<<

Draft version April 7, 2020 Typeset using LATEX twocolumn style in AASTeX61

RAPID BY THE OLIGARCHS: THE CASE FOR MASSIVE, UV-BRIGHT, STAR-FORMING WITH HIGH ESCAPE FRACTIONS

Rohan P. Naidu,1 Sandro Tacchella,1 Charlotte A. Mason,1, ∗ Sownak Bose,1 Pascal A. Oesch,2, 3 and Charlie Conroy1

1Center for Astrophysics | Harvard & Smithsonian, 60 Garden Street, Cambridge, MA 02138, USA 2Department of , Universit´ede Gen`eve,Chemin des Maillettes 51, 1290 Versoix, Switzerland 3Cosmic Dawn Center (DAWN)

(Revised Draft version April 7, 2020)

ABSTRACT The protagonists of the last great phase transition of the – cosmic reionization – remain elusive. Faint star-forming galaxies are leading candidates because they are found to be numerous and may have significant ionizing escape fractions (fesc). Here we update this picture via an empirical model that successfully predicts latest observations (e.g., the rapid drop in star-formation density (ρSFR) at z > 8). We generate an ionizing spectrum for each in our model and constrain fesc by leveraging latest measurements of the reionization timeline (e.g., Lyα damping of and galaxies at z > 7). Assuming a constant fesc across all sources at z > 6, we find +0.06 MUV< −13.5 galaxies need fesc=0.21−0.04 to complete reionization. The inferred IGM neutral fraction is [0.9, 0.5, 0.1] at z = [8.2, 6.8, 6.2] ± 0.2, i.e., the bulk of reionization transpires rapidly in 300 Myrs, driven by the z > 8 ρSFR and favored by high neutral fractions (∼60−90%) measured at z ∼ 7 − 8. Inspired by the emergent sample of Lyman Continuum (LyC) leakers spanning z ∼ 0−6.6 that overwhelmingly displays higher-than-average star-formation surface 0.4±0.1 density (ΣSFR), we propose a physically motivated model relating fesc to ΣSFR and find fesc∝ ΣSFR . Since ΣSFR falls by ∼ 2.5 dex between z = 8 and z = 0, our model explains the humble upper limits on fesc at lower and its required evolution to fesc∼ 0.2 at z > 6. Within this model, strikingly, <5% of galaxies with MUV< −18 and log(M?/M ) > 8 (the ‘oligarchs’) account for &80% of the reionization budget – a stark departure from the canonical ‘democratic’ reionization led by copious faint sources. In fact, faint sources (MUV>−16) must be relegated to a limited role in order to ensure high neutral fractions at z = 7 − 8. Shallow faint-end slopes of the UV function (αUV> −2) and/or fesc distributions skewed toward massive galaxies produce the required late and rapid reionization. We predict LyC leakers like COLA1 (z = 6.6, fesc∼ 30%, MUV= −21.5) become increasingly common towards z ∼ 6 and that the drivers of reionization do not lie hidden across the faint-end of the luminosity function, but are already known to us.

Keywords: galaxies: high- — galaxies: evolution — dark ages, reionization, first stars arXiv:1907.13130v2 [astro-ph.GA] 3 Apr 2020

Corresponding author: Rohan P. Naidu [email protected] ∗ Hubble Fellow 2 Naidu et al.

1. INTRODUCTION sitively on resolving the multi-phase ISM, and the treat- The Epoch of Reionization (EoR) marks the last great ment of small-scale processes associated with galaxy phase transition of the universe, during which vast is- formation, which, at present, are modelled in an ap- lands of neutral Hydrogen were ionized by the first proximate way (e.g., Wise et al. 2014; Ma et al. 2015; sources of light (Loeb & Barkana 2001). The protago- Trebitsch et al. 2017). nists, topology, and timeline of the EoR are intertwined However, there is a path forward: since fesc is by with our understanding of the early universe and its far the single largest uncertainty in modeling reioniza- newly born stellar populations (for a recent review, see tion, we can employ ρSFR and ξion from state-of-the-art Dayal & Ferrara 2018). Due to the rapidly fading measurements and constrain fesc against latest measure- emissivity at z > 3 (e.g., Kulkarni et al. 2019), star- ments of the timeline of reionization. The fraction of forming galaxies are favored to dominate reionization dark pixels in the Lyα and Lyβ forests provides a model- (e.g., Bouwens et al. 2015a). Bright star-forming galax- independent limit on the end of reionization (McGreer ies have not shown much promise of being effective ioniz- et al. 2015). The electron scattering (τ) re- ing sources. Until very recently, these galaxies were mea- ported by Planck Collaboration et al.(2018), much lower sured to have negligible ionizing photon escape fractions and more precise than previous measures of τ (e.g., Hin- (e.g. Steidel et al. 2018). This, combined with their ob- shaw et al. 2013), is an integrated probe of the density served rarity has meant a reservoir of ultra-faint sources of ionizing , as CMB photons scatter off of elec- far below current detection limits (modulo highly lensed trons knocked out of Hydrogen atoms. The first quasars fields) is widely invoked to drive reionization (e.g., Liv- and large Lyα surveys at z & 7 allow detailed inferences ermore et al. 2017). of the neutral fraction as a function of redshift from the Modeling reionization by star-forming galaxies is typ- magnitude of Lyα damping (e.g., Ba˜nadoset al. 2018; Mason et al. 2018). ically cast as a tale of three quantities: ρSFR, ξion, and Complementing these data on the global history of fesc (e.g., Madau et al. 1999; Robertson et al. 2015; Bouwens et al. 2015a). The cosmic star-formation rate reionization are clues about fesc on a galaxy by galaxy level. For the first time we have a robust sample of indi- density, ρSFR, provides a measure of star-formation in the early universe. It has now been tracked out to vidual star-forming galaxies securely detected in Lyman z ∼ 10, with latest studies showing an accelerating drop Continuum (LyC) spanning z ∼ 0.3 − 4 (“LyC leakers”) beyond z > 8 (Oesch et al. 2018; Ishigaki et al. 2018), (e.g., Naidu et al. 2017; Vanzella et al. 2018; Rivera- consistent with the predictions of simple models that Thorsen et al. 2019). The campaigns targeting Green link star-formation rates (SFR) to dark matter accre- Peas at z ∼ 0.3 with HST /COS have proven remark- tion (e.g., Tacchella et al. 2013, 2018; Mason et al. 2015; ably efficient (boasting a 100% success rate for LyC de- Mashian et al. 2016). The ionizing photon production tection; Izotov et al. 2016b, 2018b). Concurrently, a sample at high-z is emerging from deep rest-frame UV efficiency, ξion, is a conversion factor for how many Hy- drogen ionizing photons emerge from each episode of (e.g., Steidel et al. 2018) and imaging with HST /UVIS (e.g., Fletcher et al. 2019). Taken together, star-formation. Tight constraints on ξion can be placed using Hα measurements and some assumptions about these leakers provide hints about the galaxy properties that favor LyC leakage during the EoR. For instance, fesc, and now exist from direct spectroscopy (Nakajima et al. 2016; Matthee et al. 2017; Shivaei et al. 2018; Tang the overwhelming majority of LyC leakers are compact et al. 2019) and IRAC-excess inferences of Hα out to (e.g., Izotov et al. 2018a) and show multi-peaked Lyα z ∼ 5 (Bouwens et al. 2016a; Lam et al. 2019). (e.g., Verhamme et al. 2017). These insights can be incorporated into models of fesc that improve upon pre- The escape fraction, fesc, is the fraction of ionizing photons generated in a galaxy that evade photoelectric vious analyses that assumed a single number across the absorption and dust in the (ISM) entire galaxy population. to escape into the neutral Intergalactic Medium (IGM) Meanwhile, the bulk of observational constraints on the average LyC fesc have relied on stacking shallow and ionize it. While ρSFR and ξion will be measured ever more precisely with, e.g., the James Webb Space non-detections for individual galaxies to place stringent upper-limits of fesc< 10% out to z ∼ 4 (e.g., Siana et al. Telescope (JWST), ionizing radiation and thus fesc will never be directly measured in the EoR due to the opac- 2010; Rutkowski et al. 2017; Grazian et al. 2017; Japelj ity of the intervening neutral IGM (e.g., Fan et al. 2001; et al. 2017; Naidu et al. 2018). Taken at face value, these Inoue et al. 2014; McGreer et al. 2015). To make things studies effectively rule out an average fesc> 10% for more challenging, it is also extremely difficult to get a MUV. −20 sources and put the focus on fainter galax- ies, for which no LyC constraints exist yet, as the drivers handle on fesc through simulations since it depends sen- Reionization by the Oligarchs 3 of reionization. However, if we consider the anisotropy, summary of our findings and an outlook to the future is stochasticity, and evolution with z of fesc that recent presented in §7. simulations have brought to light (e.g., Paardekooper We use fesc to denote both the singular and plural et al. 2015; Trebitsch et al. 2017; Rosdahl et al. 2018) “escape fraction” and “escape fractions”. The volume- along with a higher CGM+IGM opacity, the limits from averaged IGM neutral fraction and ionized fraction are these studies are far less stringent and can be relaxed denoted byx ¯HI and Q¯HII = 1 − x¯HI respectively. For by factors of 2 − 5× (e.g., Fletcher et al. 2019; Steidel cosmological parameters, we adopt the following from et al. 2018). And indeed, latest studies that emphasize Planck Collaboration et al.(2018): h = 0.6772, Xp = 2 −29 2 deep spectra and photometry for individual sources find 0.75328, Ωb = 0.02241/h , ρc = 1.8787 × 10 h . All fesc∼ 10% in stacks of normal log(M?/M ) ∼ 8.5 − 10 magnitudes are in the AB system (Oke & Gunn 1983). galaxies at z ∼ 2.5 − 4 (Oesch et al. in prep. 2019; Steidel et al. 2018; Marchi et al. 2018). The emerging 2. METHODS observational picture is that average fesc of ∼10% are 2.1. An Empirical Model for Galaxy Evolution at z ≥ 4 possible in relatively bright galaxies (M < −20). UV The foundation of this work is the empirical model In parallel, early hydrodynamical simulations pro- introduced in Tacchella et al.(2018). Here we briefly duced a mixture of results: with f correlating with esc summarize it, and then describe in detail the quantities halo mass (e.g., Gnedin et al. 2008; Wise & Cen 2009), relevant to reionization. anti-correlating with halo mass (e.g., Yajima et al. 2011; Paardekooper et al. 2015; Kimm et al. 2017), and with 2.1.1. Model Description typical time-averaged values far smaller than f ∼ 20% esc The Tacchella et al.(2018) model is built on top of a (e.g., Ma et al. 2015). However, contemporary hydro- 106 Mpc3, high-resolution, N-body, dark-matter simula- dynamical simulations through a combination of feed- tion, (Sawala et al. 2016; Hellwing et al. 2016). back, binaries, turbulence, and careful modeling of the color It makes the assumption that the star-formation rate multi-phase ISM are able to produce average f in the esc (SFR) of a halo depends on the growth rate of a halo 10 − 20% range in 108 M galaxies (e.g., Ma et al. & and a star-formation efficiency that is independent of 2016; Rosdahl et al. 2018). Ma et al.(2015) and Ma redshift (see also Tacchella et al. 2013; Mason et al. et al.(2016) are particularly illustrative of this shift, 2015). The halo merger trees self-consistently give rise where these authors first wrote about the difficulty of to star-formation histories for each galaxy from which producing f > 5% due to high absorption in the birth esc spectral energy distributions (SEDs) are computed us- clouds of massive stars but then subsequently found bi- ing the Flexible Stellar Population Synthesis code ( , nary models of stellar evolution that destroyed these FSPS Conroy et al. 2009, 2010; Foreman-Mackey et al. 2014) clouds while retaining highly ionizing sources until late and the Isochrones and Stellar Tracks which incor- times could achieve f ∼ 20%. MESA esc porate the effects of rotation ( , Dotter 2016; Choi The driving impulse of this work is to unite the de- MIST et al. 2016, 2017). velopments outlined above self-consistently under the The star-formation efficiency of the model is tuned same umbrella to see what story they tell about f and esc to match the z = 4 Luminosity Function thus reionization. Our umbrella of choice is the empiri- (UVLF) and then predicts UVLFs out to z ∼ 10 consis- cal galaxy formation model by Tacchella et al.(2018) tent with the observed data (see Figure1 in this work that incorporates recent developments (e.g., cutting- and Figure 3 in Tacchella et al. 2018). The faint-end edge stellar population synthesis models) and accurately slope of the UVLF (α ) in our model steepens with predicts latest observations (e.g., the sharp drop in UV redshift and the same trend is seen in observations (e.g., ρ )(§2.1). Leaving f as a free parameter in the SFR esc Finkelstein et al. 2015; Bouwens et al. 2015b). For our equations of reionization (§2.2), we fit for it against re- best-fit Schechter function at z =[4, 6, 8, 10, 12] we cently derived constraints on reionization that we de- find α =[−1.63 ± 0.02, −1.72 ± 0.03, −1.69 ± 0.04, scribe in §2.3. Two models of f – one constant across UV esc −1.84±0.06, −1.99±0.24], consistent with recent empir- all galaxies during reionization, another dependent on ical and semi-analytical models that find α = −1.5 to star-formation surface density – are justified, set up, and UV −2.0 at z = 5 − 10 (e.g., Yung et al. 2019; Endsley et al. fit in §3 and §4. The implications of the resulting reion- 2019; Behroozi et al. 2019). Our best-fit faint-end slope ization histories – their rapid pace, the concentration of values are somewhat shallower than what is reported in the reionization budget among “oligarch” galaxies, the the recent literature from observations (Bouwens et al. path forward for observational studies – are discussed 2017; Livermore et al. 2017; Atek et al. 2018). However, in §5. We address open questions and caveats in §6.A we note the Schechter function parameters are highly 4 Naidu et al. degenerate and that our slopes are fit over a different range of MUV (we truncate at MUV < −13.5) compared to observations. Figure1 shows the actual LF from our model is in decent agreement with the latest observa- tions at z ∼ 6, down to the faintest limits that can currently be probed in the Hubble Frontier Fields. We provide UVLFs for our model out to z = 12 and χ2 com- parisons against literature data in AppendixA. Note that we compute the ionizing budget via SEDs of indi- vidual sources, and not by integrating under the UVLF. We discuss αUV in detail in §5.3. The resolution of the dark-matter simulation lim- its our model to SFR & 0.02 M /yr, corresponding to MUV . −13.5, roughly where estimates of the faint end of the UVLF begin to diverge due to magnification un- certainties of the lensing models (e.g., Bouwens et al. 2017). In all calculations we integrate down to this limit. Figure 1. The z = 6 UV Luminosity Function (UVLF). We Our fiducial model adopts a Salpeter(1955) IMF, a con- plot recent determinations using the Hubble Frontier Fields in orange (Bouwens et al. 2017), turquoise (Livermore et al. stant of Z /Z = 0.02, and the UV contin- ? 2017), and green (Atek et al. 2018). Our predicted UVLF uum dust prescription of Bouwens et al.(2014) (their (purple) agrees well with the Atek et al.(2018) and Bouwens Table 3). Swapping the MIST models for those that et al.(2017) measurements within errors, and is shallower explicitly include binaries (BPASS, Stanway & Eldridge than the Livermore et al.(2017) UVLF. While the formal 2018), or a Chabrier(2003) IMF, or a model with evolv- faint-end slope fit to our UVLF is somewhat shallower than ing metallicity make no appreciable difference to our re- what has been reported, the actual LF is in decent agreement with observations. For this comparison we correct for com- sults. Dust, the faint-end slope of the UVLF, and the 6 3 effect of changing the M cutoff are discussed in §6. pleteness to account for our box-size (10 Mpc , Sheth et al. UV 2001), and also for dust using the UV continuum prescrip- 2.1.2. The Ionizing Photon Production Efficiency (ξion) tion of Bouwens et al.(2014). Both these corrections only effect the bright-end (<−20). The Bouwens et al.(2017) and ξ provides a measure of the gross LyC photons pro- ion Livermore et al.(2017) points are adjusted by 0.15 dex to duced at a given time in a source. It is typically cast in account for differences in their mean redshift following Atek 0 terms of the rate of ionizing photons (N(H )) per unit et al.(2018). UV luminosity, usually measured at 1500A˚ (L1500):

