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

Dependence of Galaxy Stellar Properties on the Primordial Spin Factor

Jun-Sung Moon    , Jounghun Lee 11footnotetext: Corresponding author.
Abstract

We present a numerical discovery that the observable stellar properties of present galaxies retain significant dependences on the primordial density and tidal fields. Analyzing the galaxy catalogs from the IllustrisTNG 300-1 simulations, we first compute the primordial spin factor, τ\tau, defined as the mean degree of misalignments between the principal axes of the initial density and potential hessian tensors at the protogalactic sites. Then, we explore in the framework of Shannon’s information theory if and how strongly each of six stellar properties of the present galaxies, namely the stellar sizes, ages, specific star formation rates, optical colors and metallicities, share mutual information with τ\tau, measured at z=127z=127. The TNG galaxy samples are deliberately controlled to have no differences in the mass, environmental density and shear distributions and to single out net effects of τ\tau on each of the galaxy stellar properties. In the higher stellar mass range of M/(h1M)1010M_{\star}/(h^{-1}\,M_{\odot})\geq 10^{10}, significant amounts of mutual information with τ\tau are exhibited by all of the six stellar properties, while in the lower range of M/(h1M)<1010M_{\star}/(h^{-1}\,M_{\odot})<10^{10} only four of the six properties except for the specific star formation rates and colors yield significant signals of τ\tau-dependence. It is also shown that the galaxy stellar sizes, which turn out to be most robustly dependent on τ\tau regardless of MM_{\star}, follow a bimodal Gamma distribution, the physical implication of which is discussed.

1 Introduction

It was conventionally believed that the galaxy stellar properties have little connection with initial conditions of the universe as they should be mainly established through essentially stochastic processes in the subsequent evolution [1, 2, 3, 4, 5, 6]. This conventional notion implied that it would be almost impossible to probe the early universe physics with observable galaxy properties. Several recent numerical studies based on high-resolution hydrodynamic simulations, however, have indicated that this notion may be too pessimistic, demonstrating that certain observable galaxy properties associated with the total angular momenta of host dark matter (DM) halos in fact vary sensitively with initial conditions [7, 8, 9, 10, 11, 12, 13, 14].

These numerical indications are in line with the observational detection of ref. [15] that the mutual correlations among the key stellar properties of local spiral galaxies like the stellar masses, sizes, ages, metallicities, and colors appeared to be much simpler than expected, the results of which were supported by several follow-up works [16, 17, 18]. It was originally suggested by ref [15] that the simple correlation structure among the key stellar properties should be interpreted as a counter-evidence against the standard theory of hierarchical structure formation scenario in which individual galaxies undergo haphazardly different physical processes [19]. Later, ref. [20] explained that the observed simple correlation structure of galaxy stellar properties must reflect how much impact the halo angular momenta have on the galaxy intrinsic properties rather than implying any inconsistency with the hierarchical structure formation scenario.

It was revealed by several numerical and observational works that the galaxy properties like the formation epochs, sizes, surface brightnesses and morphologies are dependent not only on the total masses of DM halos but also on their total angular momenta [22, 8, 23, 24, 12, 25, 26, 20]. Besides, multiple observational studies also proved that the galaxy angular momenta still retain significant amounts of memory for the initial conditions of the early universe [27, 28, 13], as predicted by the linear tidal torque theory [29, 30, 31, 32]. As pointed out by [13], even though the halo angular momenta develop deviations from the predictions of tidal torque theory through hierarchical merging processes which indeed have an effect of modifying the directions and magnitudes of halo angular momenta from the protohalo versions [33], their connections with the initial tidal fields do not severely diminish, as the spin angular momenta of merged galaxies acquire connections with the initial tidal fields on larger scales through the process of orbital angular momenta transfer.

