This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\volnopage

20XX Vol. X No. XX, 000–000

11institutetext: School of Physics and Astronomy, Sun Yat-sen University, 2 Daxue Road, Tangjia, Zhuhai, 519082, China
22institutetext: CSST Science Center for the Guangdong-Hong Kong-Macau Greater Bay Area, Zhuhai, 519082, China; [email protected]
33institutetext: CAS Key Laboratory of Theoretical Physics, Institute of Theoretical Physics, Chinese Academy of Sciences, Beijing, 100190, China
\vs\noReceived 20XX Month Day; accepted 20XX Month Day

Quantifying the tension between cosmological models and JWST red candidate massive galaxies

Jun-Chao Wang 11    Zhi-Qi Huang *1*122    Lu Huang 33    Jianqi Liu 11
Abstract

We develop a Python tool to estimate the tail distribution of the number of dark matter halos beyond a mass threshold and in a given volume in a light-cone. The code is based on the extended Press-Schechter model and is computationally efficient, typically taking a few seconds on a personal laptop for a given set of cosmological parameters. The high efficiency of the code allows a quick estimation of the tension between cosmological models and the red candidate massive galaxies released by the James Webb Space Telescope, as well as scanning the theory space with the Markov Chain Monte Carlo method. As an example application, we use the tool to study the cosmological implication of the candidate galaxies presented in Labbé et al. (2023). The standard Λ\Lambda cold dark matter (Λ\LambdaCDM) model is well consistent with the data if the star formation efficiency can reach 0.3\sim 0.3 at high redshift. For a low star formation efficiency ϵ0.1\epsilon\sim 0.1, Λ\LambdaCDM model is disfavored at 2σ\sim 2\sigma-3σ3\sigma confidence level.

keywords:
galaxies: abundances – large-scale structure of Universe – cosmological parameters

1 Introduction

In the standard Λ\Lambda cold dark matter (Λ\LambdaCDM) picture, baryon decouples from radiation after recombination and tracks the gravitational potential of dark matter. Before the formation of early galaxies, baryon and dark matter are well mixed. The fraction of baryonic mass in a massive dark matter halo therefore approximately equals to the cosmological mean fbΩb/Ωmf_{b}\equiv\Omega_{b}/\Omega_{m}, where Ωb\Omega_{b} and Ωm\Omega_{m} are the baryon and matter abundance parameters, respectively. If the fraction of stellar mass in baryonic matter, namely the star formation efficiency (SFE) ϵ\epsilon is known, we can connect the stellar mass MM_{\star} to the halo mass MhaloM_{\rm halo} via M=ϵfbMhaloM_{\star}=\epsilon f_{b}M_{\rm halo}. This provides a way to test cosmology by estimating stellar masses of massive galaxies in a given volume. In particular, at high redshift where massive halos are predicted to be very rare, detection of very massive galaxies may pose a stringent constraint on cosmology. The red candidate massive galaxies released by the James Webb Space Telescope (JWST), for instance, have inspired a debate whether a beyond-Λ\LambdaCDM cosmology is on demand (Labbé et al. 2023; Boylan-Kolchin 2023; Lovell et al. 2023; Haslbauer et al. 2022; Chen et al. 2023; Ferrara et al. 2023; Qin et al. 2023; Wang et al. 2023c). A plenty of beyond-Λ\LambdaCDM models including alternative dark energy (Wang et al. 2023b; Adil et al. 2023; Menci et al. 2022), non-standard dark matter (Bird et al. 2023; Lin et al. 2023; Maio & Viel 2023; Dayal & Giri 2023), cosmic string (Jiao et al. 2023), primordial black holes (Yuan et al. 2023; Hütsi et al. 2023; Gouttenoire et al. 2023; Guo et al. 2023; Huang et al. 2023), and many others (Melia 2023; John & Babu Joseph 2023; Lovyagin et al. 2022; Lei et al. 2023) have been confronted with the JWST high-redshift candidate galaxies. Modification of primordial conditions may also lead to an overabundance of dark matter halos and hence more massive galaxies at high redshift (Huang 2019; Parashari & Laha 2023; Padmanabhan & Loeb 2023; Keller et al. 2023; Forconi et al. 2023; Su et al. 2023; Wang & Liu 2022; Tkachev et al. 2023).

