The Predicament of Absorption-dominated Reionization II:
Observational Estimate of the Clumping Factor at the End of Reionization

Frederick B. Davies Max-Planck-Institut für Astronomie, Königstuhl 17, D-69117 Heidelberg, Germany Sarah E. I. Bosman Institute for Theoretical Physics, Heidelberg University, Philosophenweg 12, D–69120, Heidelberg, Germany Max-Planck-Institut für Astronomie, Königstuhl 17, D-69117 Heidelberg, Germany Steven R. Furlanetto Department of Physics & Astronomy, University of California, Los Angeles, CA 90095, USA
Abstract

The history of reionization reflects the cumulative injection of ionizing photons by sources and the absorption of ionizing photons by sinks. The latter process is traditionally described in terms of a “clumping factor” which encodes the average quadratic increase in the recombination rate of dense gas within the cosmic web. The recent measurement of a short mean free path of ionizing photons from stacked quasar spectra at z6similar-to-or-equals𝑧6z\simeq 6italic_z ≃ 6 has placed the importance of sinks under increased scrutiny, but its connection to the recombination rate is not immediately obvious. Here we present analytic arguments to connect the clumping factor to the mean free path by invoking ionization equilibrium within the ionized phase of the intergalactic medium at the end of (and after) reionization. We find that the latest mean free path and hydrogen photoionization rate measurements at z=5𝑧5z=5italic_z = 56666 imply a global clumping factor C12𝐶12C\approx 12italic_C ≈ 12, much higher than previous determinations from radiation-hydrodynamic simulations of the reionization process. Similar values of C𝐶Citalic_C are also derived when applying the same procedure to observations at 2<z<52𝑧52<z<52 < italic_z < 5. Compared to the traditional assumption of C=3𝐶3C=3italic_C = 3, high-redshift galaxies must produce roughly twice as many ionizing photons (3absent3\approx 3≈ 3 photons per baryon) to reionize the universe by z6similar-to𝑧6z\sim 6italic_z ∼ 6. This additional requirement on the ionizing photon budget may help to reconcile the reionization history with JWST observations that suggest a far greater output of ionizing photons by the most distant galaxy populations.

Intergalactic medium(813), Reionization(1383)

1 Introduction

After the formation of baryons during the Big Bang, and their subsequent (re-)combination into atoms and the release of the cosmic microwave background (CMB), the hydrogen and helium in the Universe persisted in a predominantly neutral state. After the formation of the first stars and galaxies, the ionizing photons emitted by massive stars began to carve ionized bubbles into the surrounding intergalactic medium (IGM), beginning the epoch of reionization. The ionized bubbles from individual galaxies eventually merged (Furlanetto et al., 2004), filling more and more of the cosmic volume until, by z5.3similar-to𝑧5.3z\sim 5.3italic_z ∼ 5.3 (Bosman et al., 2022), the IGM was fully reionized.

The most straightforward quantitative description of reionization was put forward by Madau et al. (1999), whose “one-zone” model for the process provides valuable intuition:

dQdt=n˙ionnHQtrec,𝑑𝑄𝑑𝑡subscript˙𝑛iondelimited-⟨⟩subscript𝑛H𝑄subscript𝑡rec\frac{dQ}{dt}=\frac{\dot{n}_{\rm ion}}{\langle n_{\rm H}\rangle}-\frac{Q}{t_{% \rm rec}},divide start_ARG italic_d italic_Q end_ARG start_ARG italic_d italic_t end_ARG = divide start_ARG over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT end_ARG start_ARG ⟨ italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ⟩ end_ARG - divide start_ARG italic_Q end_ARG start_ARG italic_t start_POSTSUBSCRIPT roman_rec end_POSTSUBSCRIPT end_ARG , (1)

where Q𝑄Qitalic_Q is the ionized fraction of the IGM, and the first and second terms on the right-hand side represent the source and sink terms, respectively. The sources are represented by n˙ionsubscript˙𝑛ion\dot{n}_{\rm ion}over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT, the emissivity of ionizing photons, while the sinks are represented by trecsubscript𝑡rect_{\rm rec}italic_t start_POSTSUBSCRIPT roman_rec end_POSTSUBSCRIPT, the recombination timescale of the ionized gas. In this work, as in Madau et al. (1999), we will assume that Q𝑄Qitalic_Q represents the volume-averaged ionized fraction. While the inside-out nature of reionization implies that a mass-averaged approach may be more appropriate, in which case an additional factor of Q𝑄Qitalic_Q arises in the sink term (e.g. Chen et al. 2020), equation (1) is approximately correct in a two-phase approximation (cf. Wu et al. 2021) where the IGM is either fully ionized (xHII=1subscript𝑥HII1x_{\rm HII}=1italic_x start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT = 1) or neutral (xHII=0subscript𝑥HII0x_{\rm HII}=0italic_x start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT = 0). That is, Q𝑄Qitalic_Q represents the volume of the IGM within the ionized phase rather than a physical ionized fraction, and so it does not asymptote to a finite residual neutral fraction after reionization is complete (although see Madau 2017 for a solution to this).

Considerable effort has been undertaken to determine n˙ionsubscript˙𝑛ion\dot{n}_{\rm ion}over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT at early cosmic time. Such determinations typically involve a measurement of the UV luminosity function (LF) of galaxies at high redshifts (z>6𝑧6z>6italic_z > 6), but the connection between the UV LF and the ionizing output of galaxies is still uncertain. This connection is usually parameterized as

n˙ion=ρUVξionfesc,subscript˙𝑛ionsubscript𝜌UVsubscript𝜉ionsubscript𝑓esc\dot{n}_{\rm ion}=\rho_{\rm UV}\xi_{\rm ion}f_{\rm esc},over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT = italic_ρ start_POSTSUBSCRIPT roman_UV end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT , (2)

where ρUVsubscript𝜌UV\rho_{\rm UV}italic_ρ start_POSTSUBSCRIPT roman_UV end_POSTSUBSCRIPT is the integral over the UV LF down to some magnitude limit, ξionsubscript𝜉ion\xi_{\rm ion}italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT is the “ionizing efficiency” which represents the average (intrinsic) spectral shape of the stellar populations between the ionizing and non-ionizing UV continuum, and fescsubscript𝑓escf_{\rm esc}italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT is the escape fraction of ionizing photons from the galaxies into the IGM. While observations of galaxy nebular emission lines can constrain their ξionsubscript𝜉ion\xi_{\rm ion}italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT (e.g. Ning et al. 2023; Prieto-Lyon et al. 2023; Simmonds et al. 2023, 2024; Atek et al. 2024), direct measurements of fescsubscript𝑓escf_{\rm esc}italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT (i.e., direct detections of ionizing photons) are hindered by the high opacity of the Lyman-series forests, and thus are only possible at redshifts z4less-than-or-similar-to𝑧4z\lesssim 4italic_z ≲ 4 (e.g. Izotov et al. 2016; Vanzella et al. 2018; Ji et al. 2020; Pahl et al. 2021).

The role of sinks has been a subject of considerable debate over the years. As mentioned above, sinks of ionizing photons enter the Madau et al. (1999) formalism via trecsubscript𝑡rect_{\rm rec}italic_t start_POSTSUBSCRIPT roman_rec end_POSTSUBSCRIPT, the average recombination time of a proton in the IGM, which can be written as

trec=1CnHαHII(T),subscript𝑡rec1𝐶delimited-⟨⟩subscript𝑛Hsubscript𝛼HII𝑇t_{\rm rec}=\frac{1}{C\langle n_{\rm H}\rangle\alpha_{\rm HII}(T)},italic_t start_POSTSUBSCRIPT roman_rec end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_C ⟨ italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ⟩ italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT ( italic_T ) end_ARG , (3)

where αHIIsubscript𝛼HII\alpha_{\rm HII}italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT is the hydrogen recombination rate, and nHdelimited-⟨⟩subscript𝑛H\langle n_{\rm H}\rangle⟨ italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ⟩ is the mean cosmic hydrogen density. Crucially, as recombination is a collisional process, the rate depends on the squared density of ionized gas. It is common to summarize this dependence with the so-called “clumping factor,”

Cn2/n2,𝐶delimited-⟨⟩superscript𝑛2superscriptdelimited-⟨⟩𝑛2C\equiv\langle n^{2}\rangle/\langle n\rangle^{2},italic_C ≡ ⟨ italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_n ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (4)

such that trec=trecuniform/Csubscript𝑡recsuperscriptsubscript𝑡recuniform𝐶t_{\rm rec}=t_{\rm rec}^{\rm uniform}/Citalic_t start_POSTSUBSCRIPT roman_rec end_POSTSUBSCRIPT = italic_t start_POSTSUBSCRIPT roman_rec end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_uniform end_POSTSUPERSCRIPT / italic_C. As there is no analytic shortcut to fully determine C𝐶Citalic_C from first principles, the assumed value is typically derived from cosmological simulations.

Early cosmological hydrodynamical simulations suggested C30similar-to𝐶30C\sim 30italic_C ∼ 30 (e.g. Gnedin & Ostriker 1997), implying a dominant role of recombinations in determining the reionization history. However, this high value considered all gas particles in the simulation volume. In practice, recombinations occurring inside of galaxies are already accounted for by the fescsubscript𝑓escf_{\rm esc}italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT parameter, so care must be taken to avoid double-counting. Later works by Iliev et al. (2005, 2007) and Raičević & Theuns (2011) took this exclusion into account and found values closer to C10similar-to𝐶10C\sim 10italic_C ∼ 10, albeit using dark-matter-only N-body simulations. But the baryons are also subject to gas pressure, which smooths their distribution relative to the dark matter field. Following several works employing radiation hydrodynamic simulations of reionization (Pawlik et al., 2009; Shull et al., 2012; Finlator et al., 2012, see also McQuinn et al. 2011), a value of C2similar-to𝐶2C\sim 2italic_C ∼ 23333 is now a typical assumption in analytic reionization models in the literature. That said, more recent simulations suggest a value of C𝐶Citalic_C higher by a factor of 2similar-toabsent2\sim 2∼ 2 (e.g. Chen et al. 2020; Kannan et al. 2022).