0 between z = 4 − 10 rises by ∼ 40% (∼ 0.15 dex) as N(H ) −1 −1 −1 ξion = [s /erg s Hz ]. (1) galaxies get younger at higher redshift and their ion- L 1500 izing spectra become harder. We compare our predic- We compute ξ for each galaxy in the empirical ion tions with Bouwens et al.(2016a), who report a mean model directly from its SED by integrating the flux log (ξ ) = 25.34+0.02 (25.540.12 ) for a sample span- produced below the Lyman limit to obtain N(H0) and 10 ion −0.02 −0.12 ning z = 3.8 − 5.0 (5.0 − 5.5) using an SMC atten- then normalizing by the SED-flux at 1500A.˚ The MIST uation curve, which Reddy et al.(2018a) show to be isochrones that our SEDs are synthesized from include the appropriate curve for sub-solar metallicity popula- the effects of rotation that boost the ionizing flux pro- tions expected at z > 4. At z = 4 (z = 5) we have a duction of massive stars akin to, but not exactly like, median log (ξ ) = 25.37+0.06 (25.40+0.05) that agrees the effect of binaries (Choi et al. 2017). The harder 10 ion −0.06 −0.06 with their measurements within error-bars. At fixed red- UV spectra produced by these rotating models (or bi- shift we find ξ does not vary with mass or M . Lam naries) are the only kind that self-consistently explain ion UV et al.(2019) observe a similar invariance with bright- the strong nebular fluxes, high ionization lines, and ex- ness at z ∼ 4 − 5 for M < −17.5 galaxies. These treme line widths that are now known to be ubiquitous UV trends hold even when using BPASS templates and with at z ∼ 2.5 − 3.5 (e.g., Steidel et al. 2016; Holden et al. an evolving metallicity. While not directly dealing with 2016; Reddy et al. 2018b) and also commonly seen at the redshifts that are the focus of this work, we note z > 6 (e.g., Roberts-Borsani et al. 2016; Stark et al. that the ξ values reported for z ∼ 2 − 3 galaxies are 2015; Mainali et al. 2017). ion in broad agreement with our model assuming a linear In Figure2 we show the log 10(ξion) distribution for the galaxies in our model. We predict the median ξion Reionization by the Oligarchs 5

Figure 3. n˙ ion, the co-moving emissivity of ionizing photons (Eq.2), as a function of redshift. Shown in blue is the model from Bouwens et al.(2015a), representative of fits based on the Planck Collaboration et al.(2015) τ. Our model, here set to fesc= 0.2, MUV< −13.5, and with αUV& −2 predicts a sharp drop at z > 8 consistent with the latest HST ρSFR (Oesch et al. 2018; Ishigaki et al. 2018). At z = 10, our model produces ∼ 10× fewer ionizing photons with the difference disappearing at z . 7. This dearth of LyC in the early uni- verse, as we show later (Figure4), compresses the timeline of reionization. The MIST models with rotation produce higher ξion (solid purple) than Bouwens et al.(2016a, dashed purple; comparable to the Robertson et al. 2015 ξion), but nowhere close to bridging the gulf between the purple and blue curves. Shown in green are z ∼ 4−5 Ly-forest measurements ofn ˙ ion (Becker & Bolton 2013; Kuhlen & Faucher-Gigu`ere 2012).

Figure 2. The ionizing photon production efficiency (ξion) predicted by our model using FSPS+MIST. ξion represents the between our approach and assuming some reasonable number of ionizing photons produced in a galaxy before be- fixed value (note how close the dotted and dashed pur- ing absorbed by the ISM or attenuated by dust. Top: At ple lines are in Figure3). However, the advantage of z ∼ 4 − 5, the highest redshift at which ξ has been mea- ion our model is that it captures the diversity in ξ on a sured statistically, our model agrees within error-bars with ion the Bouwens et al.(2016a) estimate (shown in green). We galaxy by galaxy basis so that we are able to link fesc to predict evolution in ξion as the stellar populations grow older, individual galaxy properties and through the product 0 with a 40% smaller median at z = 4 than at z = 10. Bot- fesc×N(H ) probe how much each galaxy contributes tom: At fixed redshift we predict ξion does not vary with to reionization (see also §4, Figure9, and Table1). the brightness of galaxies.

2.1.3. Cosmic Star-Formation Rate Density (ρSFR) extrapolation to lower redshifts (Nakajima et al. 2016; Matthee et al. 2017; Shivaei et al. 2018). At z = 4 − 8 the Tacchella et al.(2018) model is in excellent agreement with the consensus ρSFR (e.g., Typically, reionization studies set ξion to some fixed, redshift-invariant value that lies in the locus our galax- Bouwens et al. 2015b; Finkelstein et al. 2015; Madau & ies span in Figure2. This is a reasonable assumption Dickinson 2014). At z > 8, where various measurements as evidenced by the narrow spread (∼ 0.1 dex at each diverge, the model predicts a drop in ρSFR consistent redshift across z = 4 − 10) and gradual evolution dur- with the latest HST analyses from Ishigaki et al.(2018) ing reionization (z = 6 − 10). What this means is that and Oesch et al.(2018). The sharp drop in ρSFR in when considering the total integrated ionizing output at our model comes as the bulk of halos at z > 8 begin a particular redshift, there is not much of a distinction to fall below the halo-mass corresponding to maximal 11 12 star-formation efficiency (Mh ∼ 10 − 10 M ). 6 Naidu et al.

The difference between earlier smooth power-law fits of Hydrogen (Xp), the fractional baryon density Ωb and for ρSFR at z > 4, which use steeper αUV . −2 (e.g., the critical density ρc. trec, the recombination time of −4.2 ρSFR ∝(1 + z) ; Bouwens et al. 2015b; Finkelstein ionized Hydrogen in the IGM, is given by et al. 2015; Robertson et al. 2015) and our model, which 1/t = C α (1 + (1 − X ) /4X ) hn i (1 + z)3 , predicts an accelerating decline in ρ , is as large as an rec HII B P P H SFR (5) order of magnitude at z = 10 and three orders at z = 14. where C = hn2 i/hn i2 is the clumping factor that This difference is directly reflected in the dearth of LyC HII H H models the inhomogeneity of the IGM which we set to photons available for reionization at early times. Earlier 0.76 3, and α = 2.6×10−13 T/104K cm3s−1 is the Case works (e.g., Robertson et al. 2015) had already shown B B recombination coefficient at electron temperature, T that reionization likely proceeds without significant con- that we set to 104K(Shull et al. 2012; Robertson et al. tribution from z > 10 sources. The dearth of sources 2013; Pawlik et al. 2015; Robertson et al. 2015; Sun & in our model (and thus LyC) at z ∼ 8 − 10 combined Furlanetto 2016). with other data, as we shall see in §5.1, even further The Thomson optical depth, τ, is calculated as compresses the timeline of reionization and pushes it to later times. Z z ¯ −1 02 0 2.2. Equations of Reionization τ(z) = chnHiσT feQHIIH(z) 1 + z dz , (6) 0 We closely follow the widespread approach that mod- where τ is the Thomson optical depth, c is the speed of els reionization as an interplay between ionization and light, σT is the Thomson scattering cross-section, fe is recombination (e.g., Madau et al. 1999; Robertson et al. the number of free electrons for every Hydrogen nucleus 2013). Here we outline the relevant equations. in the ionized IGM that we set to (1 + (1 − XP)/4XP ) We start with the quantity directly inherited from (Kuhlen & Faucher-Gigu`ere 2012), and H(z) is the Hub- our empirical model: the co-moving production rate of ble parameter. hydrogen-ionizing photons (n ˙ ion), i.e., the gross number We note there are caveats: for example, Case B recom- of LyC photons escaping into the IGM per unit time per bination may not be an appropriate description towards unit volume.n ˙ ion is usually computed as follows: the end of reionization when local absorption in dense clumps becomes important (Furlanetto & Oh 2005), or n˙ (z) = f ξ ρ [s−1Mpc−3], (2) ion esc ion UVdust corr CHII is bound to evolve as the universe grows more ion- ized (Pawlik et al. 2015) though its effect on reionization where ρUVdust corr is the dust-corrected UV luminosity. This UV-anchored formulation, where fesc is the es- inference is limited (Mason et al. 2019a; Bouwens et al. cape fraction of ionizing photons relative to the dust- 2015a). While we could test the effect of assuming dif- corrected observed UV, is apt for working with observa- ferent values for every individual parameter (on top of tions, where ρUV is the measured quantity around which varying the IMF, metallicity, underlying SED models, all else is based. However, in our model which is built and the truncation MUV), our guiding philosophy is to on a 106 Mpc3 simulation box, we simply sum the LyC hew to the canonical assumptions from the literature so photons, N(H0), produced by every galaxy from their all our divergent conclusions are clearly attributable to SEDs directly, reducing Equation2 to: the new data we constrain against and the models we introduce in this work. 0 The only free parameter in these equations is the es- X fescN(H ) n˙ (z) = [s−1Mpc−3]. (3) cape fraction, f . Our strategy to constrain f , and ion 106 esc esc MUV<−13.5 thus constrain reionization histories is to solve Equa- tions3-6 assuming a model for fesc and then fit the Here fesc is the escape fraction of ionizing photons relative to the total ionizing photons produced in the model parameters against the data described in the fol- galaxy. lowing section. ¯ The IGM ionized fraction, QHII, is evolved as per the 2.3. Observational Constraints on Reionization following differential equation where the first term rep- Here we enumerate state-of-the-art measurements on resents ionization, and the second recombination, the timeline of reionization that we use to constrain our ¯ f models. We briefly describe each measurement and ˙ n˙ ion QHII esc QHII = − , (4) specify how it is included in our inference. hnHi trec where hnHi = XpΩbρc is the co-moving density of Hy- 1. Thomson Optical Depth: τ = 0.0540±0.0074 from drogen, which depends on the primordial mass-fraction Planck Collaboration et al.(2018), which is a more Reionization by the Oligarchs 7