Given this existence of interdependence among the initial conditions, halo angular momenta and galaxy properties, ref. [20] developed an efficient way to quantify the effects of initial conditions on the galaxy properties in the context of Shannon’s information theory [34]. Defining a single parameter initial condition, τ\tau, as the degree of misalignments between the principal axes of the protogalaxy inertia and initial tidal tensors [21], to which the first order generations of protohalo angular momenta are critically subject [30, 31], they numerically evaluated the amounts of mutual information between τ\tau and various intrinsic properties of galactic halos as measures of the interdependences. It was found in their study that the formation epochs, spin parameters, stellar to total mass ratio and kinematic morphologies indeed share significants amounts of mutual information with this single parameter initial condition.

Nevertheless, the intrinsic properties of galactic halos considered in the work of ref. [20] were not directly observable ones. Furthermore, it was not properly taken into account that the significant amounts of mutual information detected by [20] between τ\tau and intrinsic properties of present galactic halos could be at least partially contributed by the differences in the total mass and environments, which have been well known to affect the formation epochs and evolutionary tracks of galactic halos [35, 36, 38, 39, 40, 41, 42, 43, 44]. Another difficulty in applying the methodology of ref. [20] to observations comes from the fact that it is impossible to evaluate from real data the single parameter initial condition τ\tau, which is defined in terms of the inertia tensors of protogalactic sites.

Our task here is to answer the following vital questions. Is there a more practical and feasible way to define the single parameter initial condition? Are the observable stellar properties of present galaxies also significantly dependent on the single parameter initial condition? Is it possible to single out the τ\tau-dependence of galaxy stellar properties free from the dominant effects of total mass and environments? To conduct this task, we will take the same information theoretical approach as in our prior work [20].

2 MI between the primordial spin factor and galaxy stellar properties

2.1 Primordial spin factor as the single parameter initial condition

Refer to caption
Figure 1: Mutual information (MI) values (red histogram) between six stellar properties of the TNG central galaxies at z=0z=0 and the primordial spin factor, τ\tau, measured at zi=127z_{i}=127 in three different stellar mass ranges for three different cases of the Lagrangian scale, rfr_{f} on which the initial density and potential hessian tensor fields are smoothed. Average MI values obtained from the 10001000 resamples with shuffled (S,τS,\ \tau) (blue histogram) are also plotted with errors for comparison.

To study if and how strongly the stellar properties of galactic subhalos depend on the single parameter initial condition, we utilize the particle snapshots and subhalo catalogs at z=0z=0 from the 300 Mpc volume run (TNG300-1) of the IllustrisTNG suite of cosmological hydrodynamical simulations [45, 46, 47, 48, 49, 50], which have already been released to the general public222https://www.tng-project.org/data/. The IllustrisTNG 300-1 run, performed on a period box of side length Ltot205h1MpcL_{\rm tot}\equiv 205\,h^{-1}{\rm Mpc} for the Planck Λ\LambdaCDM cosmology [51], simulated the gravitational and hydrodynamical dynamics of 250032500^{3} DM particles and equal number of baryonic cells with individual masses, 5.95.9 and 1.11.1, in unit of 107M10^{7}\,M_{\odot}, respectively, by implementing the competent AREPO code [52].

The TNG 300-1 subhalo catalogs contain the substructures of friends-of-friends (FoF) groups identified via the SUBFIND algorithm [53], providing various information on their intrinsic and stellar properties like their comoving positions (𝐱c{\bf x}_{c}), total masses (MtM_{\rm t}), stellar masses (MM_{\star}), stellar formation epochs (afa_{f\star}), specific star formation rate (sSFR), photometric magnitudes at eight different wavebands (U,B,V,K,g,r,i,zU,\ B,\ V,\ K,\ g,\ r,\ i,\ z), stellar metallicities (ZZ_{\star}) as well as the comoving positions and masses of each constituent particle ({𝐱α,mα})\{{\bf x}_{\alpha},\ m_{\alpha}\}), respectively. Here, the stellar formation epoch, afa_{f\star}, is defined as the scale factor at the mean age of constituent stars. Among the central subhalos of the TNG FoF groups, only those that contain more than 100100 stellar particles are included in our main sample of the target galaxies [54]. Throughout this paper, we will consider six different stellar properties of each target galaxy, R90R_{90\star}, R50R_{50\star}, afa_{f\star}, sSFR, grg-r colors, and ZZ_{\star}, where R90R_{90\star} and R50R_{50\star}, denote two different stellar sizes determined as the radial distances which enclose 90%90\% and 50%50\% of MM_{\star}, being called the stellar 90 and 50 percent-mass radii, respectively.