The impact of sinks is now being revisited after the measurement of a short mean free path of ionizing photons at z=6𝑧6z=6italic_z = 6 by Becker et al. (2021), and further confirmed by Zhu et al. (2023, see also ), which suggests the presence of substantially more small-scale structure in the IGM than present in traditional reionization simulations. A short mean free path can dramatically increase the number of ionizing photons required to ionize the IGM, as photons must travel long distances from ionizing sources to large-scale voids (Davies et al. 2021, henceforth Paper I, see also Cain et al. 2021). However, the quantitative connection between the short mean free path and the clumping factor is not immediately obvious (e.g. Cain et al. 2023), which has limited the extent to which the short mean free path has been taken into account by the larger reionization community.

In this work, we aim to build a stronger connection between the mean free path and clumping factor at high redshift, in an attempt to unify the description of ionizing photon sinks during the reionization epoch. We first show that the clumping factor at a given redshift can be estimated from the ionizing background intensity and mean free path. We then apply this methodology to measurements of these quantities across cosmic time, finding a nearly constant value of C𝐶Citalic_C which is several times higher than typically assumed.

We assume a Planck ΛΛ\Lambdaroman_ΛCDM cosmology (Planck Collaboration et al., 2020) with h=0.680.68h=0.68italic_h = 0.68, Ωm=0.31subscriptΩ𝑚0.31\Omega_{m}=0.31roman_Ω start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = 0.31, and Ωb=0.049subscriptΩ𝑏0.049\Omega_{b}=0.049roman_Ω start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 0.049.

2 The clumping factor and the mean free path

In this section, we will investigate the connection between the clumping factor and the mean free path of ionizing photons. But first, we must define what we mean by “clumping factor,” as its exact definition varies considerably between different works. Here we define the clumping factor to be the relevant clumping factor for solving equation (1) – i.e. the clumping factor that provides the correct globally averaged recombination rate, but where effects inside of galaxies that give rise to the escape fraction in the definition of n˙ionsubscript˙𝑛ion\dot{n}_{\rm ion}over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT are ignored. Specifically, we assume that C𝐶Citalic_C is the constant of proportionality between the true global (external) recombination rate n˙rec=nenHIIαHIIsubscript˙𝑛recdelimited-⟨⟩subscript𝑛𝑒subscript𝑛HIIsubscript𝛼HII\dot{n}_{\rm rec}=\langle n_{e}n_{\rm HII}\alpha_{\rm HII}\rangleover˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_rec end_POSTSUBSCRIPT = ⟨ italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT ⟩ and the recombination rate at the cosmic mean density,

CnenHIIαHIInenHIIαHII𝐶delimited-⟨⟩subscript𝑛𝑒subscript𝑛HIIsubscript𝛼HIIdelimited-⟨⟩subscript𝑛𝑒delimited-⟨⟩subscript𝑛HIIdelimited-⟨⟩subscript𝛼HIIC\equiv\frac{\langle n_{e}n_{\rm HII}\alpha_{\rm HII}\rangle}{\langle n_{e}% \rangle\langle n_{\rm HII}\rangle\langle\alpha_{\rm HII}\rangle}italic_C ≡ divide start_ARG ⟨ italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT ⟩ end_ARG start_ARG ⟨ italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ⟩ ⟨ italic_n start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT ⟩ ⟨ italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT ⟩ end_ARG (5)

where we assume that the fiducial recombination coefficient in the denominator is equal to the Case B recombination rate for 10,0001000010,00010 , 000 K gas111We note that αHIIB(T=10,000K)αHIIA(T=20,000K)similar-to-or-equalssuperscriptsubscript𝛼HII𝐵𝑇10000Ksuperscriptsubscript𝛼HII𝐴𝑇20000K\alpha_{\rm HII}^{B}(T=10,000\,{\rm K})\simeq\alpha_{\rm HII}^{A}(T=20,000\,{% \rm K})italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_T = 10 , 000 roman_K ) ≃ italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_T = 20 , 000 roman_K ), another common assumption in previous works., αHII=αHIIB(T=10,000K)delimited-⟨⟩subscript𝛼HIIsuperscriptsubscript𝛼HII𝐵𝑇10000K\langle\alpha_{\rm HII}\rangle=\alpha_{\rm HII}^{B}(T=10,000\,{\rm K})⟨ italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT ⟩ = italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_T = 10 , 000 roman_K ).

The most straightforward way to connect the clumping factor to the mean free path starts with the assumption of photoionization equilibrium, which should generally hold in the ionized IGM,

nHIΓHI=nenHIIαHII=Cχe(1xHI)2nH2αHII,subscript𝑛HIsubscriptΓHIdelimited-⟨⟩subscript𝑛𝑒subscript𝑛HIIsubscript𝛼HII𝐶subscript𝜒𝑒superscript1subscript𝑥HI2superscriptsubscript𝑛H2subscript𝛼HIIn_{\rm HI}\Gamma_{\rm HI}=\langle n_{e}n_{\rm HII}\alpha_{\rm HII}\rangle=C% \chi_{e}(1-x_{\rm HI})^{2}n_{\rm H}^{2}\alpha_{\rm HII},italic_n start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT = ⟨ italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT ⟩ = italic_C italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( 1 - italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT , (6)

where ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT is the photoionization rate of hydrogen and χe1.08subscript𝜒𝑒1.08\chi_{e}\approx 1.08italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ≈ 1.08 is the enhancement in the number of free electrons due to ionized helium222We assume that helium is singly-ionized at the same time as hydrogen, and that the (second) reionization of helium has not yet begun. While this assumption will become incorrect at z4less-than-or-similar-to𝑧4z\lesssim 4italic_z ≲ 4 (e.g. Worseck et al. 2019), the additional 8% boost to the electron density from the second ionization of helium is small compared to the uncertainties in the observed quantities we employ in § 3.2.. Solving for C𝐶Citalic_C, we have

C=xHInHΓHIχe(1xHI)2nH2αHII.𝐶subscript𝑥HIsubscript𝑛HsubscriptΓHIsubscript𝜒𝑒superscript1subscript𝑥HI2superscriptsubscript𝑛H2subscript𝛼HIIC=\frac{x_{\rm HI}n_{\rm H}\Gamma_{\rm HI}}{\chi_{e}(1-x_{\rm HI})^{2}n_{\rm H% }^{2}\alpha_{\rm HII}}.italic_C = divide start_ARG italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT end_ARG start_ARG italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( 1 - italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT end_ARG . (7)

The two unknowns in this expression are ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT and xHIsubscript𝑥HIx_{\rm HI}italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT. While the former can be derived from observations of the Lyα𝛼\alphaitalic_α forest, and is most sensitive to low-density gas which is unambiguously resolved in simulations, the residual neutral fraction is not so simple to derive, as it is sensitive to self-shielding and geometrical effects in dense gas (e.g. McQuinn et al. 2011; Erkal 2015).

To proceed, one can make the simplifying assumption that the neutral fraction is connected to the mean free path of ionizing photons via λmfp=(nHIσHI)1subscript𝜆mfpsuperscriptsubscript𝑛HIsubscript𝜎HI1\lambda_{\rm mfp}=(n_{\rm HI}\sigma_{\rm HI})^{-1}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT = ( italic_n start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. In this case, similar to the “effective” clumping factor in Cain et al. (2023), we have:

C=σHIΓHIλmfpχe(1xHI)2nH2αHII,𝐶subscript𝜎HIsubscriptΓHIsubscript𝜆mfpsubscript𝜒𝑒superscript1subscript𝑥HI2superscriptsubscript𝑛H2subscript𝛼HIIC=\frac{\sigma_{\rm HI}\Gamma_{\rm HI}}{\lambda_{\rm mfp}\chi_{e}(1-x_{\rm HI}% )^{2}n_{\rm H}^{2}\alpha_{\rm HII}},italic_C = divide start_ARG italic_σ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT end_ARG start_ARG italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( 1 - italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT end_ARG , (8)

where we note that both the mean free path and photoionization cross section terms implicitly represent frequency-averaged values. This expression for the mean free path, however, makes the crucial assumption that neutral hydrogen is uniformly distributed in space. In reality, the dense self-shielded gas that gives rise to optically-thick absorption should be inhomogeneous, distributed in clumps and/or in the filaments of the cosmic web (e.g. McQuinn et al. 2011).

Another way to associate the clumping factor with the mean free path was suggested by Emberson et al. (2013), who connected the attenuation of ionizing photon flux to the recombination rate. In this model, one considers the attenuation of ionizing photon flux dF𝑑𝐹dFitalic_d italic_F inside a slab of material with area dA𝑑𝐴dAitalic_d italic_A and proper width ds𝑑𝑠dsitalic_d italic_s compared to the recombinations inside said slab,

dFdA=CnenHIIαHIIdAds,𝑑𝐹𝑑𝐴𝐶subscript𝑛𝑒subscript𝑛HIIsubscript𝛼HII𝑑𝐴𝑑𝑠-dF\,dA=Cn_{e}n_{\rm HII}\alpha_{\rm HII}\,dA\,ds,- italic_d italic_F italic_d italic_A = italic_C italic_n start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_d italic_A italic_d italic_s , (9)

where the flux inside the slab is attenuated following

dFds=F(1+z)/λmfp,𝑑𝐹𝑑𝑠𝐹1𝑧subscript𝜆mfp\frac{dF}{ds}=-F(1+z)/\lambda_{\rm mfp},divide start_ARG italic_d italic_F end_ARG start_ARG italic_d italic_s end_ARG = - italic_F ( 1 + italic_z ) / italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT , (10)

where the (1+z)1𝑧(1+z)( 1 + italic_z ) term converts λmfpsubscript𝜆mfp\lambda_{\rm mfp}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT from comoving to proper units. The following expression can then be derived after solving for C𝐶Citalic_C,

C=F(1+z)λmfpχe(1xHI)2nH2αHII,𝐶𝐹1𝑧subscript𝜆mfpsubscript𝜒𝑒superscript1subscript𝑥HI2superscriptsubscript𝑛H2subscript𝛼HIIC=\frac{F(1+z)}{\lambda_{\rm mfp}\chi_{e}(1-x_{\rm HI})^{2}n_{\rm H}^{2}\alpha% _{\rm HII}},italic_C = divide start_ARG italic_F ( 1 + italic_z ) end_ARG start_ARG italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( 1 - italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT end_ARG , (11)

where we note again that the F𝐹Fitalic_F and λmfpsubscript𝜆mfp\lambda_{\rm mfp}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT terms represent values integrated over the spectrum of the ionizing background. While this expression improves upon the previous one by removing the explicit connection between the mean free path and the neutral fraction, the geometrical assumption in its derivation (i.e. the slab) introduces additional ambiguity.

