectral shape in δ Tb , a flat trough of δ Tb from ν = 72 to 85 MHz and lines connecting to the ends of the trough from ν = 65 and 92 MHz, which is in contrast with a smooth dip predicted by models in the ΛCDM.

We show our model predictions on δ Tb for a selected set of models, and compare these with the EDGES result. The selection criterion is the goodness of model fits to the PLD, and we use those minimum-χ2/ν models marked by arrows in Figure 7. We also tune fX to produce the largest absorption dip for each model but under the condition δ Tb > 0 at z ≲ 9, to (1) comply with the EDGES result with the deepest absorption possible and (2) compensate for our ignorance of the Lyα heating, which might naturally turn the 21 cm background into emission before the end of reionization (Chuzhoy & Shapiro 2007; Ciardi & Salvaterra 2007; Mittal & Kulkarni 2021 6 ) and some hints of the IGM heating at z ∼ 9 (Monsalve et al. 2017; Singh et al. 2018; Ghara et al. 2020; Mertens et al. 2020). Chuzhoy & Shapiro (2007) first showed that the Lyα-recoil heating can solely increase TK beyond TCMB before reionization is completed, correcting the estimate by Chen & Miralda-Escudé (2004). In this way, we show how far off each reionization model is from the EDGES result even when the maximum absorption is achieved in each model. One can be more inclusive in model selection because future CMB observations will probe the CMB polarization with better accuracy; nevertheless we stick to this choice here.

Different model categories show distinctive features in δ Tb (Figures 10 and 11, with the shaded regions indicating the redshift bin of the EDGES absorption trough of ∼500 mK), as follows.

  • 1.  
    The Planck-favored vanilla models show the familiar ∼180 mK absorption dip. The moment of the absorption dip and the start of the absorption due to the Lyα pumping (to be distinguished from the absorption due to collisional pumping at z ≳ 25) are delayed in the dF case compared to the F case. The dip resides at z ≃ 16 and 14.5 for the dF and F cases, respectively. The start of the absorption is at z ≃ 26 and 21 for the dF and F cases, respectively.
  • 2.  
    The Planck-favored SRI models show weaker absorption dips, with $\delta {T}_{b,\min }\simeq [-120,-140]\,\mathrm{mK}$, than the vanilla models. The dF case shows delays in the moments of the absorption dip and the start of the absorption compared to the F case, just as in the vanilla models. Compared to the vanilla models, both the dip and the start of the absorption are delayed: the dip is at z ≃ 12 (F)–13 (dF), and the absorption starts at z ≃ 18 (F)–20 (dF).
  • 3.  
    The Planck-favored SRII models show a very slowly deepening absorption "slope" during 28 ≳ z ≳ 14, which is the epoch that is about the same as the full EDGES-low observational window, before the absorption dip at z ≃ 11 − 12 occurs. The absorption dip is with $\delta {T}_{b,\min }\simeq [-90({\rm{L}}0),-120(\mathrm{LL})]\,\mathrm{mK}$, located away from the EDGES trough window. Where the EDGES trough exists, the models have the limited differential brightness temperature, δ Tb ≃ [ − 30, − 60] mK.

Several details of these results are noteworthy. (1) The ∼ 180 mK $\left|\delta {T}_{b,\min }\right|$ of the vanilla model is roughly the maximum amplitude allowed in the ΛCDM cosmology (Furlanetto 2006; Mirocha 2014; Bernardi et al. 2015). (2) The main cause for the overall delays in the dip and the start of the absorption for the SRI model relative to the vanilla model is the fact that LMACHs, which dominate the early phase of reionization, are assumed to host Population III stars. Because ${({N}_{\alpha L}/{N}_{\mathrm{ion}})}_{\mathrm{Pop}\mathrm{III}}$ is much smaller than ${({N}_{\alpha L}/{N}_{\mathrm{ion}})}_{\mathrm{Pop}\mathrm{II}}$ at a given level of x, Nα , and xα of the SRI models, while LMACHs dominate, they also become much smaller than those of the vanilla models that use the Population II SED. Such a smallness of xα then renders TS to couple only weakly to TK, bringing down the amplitude of δ Tb (Figure 10). (3) For the reason same as in detail (2), had we used the Population III SED for the vanilla model, we would have gotten delays in the dip and the start of the absorption just similar to the SRI case. Accordingly, the amplitude of the absorption dip would have been reduced. Furlanetto (2006) tested the vanilla model using both the Population II and Population III SEDs to observe this tendency. (4) The Lyα pumping efficiency of SRII in the EDGES absorption trough window is almost constant at the restricted value xα ≃ 0.1 − 0.5, and xα is bound to < 1 at z ≳ 14. This is mainly caused by the combined effect of the strongly self-regulated SFRD by the LW feedback and the smallness of ${({N}_{\alpha L}/{N}_{\mathrm{ion}})}_{\mathrm{Pop}\mathrm{III}}$, both being relevant to MHs that dominate this epoch.