precise, downward revision of τ = 0.066 ± 0.012 4. z > 7 Quasi-stellar Objects (QSOs):x ¯HI = +0.26 +0.20 (Planck Collaboration et al. 2016) and a far cry 0.48−0.26 at z = 7.09 andx ¯HI = 0.60−0.23 at from the WMAP τ = 0.088 ± 0.014 (Hinshaw z = 7.54 (Davies et al. 2018a) from the IGM Lyα et al. 2013). τ bears the imprint of free electrons damping wing signature (Miralda-Escud´e 1998) in on photons from the last-scattering surface of the the quasars ULAS J1120+0641 (Mortlock et al. CMB and provides a global, integrated, model- 2011) and ULAS J1342+0928 (Ba˜nados et al. independent constraint. The lower value of τ from 2018). This constraint arises from a similar ap- Planck Collaboration et al.(2018) allows for the proach as Mason et al.(2018) in that, detailed sharp drop in ρSFR at z > 8 (Figure3) which was empirical models (Davies et al. 2018b), IGM sim- disfavored by earlier measurements (e.g., Robert- ulations (Mesinger et al. 2011) and radiative trans- son et al. 2013). fer (Davies et al. 2016) inform the inference ofx ¯HI from quasar spectra. We adopt their conservative 2. Lyα, Lyβ dark fraction:x ¯ ≤ 0.06 ± 0.05 at HI PDF forx ¯HI (their Figure 11). z = 5.9, as per the model-independent “dark frac- tion” in Lyα and Lyβ forests of quasar spectra 5. z = 6 − 7 Lyα Emitter (LAE) Fraction: The drop (Mesinger 2010; McGreer et al. 2015). The com- in number-density of LAEs between z ∼ 6 and pletely dark pixels in the forests are either due z ∼ 7 may be interpreted as the universe un- to neutral H I in the IGM and/or astrophysical dergoing drastic evolution in neutrality between interlopers and hence provide an assumption-free these redshifts (Mesinger et al. 2015) but may also upper limit to the global neutral fraction. This be due to survey incompleteness at faint magni- constraint allows us to impose the “end of reioniza- tudes (Oyarz´unet al. 2017). Greig & Mesinger tion” in a more self-consistent fashion than abrupt (2017) conservatively quantify this as implying truncation at some redshift (often fixed to z = 6 x¯HI(z = 7) − x¯HI(z = 6) ≥ 0.4 and we adopt their (e.g., Planck Collaboration et al. 2016, 2018)). We weak half-Gaussian constraint peaked atx ¯HI = 1 adopt it as uniform forx ¯HI < 0.06 and as a half- with σ = 0.6. Note that the dark fraction con- Gaussian peaked at 0.06 with σ = 0.05 elsewhere straint at z = 5.9 along with the z ∼ 7 Lyα EWs (Greig & Mesinger 2017, §3.1). already effectively reproduce this sharp change in neutrality. 3. z ∼ 7 − 8 Lyα (EW) Distribu- +0.11 +0.05 tions:x ¯HI = 0.59−0.15 at z ∼ 7,x ¯HI = 0.88−0.10 6. z ∼ 6.6 Lyα Emitter Clustering: The observed at z ∼ 7.5, andx ¯HI > 0.76 at z ∼ 8 (Mason et al. LAE clustering function at z ∼ 6.6 (Ouchi et al. 2018; Hoag et al. 2019; Mason et al. 2019b). The 2010) when interpreted in the context of the clus- Lyα line – in particular, its EW and its velocity off- tering in detailed reionization simulations suggests set from the systemic redshift – bears the imprint x¯HI < 0.5 (Sobacchi & Mesinger 2014) which we of the neutral IGM that Mason et al.(2018) infer implement as a half-Gaussian peaked at zero with using empirical fits for the ISM, and state-of-the- σ = 0.5 (Greig & Mesinger 2017). art IGM and Lyα radiative transfer simulations (Mesinger et al. 2016). While the evolution in the We exclude some measurements: Gamma Ray Burst fraction of Lyα emitters in Lyman-break galaxies (GRB) damping spectra, while probing some of the high- (e.g., Mesinger et al. 2015) encodes the evolution est redshifts (Chornock et al. 2013; Totani et al. 2006; ofx ¯HI(z), Mason et al.(2018) show that a more Tanvir et al. 2009), preferentially arise out of low-mass competitive constraint may be derived by utiliz- halos (Savaglio et al. 2009). Given the extreme scatter ing EW distributions. We use their full posterior along the lines of sight to such halos across a patchy PDF onx ¯HI (their Figure 11) and adopt scatter IGM, bounds onx ¯HI are fated to be weak (McQuinn in the redshift at which they reportx ¯HI as per et al. 2008). Lyα-based measurements that do not vary the selection function for their sample – centred the intrinsic line-width (e.g., Ouchi et al. 2010; Inoue on z = 6.9 and σz = 0.5 (Grazian et al. 2012; et al. 2018) are likely optimistic, given the vast diver- Pentericci et al. 2014), which is consistent with sity of ISM conditions in galaxies evident in line-shapes the scatter for similarly selected z-band dropouts already seen in z ∼ 2 − 3 samples (e.g., Trainor et al. (e.g., Bouwens et al. 2015b). Mason et al.(2019b) 2015, 2016). and Hoag et al.(2019) apply the same technique at In what follows we set up and justify two models for higher redshifts and we adopt their measurements fesc that we fit against all the constraints described in in similar fashion. this Section. 8 Naidu et al.

Figure 4. Summary of our fits to Model I, in which we assume a constant fesc for all galaxies at z > 6. Top Left: The allowed fesc parameter space implied by the reionization constraints described in §2.3. The model-independent Planck τ (blue) and z = 5.9 dark fraction (pink) rule out fesc . 10%, while the z ∼ 7 Lyα profiles (orange) and z > 7 QSOs (green) are most +0.06 constraining. The resulting fesc = 0.21−0.04 during reionization requires evolution in fesc from ∼ 10% at z = 3 and ∼ 0% at z ∼ 1 (e.g., Siana et al. 2010; Steidel et al. 2018). Top Right: The evolution ofx ¯HI, the IGM neutral fraction. The most likely reionization history is tracked in purple (1 and 3σ bounds shaded). Literature inferences of the neutral fraction are plotted in green (see §2.3). In the mean, reionization starts later and proceeds faster than what earlier constraints suggested (e.g., Robertson et al. 2015, shown in blue) or what the Planck τ alone implies (green square). Bottom Left: The evolution of the Thomson Optical Depth, τ. Our model’s drop in ionizing emissivity at z > 8 (Figure3) and thus lower τ (purple) were previously disfavored by WMAP (brown strip) and earlier Planck results (grey strip). However, the latest Planck τ (green strip) allows for it. Bottom Right: The duration of reionization in redshift-space against z50, the redshift of the 50% neutral universe. We find tight bounds on both z50 and z99 − z5 combining all our constraints, while τ by itself is only sensitive to z50 (e.g., Trac 2018). The blue contours representing τ come from the τ-fesc distribution (top left panel), and are not directly +0.24 inherited from Planck – they derive z50 = 7.64 ± 0.74 while we favor even later reionization with z50 = 6.83−0.20.

3. FITTING FOR fesc MODEL I: CONSTANT fesc diversity of galaxies and the highly likely dependence of DURING REIONIZATION fesc on various galaxy properties. However, this simple model provides a useful benchmark for the “average” Here we assume the fesc of all galaxies during reion- ization to be a constant number and denote this as escape fraction that observational stacking studies com- “Model I”. Effectively, we fit for a single normaliza- pute. Further, intrinsic galaxy properties (e.g., sizes, average star-formation rates) evolve modestly between tion factor, fesc, that sets the scale of the emissivity (solid curve in Figure3). This is the common approach z = 6 − 10 where the bulk of reionization is expected to adopted in several reionization studies (e.g. Robertson occur, hence assuming a constant average is justified. et al. 2015; Ishigaki et al. 2018). Model I ignores the Reionization by the Oligarchs 9

faint-end slope of the UVLF in §6.2 and §5.3). The most Cola1 constraining measurements on fesc prove to be from the

HST/COS Peas Green A2390- damping wing analysis of quasars and Lyα EW distri- H5 Ion2 butions which both require significant neutral fractions

Ion3 at later times (¯xHI ∼ 0.5 at z ∼ 7). In this constant fesc model we make no claims about the fesc at z < 6 – our result is situated in the reion- MS1358 izing universe. fesc= 0.21 is larger than the negligible fesc measured in deep stacks at z ∼ 0 − 1 (e.g., Siana et al. 2010; Rutkowski et al. 2016), where the IGM does HDUV Sample not impede observations, and the recently established Q1549- A2218 C25 f ∼ 10% at z ∼ 2.5 − 4 (Oesch et al. in prep. 2019; Flanking esc Marchi et al. 2017; Steidel et al. 2018; Fletcher et al. 2019). To self-consistently bridge these findings of fesc∼ 0% at z ∼ 0, fesc∼ 10% at z ∼ 2, and fesc∼ 20% at z > 6 we introduce Model II below, which accounts for an evolv- ing fesc while also considering the diversity in properties of individual galaxies.

Figure 5. The redshift-evolution of star-formation rate sur- 4. FITTING FOR fesc MODEL II: fesc AS A face density (ΣSFR). Our model (purple, 1σ shaded) remark- FUNCTION OF ΣSFR ably matches the observed ΣSFR shown in blue (Shibuya et al. 2015) at z > 4 by setting only a single parameter, Here we propose a model where fesc for each galaxy λ = Rvir/Rhalo = 0.031 which defines the normalization. is solely dependent on its star-formation rate surface ΣSFR grows by 2 dex between the local universe and the b & density ΣSFR, fesc= a × ΣSFR (where a and b are free Epoch of Reionization. Motivated in part by almost all con- parameters which we fit). We justify why this is an firmed LyC leakers to date (green stars) showing higher Σ SFR apt formulation, specify how it is implemented in our than the average at their redshifts, in our Model II we link empirical model, and discuss the constraints it yields. fesc to ΣSFR (see §4). The green stars represent leakers presented by Naidu et al.(2017); de Barros et al.(2016); Vanzella et al.(2016, 2018); Shapley et al.(2016); Bian et al. 4.1. Motivation: Why ΣSFR? (2017); Matthee et al.(2018); Borthakur et al.(2014); Izo- Almost all the individual observed LyC leakers to date tov et al.(2016a,b, 2018a,b); Leethochawalit et al.(2016); spanning z ∼ 0 − 6.6 show Σ higher than the aver- Jones et al.(2013); Naidu et al. (in prep.) and their vertical SFR age Σ expected at their redshifts. We demonstrate error-bars, typically <0.2 dex, are omitted and log(ΣSFR) is SFR capped to 2 for clarity. At z ∼ 0, the aggressive bunching of this in Figure5 where we have compiled all galaxies LyC leakers in the top-left is due to the Izotov et al.(2016b) for which convincing LyC leakage is reported, and that selection that successfully targeted Green-Pea galaxies with have sizes and UV SFRs available. These include the HST /COS. Between z ∼ 2 − 4 HST /F275W, HST /F336W HST /COS sample at z . 0.3 (Borthakur et al. 2014; and ground-based UV spectrographs (e.g., Keck/LRIS) come Heckman et al. 2011; Izotov et al. 2016a,b, 2018a,b), into play. Finally at z ≥ 4 when the IGM becomes opaque the HST /F336W, HST /F275W, and ground-based UV to LyC, indirect methods must be invoked (e.g., Lyα line- profiles, covering fraction of low-ionization gas). spectrograph sources at z ∼ 2−4 (de Barros et al. 2016; Vanzella et al. 2016; Shapley et al. 2016; Bian et al. 2017; Naidu et al. 2017; Vanzella et al. 2018, Naidu et al. in We assume a uniform prior between 0 and 1 on prep.), and sources that show strong indirect hints of f and depict the resulting posteriors projected in var- esc LyC escape at z ∼ 4−6.6 (low covering fractions: Jones ious spaces in Figure4. Combining all constraints we et al. 2013; Leethochawalit et al. 2016, tightly-spaced, find f =0.21+0.06. Simply requiring reionization to be esc −0.04 double-peaked Lyα resembling the local Green-Pea sam- mostly complete by z = 5.9 via the dark fraction rules ple: Matthee et al. 2018). The SFRs are all calculated out the f 15% parameter space (upper left panel of esc. from the UV because the average Σ vs. z relation Figure4). f 15% is also disfavored by the Planck SFR esc. we calibrate our model against is derived from the UV τ. Note that as the dark fraction and τ are model- (Shibuya et al. 2015). A caveat is that the striking abun- independent constraints not much can be invoked to al- dance of sources populating the top-left corner of Figure low for f 15% (we discuss M truncation and the esc. UV 5 were selected to be Green-Pea-like (e.g., Izotov et al. 10 Naidu et al.