The major obstacle to an accurate comparison between a theory and observed galaxies is the uncertainty in the star formation efficiency ϵ\epsilon at high redshift. Galaxy formation models and low-redshift observation typically suggest ϵ0.2\epsilon\lesssim 0.2 (Guo et al. 2020; Boylan-Kolchin 2023), though ϵ\epsilon at high redshift may be significant different (Zhang et al. 2022; Qin et al. 2023). Because the UV radiation from star formation can ionize the neutral hydrogen at the reionization epoch, a larger star formation efficiency tends to accelerate the reionization process and finish it at an earlier time. Observations of the Lyman-α\alpha emitter luminosity function suggest that reionization is still on-going at z7z\gtrsim 7 (Wold et al. 2022; Goto et al. 2021), consistent with the ϵ0.2\epsilon\lesssim 0.2 picture (Minoda et al. 2023; Mobina Hosseini et al. 2023). It is however difficult to derive an upper-bound of ϵ\epsilon solely from the reionization history, due to the degeneracy between ϵ\epsilon and other astrophysical parameters.

The common practice is then to study the cosmological implication for an assumed star formation efficiency. Although most of the studies qualitatively agree, quantitative discrepancies still exist. For example, while Boylan-Kolchin (2023) and Chen et al. (2023) find that ϵ0.5\epsilon\sim 0.5 suffices to explain the Labbé et al. (2023) data in the standard Λ\LambdaCDM cosmology, Menci et al. (2022) that uses the same data claims that Λ\LambdaCDM is ruled out at 2σ2\sigma level even with ϵ=1\epsilon=1. This is not only due to different choices of mass cut and volume cut, but also because it is hard to give an accurate description of the systematic errors of redshift and stellar mass obtained with spectral energy distribution (SED) fitting. By using seven individual redshift and stellar mass measurements with different SED-fitting codes/templates, Labbé et al. (2023) estimated extremes of stellar mass and redshift of each candidate galaxy. These extremes however are insufficient for quantitative analysis, which requires precise knowledge of the probability density functions (PDFs) of the systematic errors.

The summary statistics of the statistical errors of stellar mass are often represented as MΔMinf+ΔMsupM^{+\Delta M_{\rm sup}}_{-\Delta M_{\rm inf}}, where MM is the median mass. The upper and lower errors (ΔMsup\Delta M_{\rm sup} and ΔMinf\Delta M_{\rm inf}) are defined with 1616th and 8484th percentiles. The errors in stellar mass, and similarly in redshift if only photometry information is available, are typically asymmetric and difficult to model analytically. Accurate observation-theory comparison usually requires an end-to-end comparison between high-resolution simulations and observations. On the other hand, while the summary statistics (and the extremes from different SED-fitting methods, if available) contain limited information, they are important science products of many galaxy surveys. The aim of the present work is to develop a tool that performs a preliminary scan of the theory space based on these simple products, before planning subsequent spectroscopic observation and running costly simulations.

This paper is organized as follows. Section 2 describes the physical model of the abundance of high-redshift massive galaxies and the workflow of observation-theory comparison. Section 3 uses the tool to analyze the cosmological implication of the candidate galaxies in Labbé et al. (2023). Section 4 concludes.

2 Algorithm of the Tool

Figure 1 summarizes our algorithm to do an observation-theory comparison. We consider galaxies beyond a stellar-mass threshold (M>M,cutM_{\star}>M_{\rm\star,cut}) and in a fixed comoving volume defined by the sky coverage fskyf_{\rm sky} and redshift range (zmin<z<zmaxz_{\min}<z<z_{\max}). Because both the stellar mass and redshift from SED-fitting have significant uncertainties, whether an observed galaxy is above the stellar-mass threshold and in the volume is probabilistic. The total observed galaxies above the stellar-mass threshold and in the volume, denoted as NobsN_{\rm obs}, is therefore a random variable. Due to the cosmic variance and the Poisson shot noise, the theoretical prediction of the total number of galaxies above the stellar-mass threshold and in the volume, denoted as NthN_{\rm th}, is also probabilistic.

A model is ruled out when and only when Nth<NobsN_{\rm th}<N_{\rm obs}. The tension between a cosmological model and the observed galaxies is therefore quantified by the probability P(Nth<Nobs)\mathrm{P}\left(N_{\rm th}<N_{\rm obs}\right), or more conveniently, the smallness of P(NthNobs)=1P(Nth<Nobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right)=1-\mathrm{P}\left(N_{\rm th}<N_{\rm obs}\right). For instance, P(NthNobs)=0.01\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right)=0.01 indicates that the model under consideration is ruled at 99%99\% confidence level. Here and in what follows, P()\mathrm{P}\left(\cdot\right) stands for probability (for discrete variable or an event) or probability density function (for continuous variable).