Figure 10.

Figure 10. 21 cm backgrounds (top panel) and the Lyα coupling coefficients (bottom panel) of models that fit the PLD best (those marked by arrows in Figure 7), selected from the vanilla and SRI models. The shaded regions indicate the redshift bin of the 500 mK absorption trough claimed by the EDGES observation. The EDGES data is shown by the orange, thin solid line.

Standard image High-resolution image
Figure 11.

Figure 11. 21 cm backgrounds (top panel), the Lyα coupling coefficients (middle panel), and xLW ( = JLW/JLW,th) (bottom panel) of models that fit PLD best (those marked by arrows in Figure 7), selected from SRII. In order to avoid crowdedness, we take only four models among those that match the PLD best. Regardless of the difference in values of JLW,th, MHs self-regulate star formation such that xLW ≃ 1 is maintained after JLW reaches ∼ JLW,th. This tendency continues until ACHs take over to dominate in contributing to JLW by generating stars without being hindered by the LW feedback. At the same time, this self-regulation limits xα ≲ 0.5, with additive dependence on ${f}_{\mathrm{esc}}^{{\rm{M}}}$.

Standard image High-resolution image

Compared to the signal interpreted by the EDGES team, none of the global δ Tb 's of our models can match that of EDGES in amplitude and spectral shape. Among the tested models, V-L_dF provides δ Tb closest to the EDGES data. However, this is as close as one can get to the amplitude of the EDGES absorption trough in the ΛCDM, because the case is tuned to produce the coldest TK and the largest xα , with xα ≃ 5–20 in the EDGES window, with a reionization history well under the PLD constraint. The SRI models are terrible in matching the EDGES data, mainly due to the shift of the absorption dip caused by the nature of the Population III SED of LMACH stars dominating this era. The SRII models are as bad as the SRI models in matching the EDGES data within the trough window, mainly due to the dominance of MH stars with Population III SEDs. Note that all of the cases are tuned to produce maximum possible $\left|\delta {T}_{b}\right|$ in absorption with ${f}_{X}^{{\rm{L}}}=0$ (SRI) and ${f}_{X}^{{\rm{L}}}={f}_{X}^{{\rm{M}}}=0$ (SRII) with nonzero ${f}_{X}^{{\rm{H}}}$. Any addition of nonzero X-ray heating will increase TK from these null-heating cases to reduce $\left|\delta {T}_{b}\right|$ and worsen the mismatch between the EDGES' interpretation and the theory.

The SRII models produce δ Tb that is the most peculiar in spectral shape, because across the full EDGES window (14 ≲ z ≲ 24), δ Tb (z) is almost featureless without much variation. It would even be possible that the process of foreground removal, utilizing the spectral smoothness of the foreground, from the observed signal could completely remove the true EoR (or dark ages) signal at z ≳ 14 or ν ≲ 95 MHz to yield only a null result. The limited amplitude of the absorption depth, $\left|\delta {T}_{b}(z\gtrsim 14)\right|\lesssim 60\,\mathrm{mK}$, and the featureless spectral shape designate our SRII model category as the one that disagrees with the EDGES data most. Considering the excellent agreement of many two-phase reionization models in the SRII category with the PLD polarization data, rather extreme alternative explanations to the standard model are required to explain the EDGES data in order to accept SRII. If SRII were the right model, the excess radio background or other alternatives should (1) almost solely contribute to the absorption trough of ∼500 mK because even the small (≲60 mK) absorption signal is likely to be removed in the foreground-removal process and (2) offset the possible absorption depth of δ Tb ∼ − 100 mK at z ∼ 12, in case X-ray heating is inefficient. Obviously, independent observations such as the 21 cm intensity mapping by radio interferometers will help to settle this issue.

We note that the limited amplitude and the featureless spectral shape of the global δ Tb (z) in the SRII models are a novel result, and such features are in large disagreement with other studies that also implement the LW feedback on Population III stars inside MHs to investigate its impact on δ Tb (Mirocha et al. 2018; Mirocha & Furlanetto 2019; Mebane et al. 2020; Qin et al. 2020a, 2021). Let us explain the major reason for such a discrepancy. A big difference lies in this work and others in the manner that LW feedback is implemented. As described in Section 2.2.3, we assume that there exists a threshold value of JLW such that star formation inside MHs is fully suppressed as long as JLW > JLW,th, based on the observed abrupt change in ${M}_{\min }$ (the minimum mass of star-forming halos) and ${T}_{\mathrm{vir},\min }$ (the minimum virial temperature of star-forming halos) of star-forming MHs as JLW varies across the value JLW ∼ 0.1 × 10−21 erg s−1 cm−2 Hz−1 sr−1 in Yoshida et al. (2003) and O'Shea & Norman (2008). In contrast, other studies usually adopt a prescription where ${M}_{\min }$ and consequently ${T}_{\mathrm{vir},\min }$ are smooth functions of JLW. While numerical coefficients vary somewhat in the literature (e.g., see difference between Fialkov et al. 2013 and Schauer et al. 2020), the commonly used functional form for ${M}_{\min }$ is either a redshift-independent one (advocated by Machacek et al. 2001 and Wise & Abel 2007),