We suggest a third way to conceptualize (and quantify) the connection between the mean free path and the clumping factor. A typical ionizing photon passing through the IGM will travel one mean free path before being absorbed, i.e. before ionizing a hydrogen atom. Thus a “photon photo-ionization rate,” the rate at which a given ionizing photon will ionize a hydrogen atom, can be written as Γγ=c/λmfpsubscriptΓ𝛾𝑐subscript𝜆mfp\Gamma_{\gamma}=c/\lambda_{\rm mfp}roman_Γ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT = italic_c / italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT. The space density of photoionizations is then nγΓγ=nγ×c/λmfpsubscript𝑛𝛾subscriptΓ𝛾subscript𝑛𝛾𝑐subscript𝜆mfpn_{\gamma}\Gamma_{\gamma}=n_{\gamma}\times c/\lambda_{\rm mfp}italic_n start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT roman_Γ start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT × italic_c / italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT, where nγsubscript𝑛𝛾n_{\gamma}italic_n start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT is the number density of ionizing photons. In ionization equilibrium, this rate will balance the recombination rate, i.e.

nγcλmfp=CαHIIχe(1xHI)2nH2,subscript𝑛𝛾𝑐subscript𝜆mfp𝐶subscript𝛼HIIsubscript𝜒𝑒superscript1subscript𝑥HI2superscriptsubscript𝑛𝐻2\frac{n_{\gamma}c}{\lambda_{\rm mfp}}=C\alpha_{\rm HII}\chi_{e}(1-x_{\rm HI})^% {2}n_{H}^{2},divide start_ARG italic_n start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_c end_ARG start_ARG italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT end_ARG = italic_C italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( 1 - italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (12)

Solving for C𝐶Citalic_C as above, we find

C=nγcλmfpαHIIχe(1xHI)2nH2,𝐶subscript𝑛𝛾𝑐subscript𝜆mfpsubscript𝛼HIIsubscript𝜒𝑒superscript1subscript𝑥HI2superscriptsubscript𝑛H2C=\frac{n_{\gamma}c}{\lambda_{\rm mfp}\alpha_{\rm HII}\chi_{e}(1-x_{\rm HI})^{% 2}n_{\rm H}^{2}},italic_C = divide start_ARG italic_n start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT italic_c end_ARG start_ARG italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( 1 - italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (13)

where again the nγsubscript𝑛𝛾n_{\gamma}italic_n start_POSTSUBSCRIPT italic_γ end_POSTSUBSCRIPT and λmfpsubscript𝜆mfp\lambda_{\rm mfp}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT terms represent frequency-averaged quantities.

In all three cases, the inferred clumping factor is proportional to the strength of the ionizing background (in various forms) divided by the mean free path of ionizing photons, e.g. CΓHI/λmfpproportional-to𝐶subscriptΓHIsubscript𝜆mfpC\propto\Gamma_{\rm HI}/\lambda_{\rm mfp}italic_C ∝ roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT / italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT. All three methods result in quantitatively similar values for C𝐶Citalic_C; in this work, we adopt the third method to compute C𝐶Citalic_C (i.e. equation 13), as it appears to have the fewest explicit assumptions on the nature of the distribution of neutral gas.

We have so far ignored the dependence on photon frequency of various quantities in the expressions for C𝐶Citalic_C above for the sake of clarity, but due to the steep frequency dependence of the photoionization cross-section (Verner et al., 1996), such terms could matter at the level of a factor of a few. We write the specific number density of ionizing photons nνsubscript𝑛𝜈n_{\nu}italic_n start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT as

nν=uνhν=4πcJνhν,subscript𝑛𝜈subscript𝑢𝜈𝜈4𝜋𝑐subscript𝐽𝜈𝜈n_{\nu}=\frac{u_{\nu}}{h\nu}=\frac{4\pi}{c}\frac{J_{\nu}}{h\nu},italic_n start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT = divide start_ARG italic_u start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG start_ARG italic_h italic_ν end_ARG = divide start_ARG 4 italic_π end_ARG start_ARG italic_c end_ARG divide start_ARG italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG start_ARG italic_h italic_ν end_ARG , (14)

where uνsubscript𝑢𝜈u_{\nu}italic_u start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT is the specific energy density and Jνsubscript𝐽𝜈J_{\nu}italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT is the specific angle-averaged mean intensity of the ionizing background. We then proceed to estimate C𝐶Citalic_C using the following expression:

C=[4πνHI4νHIJνhνλνdν]×1αHIIχe(1xHI)2nH2,𝐶delimited-[]4𝜋superscriptsubscriptsubscript𝜈HI4subscript𝜈HIsubscript𝐽𝜈𝜈subscript𝜆𝜈differential-d𝜈1subscript𝛼HIIsubscript𝜒𝑒superscript1subscript𝑥HI2superscriptsubscript𝑛H2C=\left[4\pi\int_{\nu_{\rm HI}}^{4\,\nu_{\rm HI}}\frac{J_{\nu}}{h\nu\lambda_{% \nu}}{\rm d}\nu\right]\times\frac{1}{\alpha_{\rm HII}\chi_{e}(1-x_{\rm HI})^{2% }n_{\rm H}^{2}},italic_C = [ 4 italic_π ∫ start_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT end_POSTSUPERSCRIPT divide start_ARG italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG start_ARG italic_h italic_ν italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG roman_d italic_ν ] × divide start_ARG 1 end_ARG start_ARG italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ( 1 - italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (15)

where νHIsubscript𝜈HI\nu_{\rm HI}italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT is the frequency of the hydrogen ionizing edge. In the following, we make the assumption that the neutral fraction is small enough that the (1xHI)1subscript𝑥HI(1-x_{\rm HI})( 1 - italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) term can be approximated as unity – that is, we compute the clumping factor relative to a fully ionized IGM. We further approximate the frequency dependencies of the mean free path and ionizing background intensity as power laws with λνναλproportional-tosubscript𝜆𝜈superscript𝜈subscript𝛼𝜆\lambda_{\nu}\propto\nu^{\alpha_{\lambda}}italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT and Jνναbproportional-tosubscript𝐽𝜈superscript𝜈subscript𝛼bJ_{\nu}\propto\nu^{-\alpha_{\rm b}}italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT - italic_α start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT end_POSTSUPERSCRIPT, leading to an analytic simplification to equation (15) as long as αb+αλ0subscript𝛼bsubscript𝛼𝜆0\alpha_{\rm b}+\alpha_{\lambda}\neq 0italic_α start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ≠ 0,

C=4πJHIhνHIλmfp[14(αb+αλ)αb+αλ]×1αHIIχenH2,𝐶4𝜋subscript𝐽HIsubscript𝜈HIsubscript𝜆mfpdelimited-[]1superscript4subscript𝛼bsubscript𝛼𝜆subscript𝛼bsubscript𝛼𝜆1subscript𝛼HIIsubscript𝜒𝑒superscriptsubscript𝑛H2C=\frac{4\pi J_{\rm HI}}{h\nu_{\rm HI}\lambda_{\rm mfp}}\left[\frac{1-4^{-(% \alpha_{\rm b}+\alpha_{\lambda})}}{\alpha_{\rm b}+\alpha_{\lambda}}\right]% \times\frac{1}{\alpha_{\rm HII}\chi_{e}n_{\rm H}^{2}},italic_C = divide start_ARG 4 italic_π italic_J start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT end_ARG start_ARG italic_h italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT end_ARG [ divide start_ARG 1 - 4 start_POSTSUPERSCRIPT - ( italic_α start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_α start_POSTSUBSCRIPT roman_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_ARG ] × divide start_ARG 1 end_ARG start_ARG italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT italic_χ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , (16)

where JHIsubscript𝐽HIJ_{\rm HI}italic_J start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT and λmfpsubscript𝜆mfp\lambda_{\rm mfp}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT are Jν(νHI)subscript𝐽𝜈subscript𝜈HIJ_{\nu}(\nu_{\rm HI})italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) and λν(νHI)subscript𝜆𝜈subscript𝜈HI\lambda_{\nu}(\nu_{\rm HI})italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ), respectively.

The frequency dependence assumptions are largely encapsulated by the term in brackets, which is a function of αb+αλsubscript𝛼𝑏subscript𝛼𝜆\alpha_{b}+\alpha_{\lambda}italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT, equal to 2222 with our assumed power-law indices333We note that at z5greater-than-or-equivalent-to𝑧5z\gtrsim 5italic_z ≳ 5, where the mean free path is short relative to the Hubble distance (the “absorption-limited” regime), we can write Jνϵνλνproportional-tosubscript𝐽𝜈subscriptitalic-ϵ𝜈subscript𝜆𝜈J_{\nu}\propto\epsilon_{\nu}\lambda_{\nu}italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ϵ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT where ϵνsubscriptitalic-ϵ𝜈\epsilon_{\nu}italic_ϵ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT is the average ionizing emissivity (Meiksin & White, 2003). The implied scaling of the emissivity is ϵνν(αb+αλ)proportional-tosubscriptitalic-ϵ𝜈superscript𝜈subscript𝛼𝑏subscript𝛼𝜆\epsilon_{\nu}\propto\nu^{-(\alpha_{b}+\alpha_{\lambda})}italic_ϵ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT - ( italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT, suggesting that our choice of αb+αλ=2subscript𝛼𝑏subscript𝛼𝜆2\alpha_{b}+\alpha_{\lambda}=2italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 2 is reasonable (e.g. Becker & Bolton 2013). Even harder ionizing spectra are also plausible for young, metal-poor stellar populations (e.g. D’Aloisio et al. 2019) which would increase our C𝐶Citalic_C estimates.. In practice, as described below in § 3.2, we will convert observational constraints on ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT to JHIsubscript𝐽HIJ_{\rm HI}italic_J start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT, which introduces an additional factor dependent on αbsubscript𝛼𝑏\alpha_{b}italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT alone. Lower (higher) values of C𝐶Citalic_C would be derived if the ionizing background is softer (harder), or if the mean free path is a stronger (weaker) function of frequency. We discuss the behavior of the frequency dependence term further in Appendix A, and note that even in the most extreme case (corresponding to αλ3subscript𝛼𝜆3\alpha_{\lambda}\approx 3italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ≈ 3) our assumptions could only overestimate C𝐶Citalic_C by a factor of two.