For each target galaxy in the main sample, we locate the initial positions, {𝐪α}\{{\bf q}_{\alpha}\}, of its constituent DM and gas particles at the initial redshift zi=127z_{i}=127, and compute their center of mass, 𝐪¯Mt1αmα𝐪α\bar{\bf q}\equiv M^{-1}_{\rm t}\sum_{\alpha}m_{\alpha}{\bf q}_{\alpha} as the protogalactic site. Then, we determine the primordial spin factor, τ(𝐪¯)\tau(\bar{\bf q}), at each protogalactic site by taking the following procedure:

  1. 1.

    Reconstruct the initial raw density contrast field, δ(𝐪)\delta({\bf q}), on 5123512^{3} grid points from the particle snapshot at ziz_{i} via the cloud-in-cell algorithm.

  2. 2.

    Compute the Fourier-space raw density contrast field, δ(𝐤)\delta({\bf k}), by using the fast Fourier transformation (FFT) code, where 𝐤{\bf k} is the Fourier-space wave vector.

  3. 3.

    Compute the density hessian tensor smoothed with a Gaussian kernel on the scale of RfR_{f} as Hij(𝐤)kikjδ(𝐤)exp(k2Rf2/2)H_{ij}({\bf k})\equiv k_{i}k_{j}\delta({\bf k})\exp\left(-k^{2}R^{2}_{f}/2\right) where k|𝐤|k\equiv|{\bf k}|.

  4. 4.

    Compute the Gaussian filtered tidal tensor as Tij(𝐤)=Hij(𝐤)/k2T_{ij}({\bf k})=H_{ij}({\bf k})/k^{2}, which is equivalent to the initial potential hessian tensor.

  5. 5.

    Conduct inverse Fourier transformations of the smoothed density hessian and tidal tensor fields into the real space counterparts, Hij(𝐪)H_{ij}({\bf q}) and Tij(𝐪)T_{ij}({\bf q}), respectively.

  6. 6.

    Find the orthonormal eigenvectors of Tij(𝐪¯)T_{ij}(\bar{\bf q}) at each protogalactic site, and express Hij(𝐪¯)H_{ij}(\bar{\bf q}) in the principal frame of Tij(𝐪¯)T_{ij}(\bar{\bf q}) via a similarity transformation

  7. 7.

    Compute the primordial spin factor, τ(rf)\tau(r_{f}), as

    τ(H122+H232+H312H112+H222+H332)1/2.\tau\equiv\left(\frac{H^{2}_{12}+H^{2}_{23}+H^{2}_{31}}{H^{2}_{11}+H^{2}_{22}+H^{2}_{33}}\right)^{1/2}\,. (2.1)

It is important to note the difference between eq. (2.1) and the original definition of τ\tau given in our prior works [20, 21]. Basically, eq.(2.1) replaces the protogalaxy inertia tensor, (Iij)(I_{ij}), by the density hessian matrix , (Hij)(H_{ij}), on the ground that (Hij)(H_{ij}) was numerically found to approximate (Iij)(I_{ij}) quite well as long as RfR_{f} is comparable to the protogalaxy Lagrangian size [10]. There are two main advantages of using (Hij)(H_{ij}) instead of (Iij)(I_{ij}) for τ\tau. First, the former can be uniquely defined unlike the latter, which depends sensitively on the subhalo identification algorithm, i.e., subhalo boundary. Second, it is in principle possible to statistically reconstruct the former from real observational data [31, 10, 27, 28], while the latter is not practically measurable.