2016b) for further follow-up, i.e., with very high ΣSFR, but it is nonetheless remarkable that the selection is so successful given the long history of LyC non-detections. While these individual LyC sources provide useful clues about the properties favoring LyC escape, they may be 0.6 b extreme outliers given their rarity. However, Marchi fesc = a SFR a = 1.6+0.3, b = 0.4+0.1 et al.(2018) find that even among normal star-forming 0.3 0.1 galaxies at z ∼ 4, the UV compact sources (which hence have higher ΣSFR) are likelier to be leaking LyC. 0.4 Independently, recent state-of-the-art hydrodynami- cal simulations have put forth the scenario of spatially b concentrated star-formation, turbulence, and feedback carving out channels in the ISM through which LyC pho- 0.2 tons can stream out of the galaxy (e.g., Ma et al. 2016; +Dark Frac. Sharma et al. 2016; Safarzadeh & Scannapieco 2016; + Neutral Frac. Trebitsch et al. 2017; Katz et al. 2018; Rosdahl et al. + Steidel+18 2018; Kakiichi & Gronke 2019; Kimm et al. 2019). For 0.0 instance, Ma et al.(2016) describe supernovae clearing 0 1 2 3 ionized channels in the ISM around stellar birth-clouds. a However, the massive stars exploding as supernovae are precisely the ones producing most of the ionizing pho- Figure 6. Relative constraining power of various data on Model II. While the global constraints on the evolution of tons. Invoking effects like binarity and rotation allow a the neutral fraction and τ (gray and orange) produce a de- significant population of UV luminous stars to survive generate surface, the fit is constrained by the Steidel et al. longer and pump LyC through the newly cleared ISM (2018) measurement (purple) of fesc that we adopt by assum- (Choi et al. 2017). The ionized channels visible in high- ing the ΣSFR for their sample follows the average relation in resolution Lyα spectra of LyC leakers (Vanzella et al. Figure5. We adopt uniform priors of 0 to 5 for a and −5 2019) support this picture. to 5 for b, allowing for a ΣSFR-fesc anti-correlation that is rejected by the evidence. The posteriors are instructive for 4.2. ΣSFR in our Empirical Model future studies: complex models of fesc only constrained by global quantities likex ¯HI and τ result in degenerate parame- We use the usual definition of ΣSFR (e.g., Shibuya ters (gray and orange above) and require measurements like et al. 2019): Steidel et al.(2018) (purple) that directly link fesc to galaxy SFR/2 properties. ΣSFR = 2 . (7) πRgal The SFR in our model is a function of the halo accre- We assume a simple power law dependence fesc= b −1 −2 tion rate and a redshift-invariant efficiency that converts a × (ΣSFR/ΣSFR,max) .ΣSFR,max = 1000 M yr kpc the halo accretion rate into an SFR. To calculate effec- is close to the value for the maximum ΣSFR that can tive radii (Rgal) for the galaxies in our model we assume be sustained without radiation pressure instabilities the angular momenta of the galaxies are a fixed fraction (Thompson et al. 2005; Heiderman et al. 2010; Hopkins of their DM halo (Mo et al. 1998). In particular, we re- et al. 2010). The scatter in λ produces a maximum −1 −2 late Rgal = λRhalo where λ is the spin parameter of the ΣSFR of typically ∼ 220 M yr kpc in our model. halo. We set λ = 0.031 to reproduce the observed ΣSFR- We fit for the coefficients a and b by summing the ion- z relation from Shibuya et al.(2015) (see Figure5). Re- izing photon contributions of each individual galaxy as markably, we are able to match the exact evolution of detailed in §2.2 and perform Bayesian inference against ΣSFR with redshift via this single parameter that only the reionization constraints from §2.3 using the dynesty sets the normalization. This is in line with Shibuya et al. nested sampling package (Speagle 2019). (2015)’s finding that the ratio of galaxy size and halo We add one additional constraint: the fesc of the Stei- size does not evolve significantly with redshift. We add del et al.(2018) sample, fesc= 0.09 ± 0.01 for a stack of log-normal scatter while assigning sizes, σlogλ = 0.22 ∼ 120 MUV < −19.5 LBGs that we assume follow the dex consistent with Kravtsov(2013); Somerville et al. average ΣSFR-z relation at z ∼ 3. While it is impossible (2018); Jiang et al.(2018). to robustly constrain fesc for individual sources at high-z due to the stochasticity of the intervening IGM, Steidel 4.3. Fitting for fesc∝ ΣSFR et al.(2018) stack in a narrow redshift bin across multi- Reionization by the Oligarchs 11

0.4 Figure 7. Summary of our fits to Model II, where we find fesc∝ΣSFR . The global reionization histories produced by Model II are very similar to those from Model I as seen through the evolution of the ionizing emissivity (top left) and the IGM neutral fraction (top right). However, the distribution of fesc among galaxies differs significantly. The bottom left panel plots ΣSFR as a function of stellar mass at z = 7 with points colored and sized according to their fesc (larger points denote higher fesc). Galaxies at log(M?/M ) ∼ 8 − 10 achieve significantly higher fesc than lower mass galaxies, though note the large scatter and that many of these galaxies are also able to attain fesc > 10%. In the bottom-right we show the evolution of the mean fesc (dashed) for the UV brightest (orange) and faintest (green) galaxies with 16th/84th and 5th/95th percentiles shaded. Faint galaxies with very low ΣSFR are limited to a mean fesc< 10% (though note the large scatter in the bottom-left panel) while the brightest galaxies are at fesc∼ 20%. The mean fesc across all galaxies (purple) remains a fairly flat ∼ 20% akin to Model I and as expected from the similar evolution inn ˙ ion seen in the top-left panel. ple lines of sight and correct for the mean IGM at that any individual source are highly uncertain due to the redshift. Further, the extremely deep spectra in their transmission along a single IGM line of sight being un- sample (∼ 10-hour exposures on a 10m telescope) show measurable. +0.3 +0.1 weak ISM lines that can be used to fine-tune models to We find a = 1.6−0.3 and b = 0.4−0.1 by deploying a match the covering fraction, correct for attenuation, and uniform prior over 0 to 5 for a and −5 to 5 for b. We produce a robust estimate of fesc. The key assumption assume a uniform prior such that 0 ≤fesc≤ 1, so our here is that the relationship between ΣSFR and fesc at best-fit relation effectively is: z ∼ 3 holds at higher-z – we argue that since fesc largely depends on the covering fraction of neutral gas at z > 3, and not dust, this is a justifiable assumption (see §6.1). We do not include any of the individual LyC leakers de- +0.1 !  0.4−0.1 picted in Figure5 in our fits because estimates of fesc for +0.3 ΣSFR fesc = min 1, 1.6−0.3 × . (8) ΣSFR,max 12 Naidu et al.

fesc∝(ΣSFR)0.4

1.6*((0.8/1000)**0.42)

0.4 Figure 8. The evolution of fesc as a function of stellar mass and redshift from Model II (fesc∝ Σ ). Top Left: The mean fesc at fixed stellar mass grows with redshift as galaxies grow more compact and star-forming though the fesc of the lowest mass galaxies remains negligible even at z = 8. The mean fesc in the highest mass bins at z ∼ 8 reaches ∼ 25% and at z = 4 it is comparable with current constraints at z = 2.5 − 4 on “normal” star-forming galaxies (green hatched region). Top Right: The fraction of galaxies with fesc above the mean during reionization (& 20%) shows similar trends. This is consistent with the current observational situation at z ∼ 2−4 (green hatched region) where a small fraction of sources like Ion2 (log(M?/M ) ∼ 9) show high fesc, even > 50%, while mean stacks (top-left) find humble estimates. We predict the fraction of Ion2 -like galaxies grows strongly at fixed mass. Bottom: fesc probability densities at z = 4 (left) and z = 8 (right) summarized in the top panels. The key features are the rightward shift of the distributions with increasing z and the high fesc tails in the right panel corresponding to the “oligarchs”. The tight posteriors are driven by the Steidel et al. physically motivated Model II (top panels of Figure7). (2018) constraint that directly links fesc to ΣSFR, while Which is to say, the evolution ofn ˙ ion(z) and average the constraints from §2.3 are useful in deciding the pos- fesc of ∼ 20% in both the models is similar during itive sign of the dependence (Figure6). We emphasize reionization. However, the way the similarn ˙ ion(z) is that in fitting for this power-law we allow for negative distributed among galaxies differs radically between the powers (i.e., an fesc-ΣSFR anti-correlation) that are re- two models in that instead of a constant fesc= 0.2 across jected by the evidence since they fail to conclude reion- all galaxies, a minority of galaxies that are more massive ization by z ∼ 6. and UV bright tend to have high ΣSFR and thus high Model I fits for a very similar evolution of the ion- escape fractions (bottom panels of Figure7). The pro- izing photon budget,n ˙ ion(z), compared to our more portion of these high ΣSFR galaxies as well as the mean Reionization by the Oligarchs 13

0.4 fesc∝(ΣSFR) fesc = 0.2 UV Luminosity Relative Number Density Stellar Mass Relative Number Density Relative Contribution to Ionizing Budget Halo Mass Redshift

Redshift

Figure 9. Which galaxies reionized the universe? The top, middle, and bottom panels show groupings of galaxies by MUV (observed), stellar mass, and halo mass respectively. The left and center columns are shaded by the relative contribution of each group to the total ionizing budget (n ˙ ion) as per our two models for fesc. The right-most column is colored by the relative number-density of each group. The black dotted lines in each panel sandwich the redshift-space when the universe is inferred to go from 90% (z90) to 10% neutral (z10). In the central panel, the reionization budget is essentially a reflection of our predicted UVLF since ξion does not vary strongly with luminosity and fesc is constant, while in the left panel the fesc∝ΣSFR model further down-weights the contribution of faint galaxies. UV-bright and massive sources represented in orange and blue dominate the reionization budget (& 50 − 80%) despite comprising . 5% of the population. This scenario, “reionization by oligarchs”, stands in stark contrast to the canonical “democratic reionization” led by copious faint sources.

ΣSFR grows with redshift driven primarily by their in- tion of reionization (z90 − z10) is tightly constrained to +0.05 creasing compactness and naturally explains the evolu- be ∆z = 3.76−0.04 (bottom right panel, Figure4). The +0.25 tion of fesc from ∼ 0% at z ∼ 0 to ∼ 10% at z ∼ 2.5 − 4 universe goes from 90% neutral at z = 8.22−0.22 to 10% +0.26 to ∼ 20% at z > 6 (Figure8). neutral at z = 6.25−0.22, in ∼ 300 Myrs (see Table2). This pace is faster than estimated by earlier studies (e.g., 5. DISCUSSION Robertson et al. 2015; Planck Collaboration et al. 2018) 5.1. Rapid Reionization at z = 6 − 8 and is driven by the sharp drop in the ionizing emissiv- ity at early times (Figure2). The high neutral fraction Both our models produce virtually identical reioniza- measurements at late times ( 50% at z ∼ 7) from Lyα tion histories (top right panel of Figure7). The mid- & +0.24 damping combined with the dark fraction requirement point of reionization (z50) is 6.83−0.20 while the dura- 14 Naidu et al.