As measurements of the ionizing background require non-zero transmission through the highly sensitive Lyα𝛼\alphaitalic_α forest (Gunn & Peterson, 1965), application of this method will only become possible at the very end of the reionization process. Thus the clumping factor during the majority of reionization cannot be constrained directly. At these earlier times, the inside-out nature of reionization should lead to higher densities in the ionized regions, and thus a higher recombination rate relative to the mean IGM (e.g. Chen et al. 2020, see also So et al. 2014). We ignore this effect for simplicity in our analytic calculations, but note that this would likely increase the effective clumping factor significantly at low ionized fractions as in Chen et al. (2020).

3 Estimating the clumping factor from IGM observations

In this section, we use observational properties of the IGM to constrain the clumping factor as defined in the previous section. Here we stress that our clumping factor is an effective quantity corresponding to the entire volume of the IGM, and not a “local” clumping factor that can be applied to the density field in simulations a posteriori (see, e.g., Raičević & Theuns 2011; Kaurov & Gnedin 2015). In particular, our definition of the clumping factor is specifically designed for use in analytic reionization calculations like equation (1) that consider the IGM as a whole (Madau et al., 1999).

Refer to caption
Figure 1: Clumping factor estimates applying equation (16) to constraints on the hydrogen photoionization rate ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT (via equation 17) and ionizing photon mean free path λmfpsubscript𝜆mfp\lambda_{\rm mfp}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT. The grey points at z<5𝑧5z<5italic_z < 5 use ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT from Becker & Bolton (2013) and λmfpsubscript𝜆mfp\lambda_{\rm mfp}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT from Worseck et al. (2014), while the black points at z>5𝑧5z>5italic_z > 5 use the corresponding constraints from Davies et al. (2024) and Zhu et al. (2023), respectively.
Refer to caption
Figure 2: Comparison between the C𝐶Citalic_C estimates from this work (black points) and from various cosmological simulations. The solid curves show fits to the simulations from Pawlik et al. (2009) in brown, Shull et al. (2012) in light blue, and Finlator et al. (2012) in pink. The shaded regions show ranges in C𝐶Citalic_C estimates from Kaurov & Gnedin (2015) in green and Chen et al. (2020) in purple, where for the latter we have set the ionized fraction to unity to mimic our assumptions. The dashed curves show simulations without photoionization heating, with the small-box adiabatic hydrodynamical simulation from Emberson et al. (2013) in blue and the N-body simulations from Mao et al. (2020) in orange.

3.1 Observations of ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT and λmfpsubscript𝜆mfp\lambda_{\rm mfp}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT

Estimating the clumping factor using the method above requires an estimate of the ionizing background intensity as well as the mean free path of ionizing photons. While most studies of the clumping factor have focused on its behavior during the reionization epoch, our formalism applies to any cosmic time where both of these quantities have been measured.

For the mean free path, at z<5𝑧5z<5italic_z < 5 we use the power-law fit to the direct measurements of quasar spectra stacked beyond the Lyman limit from Worseck et al. (2014) and references therein. At z>5𝑧5z>5italic_z > 5, we use the measurements from Zhu et al. (2023), who (following Becker et al. 2021) use a similar stacking method to Worseck et al. (2014) but additionally account for the bias due to the intense local ionizing flux from the background quasars. At all redshifts we assume a power-law frequency dependence of the mean free path of λννλαproportional-tosubscript𝜆𝜈subscriptsuperscript𝜈𝛼𝜆\lambda_{\nu}\propto\nu^{\alpha}_{\lambda}italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT, with αλ=1subscript𝛼𝜆1\alpha_{\lambda}=1italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 1. This power-law dependence approximately corresponds to a H I column density distribution function proportional to N4/3superscript𝑁43N^{-4/3}italic_N start_POSTSUPERSCRIPT - 4 / 3 end_POSTSUPERSCRIPT. This distribution shape is somewhat flatter than the distribution of lower density Lyα𝛼\alphaitalic_α forest absorbers, but consistent with some measurements and models at z=2𝑧2z=2italic_z = 26666 (Faucher-Giguère et al., 2009; Songaila & Cowie, 2010). We note that this connection to the column density distribution is only an approximation, as the shape is likely much more complicated in the relevant range of H I column densities (e.g. Haardt & Madau 2012; O’Meara et al. 2013; Prochaska et al. 2014).

For the ionizing background, at z<5𝑧5z<5italic_z < 5 we adopt the measurements of ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT from Becker & Bolton (2013), who calibrated a suite of hydrodynamical simulations to the mean transmitted flux of Lyα𝛼\alphaitalic_α derived from a stacking analysis of SDSS quasars (Becker et al., 2013), and comprehensively accounted for various sources of systematic uncertainty. At z>5𝑧5z>5italic_z > 5, we use the constraints on ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT from Davies et al. (2024), who fit a fluctuating ionizing background model (Davies & Furlanetto, 2016) to the Lyα𝛼\alphaitalic_α forest opacity distributions from Bosman et al. (2022). To determine the specific intensity at the hydrogen-ionizing edge JHIsubscript𝐽HIJ_{\rm HI}italic_J start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT required by equation (16), we assume that the spectrum of the hydrogen-ionizing background (i.e. νHI<ν<4νHIsubscript𝜈HI𝜈4subscript𝜈HI\nu_{\rm HI}<\nu<4\,\nu_{\rm HI}italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT < italic_ν < 4 italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT) is described by a power-law shape Jνναbproportional-tosubscript𝐽𝜈superscript𝜈subscript𝛼𝑏J_{\nu}\propto\nu^{-\alpha_{b}}italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT - italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUPERSCRIPT with αb=1.0subscript𝛼𝑏1.0\alpha_{b}=1.0italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 1.0. This spectral shape is consistent with an intrinsic ionizing emissivity proportional to ν2superscript𝜈2\nu^{-2}italic_ν start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT (cf. Becker & Bolton 2013; D’Aloisio et al. 2019) filtered through the absorber distribution giving rise to λνν1proportional-tosubscript𝜆𝜈superscript𝜈1\lambda_{\nu}\propto\nu^{1}italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT as assumed above. We then determine the corresponding JHIsubscript𝐽HIJ_{\rm HI}italic_J start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT by requiring that the observed ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT is reproduced by

ΓHI=4πνHI4νHIJHI(ν/νHI)αbhνσHI(ν)dν,subscriptΓHI4𝜋superscriptsubscriptsubscript𝜈HI4subscript𝜈HIsubscript𝐽HIsuperscript𝜈subscript𝜈HIsubscript𝛼𝑏𝜈subscript𝜎HI𝜈differential-d𝜈\Gamma_{\rm HI}=4\pi\int_{\nu_{\rm HI}}^{4\,\nu_{\rm HI}}\frac{J_{\rm HI}(\nu/% \nu_{\rm HI})^{-\alpha_{b}}}{h\nu}\sigma_{\rm HI}(\nu){\rm d}\nu,roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT = 4 italic_π ∫ start_POSTSUBSCRIPT italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT end_POSTSUPERSCRIPT divide start_ARG italic_J start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ( italic_ν / italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_h italic_ν end_ARG italic_σ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ( italic_ν ) roman_d italic_ν , (17)

where σHI(ν)subscript𝜎HI𝜈\sigma_{\rm HI}(\nu)italic_σ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ( italic_ν ) is the hydrogen photoionization cross-section from Verner et al. (1996).

We note that the mean free path measurements from Zhu et al. (2023) are derived assuming specific values of the hydrogen photoionization rate (and its uncertainty) from Gaikwad et al. (2023). To ensure self-consistency, we recompute the mean free path and corresponding uncertainties using the Davies et al. (2024) constraints on ΓHI(z)subscriptΓHI𝑧\Gamma_{\rm HI}(z)roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ( italic_z ), but note that this does not make a substantial difference to our results.

3.2 Estimates of the effective clumping factor

With the ionizing background strength and mean free path in hand, we can now proceed to compute the clumping factor following Section 2. Specifically, we evaluate equation (16) using the mean free path measured by Worseck et al. (2014) (i.e. the power-law fit from z=2𝑧2z=2italic_z = 25555) and Zhu et al. (2023), and the ionizing background intensity implied by the photoionization rate measurements of Becker & Bolton (2013) and Davies et al. (2024), with assumed frequency dependencies λννproportional-tosubscript𝜆𝜈𝜈\lambda_{\nu}\propto\nuitalic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν and Jνν1proportional-tosubscript𝐽𝜈superscript𝜈1J_{\nu}\propto\nu^{-1}italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

We show the resulting estimates of C𝐶Citalic_C from z=2𝑧2z=2italic_z = 26666 in Figure 1. The error bars at z<5𝑧5z<5italic_z < 5 include only the uncertainty in ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT from Becker & Bolton (2013), while at z>5𝑧5z>5italic_z > 5 they include both the uncertainty in ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT from Davies et al. (2024) and in λmfpsubscript𝜆mfp\lambda_{\rm mfp}italic_λ start_POSTSUBSCRIPT roman_mfp end_POSTSUBSCRIPT from Zhu et al. (2023). We find a remarkably constant value of C10similar-to𝐶10C\sim 10italic_C ∼ 1015151515 across the entire redshift range, with an average value of C12𝐶12C\approx 12italic_C ≈ 12 at z=5𝑧5z=5italic_z = 56666, well above the simulation-calibrated prescriptions often used in the literature (C3similar-to𝐶3C\sim 3italic_C ∼ 3). While the short mean free path at z=6𝑧6z=6italic_z = 6 suggests a rather high value C17similar-to𝐶17C\sim 17italic_C ∼ 17, its corresponding uncertainty is large enough to be consistent with all lower redshifts.