The non-negligible value of τ\tau is the sole initial condition for the first order generation of protogalaxy angular momenta [30]. From here on, this single parameter initial condition, τ\tau, will be called the primordial spin factor, and its mutual information with the six stellar properties of the TNG galaxies at z=0z=0 will be determined for three different cases of the smoothing scale, rfRf/(h1Mpc)=0.5r_{f}\equiv R_{f}/(h^{-1}{\rm Mpc})=0.5, 11, and 22, in sections 2.2-3.

Refer to caption
Figure 2: Same as figure 1 but from the MtM_{\rm t}-controlled sample where possible effects of mass differences are all reduced to a negligible level.

2.2 Mutual information from the original galaxy sample

To evaluate the mutual information (MI) between a galaxy stellar property (representatively say, SS) and the primordial spin factor (τ\tau), we first divide two dimensional configuration space spanned by SS and τ\tau into a total number of NpixelN_{\rm pixel} pixels of small area. Let NgN_{g} denote the total number of the selected target galaxies in the main sample. Let also n(a,b)n(a,b) denote the number of those target galaxies whose values of (S,τ)(S,\tau) belong to the (a,b)(a,b)th pixel. The MI(S,τ){\rm MI}(S,\tau), can be computed as

MI(S,τ)=a=1Nab=1NbP(Sa,τb)logP(Sa,τb)P(Sa)P(τb),{\rm MI}(S,\tau)=\sum_{a=1}^{N_{a}}\sum_{b=1}^{N_{b}}P(S_{a},\tau_{b})\log\frac{P(S_{a},\tau_{b})}{P(S_{a})P(\tau_{b})}\,, (2.2)

Here, (Sa,τb)(S_{a},\tau_{b}) represent the median of (S,τ)(S,\tau) values in the (a,b)(a,b)th pixel, while P(Sa,τb)n(a,b)/NgP(S_{a},\tau_{b})\equiv n(a,b)/N_{\rm g}, P(Sa)b=1Nbn(a,b)/NgP(S_{a})\equiv\sum_{b=1}^{N_{b}}\,n(a,b)/N_{\rm g}, and P(τb)a=1Nan(a,b)/NgP(\tau_{b})\equiv\sum_{a=1}^{N_{\rm a}}\,n(a,b)/N_{\rm g}, where NaN_{a} and NbN_{b} denote the total numbers of bins into which the ranges of SS and τ\tau values are divided, satisfying NaNb=NpixelN_{a}N_{b}=N_{\rm pixel}. A larger amount of MI(S,τ){\rm MI}(S,\tau) translates into the existence of a stronger connection between SS and τ\tau. To evaluate the statistical significance of MI values, we create 10001000 resamples of the target galaxies by randomly shuffling their SS and τ\tau values, and compute the average and standard deviation over the 10001000 resamples for each property.

Given the existence of strong correlations between SS and MM_{\star} [55, 56, 57], we classify the main sample into low-, intermediate-, and high-mass galaxies corresponding to three stellar mass ranges of 9m<9.59\leq m_{\star}<9.5, 9.5m<109.5\leq m_{\star}<10 and 10m<1110\leq m_{\star}<11 with mM/(h1M)m_{\star}\equiv M_{\star}/(h^{-1}M_{\odot}), respectively, and separately compute the mean amounts of MI(S,τ){\rm MI}(S,\tau) by eq. (2.2) in each mm_{\star}-range. Figure 1 plots the amounts of MI (red histogram) from the original sample, and average MI over the shuffled resamples with one standard deviation errors (blue histogram) versus the six stellar properties in the three different ranges of mm_{\star} for the three different case of rfr_{f}. As can be seen, for the case of rf=0.5r_{f}=0.5, all of the six stellar properties exhibit statistically significant signals of MI in all of the three stellar mass ranges except for the stellar metallicities that show negligibly low MI for the case of high-mass galaxies. The amounts of MI increase as mm_{\star} increases and that the largest amount of MI is produced by R90R_{90\star} in all of the three mm_{\star} ranges. The low and intermediate-mass galaxies exhibit an overall trend that the amounts and statistical significances of MI are lower for the case of larger rfr_{f}. Whereas, the high-mass galaxies yield the least significant amounts of MI on the scale of rf=1r_{f}=1 rather than on rf=2r_{f}=2, which implies that the initial density and tidal field defined on the larger scales (rf2r_{f}\geq 2) also contribute to the establishments of the high-mass galaxy stellar properties.