Table 1. Which galaxies reionized the universe? Mean properties weighted by contribution to reionization (n ˙ ion) for Model II (Model I).

Parameter z = 6 z = 7 z = 8 z = 9 z = 10 0.42 fesc = 1.62 × ΣSFR (fesc=0.21)

fesc 0.24 (0.21) 0.26 (0.21) 0.24 (0.21) 0.21 (0.21) 0.19 (0.21) a MUV -19.7 (-18.9) -19.5 (-18.6) -19.2 (-18.2) -18.7 (-17.8) -18.3 (-17.3)

log(M?/M ) 9.5 (9.2) 9.2 (8.9) 8.9 (8.6) 8.6 (8.4) 8.3 (8.0)

log(Mhalo/M ) 11.1 (10.9) 10.9 (10.7) 10.7 (10.6) 10.5 (10.4) 10.4 (10.2) −1 SFR/M yr 15.3 (9.3) 12.1 (6.5) 7.7 (4.0) 3.5 (2.1) 2.0 (1.0) −1 −2 ΣSFR/M yr kpc 17.7 (8.2) 23.2 (9.7) 17.6 (8.1) 11.6 (5.5) 9.3 (4.3)

Rgal/kpc 0.59 (0.62) 0.46 (0.48) 0.38 (0.39) 0.33 (0.34) 0.27 (0.27)

(a) This is the observed MUV, i.e., we have applied attenuation to the intrinsic MUV as described in §2.1.

Table 2. A Rapidly Reionizing Universe never exceeds an average of ∼ 25% even in the highest mass bin (log(M?/M ) = 9 − 10). At z ∼ 4 we predict Redshiftx ¯HI (Model I)x ¯HI (Model II) a mean fesc of ∼ 10% with ∼ 10% of sources displaying +0.05 +0.08 6 0.00−0.00 0.01−0.01 fesc> 20% for galaxies at log(M?/M ) = 9 − 10. This +0.08 +0.03 is a faithful representation of the current observational 7 0.58−0.12 0.64−0.04 +0.03 +0.01 situation at z ∼ 2 − 4 where stacks produce average 8 0.87−0.04 0.89−0.01 fesc of ∼ 10% (Marchi et al. 2017; Steidel et al. 2018; 9 0.95+0.01 0.97+0.01 −0.01 −0.01 Oesch et al. in prep. 2019) while a small fraction of +0.01 +0.01 10 0.99−0.01 0.99−0.01 sources show fesc > 20%, even reaching fesc> 50% (e.g., Naidu et al. 2017; de Barros et al. 2016; Vanzella et al. 2018) for galaxies in a similar mass range. In fact, in the for the end of reionization by z ∼ 5.9 favor this rapid Steidel et al.(2018) sample at z ∼ 3 (which we approx- pace. A corollary of this timeline is that efforts to under- imately compare with our predictions at z ∼ 4) 10/124 stand the sources of reionization do not have to probe sources (∼ 8%) show significant LyC leakage while the the highest redshifts since more than half of the process stacked mean is ∼ 10% – the agreement in the frac- occurs between z = 6 − 7. tion of sources with high fesc is noteworthy since we fit our model against the mean f and the fraction 5.2. The ‘Oligarchs’ that Reionized the Universe esc of fesc> 20% galaxies is a genuine prediction (top-right In both our models, especially in Model II, we find panel of Figure8). the reionization budget (n ˙ ion) is concentrated in a ultra- In Model I this distribution of the ionizing budget is a minority of galaxies with high ΣSFR at MUV< −18, direct reflection of the shape of the UVLF arising from log(M?/M ) > 8 (see Figure9 and Table1). In Model our model since ξion does not vary with MUV (Figure2) II less than 5% of galaxies constitute & 80% of the reion- and all galaxies have the same fesc. Steeper αUV keeping ization budget. Adopting a popular income inequal- all else same, will lead to a lower average fesc, larger ity measure from Macroeconomics, the Gini coefficient luminosity densities at early times, a less oligarchic dis- (Gini 1912), we find then ˙ ion distributions for Model II tribution, and possibly tension with current constraints at z =[6, 7, 8] have Gini coefficients of [0.93, 0.92, 0.90]. favoring late and rapid reionization (see §5.3). However, For reference, a distribution comprised of equal numbers in Model II, MUV > −17 galaxies are limited to very low has a Gini coefficient ∼ 0 and a distribution where all fesc and they constitute a negligible portion of the reion- the density is held by a single member has a coefficient ization budget. Truncating as high as MUV = −17 has of ∼ 0.99. Drawing again from the language of wealth no effect on the model parameters shown in Figure6, concentration, we christen this ultra-minority of galaxies i.e., the ionizing emissivity between MUV = −13.5 to that dominate reionization the “oligarchs”. −17 is severely down-weighted by assigning a low fesc. In Figure8 we show the occurrence of the oligarchs Even with steeper αUV, we expect the oligarch scenario grows with redshift as ΣSFR increases and galaxies be- to hold since Model II has the flexibility to ensure late come more compact and star-forming. Consequently, reionization as required by the constraints by setting the mean fesc also grows with redshift but note that it Reionization by the Oligarchs 15 fesc ∼ 0 for the numerous faint galaxies with low ΣSFR. et al. 2019), in tension with the damping wing mea- We discuss αUV further in §5.3. surements (¯xHI ∼ 90%). Using a constant fesc across all galaxies with αUV≤ −2 like in Ishigaki et al.(2018) and 5.3. “Democratic” Reionization by Faint Galaxies and Robertson et al.(2015) requires integrating to −MUV= the Faint-End Slope of the UVLF in a Rapidly 11 − 13 and still makes for too low of a neutral fraction Reionizing Universe (∼ 60 − 70%) at z = 8. Simply lowering the constant f in these models would delay reionization but then Reionization by oligarchs stands in sharp contrast to esc it would not conclude by z ∼ 6 – raising the f while “democratic” reionization that is dominated by copi- esc lowering the M (trunc.) and/or shallower α are ous faint sources that lie at M > −18 and might po- UV UV UV needed. We illustrate this in the right panel of Fig- tentially have high escape fractions (e.g., Oesch et al. ure 10, where we assume Schechter parameters from 2009; Bouwens et al. 2011; Wise et al. 2014; Atek Finkelstein et al.(2019), ξ from this work, and a con- et al. 2015; Anderson et al. 2017; Livermore et al. 2017; ion stant f to evaluate how likely various combinations of Finkelstein et al. 2019). Faint galaxies emerged as esc f and M truncation are (as per constraints from the candidate-leaders of reionization because the steep esc UV §2.3). A truncation M of ≤ −15, implying a limited slopes (α ≤ −2 at z > 6) of the UVLF measured UV UV role for fainter galaxies is favored by the constraints. after the installation of HST /WFC3 implied they dom- The general feature ofn ˙ dominated by faint galax- inated the luminosity density (e.g., Bouwens et al. 2012). ion ies in the models discussed above is thatn ˙ is already The τ measurements from WMAP-9 (0.089 ± 0.014, ion high at z = 10, resulting in smooth and early reioniza- z = 10.5 ± 1.1) and Planck Collaboration et al.(2016) 50 tion (Figures3 and 10). (0.066±0.013, z = 8.8±1.3) required significant reion- 50 On the other hand, the required late and rapid reion- ization at z > 8 and hence large contributions towards ization is naturally produced by shallower (α ≥ −2) the ionizing emissivity from faint galaxies (e.g., Robert- UV faint-end slopes (Model I, Model II), distributions of son et al. 2013, 2015; Bouwens et al. 2015a). Concur- f highly skewed toward brighter galaxies (Model II), rently, the very low f reported for bright star-forming esc esc and/or a sharp drop in the z > 8 ρ in models linking galaxies out to z ∼ 4 (see §1) and the sharply dropping SFR star-formation to dark matter accretion (e.g. Model I, AGN luminosity function (Kulkarni et al. 2019) further Model II, Mason et al. 2015; Mashian et al. 2016). Trun- shifted the spotlight onto faint star-forming galaxies. cating at M = −16 (−17) in Model I (Model II) pro- However, the recent constraints on neutral fractions UV duces no change in the reionization histories and model detailed in §2.3 and the latest Planck τ favor late, parameters, i.e., the ionizing emissivity requires no con- rapid reionization between z = 6 − 8 (we calculate tributions from M > −16 (−17) galaxies (Figures9 z = 6.83+0.24 for Model I) i.e., high emissivity from UV 50 −0.20 and 11). M > −16 galaxies are rare in the early uni- faint galaxies at z > 8 is no longer required. This, and UV verse and their appearance causesn ˙ to rise steeply by the high average f measured even for more massive, ion esc more than a dex between z = 6 − 10 (Figure3). Thus, M < −18 galaxies allow for reionization by the oli- UV while the faint-end slope of the UVLF may indeed be garchs. At z > 8,n ˙ must be low enough for the ion extremely steep, galaxies at M > −16 must play only universe to remain significantly neutral ( 90%), and UV & a minimal role in order to achieve rapid and late reion- between z = 8 − 6 it must rise sharply to complete ization. Model II explains this as very low fesc occurring reionization. Since ξion evolves modestly with redshift in these abundant albeit low ΣSFR galaxies. and across MUV (see Figure2), ˙nion effectively depends on ρSFR (αUV, MUV truncation) and fesc. 5.4. Observing the Oligarchs in Action: Promising Latest studies report αUV . −2 at z ≥ 6, albeit with significant uncertainties, that grows steeper with red- Hints and Future Prospects shift at a rate dα/dz ∼ −0.1 (e.g., Finkelstein et al. 2015; The luminous Lyα emitter, COLA1 at z = 6.6 (fesc∼ 2 Livermore et al. 2017; Bouwens et al. 2017; Atek et al. 30%, MUV= −21.5, ΣSFR = 100 M /yr/kpc ) is a 2018; Ishigaki et al. 2018; Oesch et al. 2018). We com- poster-child oligarch (Hu et al. 2016; Matthee et al. pare our reionization histories with models that assume 2018). It displays double-peaked Lyα with a low-peak these steep slopes and model ρSFR based on Schechter separation reminiscent of local LyC leakers (Verhamme parameters extrapolated from z < 10 fits in Figure et al. 2017). More statistically, Songaila et al.(2018); Hu 10. Assuming αUV< −2 and setting fesc preferentially et al.(2016); Matthee et al.(2018) find luminous Ly α higher in the faintest galaxies requires integration down emitters at z ∼ 7 have line profiles that are broader to MUV= −10 and reionizes large volumes of the z > 8 (while not being Active Galactic Nuclei (AGN)) and universe, reachingx ¯HI ∼ 40% at z = 8 (Finkelstein more complex than their lower luminosity counterparts 16 Naidu et al.