We note that at z4less-than-or-similar-to𝑧4z\lesssim 4italic_z ≲ 4 we expect that our assumption of local photoionization equilibrium becomes increasingly incorrect. The mean free path at this time is long enough that the local source approximation is no longer valid (see, e.g., the discussion in Becker & Bolton 2013), so the photons being absorbed at a given epoch were emitted at a substantially earlier time. In addition, as the mean free path increases an increasing fraction of ionizing photons will redshift below νHIsubscript𝜈HI\nu_{\rm HI}italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT before encountering a hydrogen atom. Neglecting these effects is likely the cause for the upturn in our C𝐶Citalic_C estimates at z<3𝑧3z<3italic_z < 3 visible in Figure 1.

3.3 Comparison to simulations

In Figure 2, we compare our estimates of C𝐶Citalic_C to various determinations of C𝐶Citalic_C-like quantities in the literature. The solid curves from Pawlik et al. (2009), Shull et al. (2012), and Finlator et al. (2012) represent the basis behind the commonly-assumed values of C=2𝐶2C=2italic_C = 23333, with the ranges of estimates from more recent simulations by Kaurov & Gnedin (2015), Chen et al. (2020), and Kannan et al. (2022) shown as shaded regions with moderately higher values up to C5similar-to𝐶5C\sim 5italic_C ∼ 5 at z=5𝑧5z=5italic_z = 56666. All of these works compute the clumping factor in different ways, but in principle they are all designed to fulfill the same role: to quantify the effect of the sink term on the progression of reionization in equation (1).

The dashed curves in Figure 2 show clumping factors measured from simulations without any contribution from photoionization heating, with Emberson et al. (2013) employing small-volume adiabatic hydrodynamical simulations and Mao et al. (2020, see also ) using a combination of small and large N-body simulations. These curves can be considered to be theoretical “maximum” values for C𝐶Citalic_C, and our estimates lie comfortably below them.

Why, then, do we recover such a large value for C𝐶Citalic_C compared to the commonly-accepted value from simulations? In the absence of an unforeseen source of bias in our approach, it is possible that the C𝐶Citalic_C measured in simulations is not directly comparable to our value of C𝐶Citalic_C due to a difference in definition. Simulations are careful to compute C𝐶Citalic_C without including dense gas inside of halos, to avoid double-counting this gas which could be responsible for the galactic escape fraction. This is typically implemented as a density threshold, e.g. C100subscript𝐶100C_{100}italic_C start_POSTSUBSCRIPT 100 end_POSTSUBSCRIPT from Pawlik et al. (2009) is measured from gas with overdensity Δ<100Δ100\Delta<100roman_Δ < 100. Other works include cuts on the temperature and ionization state of the gas (e.g. Finlator et al. 2012; Kaurov & Gnedin 2015).

Refer to caption
Figure 3: Reionization histories (upper panel) computed using equation (1) with C=3𝐶3C=3italic_C = 3 (purple) and C=12𝐶12C=12italic_C = 12 (black) and the cumulative number of ionizing photons per baryon (lower panel). We integrate the UV LFs from Bouwens et al. (2021, solid) and Finkelstein & Bagley (2022, dot-dashed) down to MUV=13subscript𝑀UV13M_{\rm UV}=-13italic_M start_POSTSUBSCRIPT roman_UV end_POSTSUBSCRIPT = - 13 to compute the UV luminosity density, and multiply by fixed values of fescξionsubscript𝑓escsubscript𝜉ionf_{\rm esc}\xi_{\rm ion}italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT given in the legend to calibrate n˙ionsubscript˙𝑛ion\dot{n}_{\rm ion}over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT to reach Q=0.9𝑄0.9Q=0.9italic_Q = 0.9 at z=5.9𝑧5.9z=5.9italic_z = 5.9, which then finishes reionization at z5.7𝑧5.7z\approx 5.7italic_z ≈ 5.7 (vertical dashed line). While our estimate of C=12𝐶12C=12italic_C = 12 does not affect the reionization history compared to the classical assumption of C=3𝐶3C=3italic_C = 3 in this scenario, it requires a factor 2similar-toabsent2\sim 2∼ 2 higher total photon output from galaxies to finish reionization.

But this dense gas masking ignores the crucial possibility that dense halo gas can also be illuminated from the outside by the UV background, and potentially make up a substantial fraction of the opacity to ionizing photons streaming through the IGM. The mean free path measured in stacked quasar spectra (e.g. Prochaska et al. 2009; Worseck et al. 2014; Becker et al. 2021; Zhu et al. 2023) or from the H I column density distribution (Rudie et al., 2013; Prochaska et al., 2014), includes encounters of ionizing photons with all gas without any regard for whether it is associated with a galaxy.

Recent simulations by Cain et al. (2023) which take into account the effect of IGM small-scale (similar-to\sim kpc) structure and its relaxation dynamics after reionization heating (Park et al., 2016; D’Aloisio et al., 2020; Chan et al., 2024), have found that applying a clumping factor C=5𝐶5C=5italic_C = 5 to their coarse 1 Mpc-resolution simulation provides a decent match to their more sophisticated sink modeling. On the surface, this value is substantially lower than our estimates, but recall that our C𝐶Citalic_C is defined globally – this distinction is important, because the locally-defined C𝐶Citalic_C can be much smaller than the global one (Raičević & Theuns, 2011). In fact, the scale-dependence of the clumping factor found by Kaurov & Gnedin (2015) suggests that the global C𝐶Citalic_C is 2similar-toabsent2\sim 2∼ 2 times larger than the local C𝐶Citalic_C on 1similar-toabsent1\sim 1∼ 1 Mpc scales, implying that our estimate for the (global) C𝐶Citalic_C is reasonably consistent with the model from Cain et al. (2023).

4 Implications for reionization

Fundamentally, the purpose of estimating this particular definition of the clumping factor is to explore what implications the short mean free path of ionizing photons at z=6𝑧6z=6italic_z = 6 has for the reionization history, and particularly, for the requirements on the number of ionizing photons that must have been emitted to complete the process. The semi-numerical simulations in Paper I examined this in the context of the way a short mean free path inhibits the ionization of the last remaining voids; here, instead, we consider solely the effect of the additional recombinations inside of ionized gas from the large clumping factor implied by IGM observations at z=5𝑧5z=5italic_z = 56666 as shown above. While analytically convenient, this choice comes at the expense of neglecting the effect of the spatial offset between ionizing sources and the last patches of neutral gas at the end of reionization (Paper I; Davies & Furlanetto 2022); we leave a closer look at that effect to future work. In this section we will consider the impact of our high value of C=12𝐶12C=12italic_C = 12 on reionization calculations involving equation (1).

Refer to caption
Figure 4: Similar to the upper panel of Figure 3 but keeping fescξionsubscript𝑓escsubscript𝜉ionf_{\rm esc}\xi_{\rm ion}italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT fixed to 1024.8superscript1024.810^{24.8}10 start_POSTSUPERSCRIPT 24.8 end_POSTSUPERSCRIPT erg/Hz, and including the case of a uniform IGM (C=1𝐶1C=1italic_C = 1, grey).
Refer to caption
Figure 5: Required UV luminosity density from galaxies in order to maintain the Universe fully, 50%percent5050\%50 %, and 25%percent2525\%25 % ionized (thick to thin dashed lines). Assuming an effective ionization production f\textescξ\textion=1024.8subscript𝑓\text𝑒𝑠𝑐subscript𝜉\text𝑖𝑜𝑛superscript1024.8f_{\text{esc}}\xi_{\text{ion}}=10^{24.8}italic_f start_POSTSUBSCRIPT italic_e italic_s italic_c end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT italic_i italic_o italic_n end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 24.8 end_POSTSUPERSCRIPT erg/Hz and a clumping factor C=12𝐶12C=12italic_C = 12, the integrated light from galaxies down to M\textUV=13subscript𝑀\text𝑈𝑉13M_{\text{UV}}=-13italic_M start_POSTSUBSCRIPT italic_U italic_V end_POSTSUBSCRIPT = - 13 is capable of maintaining a fully ionized the Universe at z=6𝑧6z=6italic_z = 6, but only <50%absentpercent50<50\%< 50 % at z=8𝑧8z=8italic_z = 8. The results depend only weakly on the choice of UV LF, Bouwens et al. (2021) (solid) or Finkelstein & Bagley (2022) (long-dashed).

We must first consider our model for the sources, i.e. n˙ionsubscript˙𝑛ion\dot{n}_{\rm ion}over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT (equation 2). We compute ρUV(z)subscript𝜌UV𝑧\rho_{\rm UV}(z)italic_ρ start_POSTSUBSCRIPT roman_UV end_POSTSUBSCRIPT ( italic_z ) from the evolving UV LF parameterizations by Bouwens et al. (2021) and Finkelstein & Bagley (2022), extrapolating up to z=15𝑧15z=15italic_z = 15, and integrating down to a fiducial limiting UV magnitude of MUV=13subscript𝑀UV13M_{\rm UV}=-13italic_M start_POSTSUBSCRIPT roman_UV end_POSTSUBSCRIPT = - 13. The redshift evolution of the Finkelstein & Bagley (2022) LF includes a strongly evolving suppression at the faint end leading to a rapid decline in the UV luminosity density at the upper end of (and beyond) their fitting range at z9greater-than-or-equivalent-to𝑧9z\gtrsim 9italic_z ≳ 9. We thus extrapolate to higher redshift with a double-power-law fit to the evolution at 5<z<8.55𝑧8.55<z<8.55 < italic_z < 8.5, although we note that this makes little difference to our main results.

We can now explore the consequences for the reionization history, integrating equation (1) across cosmic time. We adopt clumping factors of C=3𝐶3C=3italic_C = 3, representing the traditional approach, and C=12𝐶12C=12italic_C = 12, as determined in this work. We then tune the product of the ionizing escape fraction and ionizing efficiency fescξionsubscript𝑓escsubscript𝜉ionf_{\rm esc}\xi_{\rm ion}italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT in each case to reach a neutral fraction of 10%percent1010\%10 % at z=5.9𝑧5.9z=5.9italic_z = 5.9, consistent with the Lyα𝛼\alphaitalic_α forest dark pixel constraint from McGreer et al. (2015) and with the model from Paper I, leading to a late end to reionization consistent with the most recent constraints from the Lyα𝛼\alphaitalic_α forest (Zhu et al., 2021, 2022, 2024; Bosman et al., 2022; Spina et al., 2024).