For a given cosmology, P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right) is evaluated in three steps.

  • 1.

    For Nobs=0,1,2,N_{\rm obs}=0,1,2,\ldots compute the distribution function P(Nobs)\mathrm{P}\left(N_{\rm obs}\right) from summary statistics of observed galaxies and user-specified assumptions about the functional form of probability density functions of stellar mass and redshift.

  • 2.

    For Nth=0,1,2,N_{\rm th}=0,1,2,\ldots compute the distribution function P(Nth)\mathrm{P}\left(N_{\rm th}\right) from the cosmological model and a user-specified star formation efficiency.

  • 3.

    Compute

    P(NthNobs)=Nobs=0P(Nobs)Nth=NobsP(Nth).\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right)=\sum_{N_{\rm obs}=0}^{\infty}\mathrm{P}\left(N_{\rm obs}\right)\sum_{N_{\rm th}=N_{\rm obs}}^{\infty}\mathrm{P}\left(N_{\rm th}\right).
SFEgalaxy M,cutM_{\rm\star,cut}cosmologyfb=Ωb/Ωmf_{b}=\Omega_{b}/\Omega_{m}volume cutobserved galaxiesmodel of error PDFshalo McutM_{\rm cut}P(Nth)\mathrm{P}\left(N_{\rm th}\right)P(Nobs)\mathrm{P}\left(N_{\rm obs}\right)P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right)
Figure 1: Workflow of observation-theory comparison

Explicit evaluation of P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right) requires perfect knowledge about the PDFs of statistical/systematic errors of the stellar masses and redshifts of the candidate galaxies. While the PDFs of statistical errors can be recovered by going back to the SED-fitting process, it is not easy to give an accurate description of the PDFs of systematic errors, which often dominate the error budget. Without further knowledge about the systematic errors, we simply model the PDFs with a few generic distribution functions: skew normal, skew triangular, and skew rectangular, whose definitions are given in the Appendix. The code is organized in such a way that the users can easily add their own distribution functions. We also allow the users to turn off systematic errors by using Dirac delta distribution. In the case when systematic errors are given in the form of extremes from different SED-fitting methods, we match the extremes with exact upper/lower bounds of skew triangular/rectangular distributions or 2σ2\sigma bounds of skew normal distribution. Finally, the stellar mass MM_{\star} and redshift zz from different SED-fitting methods are typically positively correlated. We test the impact of MM_{\star}-zz correlation by forcing (MM,median)(zzmedian)0(M_{\star}-M_{\rm\star,median})(z-z_{\rm median})\geq 0 in the joint distribution of MM_{\star} and zz, and find that the correlation does not make a qualitative difference.

The evaluation of P(Nth)\mathrm{P}\left(N_{\rm th}\right) involves the extended Press-Schechter ellipsoidal collapse model (Press & Schechter 1974; Sheth et al. 2001; Sheth & Tormen 2002), where the comoving halo number density per mass interval, namely the halo mass function (at a fixed redshift zz) is given by

dndM=f(ν)ρmMdlnσdM.\frac{\mathrm{d}n}{\mathrm{d}M}=-f(\nu)\frac{\rho_{m}}{M}\frac{\mathrm{d}\ln\sigma}{\mathrm{d}M}. (1)

The ρm\rho_{m} on the right-hand side is the comoving average background matter density and σ\sigma is the mass fluctuation at scale R=(3M4πρm)1/3R=\left(\frac{3M}{4\pi\rho_{m}}\right)^{1/3}, given by

σ2=12π20k2Pm(k,z)W2(kR)dk,\sigma^{2}=\frac{1}{2\pi^{2}}\int_{0}^{\infty}k^{2}\mathrm{P}_{m}(k,z)\mathrm{W}^{2}(kR)\,\mathrm{d}k, (2)

where kk is the comoving wavenumber, and Pm(k,z)\mathrm{P}_{m}(k,z) is the linear matter power spectrum at redshift zz. For the filter function W(kR)\mathrm{W}(kR), we take the form of top-hat

W(x)=3(sinxxcosx)x3.\mathrm{W}(x)=\frac{3(\sin x-x\cos x)}{x^{3}}. (3)

The simulation calibrated factor f(ν)f(\nu) is given by

f(ν)=2A[1+1(aν2)q](aν22π)1/2exp(aν22),f(\nu)=2A\left[1+\frac{1}{(a\nu^{2})^{q}}\right]\left(\frac{a\nu^{2}}{2\pi}\right)^{1/2}\exp(-\frac{a\nu^{2}}{2}), (4)

with A=0.322A=0.322, a=0.707a=0.707, q=0.3q=0.3 and ν=δcσ(R,z)\nu=\frac{\delta_{c}}{\sigma(R,z)}, where parameter δc=1.686\delta_{c}=1.686 corresponds to the critical linear overdensity. We ignore the tiny difference in δc\delta_{c} for different cosmological models, as we are only interested in the models that are close to the concordance Λ\LambdaCDM model.