Figure 10. Comparison with reionization by faint galaxies. Left: In turquoise we plotx ¯HI(z) from Finkelstein et al.(2019) who explore reionization dominated by MUV> −15 galaxies with steep faint-end slopes (α < −2) and the highest fesc occurring in the least massive galaxies by integrating down to MUV = −10. In gold we plotx ¯HI(z) from Ishigaki et al.(2018) who assume a constant fesc and αUV < −2 to find fesc= 17% and MUV(trunc) = −11 in order to complete reionization by z = 6. Both these models ionize a large volume of the universe at early times, in tension with Lyα damping wing constraints (green stars and pentagons). On the other hand, the shallower faint-end slopes (αUV> −2) and fesc distributions highly skewed toward bright galaxies in our models ensure rapid, late reionization (purple curves). Right: Assuming Schechter parameters from Finkelstein et al.(2019), a constant fesc across all galaxies, and ξion from this work, we show the likelihood of various combinations of fesc and MUV-truncation arising from the constraints in §2.3. When the ionizing emissivity is dominated by faint galaxies (MUV> −15), even with very low fesc, early reionization occurs, and such scenarios are disfavored compared to those starring brighter galaxies. with two sources in a sample of seven showing blue et al.(2018) find Ly α emission arising only from three wings despite a highly neutral IGM (e.g., Mason et al. UV-bright galaxies among the dropouts while all their 2018). We speculate these galaxies are oligarchs with faint galaxies are undetected in Lyα despite Lyα EWs high escape fractions that are able to carve out their generally anti-correlating with brightness. We speculate own ionized bubbles perhaps allowing for their whole the bright oligarchs with high fesc have reionized their line profiles (including blue wings and peaks) to es- immediate surroundings rendering them transparent to cape unattenuated by neutral gas. The lower luminosity, Lyα while the fainter sources lie just outside these ion- low fesc sources have narrow, less complex Lyα profiles ized bubbles. With JWST ’s planned censuses at high-z that are perhaps truncated by the neutral gas surround- more such ionized overdensities at z > 6 will come into ing them. High-resolution (R > 4500) Lyα surveys view and deep follow-up spectroscopy that reveals fea- with well-defined selection and completeness functions tures of LyC fesc (e.g., multi-peaked Lyα) will help test at z ∼ 0 − 6 will help test if these complex Lyα profiles if they are indeed powered by oligarchs. that have been linked to ionized channels and thus LyC Our proposed scenario also has strong implications for fesc (Vanzella et al. 2019; Rivera-Thorsen et al. 2019; the topology of reionization, with the distribution of Herenz et al. 2017) grow more common with redshift ionized bubble sizes and the patchiness resulting from and with galaxy properties like ΣSFR. Since we do not our model likely lying somewhere intermediate between expect fesc to evolve appreciably between z ∼ 6 − 8 as AGN-driven reionization (e.g., Kulkarni et al. 2017) ΣSFR flattens (Figure5), such a survey can be limited and reionization by widely distributed, numerous faint to z < 6 where the IGM transmission is higher and Lyα sources (e.g., Geil et al. 2016). Upcoming 21cm surveys is easily observable. will provide a strong test of this prediction (e.g., Hutter Another intriguing observation is that of an overden- et al. 2019a,b; Seiler et al. 2019). Our empirical model sity of 17 HST dropouts at z ∼ 7. In an extremely also tracks the spatial distribution of galaxies and this long integration (22.5 hrs on VLT/VIMOS), Castellano information can be coupled with models for fesc to pro- Reionization by the Oligarchs 17 duce more quantitative predictions. We defer this to future work.

5.5. Related Work: the fesc-ΣSFR Connection

Heckman et al.(2001) explicitly link fesc to a criti- 2 cal value of ΣSFR ∼ 0.1M /yr/kpc above which they observe the occurrence of strong winds becomes com- mon in star-burst galaxies. They hypothesize that these winds are responsible for LyC leakage. Sharma et al. (2016) adopt this idea in the EAGLE simulations (Schaye et al. 2015; Crain et al. 2015) setting fesc= 0.2 for star- forming regions above the critical ΣSFR and averaging these in each galaxy to find an emissivity consistent with Bouwens et al.(2015a). The fesc= 0.2 hard upper-limit was motivated by the lack of LyC detections, prior to the recent LyC renaissance detailed in §1. Figure 11. The effect of M truncation on our We do not assume a threshold or an upper bound em- UV fesc posterior from Model I. fesc converges as we go to lower boldened by recent discoveries (see Figure5) and em- MUV. We do not expect any significant change from exten- pirically constrain an fesc-ΣSFR dependence. We use sions to even fainter magnitudes as the fractional increase in a prior that allows for no relation (b = 0) or an anti- the cumulativen ˙ ion becomes negligible at MUV> −16. This correlation (b < 0) and fit against the latest reionization is expected since our model produces αUV> −2, and thus constraints. Note that the Sharma et al.(2016) prescrip- convergent cumulative luminosity densities, during the EoR. tion, even though it invokes ΣSFR , ends up effectively +0.06 similar to our Model I that fits fesc=0.21−0.04, since es- 2016b; Fudamoto et al. 2017) though significant uncer- sentially all galaxies at z ∼ 6 − 8 at MUV < −13.5 have tainties persist (e.g., Casey et al. 2018). Our attenuation −1 −2 ΣSFR> 0.1M yr kpc (see bottom left panel in Fig- prescription bears this out – for instance, on average, an ure7) and have sizes on the order of 1 kpc (their beam MUV =[−22,−20,−18] galaxy at z = 4 has AUV=[1.51, size). Furthermore, Sharma et al.(2016) find EAGLE 1.05, 0.6] and a z = 8 galaxy has AUV=[1.40, 0.6, galaxies at MUV < −18 produce ∼ 50% of the reioniza- 0.0]. The key physical picture motivating Model II, tion budget. Our budget is far more oligarchic (> 80%), that of spatially concentrated star-formation, winds, and our reionization far more rapid (z ∼ 6−8; see §5.1). and feedback carving out ionized gas channels is inti- mately linked with fcov and thus needs no extra dust 6. CAVEATS AND OPEN QUESTIONS parameter, since fcov essentially determines fesc. How- ever, we note that not explicitly modeling dust in Model 6.1. On Dust and fesc at z > 6 II prevents a simple extrapolation of f using our fit In our Model I, f implicitly folds in the role of esc esc power-law to z ∼ 0 where attenuation of LyC by dust dust and all other processes that may curtail LyC leak- is highly significant though the qualitative picture and age. This is not the case in Model II in which f is esc general trend stands. solely a function of ΣSFR. Further, we adapt the Steidel et al.(2018) measurement to constrain Model II assum- 6.2. Effect of MUV Truncation ing that the relationship between ΣSFR and fesc at z ∼ 3 carries over to higher-z, unmodulated by dust. We jus- We have limited all our calculations to galaxies with −1 tify these assumptions here. SFR > 0.02 M yr which corresponds approximately At z ∼ 3 already, deep stacks of typical LBGs show to MUV(observed)< −13.5. This limitation arises from that it is not dust, but photoelectric absorption in the the resolution of the dark matter simulations our model ISM that dominates the attenuation of LyC photons to is built on as well as the significant uncertainties around the extent that the UVLF fainter than this magnitude (Livermore et al. 2017; Bouwens et al. 2017; Atek et al. 2018). What effect fesc ≈ 1 − fcov does this truncation have on our results? where fcov is the H I covering fraction (Reddy et al. In Model I extending to fainter magnitudes adds to the 2016a,b; Steidel et al. 2018). Moving into the EoR, this ionizing emissivity and should lower the average fesc we approximation is likely even better since the dust atten- report. However, since our model has αUV> −2 during uation at z > 6 appears to be lower (Bouwens et al. reionization we expect this lowering to become negligible 18 Naidu et al. at MUV> −13.5, since the differential change to the bulk at z = 6 (e.g., Kulkarni et al. 2019, though see Gial- n˙ ion asymptotes to zero. We explore this by shifting the longo et al. 2015; Boutsia et al. 2018). In a companion limiting magnitude brighter (see Figure 11). We find our work we deploy a similar framework, but assume noth- fesc solution is essentially converged at MUV< −16 since ing about the underlying ionizing population and fit a moving down to MUV< −13.5 produces no appreciable non-parametric model to recovern ˙ ion(z)(Mason et al. change. Moving further down to MUV< −11 should 2019b). A sharp drop inn ˙ ion(z) at z = 6−8 fully consis- make an even smaller difference especially if the UVLF tent with the galaxy-only models presented in this work turns over around these magnitudes due to inefficient is recovered. star-formation and photo-evaporation in low-mass halos A related issue is whether our models overrun the (e.g., Gnedin 2016). constraints on ionizing emissivity at lower-z plotted in In Model II the majority of the low-luminosity galax- the top-left panel of our Figure7 when combined with ies have extremely low fesc (see bottom panels of Figure the AGN emissivity at lower redshifts (Becker & Bolton 8) to go along with their lower LyC output so the exclu- 2013; Kuhlen & Faucher-Gigu`ere 2012). Note that in sion of MUV> −13.5 galaxies or the faint-end slope have Model I we make no claims about fesc at z < 6 and fit a negligible impact on model parameters. We have verified ∼ 20% fesc during the short window when reionization that even truncating as high as MUV< −17 produces transpires. In Model II we have an evolving fesc that very similar reionization histories to those reported here. falls from ∼ 20% at z > 6 to ∼ 10% (∼ 4%) at z = 3 (z = 2). This causesn ˙ ion to flatten and turn-over at 6.3. Model Dependent Constraints lower-z so that AGN can dominate the emissivity – we The Lyα damping measurements for z > 7 quasars begin to see this in the top-left panel of Figure7 (dotted and galaxies prove to be the most constraining for purple curve). our Model I. However, these are model dependent con- 6.5. Cosmic Variance and Completeness straints in that their reportedx ¯HI is a product of sev- eral assumptions, e.g., about how Lyα is processed by Due to the finite volume of the N-body simulation on the ISM at high-z or about the intrinsic spectrum of which our empirical model is built (106 Mpc3), we miss reionization epoch quasars. These assumptions while some of the most massive halos (see Figure 20 and Ap- reasonable are yet to be tested (e.g., see §2 in Mason pendix B in Tacchella et al. 2018). At z > 6 these halos et al. 2018). In Model I the model-independent τ and also tend to be the most star-forming and UV bright. dark fraction by themselves are unable to zero in on an Comparing to an analytical halo mass function (Sheth fesc solution, but they rule out fesc. 15% and so fa- et al. 2001) and applying a completeness correction pro- vor rapid reionization histories. In Model II the Lyα duces a . 0.3 dex difference at the brightest end of the damping measurements are unable to collapse the pos- UVLF at z = 10. The correction is smaller at lower red- terior much further beyond τ and the dark fraction com- shifts, where the bulk of reionization occurs and hence bined, and the Steidel et al.(2018) measurement proves the magnitude of the effect is likely small. For Model I, crucial (see Figure6). This measurement depends on the mean fesc estimated would slightly shrink due to the several assumptions e.g., stellar population model pre- missing luminosity density. For Model II, extrapolating dictions at < 912A,˚ the IGM+CGM transmission func- from the trends shown in Figure8 for the proportion tions, and details of the “hole” and “screen” models for of oligarchs as a function of galaxy mass at z > 6, in- fesc developed in their work. These are all sources of sys- cluding these massive halos would make the reionization tematic uncertainty on the reported absolute fesc. We budget even more oligarchic. An update to the empirical have tested that the sign of the power (> 0) recovered model using a larger box that also self-consistently in- for the fesc-ΣSFR dependence is not sensitive to the ex- cludes AGN is currently under preparation. The larger act scale of their reported fesc as long as fesc> 0, and box will also allow us to address cosmic variance by re- the dark fraction measurement that ensures the timely sampling smaller volumes and computing the resulting conclusion of reionization is used. scatter introduced in fesc.