The resulting reionization histories are shown in the top panel of Figure 3. At this fixed endpoint of reionization, and with our particular models for n˙ion(z)subscript˙𝑛ion𝑧\dot{n}_{\rm ion}(z)over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT ( italic_z ), increasing the clumping factor from C=3𝐶3C=3italic_C = 3 to C=12𝐶12C=12italic_C = 12 has a negligible effect on the reionization history at earlier times. In the lower panel of Figure 3, we show the corresponding integrated number of ionizing photons per baryon. Assuming C=3𝐶3C=3italic_C = 3 requires 1.31.31.31.31.51.51.51.5 photons per baryon to complete reionization, while with C=12𝐶12C=12italic_C = 12 the number doubles to 2.52.52.52.53.03.03.03.0 photons per baryon. This elevated photon budget is nevertheless still slightly below the nominal range from Paper I, but it is very similar to the dynamic sink radiative transfer models of Cain et al. (2021). This number is also consistent with the total number of recombinations at the end of reionization (i.e. the number of emitted photons per baryon minus one) in the CROC radiation hydrodynamical simulations (Gnedin, 2022, 2024).

More generally, if we remove the restriction on the endpoint of reionization, we can explore how different values of C𝐶Citalic_C change the reionization history at fixed n˙ion(z)subscript˙𝑛ion𝑧\dot{n}_{\rm ion}(z)over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT ( italic_z ). In Figure 4, we show reionization histories similar to Figure 3 but with a fixed fescξion=1024.8subscript𝑓escsubscript𝜉ionsuperscript1024.8f_{\rm esc}\xi_{\rm ion}=10^{24.8}italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 24.8 end_POSTSUPERSCRIPT erg/Hz, corresponding to e.g. a model with ξion=1025.8subscript𝜉ionsuperscript1025.8\xi_{\rm ion}=10^{25.8}italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 25.8 end_POSTSUPERSCRIPT erg/Hz, consistent with a recent determination for UV-faint galaxies with JWST (Atek et al., 2024), and fesc=0.1subscript𝑓esc0.1f_{\rm esc}=0.1italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT = 0.1, consistent with direct measurements of Lyman continuum photons from Lyman-break galaxies at z3similar-to𝑧3z\sim 3italic_z ∼ 3 (Pahl et al., 2021). While the conventional assumption of C=3𝐶3C=3italic_C = 3 only modestly postpones the end of reionization by Δz0.3similar-toΔ𝑧0.3\Delta z\sim 0.3roman_Δ italic_z ∼ 0.3 relative to a uniform IGM (C=1𝐶1C=1italic_C = 1), our fiducial C=12𝐶12C=12italic_C = 12 delays its completion by Δz1greater-than-or-equivalent-toΔ𝑧1\Delta z\gtrsim 1roman_Δ italic_z ≳ 1.

Next, we examine the commonly-used criterion for reionization to remain complete, defined by setting Q=1𝑄1Q=1italic_Q = 1 and dQ/dt=0𝑑𝑄𝑑𝑡0dQ/dt=0italic_d italic_Q / italic_d italic_t = 0 in equation (1):

n˙ion,critCnH2αHII.subscript˙𝑛ioncrit𝐶superscriptdelimited-⟨⟩subscript𝑛H2subscript𝛼HII\dot{n}_{\rm ion,crit}\geq C\langle n_{\rm H}\rangle^{2}\alpha_{\rm HII}.over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion , roman_crit end_POSTSUBSCRIPT ≥ italic_C ⟨ italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT . (18)

We note that this expression can be re-stated as a criterion that reionization progresses at a given value of the ionized fraction (e.g. Cullen et al. 2024),

n˙ion,crit(Q)QCnH2αHII=Q×n˙ion,crit.subscript˙𝑛ioncrit𝑄𝑄𝐶superscriptdelimited-⟨⟩subscript𝑛H2subscript𝛼HII𝑄subscript˙𝑛ioncrit\dot{n}_{\rm ion,crit}(Q)\geq QC\langle n_{\rm H}\rangle^{2}\alpha_{\rm HII}=Q% \times\dot{n}_{\rm ion,crit}.over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion , roman_crit end_POSTSUBSCRIPT ( italic_Q ) ≥ italic_Q italic_C ⟨ italic_n start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT roman_HII end_POSTSUBSCRIPT = italic_Q × over˙ start_ARG italic_n end_ARG start_POSTSUBSCRIPT roman_ion , roman_crit end_POSTSUBSCRIPT . (19)

i.e. for the ionized fraction to increase with time, the number of new ionizations must be larger than the number of recombinations within the ionized phase of the IGM.

In Figure 5, we compare the critical values of ionizing photon emissivity for ionized fractions of 25%–100% at z=6𝑧6z=6italic_z = 68888 with the corresponding emissivity calculated from the UV LFs versus the UV magnitude integration limit. As in Figure 4, we assume fescξion=1024.8subscript𝑓escsubscript𝜉ionsuperscript1024.8f_{\rm esc}\xi_{\rm ion}=10^{24.8}italic_f start_POSTSUBSCRIPT roman_esc end_POSTSUBSCRIPT italic_ξ start_POSTSUBSCRIPT roman_ion end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 24.8 end_POSTSUPERSCRIPT erg/Hz. Under this assumption, galaxies at z=6𝑧6z=6italic_z = 6 can maintain reionization provided that ionizing photons escape from galaxies as faint as MUV14similar-tosubscript𝑀UV14M_{\rm UV}\sim-14italic_M start_POSTSUBSCRIPT roman_UV end_POSTSUBSCRIPT ∼ - 14, while at z=7𝑧7z=7italic_z = 7 and z=8𝑧8z=8italic_z = 8 this would only be sufficient to continue reionizing the universe at ionized fractions of 50%percent5050\%50 % and 25%percent2525\%25 %, respectively.

5 Summary & Conclusion

In this work, we have explored the implications of the short mean free path of ionizing photons at z6𝑧6z\approx 6italic_z ≈ 6 (Becker et al., 2021; Zhu et al., 2023) for the recombination rate in the intergalactic medium as a whole, quantified by the clumping factor C𝐶Citalic_C. We first build an analytic connection between the mean free path and the recombination rate with minimal assumptions. The number density of ionizing photons in the optically-thin IGM can be derived from the hydrogen photoionization rate ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT measured from the Lyα𝛼\alphaitalic_α forest, and the rate at which these photons perform an ionization can be derived from the photon lifetime implied by the typical distance they travel before being absorbed, i.e. the mean free path. A global value for C𝐶Citalic_C can then be estimated by comparing this rate to the recombination rate expected for a uniform IGM at the cosmic mean density.

We find a characteristic value of C12𝐶12C\approx 12italic_C ≈ 12 at z=5𝑧5z=5italic_z = 56666 that is well in excess of the C=3𝐶3C=3italic_C = 3 assumption commonly made in the literature based on cosmological radiation-hydrodynamics simulations. Surprisingly, this elevated value persists to later times, consistent with non-evolution down to z2.5similar-to𝑧2.5z\sim 2.5italic_z ∼ 2.5. We tentatively attribute our higher value of C𝐶Citalic_C to the way in which simulation analyses explicitly neglect dense gas within galaxy halos. While such an exclusion appears necessary to avoid double-counting the gas responsible for the galactic escape fraction, it ignores the fact that this dense gas can still absorb external photons streaming through the IGM, and thus play an important role in determining the total budget of recombinations.

Compared to the typical assumption of C=3𝐶3C=3italic_C = 3, we find that late-ending reionization histories with C=12𝐶12C=12italic_C = 12 require roughly twice as many ionizing photons to complete the process at z6less-than-or-similar-to𝑧6z\lesssim 6italic_z ≲ 6. However, recent observations of the ionizing efficiency of z>6𝑧6z>6italic_z > 6 galaxies from JWST (e.g. Simmonds et al. 2024) and scaling relations for the ionizing escape fraction from low-redshift Lyman continuum leakers (e.g. Chisholm et al. 2022) imply a tremendous excess in the ionizing photon budget (Muñoz et al., 2024). Due to the difference in our assumed recombination coefficient, the recombination rate in our fiducial model with C=12𝐶12C=12italic_C = 12 is comparable to that of the C=20𝐶20C=20italic_C = 20 model explored by Muñoz et al. (2024) in which reionization still ends quite early at z7.5similar-to𝑧7.5z\sim 7.5italic_z ∼ 7.5.

We note also that the clumping factor may not provide a complete picture of the number of photons required to complete the reionization process. As shown in Paper I, the fact that the ionizing sources and neutral islands are physically offset from one another implies a large degree of attenuation, requiring 6similar-toabsent6\sim 6∼ 6 photons per baryon to reach a neutral fraction of xHI10%similar-tosubscript𝑥HIpercent10x_{\rm HI}\sim 10\%italic_x start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ∼ 10 % at z6similar-to𝑧6z\sim 6italic_z ∼ 6; about a factor of two higher than our fiducial model here with C=12𝐶12C=12italic_C = 12. It is possible that both a large recombination rate and a consideration of the physical offset are required to reconcile the copious ionizing photon production of the first galaxies with current constraints on the reionization history.

The manuscript was completed following productive discussions with Girish Kulkarni, Laura Keating, Anson D’Aloisio, and Christopher Cain at the NORDITA workshop programme “Cosmic Dawn at High Latitudes”. SEIB is supported by the Deutsche Forschungsgemeinschaft (DFG) under Emmy Noether grant number BO 5771/1-1.