For a massive galaxy with stellar mass M>M,cutM_{\star}>M_{\rm\star,cut}, the mass of its host halo must exceed Mcut=M,cut/(ϵfb)M_{\rm cut}=M_{\rm\star,cut}/(\epsilon f_{b}). The halo mass function Eq. (1) allows to compute the ensemble average of the mean number of such halos

Nth=4πfskyMcutdMzminzmaxdndMdVdzdΩdz,\langle N_{\rm th}\rangle=4\pi f_{\rm sky}\int_{M_{\rm cut}}^{\infty}\mathrm{d}M\int_{z_{\min}}^{z_{\max}}\frac{\mathrm{d}n}{\mathrm{d}M}\frac{\mathrm{d}V}{\mathrm{d}z\mathrm{d}\Omega}\,\mathrm{d}z, (5)

where dVdzdΩ\frac{\mathrm{d}V}{\mathrm{d}z\mathrm{d}\Omega} is comoving volume per redshift interval per solid angle. Here we have equalized the number of massive galaxies and the number of their host halos, assuming each massive halo has a central galaxy.

For the purpose of using rare objects to study cosmological tensions, we are only interested in the NthO(1)\langle N_{\rm th}\rangle\lesssim O(1) case, where NthN_{\rm th} approximately follows a Poisson distribution

P(Nth)eλλNthNth!,\mathrm{P}\left(N_{\rm th}\right)\sim e^{-\lambda}\frac{\lambda^{N_{\rm th}}}{N_{\rm th}!}, (6)

where λNth\lambda\approx\langle N_{\rm th}\rangle is given in Eq. (5).

To obtain a more accurate result, we integrate λ\lambda on the right-hand side of Eq. (6) over a Gaussian distribution, giving

P(Nth)=12πσλe(λNth)22σλ2eλλNthNth!dλ.\mathrm{P}\left(N_{\rm th}\right)=\int_{-\infty}^{\infty}\frac{1}{\sqrt{2\pi}\sigma_{\lambda}}e^{-\frac{(\lambda-\langle N_{\rm th}\rangle)^{2}}{2\sigma_{\lambda}^{2}}}e^{-\lambda}\frac{\lambda^{N_{\rm th}}}{N_{\rm th}!}\,\mathrm{d}\lambda. (7)

The beyond-Poisson cosmic variance from the large scale structure, σλ\sigma_{\lambda}, can be read out from the online emulator at https://www.ph.unimelb.edu.au/~mtrenti/cvc/CosmicVariance.html (Trenti & Stiavelli 2008). For a pencil-like survey volume, we find the cosmic variance approximately equal to the product of the linear result, which equals to the product of linear halo bias and matter cosmic variance, and a fudge factor 1.4\approx 1.4 that accounts for the nonlinear correction. Since the cosmic-variance correction is a subdominant effect111Our result does not contradict that of Chen et al. (2023), which claims that cosmic variance taken from their simulations is a dominant factor. Our “cosmic variance” here, as well as that defined in Trenti & Stiavelli (2008), refers to the variance due to power spectrum uncertainties caused by finite sampling and does not include Poisson shot noise, while “cosmic variance” taken from simulations include both effects. , this approximation suffices and allows us to make our tool self-contained.

The algorithm is realized with a Python tool publicly available at http://zhiqihuang.top/codes/massivehalo.htm.

3 Applying to JWST data

We now apply the tool to the 13 candidate galaxies in Labbé et al. (2023). The summary statistics of stellar mass and redshift are given in the Extended Data Table 2 of Labbé et al. (2023), as well as in the online repository given in Sec. LABEL:sec:data. Stellar mass and redshift values include two uncertainties: ±(ran)±(sys)\pm(\text{ran})\pm(\text{sys}), where the random uncertainty (ran) corresponds to the 1616th and 8484th percentiles of the combined posterior distribution, and the systematic uncertainty (sys) corresponds to the extreme values from seven different methods (EAZY, Prospector, and Bagpipes with 5 variations, including dust, SFH, age prior, and SNR limit). In this study, we update the data of three candidate galaxies (L23-13050, L23-35300 and L23-39575) with the information from spectroscopic follow-up observations (Kocevski et al. 2023; Pérez-González et al. 2023; Fujimoto et al. 2023). The updated data is also available in the online repositary given in Sec. LABEL:sec:data.