6.4. Ionizing Emissivity at z < 6 and the Role of AGN 7. SUMMARY In this work we have focused on reionization driven by Using an empirical model that accurately predicts ob- galaxies and fit for parameters that show them satisfying servations (e.g., the sharp drop in ρSFR at z > 8, UVLFs all constraints outlined in §2.3 without invoking AGN. at z > 4, ξion at z ∼ 4 − 5, the z-evolution of ΣSFR) and This is supported by latest determinations of AGN lumi- leveraging recent measurements of the timeline of reion- nosity functions that limit their contribution to < 5% ization (e.g.,x ¯HI at z & 7, the Planck τ) we constrain Reionization by the Oligarchs 19 fesc, the most uncertain parameter in reionization cal- the universe at z = 7 − 8, in tension with Lyα culations. Deploying two models – one assuming a con- damping constraints that require a 60 − 90% neu- stant fesc across all galaxies during reionization (Model tral universe at these redshifts. Shallower faint- I), another linking fesc to ΣSFR (Model II) – we find the end slopes (αUV > −2) and/or fesc distributions following: skewed toward massive galaxies like in our models ensure high neutral fractions at late times while • In both our models, M <−13.5 star-forming UV also completing reionization by z = 6. Concur- galaxies need an average f ∼20% at z > 6 to con- esc rently, the motivation for excluding galaxies at clude reionization, a factor of only 2 higher than M < −18 with high f as the protagonists the recently measured f ∼ 10% at z ∼ 2.5 − 4. UV esc esc of reionization has grown weaker as the observa- [Figures4,7] tional picture has shifted to these galaxies being • Our Model II explains this evolution in fesc by able to produce fesc ∼ 10% at z = 2.5 − 4. [Figure appealing to ΣSFR that decreases by ∼2.5 dex be- 5, 10] tween z = 8 (f ∼ 0.2) and z = 0 (f ∼ 0) by esc esc Our predictions are eminently testable since the oli- fitting f ∝ (Σ )0.4. This f -Σ connection esc SFR esc SFR garchs are bright, currently observable galaxies. Deep is inspired by the newly emerging sample of LyC Lyα surveys at high-resolution (R > 4500) spanning leakers that show higher Σ than the average SFR z ∼ 0 − 6 should show a growing incidence of galaxies galaxy population at their redshifts. Latest hydro- with multi-peaked Lyα at higher-z. These peaks rep- dynamical simulations qualitatively support the resent ionized channels for LyC escape, as seen in the idea of spatially concentrated star-formation blow- z 4 LyC leakers. Upcoming 21 cm experiments should ing channels in the ISM through which LyC pro- . infer a bubble size distribution with a high Gini coeffi- duced by long-lived, rotating, binary stars escapes. cient as the first ionization fronts form predominantly [Figures5,6] around the oligarchs. • The universe goes from 90% neutral to 10% neu- tral in a short span of ∼ 300 Myrs between We are grateful to the referee, Brant Robertson, for z ∼ 6 − 8, and favored by Lyα damping mea- a thorough report that greatly improved the clarity of surements requiring a & 50% neutral universe at this work. It is a pleasure to acknowledge illuminat- z ∼ 7. This conclusion stands even considering ing discussions with Steve Finkelstein, Jorryt Matthee, only model-independent constraints (τ, dark frac- Daniel Schaerer, David Sobral, Irene Shivaei, Ben John- tion) that rule out fesc. 15%. [Figure4, Table son, and Chuck Steidel that enriched this work. We 2] thank Daniel Eisenstein, Avi Loeb, and Lars Hernquist for participating in a robust discussion of these results • The bulk of the reionization budget (∼50% in during RPN’s research exam. We thank Steve Finkel- Model I, ∼80% in Model II) is concentrated stein for sharing the posteriors from Finkelstein et al. among a small number (<5%) of galaxies (the (2019) featured in Figure 10. We are grateful to Eros “oligarchs”). This is due to the faint-end slopes Vanzella for sharing their data on Ion3 and for useful of the UVLF (α > −2) in our model and the UV comments on Naidu et al.(2018) that sparked parts of distribution of f skewed toward high Σ , esc SFR this work. RPN thanks Michelle L. Peters for a hilarious massive galaxies. The oligarchs are compact proofreading session that highlighted the ludicrous ver- (R ∼ 0.5 kpc), have higher Σ than average gal SFR biage we deploy in the astrophysics community. RPN (∼10−20 M yr−1kpc−2), are relatively massive thanks the organizers and attendees of IAU Symposium (log(M /M )>8) and are UV bright (M <−18). ? UV 352 “Uncovering early galaxy evolution in the ALMA The fraction of these oligarchs grows with redshift, and JWST era” held in Viana do Castelo for provid- while keeping the average f to ∼20%. Extrapo- esc ing much-needed affirmation that he is not alone in the lating to z ∼ 3 − 4 we match the current situation often testing endeavour that is working on f . where a small fraction of galaxies ( 10%) display esc . RPN gratefully acknowledges an Ashford Fellowship f > 0.2, some even exceeding 50%, while the esc and Peirce Fellowship granted by Harvard University. average f stays at ∼10%. [Figures8,9, Table esc ST is supported by the Smithsonian Astrophysical Ob- 1] servatory through the CfA Fellowship. CAM acknowl- • Faint galaxies are disfavored to drive reionization. edges support by NASA Headquarters through the When faint galaxies with steep αUV < −2 dom- NASA Hubble Fellowship grant HST-HF2-51413.001- inate the emissivity, they ionize large volumes of A awarded by the Space Telescope Science Institute, 20 Naidu et al. which is operated by the Association of Universities for Danish National Research Foundation. CC acknowl- Research in Astronomy, Inc., for NASA, under contract edges funding from the Packard foundation. NAS5-26555. SB acknowledges Harvard University’s Software: IPython (P´erez& Granger 2007), ITC Fellowship. PAO acknowledges support from the matplotlib (Hunter 2007), numpy (Oliphant 2015), Swiss National Science Foundation through the SNSF scipy (Jones et al. 2001--), jupyter (Kluyver Professorship grant 157567 “Galaxy Build-up at Cos- et al. 2016), seaborn (Waskom et al. 2018), FSPS mic Dawn”. The Cosmic Dawn Center is funded by the (Conroy et al. 2009, 2010), dynesty (Speagle 2019), python-FSPS (Foreman-Mackey et al. 2014)

REFERENCES

Anderson, L., Governato, F., Karcher, M., Quinn, T., & Chornock, R., Berger, E., Fox, D. B., et al. 2013, ApJ, 774, Wadsley, J. 2017, MNRAS, 468, 4077 26 Atek, H., Richard, J., Kneib, J.-P., & Schaerer, D. 2018, Conroy, C., Gunn, J. E., & White, M. 2009, ApJ, 699, 486 MNRAS, 479, 5184 Conroy, C., White, M., & Gunn, J. E. 2010, ApJ, 708, 58 Atek, H., Richard, J., Jauzac, M., et al. 2015, ApJ, 814, 69 Crain, R. A., Schaye, J., Bower, R. G., et al. 2015, Ba˜nados,E., Venemans, B. P., Mazzucchelli, C., et al. 2018, MNRAS, 450, 1937 Nature, 553, 473 Davies, F. B., Furlanetto, S. R., & McQuinn, M. 2016, Becker, G. D., & Bolton, J. S. 2013, MNRAS, 436, 1023 MNRAS, 457, 3006 Behroozi, P., Wechsler, R. H., Hearin, A. P., & Conroy, C. Davies, F. B., Hennawi, J. F., Ba˜nados,E., et al. 2018a, 2019, MNRAS, 1134 ApJ, 864, 143 Bian, F., Fan, X., McGreer, I., Cai, Z., & Jiang, L. 2017, —. 2018b, ApJ, 864, 142 ApJL, 837, L12 Dayal, P., & Ferrara, A. 2018, PhR, 780, 1 Borthakur, S., Heckman, T. M., Leitherer, C., & Overzier, de Barros, S., Vanzella, E., Amor´ın,R., et al. 2016, A&A, R. A. 2014, Science, 346, 216 585, A51 Dotter, A. 2016, The Astrophysical Journal Supplement Boutsia, K., Grazian, A., Giallongo, E., Fiore, F., & Series, 222, 8 Civano, F. 2018, ApJ, 869, 20 Endsley, R., Behroozi, P., Stark, D. P., et al. 2019, arXiv Bouwens, R. J., Illingworth, G. D., Oesch, P. A., et al. e-prints, arXiv:1907.02546 2015a, ApJ, 811, 140 Fan, X., Narayanan, V. K., Lupton, R. H., et al. 2001, AJ, Bouwens, R. J., Oesch, P. A., Illingworth, G. D., Ellis, 122, 2833 R. S., & Stefanon, M. 2017, ApJ, 843, 129 Finkelstein, S. L., Ryan, Russell E., J., Papovich, C., et al. Bouwens, R. J., Smit, R., Labb´e,I., et al. 2016a, ApJ, 831, 2015, ApJ, 810, 71 176 Finkelstein, S. L., D’Aloisio, A., Paardekooper, J.-P., et al. Bouwens, R. J., Illingworth, G. D., Oesch, P. A., et al. 2019, ApJ, 879, 36 2011, ApJ, 737, 90 Fletcher, T. J., Tang, M., Robertson, B. E., et al. 2019, —. 2012, ApJ, 754, 83 ApJ, 878, 87 —. 2014, ApJ, 793, 115 Foreman-Mackey, D., Sick, J., & Johnson, B. 2014, —. 2015b, ApJ, 803, 34 python-fsps: Python bindings to FSPS (v0.1.1), Bouwens, R. J., Aravena, M., Decarli, R., et al. 2016b, ApJ, doi:10.5281/zenodo.12157 833, 72 Fudamoto, Y., Oesch, P. A., Schinnerer, E., et al. 2017, Casey, C. M., Zavala, J. A., Spilker, J., et al. 2018, ApJ, MNRAS, 472, 483 862, 77 Furlanetto, S. R., & Oh, S. P. 2005, MNRAS, 363, 1031 Castellano, M., Pentericci, L., Vanzella, E., et al. 2018, Geil, P. M., Mutch, S. J., Poole, G. B., et al. 2016, ApJ, 863, L3 MNRAS, 462, 804 Chabrier, G. 2003, Publications of the Astronomical Giallongo, E., Grazian, A., Fiore, F., et al. 2015, A&A, 578, Society of the Pacific, 115, 763 A83 Choi, J., Conroy, C., & Byler, N. 2017, ApJ, 838, 159 Gini, C. 1912, Variabilit`ae mutabilit`a Choi, J., Dotter, A., Conroy, C., et al. 2016, ApJ, 823, 102 Gnedin, N. Y. 2016, ApJ, 825, L17 Reionization by the Oligarchs 21

Gnedin, N. Y., Kravtsov, A. V., & Chen, H. 2008, ApJ, Jones, E., Oliphant, T., Peterson, P., et al. 2001–, SciPy: 672, 765 Open source scientific tools for Python Grazian, A., Castellano, M., Fontana, A., et al. 2012, A&A, Jones, T. A., Ellis, R. S., Schenker, M. A., & Stark, D. P. 547, A51 2013, ApJ, 779, 52 Grazian, A., Giallongo, E., Paris, D., et al. 2017, A&A, 602, Kakiichi, K., & Gronke, M. 2019, arXiv e-prints, A18 arXiv:1905.02480 Greig, B., & Mesinger, A. 2017, MNRAS, 465, 4838 Katz, H., Kimm, T., Haehnelt, M., et al. 2018, MNRAS, Heckman, T. M., Sembach, K. R., Meurer, G. R., et al. 478, 4986 2001, ApJ, 558, 56 Kimm, T., Blaizot, J., Garel, T., et al. 2019, MNRAS, 486, Heckman, T. M., Borthakur, S., Overzier, R., et al. 2011, 2215 ApJ, 730, 5 Kimm, T., Katz, H., Haehnelt, M., et al. 2017, MNRAS, Heiderman, A., Evans, Neal J., I., Allen, L. E., Huard, T., 466, 4826 & Heyer, M. 2010, ApJ, 723, 1019 Kluyver, T., Ragan-Kelley, B., P´erez,F., et al. 2016, in Hellwing, W. A., Frenk, C. S., Cautun, M., et al. 2016, Positioning and Power in Academic Publishing: Players, MNRAS, 457, 3492 Agents and Agendas, ed. F. Loizides & B. Schmidt, IOS Herenz, E. C., Hayes, M., Papaderos, P., et al. 2017, A&A, Press, 87 – 90 606, L11 Kravtsov, A. V. 2013, ApJ, 764, L31 Hinshaw, G., Larson, D., Komatsu, E., et al. 2013, The Kuhlen, M., & Faucher-Gigu`ere,C.-A. 2012, MNRAS, 423, Astrophysical Journal Supplement Series, 208, 19 862 Hoag, A., Bradaˇc,M., Huang, K., et al. 2019, ApJ, 878, 12 Kulkarni, G., Choudhury, T. R., Puchwein, E., & Haehnelt, Holden, B. P., Oesch, P. A., Gonz´alez,V. G., et al. 2016, M. G. 2017, MNRAS, 469, 4283 ApJ, 820, 73 Kulkarni, G., Worseck, G., & Hennawi, J. F. 2019, Hopkins, P. F., Murray, N., Quataert, E., & Thompson, MNRAS, 1429 T. A. 2010, MNRAS, 401, L19 Lam, D., Bouwens, R. J., Labbe, I., et al. 2019, arXiv Hu, E. M., Cowie, L. L., Songaila, A., et al. 2016, ApJ, 825, e-prints, arXiv:1902.02786 L7 Leethochawalit, N., Jones, T. A., Ellis, R. S., Stark, D. P., Hunter, J. D. 2007, Computing In Science & Engineering, & Zitrin, A. 2016, ApJ, 831, 152 9, 90 Livermore, R. C., Finkelstein, S. L., & Lotz, J. M. 2017, Hutter, A., Watkinson, C. A., Seiler, J., et al. 2019a, arXiv ApJ, 835, 113 e-prints, arXiv:1907.04342 Loeb, A., & Barkana, R. 2001, ARA&A, 39, 19 Hutter, A., Dayal, P., Malhotra, S., et al. 2019b, in BAAS, Ma, X., Hopkins, P. F., Kasen, D., et al. 2016, MNRAS, Vol. 51, 57 459, 3614 Inoue, A. K., Shimizu, I., Iwata, I., & Tanaka, M. 2014, Ma, X., Kasen, D., Hopkins, P. F., et al. 2015, MNRAS, MNRAS, 442, 1805 453, 960 Inoue, A. K., Hasegawa, K., Ishiyama, T., et al. 2018, Madau, P., & Dickinson, M. 2014, ARA&A, 52, 415 Publications of the Astronomical Society of Japan, 70, 55 Madau, P., Haardt, F., & Rees, M. J. 1999, ApJ, 514, 648 Ishigaki, M., Kawamata, R., Ouchi, M., et al. 2018, ApJ, Mainali, R., Kollmeier, J. A., Stark, D. P., et al. 2017, ApJ, 854, 73 836, L14 Izotov, Y. I., Orlitov´a,I., Schaerer, D., et al. 2016a, Nature, Marchi, F., Pentericci, L., Guaita, L., et al. 2017, A&A, 529, 178 601, A73 Izotov, Y. I., Schaerer, D., Thuan, T. X., et al. 2016b, —. 2018, A&A, 614, A11 MNRAS, 461, 3683 Mashian, N., Oesch, P. A., & Loeb, A. 2016, MNRAS, 455, Izotov, Y. I., Schaerer, D., Worseck, G., et al. 2018a, 2101 MNRAS, 474, 4514 Mason, C. A., Naidu, R. P., Tacchella, S., & Leja, J. 2019a, Izotov, Y. I., Worseck, G., Schaerer, D., et al. 2018b, arXiv e-prints, arXiv:1907.11332 MNRAS, 478, 4851 Mason, C. A., Trenti, M., & Treu, T. 2015, ApJ, 813, 21 Japelj, J., Vanzella, E., Fontanot, F., et al. 2017, MNRAS, Mason, C. A., Treu, T., Dijkstra, M., et al. 2018, ApJ, 856, 468, 389 2 Jiang, F., Dekel, A., Kneller, O., et al. 2018, ArXiv Mason, C. A., Fontana, A., Treu, T., et al. 2019b, MNRAS, e-prints, arXiv:1804.07306 485, 3947 22 Naidu et al.