Appendix A Systematic variation with frequency dependence assumptions

The derivation of the effective global clumping factor in § 2 relies on two assumptions of frequency dependence: the spectral index of the UV background intensity (Jνναbproportional-tosubscript𝐽𝜈superscript𝜈subscript𝛼𝑏J_{\nu}\propto\nu^{-\alpha_{b}}italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT - italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_POSTSUPERSCRIPT) and the mean free path (λνναλproportional-tosubscript𝜆𝜈superscript𝜈subscript𝛼𝜆\lambda_{\nu}\propto\nu^{\alpha_{\lambda}}italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT). These two indices are not completely independent – at higher redshifts, the connection can be described via the absorption-limited approximation Jνϵνλνproportional-tosubscript𝐽𝜈subscriptitalic-ϵ𝜈subscript𝜆𝜈J_{\nu}\propto\epsilon_{\nu}\lambda_{\nu}italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ϵ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT (Meiksin & White, 2003), where ϵνναϵproportional-tosubscriptitalic-ϵ𝜈superscript𝜈subscript𝛼italic-ϵ\epsilon_{\nu}\propto\nu^{-\alpha_{\epsilon}}italic_ϵ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT - italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the specific ionizing emissivity of the sources. There is a further frequency dependence in our conversion from the constraints on ΓHIsubscriptΓHI\Gamma_{\rm HI}roman_Γ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT from the Lyα𝛼\alphaitalic_α forest to JHIsubscript𝐽HIJ_{\rm HI}italic_J start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT, which involves an integral over the hydrogen photoionization cross-section (equation 17). Approximating the cross-section as σHIν3proportional-tosubscript𝜎HIsuperscript𝜈3\sigma_{\rm HI}\propto\nu^{-3}italic_σ start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT ∝ italic_ν start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, the combined frequency dependent term can be written as

C(αb,αλ)αb+314αb+314αb+αλαb+αλ,proportional-to𝐶subscript𝛼𝑏subscript𝛼𝜆subscript𝛼𝑏31superscript4subscript𝛼𝑏31superscript4subscript𝛼𝑏subscript𝛼𝜆subscript𝛼𝑏subscript𝛼𝜆C(\alpha_{b},\alpha_{\lambda})\propto\frac{\alpha_{b}+3}{1-4^{\alpha_{b}+3}}% \frac{1-4^{\alpha_{b}+\alpha_{\lambda}}}{\alpha_{b}+\alpha_{\lambda}},italic_C ( italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT , italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ) ∝ divide start_ARG italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + 3 end_ARG start_ARG 1 - 4 start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + 3 end_POSTSUPERSCRIPT end_ARG divide start_ARG 1 - 4 start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT end_ARG , (A1)

although in practice we integrate over the full form of the photoionization cross-section from Verner et al. (1996).

Refer to caption
Figure 6: Effect of different frequency dependence assumptions on the estimated clumping factor relative to our fiducial choice of αb=1subscript𝛼𝑏1\alpha_{b}=1italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT = 1 and αλ=1subscript𝛼𝜆1\alpha_{\lambda}=1italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 1 (red cross). The thick curve shows the variation with αbsubscript𝛼𝑏\alpha_{b}italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT at fixed αλ=1subscript𝛼𝜆1\alpha_{\lambda}=1italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 1, with thin curves showing αλ=0subscript𝛼𝜆0\alpha_{\lambda}=0italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT = 03333 (top to bottom) in steps of 0.50.50.50.5. The dashed colored lines show fixed values of αϵ=αb+αλsubscript𝛼italic-ϵsubscript𝛼𝑏subscript𝛼𝜆\alpha_{\epsilon}=\alpha_{b}+\alpha_{\lambda}italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT, where our fiducial choice corresponds to αϵ=2subscript𝛼italic-ϵ2\alpha_{\epsilon}=2italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = 2.

In Figure 6, we show the full frequency dependence of the clumping factor calculation as a function of αbsubscript𝛼𝑏\alpha_{b}italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT for different values of αλsubscript𝛼𝜆\alpha_{\lambda}italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT, taking into account the frequency dependencies in equation (15) and equation (17). Constant values of αϵ=αb+αλsubscript𝛼italic-ϵsubscript𝛼𝑏subscript𝛼𝜆\alpha_{\epsilon}=\alpha_{b}+\alpha_{\lambda}italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT + italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT are indicated, showing the interplay between these three quantities and the estimated clumping factor. Values of C𝐶Citalic_C as low as half of our fiducial estimates are possible only if the mean free path increases sharply with frequency.

We note that the column density distribution assumed in modern cosmological radiative transfer calculations is not a single power-law, but instead described by a piecewise series of power-laws surrounding the most relevant H I column densities within ±2plus-or-minus2\pm 2± 2 dex of the NLLS=1017.2subscript𝑁LLSsuperscript1017.2N_{\rm LLS}=10^{17.2}italic_N start_POSTSUBSCRIPT roman_LLS end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 17.2 end_POSTSUPERSCRIPT cm-2 (Haardt & Madau, 2012; Puchwein et al., 2019; Faucher-Giguère, 2020). In these models, the slope of the distribution at NLLSsubscript𝑁LLSN_{\rm LLS}italic_N start_POSTSUBSCRIPT roman_LLS end_POSTSUBSCRIPT can be very steep; for example, Puchwein et al. (2019) adopt f(N)N1.95proportional-to𝑓𝑁superscript𝑁1.95f(N)\propto N^{-1.95}italic_f ( italic_N ) ∝ italic_N start_POSTSUPERSCRIPT - 1.95 end_POSTSUPERSCRIPT in the range 1016superscript101610^{16}10 start_POSTSUPERSCRIPT 16 end_POSTSUPERSCRIPT1018superscript101810^{18}10 start_POSTSUPERSCRIPT 18 end_POSTSUPERSCRIPT cm-2. Analytically, this should result in an extremely steep dependence of the mean free path with frequency, with αλ3similar-tosubscript𝛼𝜆3\alpha_{\lambda}\sim 3italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ∼ 3. Instead, we find that the (no longer power-law) frequency dependence arising from integrating over their column density distribution model at z=6𝑧6z=6italic_z = 6 is better approximated by αλ1.6similar-tosubscript𝛼𝜆1.6\alpha_{\lambda}\sim 1.6italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT ∼ 1.6 close to νHIsubscript𝜈HI\nu_{\rm HI}italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT, turning over to αλ<1subscript𝛼𝜆1\alpha_{\lambda}<1italic_α start_POSTSUBSCRIPT italic_λ end_POSTSUBSCRIPT < 1 above 2νHI2subscript𝜈HI2\nu_{\rm HI}2 italic_ν start_POSTSUBSCRIPT roman_HI end_POSTSUBSCRIPT. Self-consistently adopting this parameterization for λνsubscript𝜆𝜈\lambda_{\nu}italic_λ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT in our estimation of the clumping factor in equation (15), including the resulting non-power-law shape of Jνsubscript𝐽𝜈J_{\nu}italic_J start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT, would reduce our C𝐶Citalic_C from 12 to 10. However, a modest hardening of the source spectrum to αϵ=1.2subscript𝛼italic-ϵ1.2\alpha_{\epsilon}=1.2italic_α start_POSTSUBSCRIPT italic_ϵ end_POSTSUBSCRIPT = 1.2 would return C𝐶Citalic_C to 12.