Unless otherwise specified, we choose the stellar-mass threshold to be M,cut=1010MM_{\rm\star,cut}=10^{10}M_{\odot} and the redshift range to be zmin=7z_{\min}=7 and zmax=10z_{\rm max}=10, and we assume the statistical errors follow a skew normal distribution. The data covers a sky area 38arcmin238\,\mathrm{arcmin}^{2}, which translates to fsky=2.56×107f_{\rm sky}=2.56\times 10^{-7}. We take the cosmology to be the Planck best-fit Λ\LambdaCDM model (Aghanim et al. 2020) with Ωm=0.3158\Omega_{m}=0.3158, Ωb=0.049389\Omega_{b}=0.049389, primordial spectrum index ns=0.96605n_{s}=0.96605, matter fluctuation parameter σ8=0.812\sigma_{8}=0.812, and Hubble constant H0=67.32km/s/MpcH_{0}=67.32\,\mathrm{km/s/Mpc}. Figure 2 shows the dependence of P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right) on the star formation efficiency and the model of the PDFs of systematic errors. For ϵ0.3\epsilon\gtrsim 0.3 that is fairly possible at the early epoch of galaxy formation (Zhang et al. 2022; Qin et al. 2023), the data is well consistent with Λ\LambdaCDM model even when the systematic errors are ignored (dotted blue line in Figure 2). For the most conservative ϵ0.1\epsilon\sim 0.1 case, Λ\LambdaCDM model is slightly disfavored by the data. For different forms of the PDFs of the systematic errors, the statistical significance of the tension varies between 2σ\sim 2\sigma and 3σ\sim 3\sigma.

Refer to caption
Figure 2: Dependence of P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right) on the star formation efficiency and the model of the PDF of systematic errors.

It has been shown that dynamic dark energy model with equation of state w(a)=w0+wa(1a)w(a)=w_{0}+w_{a}(1-a) (Chevallier & Polarski 2001; Linder 2003), where aa is the scale factor, may alleviate the tension between cosmology and the Labbé et al. (2023) data. Since the tension is only significant when the star formation efficiency ϵ\epsilon is 0.1\lesssim 0.1, we hereafter fix ϵ=0.1\epsilon=0.1 and investigate whether the dynamic dark energy model is well consistent with the data. For simplicity, we also assume that the systematic errors follow the skew triangular distribution. We take cosmological parameters from the Markov chains that are sampled with Planck+BAO+SNe likelihood. Here Planck refers to the TTTEEE+lowl+LowE likelihood of the cosmic microwave background maps measured by the Planck satellite (Aghanim et al. 2020); BAO refers to the baryon acoustic oscillations data that are summarized in  Alam et al. (2017); SNe stands for the Pantheon supernova catalog (Scolnic et al. 2018). We download the chains from Planck Legacy Archive (https://pla.esac.esa.int/pla/) and, for better visual effect, thin the chains by a factor of 5. For each set of cosmological parameters of the thinned chains, we compute P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right) value and show it in Figure 3 as a dot, whose color represents the value of P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right) and coordinates indicate the dark energy parameters (w0,wa)(w_{0},w_{a}). The variation of the dynamic dark energy equation of state in the range that is allowed by Planck+BAO+SNe does lead to a variation of P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right), however, not at a very significant level that would make the result qualitatively different.

Refer to caption
Figure 3: P(NthNobs)\mathrm{P}\left(N_{\rm th}\geq N_{\rm obs}\right) for w0w_{0}-waw_{a} cosmology. The star formation efficiency is fixed to be 0.10.1. Systematic errors are assumed to follow skew triangular distribution. For each point, the cosmological parameters are drawn from thinned Planck+BAO+SNe Markov chains. The solid black lines are the marginalized 68.3%68.3\% and 95.4%95.4\% confidence contours for the Planck+BAO+SNe likelihood.

4 Discussion and Conclusions

In this work, we present and make publicly available an efficient numeric tool that can quickly estimate the tension between cosmological models and observed high-redshift massive galaxies. Our method is based on the statistics of the total numbers of massive halos, whereas the previous works are all based on the statistics of the cumulative stellar mass ρ(>M)\rho_{\star}(>M_{\star}) (Labbé et al. 2023; Menci et al. 2022; Boylan-Kolchin 2023; Wang et al. 2023b; Chen et al. 2023). The advantage of using the number of halos as a fundamental variable is that the dominating Poisson shot noise contribution can be explicitly formulated.