Equation (33)

where J21JLW/(10−21 erg s−1 cm−2 Hz−1 sr−1), or a redshift-dependent one (Fialkov et al. 2013; Visbal et al. 2020; Qin et al. 2021),

Equation (34)

where f(vbc; z) ≥ 1 is an additional factor accounting for the suppression of star formation due to the baryon–dark matter streaming velocity (Tseliakhovich & Hirata 2010; Ahn 2016). If one uses one of Equations (33) and (34), the effective JLW,th is much larger than J21 = 0.1 and thus star formation inside MHs will become much stronger than our prescription at a given JLW. Therefore, studies adopting Equation (34), e.g., Fialkov et al. (2013), Qin et al. (2021), and Visbal et al. (2020), find Lyα intensity much stronger than our prediction, xα (z ≳ 14) ≲ 0.5, in SRII models. Consequently, these studies find that MH stars cause the absorption dip to occur much earlier at z ∼ 16–18 than in cases without MH stars (Mebane et al. 2020; Qin et al. 2020a) and with amplitude easily reaching $\left|\delta {T}_{b}\right|\gtrsim 100\,\mathrm{mK}$. However, we stress that if one focuses on ${T}_{\mathrm{vir},\min }$, O'Shea & Norman (2008) clearly shows ${T}_{\mathrm{vir},\min }\simeq 8000\,{\rm{K}}$ when J21 ≃ 0.1. If we take this result and extrapolate to any other redshift, one can instead conclude that JLW,th should lie around J21 = 0.1. In this paper, we parameterize JLW,th but to a limited value of JLW,th,21 ≤ 0.1, respecting the results of Yoshida et al. (2003) and O'Shea & Norman (2008). Of course, one cannot exclude one LW feedback scheme against another at the moment, because there is practically no observational constraint on ultra-high-z (z ≳ 14) radiation sources.

The success of our SRII models in producing two-stage reionization models, which are somewhat favored by PLD against the vanilla and SRI models, can be taken as a hint that SRII models may represent the reality, including how the LW feedback operates in nature. If this were true, the gross disagreement of the global δ Tb (z) of SRII models with the EDGES result would be hardly conceivable in the standard model, or the EDGES result could be an incorrect claim. Further study is warranted.

4. Summary and Conclusion

We studied three types of reionization models semi-analytically: the vanilla, SRI, and SRII models. The SRI and SRII models implement the negative feedback effects on star formation inside LMACHs (SRI, SRII) and MHs (SRII only), namely the Jeans-mass filtering due to photoionization (LMACH, MH) and the LW feedback (MH). As long as LMACHs and MHs host Population III stars and dominate the era of z ≳ 14, we find that δ Tb is constrained to −50 mK ≲ δ Tb < 0 in both the SRI and SRII models, which is in stark contrast with the ∼500 mK absorption trough claimed by EDGES. δ Tb 's predicted by SRII models are almost featureless in their spectral shape for z ≳ 14 (ν ≲ 95 MHz) due to the strong self-regulation of star formation inside MHs by the LW background built up my MHs.

At this stage, with the PLD being the most accurate large-scale-CMB anisotropy observation, we find that all three models can provide acceptable reionization scenarios. In particular, we find that SRII models with substantial high-redshift (z ≳ 15) ionization tails are as favored as those models without such tails, if corresponding ${C}_{l}^{\mathrm{EE}}$ is analyzed against the ${C}_{l}^{\mathrm{EE}}$ of the PLD. SRII models have the two-stage ionization feature, the high-redshift slow-ionization stage, and the low-redshift fast-ionization stage, which seems more favored by PLD than tailless reionization models, even though this tendency is largely uncertain in PLD. In conclusion, SRII models with substantial high-z ionization tails should NOT be ruled out as claimed by Planck Collaboration et al. (2020). Our ${C}_{l}^{\mathrm{EE}}$-only analysis favors models with τes ≃ 0.055–0.064, which is relatively larger than the ${C}_{l}^{\mathrm{EE}}$-only inference by Planck Collaboration et al. (2020), ${\tau }_{\mathrm{es}}={0.0504}_{-0.0079}^{+0.0050}$. This issue needs to be investigated further.