References

  • Atek et al. (2024) Atek, H., Labbé, I., Furtak, L. J., et al. 2024, Nature, 626, 975, doi: 10.1038/s41586-024-07043-6
  • Becker & Bolton (2013) Becker, G. D., & Bolton, J. S. 2013, MNRAS, 436, 1023, doi: 10.1093/mnras/stt1610
  • Becker et al. (2021) Becker, G. D., D’Aloisio, A., Christenson, H. M., et al. 2021, MNRAS, 508, 1853, doi: 10.1093/mnras/stab2696
  • Becker et al. (2013) Becker, G. D., Hewett, P. C., Worseck, G., & Prochaska, J. X. 2013, MNRAS, 430, 2067, doi: 10.1093/mnras/stt031
  • Bianco et al. (2021) Bianco, M., Iliev, I. T., Ahn, K., et al. 2021, MNRAS, 504, 2443, doi: 10.1093/mnras/stab787
  • Bosman (2021) Bosman, S. E. I. 2021, arXiv e-prints, arXiv:2108.12446. https://arxiv.org/abs/2108.12446
  • Bosman et al. (2022) Bosman, S. E. I., Davies, F. B., Becker, G. D., et al. 2022, MNRAS, 514, 55, doi: 10.1093/mnras/stac1046
  • Bouwens et al. (2021) Bouwens, R. J., Oesch, P. A., Stefanon, M., et al. 2021, AJ, 162, 47, doi: 10.3847/1538-3881/abf83e
  • Cain et al. (2021) Cain, C., D’Aloisio, A., Gangolli, N., & Becker, G. D. 2021, ApJ, 917, L37, doi: 10.3847/2041-8213/ac1ace
  • Cain et al. (2023) Cain, C., D’Aloisio, A., Gangolli, N., & McQuinn, M. 2023, MNRAS, 522, 2047, doi: 10.1093/mnras/stad1057
  • Chan et al. (2024) Chan, T. K., Benítez-Llambay, A., Theuns, T., Frenk, C., & Bower, R. 2024, MNRAS, 528, 1296, doi: 10.1093/mnras/stae114
  • Chen et al. (2020) Chen, N., Doussot, A., Trac, H., & Cen, R. 2020, ApJ, 905, 132, doi: 10.3847/1538-4357/abc890
  • Chisholm et al. (2022) Chisholm, J., Saldana-Lopez, A., Flury, S., et al. 2022, MNRAS, 517, 5104, doi: 10.1093/mnras/stac2874
  • Cullen et al. (2024) Cullen, F., McLeod, D. J., McLure, R. J., et al. 2024, MNRAS, 531, 997, doi: 10.1093/mnras/stae1211
  • D’Aloisio et al. (2019) D’Aloisio, A., McQuinn, M., Maupin, O., et al. 2019, ApJ, 874, 154, doi: 10.3847/1538-4357/ab0d83
  • D’Aloisio et al. (2020) D’Aloisio, A., McQuinn, M., Trac, H., Cain, C., & Mesinger, A. 2020, ApJ, 898, 149, doi: 10.3847/1538-4357/ab9f2f
  • Davies et al. (2021) Davies, F. B., Bosman, S. E. I., Furlanetto, S. R., Becker, G. D., & D’Aloisio, A. 2021, ApJ, 918, L35, doi: 10.3847/2041-8213/ac1ffb
  • Davies & Furlanetto (2016) Davies, F. B., & Furlanetto, S. R. 2016, MNRAS, 460, 1328, doi: 10.1093/mnras/stw931
  • Davies & Furlanetto (2022) —. 2022, MNRAS, 514, 1302, doi: 10.1093/mnras/stac1005
  • Davies et al. (2024) Davies, F. B., Bosman, S. E. I., Gaikwad, P., et al. 2024, ApJ, 965, 134, doi: 10.3847/1538-4357/ad1d5d
  • Emberson et al. (2013) Emberson, J. D., Thomas, R. M., & Alvarez, M. A. 2013, ApJ, 763, 146, doi: 10.1088/0004-637X/763/2/146
  • Erkal (2015) Erkal, D. 2015, MNRAS, 451, 904, doi: 10.1093/mnras/stv980
  • Faucher-Giguère (2020) Faucher-Giguère, C.-A. 2020, MNRAS, 493, 1614, doi: 10.1093/mnras/staa302
  • Faucher-Giguère et al. (2009) Faucher-Giguère, C.-A., Lidz, A., Zaldarriaga, M., & Hernquist, L. 2009, ApJ, 703, 1416, doi: 10.1088/0004-637X/703/2/1416
  • Finkelstein & Bagley (2022) Finkelstein, S. L., & Bagley, M. B. 2022, ApJ, 938, 25, doi: 10.3847/1538-4357/ac89eb
  • Finlator et al. (2012) Finlator, K., Oh, S. P., Özel, F., & Davé, R. 2012, MNRAS, 427, 2464, doi: 10.1111/j.1365-2966.2012.22114.x
  • Furlanetto et al. (2004) Furlanetto, S. R., Zaldarriaga, M., & Hernquist, L. 2004, ApJ, 613, 1, doi: 10.1086/423025
  • Gaikwad et al. (2023) Gaikwad, P., Haehnelt, M. G., Davies, F. B., et al. 2023, MNRAS, 525, 4093, doi: 10.1093/mnras/stad2566
  • Gnedin (2022) Gnedin, N. Y. 2022, ApJ, 937, 17, doi: 10.3847/1538-4357/ac8d0a
  • Gnedin (2024) —. 2024, ApJ, 963, 150, doi: 10.3847/1538-4357/ad298e
  • Gnedin & Ostriker (1997) Gnedin, N. Y., & Ostriker, J. P. 1997, ApJ, 486, 581, doi: 10.1086/304548
  • Gunn & Peterson (1965) Gunn, J. E., & Peterson, B. A. 1965, ApJ, 142, 1633, doi: 10.1086/148444
  • Haardt & Madau (2012) Haardt, F., & Madau, P. 2012, ApJ, 746, 125, doi: 10.1088/0004-637X/746/2/125
  • Iliev et al. (2007) Iliev, I. T., Mellema, G., Shapiro, P. R., & Pen, U.-L. 2007, MNRAS, 376, 534, doi: 10.1111/j.1365-2966.2007.11482.x
  • Iliev et al. (2005) Iliev, I. T., Scannapieco, E., & Shapiro, P. R. 2005, ApJ, 624, 491, doi: 10.1086/429083
  • Izotov et al. (2016) Izotov, Y. I., Schaerer, D., Thuan, T. X., et al. 2016, MNRAS, 461, 3683, doi: 10.1093/mnras/stw1205
  • Ji et al. (2020) Ji, Z., Giavalisco, M., Vanzella, E., et al. 2020, ApJ, 888, 109, doi: 10.3847/1538-4357/ab5fdc
  • Kannan et al. (2022) Kannan, R., Garaldi, E., Smith, A., et al. 2022, MNRAS, 511, 4005, doi: 10.1093/mnras/stab3710
  • Kaurov & Gnedin (2015) Kaurov, A. A., & Gnedin, N. Y. 2015, ApJ, 810, 154, doi: 10.1088/0004-637X/810/2/154
  • Madau (2017) Madau, P. 2017, ApJ, 851, 50, doi: 10.3847/1538-4357/aa9715
  • Madau et al. (1999) Madau, P., Haardt, F., & Rees, M. J. 1999, ApJ, 514, 648, doi: 10.1086/306975
  • Mao et al. (2020) Mao, Y., Koda, J., Shapiro, P. R., et al. 2020, MNRAS, 491, 1600, doi: 10.1093/mnras/stz2986
  • McGreer et al. (2015) McGreer, I. D., Mesinger, A., & D’Odorico, V. 2015, MNRAS, 447, 499, doi: 10.1093/mnras/stu2449
  • McQuinn et al. (2011) McQuinn, M., Oh, S. P., & Faucher-Giguère, C.-A. 2011, ApJ, 743, 82, doi: 10.1088/0004-637X/743/1/82
  • Meiksin & White (2003) Meiksin, A., & White, M. 2003, MNRAS, 342, 1205, doi: 10.1046/j.1365-8711.2003.06624.x
  • Muñoz et al. (2024) Muñoz, J. B., Mirocha, J., Chisholm, J., Furlanetto, S. R., & Mason, C. 2024, arXiv e-prints, arXiv:2404.07250, doi: 10.48550/arXiv.2404.07250
  • Ning et al. (2023) Ning, Y., Cai, Z., Jiang, L., et al. 2023, ApJ, 944, L1, doi: 10.3847/2041-8213/acb26b
  • O’Meara et al. (2013) O’Meara, J. M., Prochaska, J. X., Worseck, G., Chen, H.-W., & Madau, P. 2013, ApJ, 765, 137, doi: 10.1088/0004-637X/765/2/137
  • Pahl et al. (2021) Pahl, A. J., Shapley, A., Steidel, C. C., Chen, Y., & Reddy, N. A. 2021, MNRAS, 505, 2447, doi: 10.1093/mnras/stab1374
  • Park et al. (2016) Park, H., Shapiro, P. R., Choi, J.-h., et al. 2016, ApJ, 831, 86, doi: 10.3847/0004-637X/831/1/86
  • Pawlik et al. (2009) Pawlik, A. H., Schaye, J., & van Scherpenzeel, E. 2009, MNRAS, 394, 1812, doi: 10.1111/j.1365-2966.2009.14486.x
  • Planck Collaboration et al. (2020) Planck Collaboration, Aghanim, N., Akrami, Y., et al. 2020, A&A, 641, A6, doi: 10.1051/0004-6361/201833910
  • Prieto-Lyon et al. (2023) Prieto-Lyon, G., Strait, V., Mason, C. A., et al. 2023, A&A, 672, A186, doi: 10.1051/0004-6361/202245532
  • Prochaska et al. (2014) Prochaska, J. X., Madau, P., O’Meara, J. M., & Fumagalli, M. 2014, MNRAS, 438, 476, doi: 10.1093/mnras/stt2218
  • Prochaska et al. (2009) Prochaska, J. X., Worseck, G., & O’Meara, J. M. 2009, ApJ, 705, L113, doi: 10.1088/0004-637X/705/2/L113
  • Puchwein et al. (2019) Puchwein, E., Haardt, F., Haehnelt, M. G., & Madau, P. 2019, MNRAS, 485, 47, doi: 10.1093/mnras/stz222
  • Raičević & Theuns (2011) Raičević, M., & Theuns, T. 2011, MNRAS, 412, L16, doi: 10.1111/j.1745-3933.2010.00993.x
  • Rudie et al. (2013) Rudie, G. C., Steidel, C. C., Shapley, A. E., & Pettini, M. 2013, ApJ, 769, 146, doi: 10.1088/0004-637X/769/2/146
  • Satyavolu et al. (2023) Satyavolu, S., Kulkarni, G., Keating, L. C., & Haehnelt, M. G. 2023, MNRAS, 521, 3108, doi: 10.1093/mnras/stad729
  • Shull et al. (2012) Shull, J. M., Harness, A., Trenti, M., & Smith, B. D. 2012, ApJ, 747, 100, doi: 10.1088/0004-637X/747/2/100
  • Simmonds et al. (2023) Simmonds, C., Tacchella, S., Maseda, M., et al. 2023, MNRAS, 523, 5468, doi: 10.1093/mnras/stad1749
  • Simmonds et al. (2024) Simmonds, C., Tacchella, S., Hainline, K., et al. 2024, MNRAS, 527, 6139, doi: 10.1093/mnras/stad3605
  • So et al. (2014) So, G. C., Norman, M. L., Reynolds, D. R., & Wise, J. H. 2014, ApJ, 789, 149, doi: 10.1088/0004-637X/789/2/149
  • Songaila & Cowie (2010) Songaila, A., & Cowie, L. L. 2010, ApJ, 721, 1448, doi: 10.1088/0004-637X/721/2/1448
  • Spina et al. (2024) Spina, B., Bosman, S. E. I., Davies, F. B., Gaikwad, P., & Zhu, Y. 2024, arXiv e-prints, arXiv:2405.12273, doi: 10.48550/arXiv.2405.12273
  • Vanzella et al. (2018) Vanzella, E., Nonino, M., Cupani, G., et al. 2018, MNRAS, 476, L15, doi: 10.1093/mnrasl/sly023
  • Verner et al. (1996) Verner, D. A., Ferland, G. J., Korista, K. T., & Yakovlev, D. G. 1996, ApJ, 465, 487
  • Worseck et al. (2019) Worseck, G., Davies, F. B., Hennawi, J. F., & Prochaska, J. X. 2019, ApJ, 875, 111, doi: 10.3847/1538-4357/ab0fa1
  • Worseck et al. (2014) Worseck, G., Prochaska, J. X., O’Meara, J. M., et al. 2014, MNRAS, 445, 1745, doi: 10.1093/mnras/stu1827
  • Wu et al. (2021) Wu, X., McQuinn, M., Eisenstein, D., & Iršič, V. 2021, MNRAS, 508, 2784, doi: 10.1093/mnras/stab2815
  • Zhu et al. (2021) Zhu, Y., Becker, G. D., Bosman, S. E. I., et al. 2021, ApJ, 923, 223, doi: 10.3847/1538-4357/ac26c2
  • Zhu et al. (2022) —. 2022, ApJ, 932, 76, doi: 10.3847/1538-4357/ac6e60
  • Zhu et al. (2023) Zhu, Y., Becker, G. D., Christenson, H. M., et al. 2023, ApJ, 955, 115, doi: 10.3847/1538-4357/aceef4
  • Zhu et al. (2024) Zhu, Y., Becker, G. D., Bosman, S. E. I., et al. 2024, MNRAS, doi: 10.1093/mnrasl/slae061