We update the parameters of some of the candidate galaxies in Labbé et al. (2023) with recent spectroscopic follow-up observations. For a low star formation efficiency ϵ0.1\epsilon\sim 0.1, the data remains in 2σ\gtrsim 2\sigma tension with the concordance Λ\LambdaCDM cosmology, and the tension is insensitive to the dark energy equation of state and other cosmological parameters, provided that the cosmological parameters are confined by the Planck+BAO+SNe likelihood. We also verify that the updated spectroscopic redshift information slightly reduces the tension between Λ\LambdaCDM model and the JWST candidate galaxies. For instance, assuming a skew normal distribution of both systematic and statistical errors and star formation efficiency ϵ=0.1\epsilon=0.1, we find Planck best-fit Λ\LambdaCDM model is ruled out at 2.59σ2.59\sigma level with the original data from Labbé et al. (2023) and 2.49σ2.49\sigma level with the updated data.

Although the algorithm in Figure 1 captures the main physics, there are subtleties beyond the scope of the present work. First, it is known that the extended Press-Schechter halo mass functions at high redshift are 20%-50% higher than those obtained from N-body simulations (Reed et al. 2003; Despali et al. 2016; Shirasaki et al. 2021; Wang et al. 2022). Secondly, backsplash halos, whose dark matter has left the host halo while its gas may remain in the host halo (Balogh et al. 2000; Wang et al. 2009; Wetzel et al. 2014; Diemer 2021; Wang et al. 2023a), may make the interpretation of star formation efficiency ambiguous in the real world and may alter the statistical significance of the cosmological tension (Chen et al. 2023). Finally, we have ignored catastrophic redshift errors which, despite being rare, might have a significant impact on the counting of rare objects. We leave further studies on these nontrivial subjects as our future work.

Acknowledgements.
This work is supported by National key R&D Program of China (Grant No. 2020YFC2201600), the National Natural Science Foundation of China (NSFC) under Grant No. 12073088, and National SKA Program of China No. 2020SKA0110402.

Appendix A Modelling the Distribution Functions of Stellar Mass and Redshift

Here we introduce the distribution functions that are built in the tool.

A.1 skew normal distribution

The standardized skew normal distribution with parameter aa and of a random variable xx is given by the PDF

𝒩(x;a)=12πex2/2[1+erf(ax/2)].\mathcal{N}(x;a)=\frac{1}{\sqrt{2\pi}}e^{-x^{2}/2}\left[1+\mathrm{erf}\,\left(ax/\sqrt{2}\right)\right]. (8)

In the general case, the skew normal distribution is obtained by rescaling and translating xx,

N(x;a,ζ,ω)=𝒩(xζω;a).\mathrm{N}(x;a,\zeta,\omega)=\mathcal{N}\left(\frac{x-\zeta}{\omega};a\right). (9)

For a given summary statistics, the parameters a,ζ,ωa,\zeta,\omega can be fixed by matching the 1616th and 8484th percentiles and the median. The skew normal distribution (9) cannot describe a very skewed distribution with the ratio between two errors exceeding 1.551.55. In this extreme case, we replace the linear transformation on xx in (8) with a nonlinear one. Since this procedure is quite complicated and does not have much to do with the scientific result in the present work, we refer interested readers to the publicly available source code http://zhiqihuang.top/codes/fitskew.py.

A.2 skew triangular and skew rectangular

For parameters a>0a>0, b>0b>0 and μ\mu, the skew triangular distribution of a random variable xx is defined as

T(x;μ,a,b)={xμ+aa2, if μa<xμ;μ+bxb2, if μ<x<μ+b;0, else.\mathrm{T}(x;\mu,a,b)=\left\{\begin{array}[]{ll}\frac{x-\mu+a}{a^{2}},&\text{ if }\mu-a<x\leq\mu;\\ \frac{\mu+b-x}{b^{2}},&\text{ if }\mu<x<\mu+b;\\ 0,&\text{ else.}\end{array}\right. (10)

The skew rectangular distribution is defined as

R(x;μ,a,b)={12a, if μa<xμ;12b, if μ<x<μ+b;0, else.\mathrm{R}(x;\mu,a,b)=\left\{\begin{array}[]{ll}\frac{1}{2a},&\text{ if }\mu-a<x\leq\mu;\\ \frac{1}{2b},&\text{ if }\mu<x<\mu+b;\\ 0,&\text{ else.}\end{array}\right. (11)