Matthee, J., Sobral, D., Best, P., et al. 2017, MNRAS, 465, Planck Collaboration, Aghanim, N., Akrami, Y., et al. 3637 2018, arXiv e-prints, arXiv:1807.06209 Matthee, J., Sobral, D., Gronke, M., et al. 2018, A&A, 619, Reddy, N. A., Steidel, C. C., Pettini, M., & Bogosavljevi´c, A136 M. 2016a, ApJ, 828, 107 McGreer, I. D., Mesinger, A., & D’Odorico, V. 2015, Reddy, N. A., Steidel, C. C., Pettini, M., Bogosavljevic, M., MNRAS, 447, 499 & Shapley, A. 2016b, ArXiv e-prints, arXiv:1606.03452 McQuinn, M., Lidz, A., Zaldarriaga, M., Hernquist, L., & Reddy, N. A., Oesch, P. A., Bouwens, R. J., et al. 2018a, Dutta, S. 2008, MNRAS, 388, 1101 ApJ, 853, 56 Mesinger, A. 2010, MNRAS, 407, 1328 Reddy, N. A., Shapley, A. E., Sanders, R. L., et al. 2018b, Mesinger, A., Aykutalp, A., Vanzella, E., et al. 2015, ArXiv e-prints, arXiv:1811.11767 MNRAS, 446, 566 Rivera-Thorsen, T. E., Dahle, H., Chisholm, J., et al. 2019, Mesinger, A., Furlanetto, S., & Cen, R. 2011, MNRAS, 411, arXiv e-prints, arXiv:1904.08186 955 Roberts-Borsani, G. W., Bouwens, R. J., Oesch, P. A., Mesinger, A., Greig, B., & Sobacchi, E. 2016, MNRAS, 459, et al. 2016, ApJ, 823, 143 2342 Robertson, B. E., Ellis, R. S., Furlanetto, S. R., & Dunlop, Miralda-Escud´e,J. 1998, ApJ, 501, 15 J. S. 2015, ApJL, 802, L19 Mo, H. J., Mao, S., & White, S. D. M. 1998, MNRAS, 295, Robertson, B. E., Furlanetto, S. R., Schneider, E., et al. 2013, ApJ, 768, 71 319 Rosdahl, J., Katz, H., Blaizot, J., et al. 2018, MNRAS, 479, Mortlock, D. J., Warren, S. J., Venemans, B. P., et al. 2011, 994 Nature, 474, 616 Rutkowski, M. J., Scarlata, C., Haardt, F., et al. 2016, Naidu, R. P., Forrest, B., Oesch, P. A., Tran, K.-V. H., & ApJ, 819, 81 Holden, B. P. 2018, MNRAS, 478, 791 Rutkowski, M. J., Scarlata, C., Henry, A., et al. 2017, Naidu, R. P., Oesch, P. A., Reddy, N., et al. 2017, ApJ, ApJL, 841, L27 847, 12 Safarzadeh, M., & Scannapieco, E. 2016, ApJL, 832, L9 Nakajima, K., Ellis, R. S., Iwata, I., et al. 2016, ApJL, 831, Salpeter, E. E. 1955, ApJ, 121, 161 L9 Savaglio, S., Glazebrook, K., & Le Borgne, D. 2009, ApJ, Oesch, P. A., Bouwens, R. J., Illingworth, G. D., Labb´e,I., 691, 182 & Stefanon, M. 2018, ApJ, 855, 105 Sawala, T., Frenk, C. S., Fattahi, A., et al. 2016, MNRAS, Oesch, P. A., Carollo, C. M., Stiavelli, M., et al. 2009, ApJ, 457, 1931 690, 1350 Schaye, J., Crain, R. A., Bower, R. G., et al. 2015, Oesch et al. in prep. 2019 MNRAS, 446, 521 Oke, J. B., & Gunn, J. E. 1983, ApJ, 266, 713 Seiler, J., Hutter, A., Sinha, M., & Croton, D. 2019, Oliphant, T. E. 2015, Guide to NumPy (Continuum Press) MNRAS, 1578 Ouchi, M., Shimasaku, K., Furusawa, H., et al. 2010, ApJ, Shapley, A. E., Steidel, C. C., Strom, A. L., et al. 2016, 723, 869 ApJL, 826, L24 Oyarz´un,G. A., Blanc, G. A., Gonz´alez,V., Mateo, M., & Sharma, M., Theuns, T., Frenk, C., et al. 2016, MNRAS, Bailey, John I., I. 2017, ApJ, 843, 133 458, L94 Paardekooper, J.-P., Khochfar, S., & Dalla Vecchia, C. Sheth, R. K., Mo, H. J., & Tormen, G. 2001, MNRAS, 323, 2015, MNRAS, 451, 2544 1 Pawlik, A. H., Schaye, J., & Dalla Vecchia, C. 2015, Shibuya, T., Ouchi, M., & Harikane, Y. 2015, The MNRAS, 451, 1586 Astrophysical Journal Supplement Series, 219, 15 Pentericci, L., Vanzella, E., Fontana, A., et al. 2014, ApJ, Shibuya, T., Ouchi, M., Harikane, Y., & Nakajima, K. 793, 113 2019, ApJ, 871, 164 P´erez,F., & Granger, B. E. 2007, Computing in Science Shivaei, I., Reddy, N. A., Siana, B., et al. 2018, ApJ, 855, and Engineering, 9, 21 42 Planck Collaboration, Ade, P. A. R., Aghanim, N., et al. Shull, J. M., Harness, A., Trenti, M., & Smith, B. D. 2012, 2015, ArXiv e-prints, arXiv:1502.01589 ApJ, 747, 100 Planck Collaboration, Adam, R., Aghanim, N., et al. 2016, Siana, B., Teplitz, H. I., Ferguson, H. C., et al. 2010, ApJ, A&A, 596, A108 723, 241 Reionization by the Oligarchs 23

Sobacchi, E., & Mesinger, A. 2014, MNRAS, 440, 1662 Totani, T., Kawai, N., Kosugi, G., et al. 2006, Publications Somerville, R. S., Behroozi, P., Pandya, V., et al. 2018, of the Astronomical Society of Japan, 58, 485 MNRAS, 473, 2714 Trac, H. 2018, ApJ, 858, L11 Trainor, R. F., Steidel, C. C., Strom, A. L., & Rudie, G. C. Songaila, A., Hu, E. M., Barger, A. J., et al. 2018, ApJ, 2015, ApJ, 809, 89 859, 91 Trainor, R. F., Strom, A. L., Steidel, C. C., & Rudie, G. C. Speagle, J. S. 2019, arXiv e-prints, arXiv:1904.02180 2016, ApJ, 832, 171 Stanway, E. R., & Eldridge, J. J. 2018, MNRAS, 479, 75 Trebitsch, M., Blaizot, J., Rosdahl, J., Devriendt, J., & Stark, D. P., Richard, J., Charlot, S., et al. 2015, MNRAS, Slyz, A. 2017, MNRAS, 470, 224 450, 1846 Vanzella, E., de Barros, S., Vasei, K., et al. 2016, ApJ, 825, 41 Steidel, C. C., Bogosavljevi´c,M., Shapley, A. E., et al. Vanzella, E., Nonino, M., Cupani, G., et al. 2018, MNRAS, 2018, ApJ, 869, 123 476, L15 Steidel, C. C., Strom, A. L., Pettini, M., et al. 2016, ApJ, Vanzella, E., Caminha, G. B., Calura, F., et al. 2019, arXiv 826, 159 e-prints, arXiv:1904.07941 Sun, G., & Furlanetto, S. R. 2016, MNRAS, 460, 417 Verhamme, A., Orlitov´a,I., Schaerer, D., et al. 2017, A&A, Tacchella, S., Bose, S., Conroy, C., Eisenstein, D. J., & 597, A13 Johnson, B. D. 2018, ApJ, 868, 92 Waskom, M., Botvinnik, O., O’Kane, D., et al. 2018, mwaskom/seaborn: v0.9.0 (July 2018), Tacchella, S., Trenti, M., & Carollo, C. M. 2013, ApJL, doi:10.5281/zenodo.1313201 768, L37 Wise, J. H., & Cen, R. 2009, ApJ, 693, 984 Tang, M., Stark, D. P., Chevallard, J., & Charlot, S. 2019, Wise, J. H., Demchenko, V. G., Halicek, M. T., et al. 2014, MNRAS, 489, 2572 MNRAS, 442, 2560 Tanvir, N. R., Fox, D. B., Levan, A. J., et al. 2009, Nature, Yajima, H., Choi, J.-H., & Nagamine, K. 2011, MNRAS, 461, 1254 412, 411 Thompson, T. A., Quataert, E., & Murray, N. 2005, ApJ, Yung, L. Y. A., Somerville, R. S., Finkelstein, S. L., Popping, G., & Dav´e,R. 2019, MNRAS, 483, 2983 630, 167 24 Naidu et al.

APPENDIX

A. UV LUMINOSITY FUNCTIONS In addition to Figure1, where we compare against literature data at z = 6 (the redshift at which the faint-end of the UVLF has been measured), we provide UVLFs predicted by our empirical model for z = 7 − 12. We also quantify agreement between our model and observational data in Table3. In general, our predicted UVLFs show good agreement with the data, both at the bright and faint ends. The only tension is at z = 6 with Atek et al.(2018) at MUV < −16 where they estimate several more galaxies than predicted by our model – we find better agreement with Bouwens et al.(2015b) and Livermore et al.(2017) in the same luminosity regime, while also noting that the excess of brighter galaxies would further support this paper’s point of view about the reionization budget.

Figure 12. UVLFs from the Tacchella et al.(2018) model compared with data from Bouwens et al.(2015b)( z = 7 − 9) and Oesch et al.(2018)( z = 10).

Table 3. Comparison of Model UVLFs against literature data.

2 2 Redshift χ (MUV < −16) χ (MUV > −16) Source

6 0.7 0.5 Bouwens et al.(2017) 9.9 0.1 Atek et al.(2018) 1.5 1.7 Livermore et al.(2017) 7 1.4 − Bouwens et al.(2015a) 8 1.3 − Bouwens et al.(2015a) 10 0.9 − Oesch et al.(2018)

2 2 1 n (xi−µi) (a) χ is reported as n−1 Σi=1 2 , where x and µ represent the data and model values respectively, and σ is the error σi on x. When reported errors are non-Gaussian, we set σ to half the difference between the 84th and 16th percentile.