2.3 Mutual information from controlled galaxy samples

mm_{\star} Ng(rf=0.5)N_{g}(r_{f}=0.5) Ng(rf=1)N_{g}(r_{f}=1) Ng(rf=2)N_{g}(r_{f}=2)
[9.0,9.5][9.0,9.5] 2377823778 2559625596 2491224912
[9.5,10.0][9.5,10.0] 1633816338 1628416284 1742417424
[10.0,11.0][10.0,11.0] 1432214322 1685416854 1332613326
Table 1: Total number of the galaxies in the controlled sample for three different cases of rfr_{f} in three different stellar mass ranges.
Refer to caption
Figure 3: Same as figure 1 but from the MtM_{\rm t} and δnl\delta_{\rm nl} controlled sample in which any possible effects caused by differences in total masses and environmental densities are all reduced to a negligible level.

The galaxy stellar properties are well known to be a function of the total mass, MtM_{\rm t}, of its host DM halo [36, 39]. Furthermore, the DM halos with different total masses were found to have different τ\tau values [20], which implies that the MI signals displayed in figure 1 could be at least partially contributed by the differences in MtM_{\rm t} among the galaxies belonging to different SS-τ\tau pixels. To nullify a possible effect of mass difference on MI(S,τ){\rm MI}(S,\tau), it is necessary to eliminate MtM_{\rm t} differences depending on the τ\tau values. Binning the ranges of logmt\log m_{\rm t} and τ\tau, we count the number of galaxies whose values of mtm_{\rm t} fall in each bin. Then, we look for a τ\tau bin which yields the lowest galaxy number (say, nminn_{\rm min}) at a given mtm_{\rm t} bin. Selecting only nmin(logmt)n_{\rm min}(\log m_{\rm t}) galaxies from each of the τ\tau bins, we end up having a controlled galaxy sample where the effect of mass differences among different pixels disappear. Using this controlled sample, we recalculate the MI values by following the same procedure described in section 2.2.

Figure 2 shows the same as figure 1 but from the mtm_{\rm t}-controlled sample. As can be seen, although the controlled sample tends to yield slightly lower amounts of MI between the galaxy stellar properties and primordial spin factors, the signals are still quite significant for the case of rf=0.5r_{f}=0.5, confirming that mtm_{\rm t}-difference among the galaxies in the uncontrolled sample contribute at most only partially to the MI values. Note that the stellar ages and colors of the high-mass galaxies and the two stellar sizes of the low and intermediate-mass galaxies from the mtm_{\rm t} controlled sample yield even larger and more significant amounts of MI for the cases of rf=0.5r_{f}=0.5 and 22, respectively, than those from the original uncontrolled sample. This result implies that the mtm_{\rm t} dependences of these stellar properties in fact played the role of veiling their connections with the primordial spin factors.

Refer to caption
Figure 4: Same as figure 1 but from the MtM_{\rm t}, δnl\delta_{\rm nl} and qq-controlled sample in which any possible effects caused by differences in total masses, environmental densities and shears are all reduced to a negligible level.