References

  • Adil et al. (2023) Adil, S. A., Mukhopadhyay, U., Sen, A. A., & Vagnozzi, S. 2023, arXiv e-prints, arXiv:2307.12763
  • Aghanim et al. (2020) Aghanim, N., Akrami, Y., Ashdown, M., et al. 2020, Astronomy & Astrophysics, 641, A6
  • Alam et al. (2017) Alam, S., Ata, M., Bailey, S., et al. 2017, MNRAS, 470, 2617
  • Balogh et al. (2000) Balogh, M. L., Navarro, J. F., & Morris, S. L. 2000, ApJ, 540, 113
  • Bird et al. (2023) Bird, S., Chang, C.-F., Cui, Y., & Yang, D. 2023, arXiv e-prints, arXiv:2307.10302
  • Boylan-Kolchin (2023) Boylan-Kolchin, M. 2023, Nature Astronomy, 7, 731
  • Chen et al. (2023) Chen, Y., Mo, H. J., & Wang, K. 2023, arXiv e-prints, arXiv:2304.13890
  • Chevallier & Polarski (2001) Chevallier, M., & Polarski, D. 2001, International Journal of Modern Physics D, 10, 213
  • Dayal & Giri (2023) Dayal, P., & Giri, S. K. 2023, arXiv e-prints, arXiv:2303.14239
  • Despali et al. (2016) Despali, G., Giocoli, C., Angulo, R. E., et al. 2016, MNRAS, 456, 2486
  • Diemer (2021) Diemer, B. 2021, ApJ, 909, 112
  • Ferrara et al. (2023) Ferrara, A., Pallottini, A., & Dayal, P. 2023, MNRAS, 522, 3986
  • Forconi et al. (2023) Forconi, M., Ruchika, Melchiorri, A., Mena, O., & Menci, N. 2023, arXiv e-prints, arXiv:2306.07781
  • Fujimoto et al. (2023) Fujimoto, S., Haro, P. A., Dickinson, M., et al. 2023, The Astrophysical Journal Letters, 949, L25
  • Goto et al. (2021) Goto, H., Shimasaku, K., Yamanaka, S., et al. 2021, ApJ, 923, 229
  • Gouttenoire et al. (2023) Gouttenoire, Y., Trifinopoulos, S., Valogiannis, G., & Vanvlasselaer, M. 2023, arXiv e-prints, arXiv:2307.01457
  • Guo et al. (2020) Guo, Q., Hu, H., Zheng, Z., et al. 2020, Nature Astronomy, 4, 246
  • Guo et al. (2023) Guo, S.-Y., Khlopov, M., Liu, X., et al. 2023, arXiv e-prints, arXiv:2306.17022
  • Haslbauer et al. (2022) Haslbauer, M., Kroupa, P., Zonoozi, A. H., & Haghi, H. 2022, ApJ, 939, L31
  • Huang et al. (2023) Huang, H.-L., Cai, Y., Jiang, J.-Q., Zhang, J., & Piao, Y.-S. 2023, arXiv e-prints, arXiv:2306.17577
  • Huang (2019) Huang, Z. 2019, Phys. Rev. D, 99, 103537
  • Hütsi et al. (2023) Hütsi, G., Raidal, M., Urrutia, J., Vaskonen, V., & Veermäe, H. 2023, Phys. Rev. D, 107, 043502
  • Jiao et al. (2023) Jiao, H., Brandenberger, R., & Refregier, A. 2023, arXiv e-prints, arXiv:2304.06429
  • John & Babu Joseph (2023) John, M. V., & Babu Joseph, K. 2023, arXiv e-prints, arXiv:2306.04577
  • Keller et al. (2023) Keller, B. W., Munshi, F., Trebitsch, M., & Tremmel, M. 2023, ApJ, 943, L28
  • Kocevski et al. (2023) Kocevski, D. D., Onoue, M., Inayoshi, K., et al. 2023, arXiv preprint arXiv:2302.00012
  • Labbé et al. (2023) Labbé, I., van Dokkum, P., Nelson, E., et al. 2023, Nature, 616, 266
  • Lei et al. (2023) Lei, L., Zu, L., Yuan, G.-W., et al. 2023, arXiv e-prints, arXiv:2305.03408
  • Lin et al. (2023) Lin, H., Gong, Y., Yue, B., & Chen, X. 2023, arXiv e-prints, arXiv:2306.05648
  • Linder (2003) Linder, E. V. 2003, Physical Review Letters, 90, 091301