In light of the peculiarity of δ Tb but the good agreement with PLD-${C}_{l}^{\mathrm{EE}}$ and the hint of two-stage reionization, we stress that SRII models should be considered more seriously. Such a disagreement with the EDGES data requires a change in the standard ΛCDM model or the "interpretation" of the EDGES data. Even though many statistical analyses have already been carried out by accepting the full result (e.g., Mebane et al. 2020) or a part of the result (Qin et al. 2020a; only the redshift window of the absorption trough is taken as the possible location for the absorption dip) of EDGES, we question the foreground-removal scheme of the EDGES team, as Hills et al. (2018) and Tauscher et al. (2020) did. The claimed signal is indeed a result of extracting an arbitrary smooth signal from the residual signal (Figure 1(b) in Bowman et al. 2018) after the foreground removal. There is no guarantee that a combination of log-power-law spectral curves could completely remove the foreground, and such an additional arbitrary removal seems even more dubious. It would be even harder to probe the EoR signal if nature were indeed described by our SRII models, because the featureless spectral shape at z ≳ 14 of the models is going to be removed by any foreground-removal scheme utilizing the spectral smoothness. In this regard, probing the dipole anisotropy of δ Tb seems very promising (Deshpande 2018; Trombetti et al. 2021), because the dipole moment is caused by the monopole moment seen by an observer with a peculiar motion and thus can be an independent probe of the global δ Tb .

It is therefore crucial to carry out higher-precision CMB observations and the 21 cm intensity mapping to further constrain the reionization history. A superb CMB polarization apparatus limited only by the cosmic variance (for all 2 ≤ l ≲ 200), LiteBIRD (Lite (Light) satellite for the studies of B-mode polarization and Inflation from cosmic background Radiation Detection), is under way, which will significantly sharpen the constraint on the reionization history. Many different reionization histories can share the same τes and zov (see, e.g., Figure 4(b) in Ahn et al. 2012), and breaking this degeneracy seems elusive with the PLD at this low-τes era but will become more feasible with such a high-precision CMB observation. The boost of ${C}_{8\lesssim l\lesssim 24}^{\mathrm{EE}}$ of SRII models relative to that of the vanilla and SRI models is of particular interest. The 21 cm intensity mapping will be able to probe large-scale fluctuation of HI density when large H ii regions are produced during EoR. The high-redshift regime with z ≳ 15 is more in the dark ages without too much ionization in typical reionization scenarios, while some SRII models produce a substantial amount of ionization. Nevertheless, SRII models predict only small-size H ii regions by MHs (Ahn et al. 2012) that might not be imaged individually by radio interferometers. If future CMB observations favored two-stage reionization but radio interferometry did not reveal any noticeable H ii bubbles at high redshift, SRII models would be the strongest candidate to explain both. We will investigate detailed observational prospects in the future.

This work was supported by Korea NRF grant NRF-2016R1D1A1B04935414 and a research grant from Chosun University (2016).

Footnotes

  • 3  

    If the particle-splitting scheme or the adaptive mesh refinement scheme is used, one may in principle recover the time-varying mass criterion for MH and ACH determination.

  • 4  

    Reionization histories of SRII shown in Figures 45 reach smaller values of the midway peak x at z ≃ 15, and the recombination is stronger than matching models of Ahn et al. (2012). The individual H ii regions created by MHs were too small to be numerically resolved in the simulation of Ahn et al. (2012), and thus the grid cells with given resolution were partially ionized before ACHs emerged. Recombination rate per hydrogen in each grid cell was calculated as α C(nH + nHe)x2, even though the rate should have been α C(nH + nHe)x instead (as in Equation (5)) because UV-driven H ii regions are practically fully ionized and surrounded by the neutral IGM. We experimentally calculated x(t) after changing the sink term in Equation (5) to α C(nH + nHe)x2, and could recover the global ionization histories of Ahn et al. (2012) with matching parameters. Therefore, the quantitative predictions of Ahn et al. (2012) need to be modified to some extent or considered as models that have a smoother transition of stars from Population III to Population II than the SRII models studied here.

  • 5  

    ${\tilde{D}}_{l}^{\mathrm{EE}}$ and σl are from "COM_PowerSpect_CMB-EE-full_R3.01.txt," downloadable from the Planck Legacy Archive (https://pla.esac.esa.int).

  • 6  

    We note that Ghara & Mellema (2020) claims that the Lyα heating is efficient enough to render δ Tb > 0 before the end of reionization. Even though this claim agrees with that of Chuzhoy & Shapiro (2007) qualitatively, the heating rate by Ghara & Mellema (2020) is wrongfully overestimated and should be reduced by about an order of magnitude as clarified by Mittal & Kulkarni (2021). Except for the assumed SFRD and the SED, Mittal & Kulkarni (2021) is basically identical to and a reproduction of Chuzhoy & Shapiro (2007).

Please wait… references are loading.
10.3847/1538-4357/abf3bf