Recalling that the galaxy stellar properties also exhibit strong variations with environmental density contrasts, δnl\delta_{\rm nl} [37, 38], we further control the target galaxy sample to have identical joint distributions of (mt,δnl)(m_{\rm t},\delta_{\rm nl}), where the local density contrast field δnl\delta_{\rm nl} is reconstructed by applying the cloud-in-cell method to the TNG particle snapshot at z=0z=0 in the same manner used for the reconstruction of the initial density field. Figure 3 shows the same as figure 1 but from the mtm_{\rm t} and δnl\delta_{\rm nl}-controlled sample. As can be seen, no drastic change is witnessed after the δnl\delta_{\rm nl} differences are eliminated. In all of the three mm_{\star} ranges, the amounts of MI between R90R_{90\star} and τ\tau are substantially reduced for the case of rf=0.5r_{f}=0.5, while for the low-mass galaxies, the statistical significances of MI between R90R_{90\star} and τ\tau is enhanced for the case of rf=2r_{f}=2.

To be as scrupulous and conservative as possible in detecting a signal of net τ\tau-dependences of the galaxy stellar properties, we control the galaxy sample even more strictly to have identical joint distributions of (mt,δnl,q)(m_{\rm t},\delta_{\rm nl},q) among different (S,τ)(S,\tau) pixels, where qq is the environmental shear [58] on which the galaxy properties were shown to depend [40, 41, 44], defined as [43]:

q{12[(ϱ1ϱ2)2+(ϱ2ϱ3)2+(ϱ3ϱ1)2]}1/2,q\equiv\bigg{\{}\frac{1}{2}\left[(\varrho_{1}-\varrho_{2})^{2}+(\varrho_{2}-\varrho_{3})^{2}+(\varrho_{3}-\varrho_{1})^{2}\right]\bigg{\}}^{1/2}\,, (2.3)

where {ϱ1,ϱ2,ϱ3}\{\varrho_{1},\ \varrho_{2},\ \varrho_{3}\} are three eigenvalues of the local tidal field measured at z=0z=0. The local tidal field is reconstructed for the determination of qq from δnl\delta_{\rm nl} in the same manner used to construct (Tij)(T_{ij}) from δ\delta, under the assumption that the magnitude of vorticity is negligible on the scale rf0.5r_{f}\geq 0.5. Table 1 lists the numbers of the target galaxies included in the controlled samples for the three cases of rfr_{f} in the three mm_{\star}-ranges. As can be read, the sizes of the controlled samples depend on rfr_{f}, since the τ\tau values for a galaxy varies with rfr_{f}.

Figure 4 shows the same as figure 1 but from the mtm_{\rm t}, δnl\delta_{\rm nl} and qq-controlled sample. As can be seen, the strictly controlled sample still yield significant signals of MI between the galaxy stellar properties and primordial spin factors, despite that possible effects of MtM_{\rm t}, δnl\delta_{\rm nl} and qq are all eliminated. It is interesting to notice that it varies with mm_{\star} which stellar property among the six has the largest amounts of MI(τ){\rm MI}(\tau) and how many stellar properties yield statistically significant MI signals. The overall trends are that for the case of rf=0.5r_{f}=0.5, the more stellar properties yield stronger signals, as mm_{\star} increases, while R90R_{90\star} is mm_{\star}-independent signals of MI. Another interesting aspect of the results shown in figure 4 is that the significant signals of MI(τ){\rm MI}(\tau) seem to be contained in the two stellar sizes of the low and intermediate-mass galaxies, and in grg-r colors of the high-mass galaxies for the case of rf=1r_{f}=1 and 22, respectively. The results shown in figure 4 compellingly demonstrates that the stellar properties of present galaxies have significant net τ\tau dependences, independent of total mass and environments, whose establishments should be ascribed to the multi-scale influences of initial density and tidal fields at the protogalactic stages.