  • Lovell et al. (2023) Lovell, C. C., Harrison, I., Harikane, Y., Tacchella, S., & Wilkins, S. M. 2023, MNRAS, 518, 2511
  • Lovyagin et al. (2022) Lovyagin, N., Raikov, A., Yershov, V., & Lovyagin, Y. 2022, Galaxies, 10, 108
  • Maio & Viel (2023) Maio, U., & Viel, M. 2023, A&A, 672, A71
  • Melia (2023) Melia, F. 2023, MNRAS, 521, L85
  • Menci et al. (2022) Menci, N., Castellano, M., Santini, P., et al. 2022, ApJ, 938, L5
  • Minoda et al. (2023) Minoda, T., Yoshiura, S., & Takahashi, T. 2023, arXiv e-prints, arXiv:2304.09474
  • Mobina Hosseini et al. (2023) Mobina Hosseini, S., Soleimanpour Salmasi, B., Sajad Tabasi, S., & Firouzjaee, J. T. 2023, arXiv e-prints, arXiv:2306.12954
  • Padmanabhan & Loeb (2023) Padmanabhan, H., & Loeb, A. 2023, arXiv e-prints, arXiv:2306.04684
  • Parashari & Laha (2023) Parashari, P., & Laha, R. 2023, arXiv e-prints, arXiv:2305.00999
  • Pérez-González et al. (2023) Pérez-González, P. G., Barro, G., Annunziatella, M., et al. 2023, The Astrophysical Journal Letters, 946, L16
  • Press & Schechter (1974) Press, W. H., & Schechter, P. 1974, The Astrophysical Journal, 187, 425
  • Qin et al. (2023) Qin, Y., Balu, S., & Wyithe, J. S. B. 2023, arXiv e-prints, arXiv:2305.17959
  • Reed et al. (2003) Reed, D., Gardner, J., Quinn, T., et al. 2003, MNRAS, 346, 565
  • Scolnic et al. (2018) Scolnic, D. M., Jones, D. O., Rest, A., et al. 2018, ApJ, 859, 101
  • Sheth et al. (2001) Sheth, R. K., Mo, H., & Tormen, G. 2001, Monthly Notices of the Royal Astronomical Society, 323, 1
  • Sheth & Tormen (2002) Sheth, R. K., & Tormen, G. 2002, Monthly Notices of the Royal Astronomical Society, 329, 61
  • Shirasaki et al. (2021) Shirasaki, M., Ishiyama, T., & Ando, S. 2021, ApJ, 922, 89
  • Su et al. (2023) Su, B.-Y., Li, N., & Feng, L. 2023, arXiv e-prints, arXiv:2306.05364
  • Tkachev et al. (2023) Tkachev, M. V., Pilipenko, S. V., Mikheeva, E. V., & Lukash, V. N. 2023, arXiv:2307.13774
  • Trenti & Stiavelli (2008) Trenti, M., & Stiavelli, M. 2008, ApJ, 676, 767
  • Wang & Liu (2022) Wang, D., & Liu, Y. 2022, arXiv e-prints, arXiv:2301.00347
  • Wang et al. (2009) Wang, H., Mo, H. J., & Jing, Y. P. 2009, MNRAS, 396, 2249
  • Wang et al. (2023a) Wang, K., Peng, Y., & Chen, Y. 2023a, MNRAS, 523, 1268
  • Wang et al. (2023b) Wang, P., Su, B.-Y., Zu, L., Yang, Y., & Feng, L. 2023b, arXiv e-prints, arXiv:2307.11374
  • Wang et al. (2022) Wang, Q., Gao, L., & Meng, C. 2022, MNRAS, 517, 6004
  • Wang et al. (2023c) Wang, Y.-Y., Lei, L., Yuan, G.-W., & Fan, Y.-Z. 2023c, arXiv e-prints, arXiv:2307.12487
  • Wetzel et al. (2014) Wetzel, A. R., Tinker, J. L., Conroy, C., & van den Bosch, F. C. 2014, MNRAS, 439, 2687
  • Wold et al. (2022) Wold, I. G. B., Malhotra, S., Rhoads, J., et al. 2022, ApJ, 927, 36
  • Yuan et al. (2023) Yuan, G.-W., Lei, L., Wang, Y.-Z., et al. 2023, arXiv e-prints, arXiv:2303.09391
  • Zhang et al. (2022) Zhang, Z., Wang, H., Luo, W., et al. 2022, A&A, 663, A85