2.4 A bimodal Γ\Gamma-distribution of galaxy stellar sizes

Refer to caption
Figure 5: Probability density function of the galaxy stellar sizes with bootstrap errors (filled circles) compared with a bimodal model (thick solid lines) consisting of two different Γ\Gamma distributions (thin solid lines) in each of three stellar mass ranges.
mm_{\star} (k1,θ1)k_{1},\ \theta_{1}) ξ\xi (k2,θ2)k_{2},\ \theta_{2})
[9.0,9.5][9.0,9.5] (15.57±5.11, 1.00±0.05)(15.57\pm 5.11,\ 1.00\pm 0.05) 0.38±0.090.38\pm 0.09 (36.14±2.39, 0.32±0.12)(36.14\pm 2.39,\ 0.32\pm 0.12)
[9.5,10.0][9.5,10.0] (12.10±2.14, 1.37±0.19)(12.10\pm 2.14,\ 1.37\pm 0.19) 0.32±0.100.32\pm 0.10 (30.17±5.70, 0.38±0.07)(30.17\pm 5.70,\ 0.38\pm 0.07)
[10.0,11.0][10.0,11.0] (4.70±2.79, 5.75±0.25)(4.70\pm 2.79,\ 5.75\pm 0.25) 0.36±0.110.36\pm 0.11 (12.86±1.20, 1.09±1.04)(12.86\pm 1.20,\ 1.09\pm 1.04)
Table 2: Two sets of two best-fitparameters of two Γ\Gamma modes of the bimodal model for p(R90)p(R_{90\star}) in three stellar mass ranges.

Now that among the 90 percent-mass radius, R90R_{90\star}, is found to yield the most robust signals of MI(τ){\rm MI}(\tau) in all of the three stellar mass ranges, it should be worth investigating what probability density distribution this stellar property follows. Given that the galaxy stellar sizes are strongly correlated with the halo spin parameter, λ\lambda [8], and that the probability density function of λ\lambda was found to be approximated much better by a Γ\Gamma-distribution with two characteristic parameters [20] rather than by the conventional log-normal model [59], we speculate that R90R_{90\star} may also be well modeled by a Γ\Gamma distribution. To verify this speculation, we determine the distribution of R90R_{90\star} by taking the following steps. Divide the range of R90R_{90\star} into multiple differential intervals of equal size, Δ90\Delta_{90\star}. Count the numbers, ngn_{\rm g}, of the galaxies whose values of R90R_{90\star} fall in each interval, and then determine the probability density, p(R90)p(R_{90\star}), at each differential interval as ng/(NgΔ90)n_{\rm g}/(N_{\rm g}\Delta_{90\star}).

Figure 5 plots p(R90)p(R_{90\star}) for the three cases of mm_{\star} range, revealing that p(R90)p(R_{90\star}) has a long tail in the large size section (R90>R90,cR_{90\star}>R_{90\star,c}), but drops rapidly in the opposite section (R90R90,cR_{90\star}\leq R_{90\star,c}) with size threshold in the range of 15R90,c/kpc2015\leq R_{90\star,c}/{\rm kpc}\leq 20. Comparing this numerically determined p(R90)p(R_{90\star}) to a Γ\Gamma distribution, we find that unlike the case of p(λ)p(\lambda), a single Γ\Gamma distribution fails to match simultaneously the stellar size distribution p(R90)p(R_{90\star}) in the whole range of R90R_{90\star}. Instead, a separate treatment of two ranges, R90R90,cR_{90\star}\leq R_{90\star,c} and R90>R90,cR_{90\star}>R_{90\star,c}, in fitting p(R90)p(R_{90\star}) to a single Γ\Gamma distribution turns out to work well, yielding two different sets of characteristic parameters. In other words, the stellar size distribution, p(R90)p(R_{90\star}), turns out to have two different Γ\Gamma modes, and this fitting result leads us to discover that the following bimodal Gamma distribution describes quite well p(R90)p(R_{90\star}) in the whole range of R90R_{90\star}:

p(R90)\displaystyle p(R_{90\star}) =\displaystyle= ξ×Γ(R90;k1,θ1)+(1ξ)×Γ(R90;k2,θ2),\displaystyle\xi\times\Gamma\left(R_{90\star};k_{1},\theta_{1}\right)+(1-\xi)\times\Gamma\left(R_{90\star};k_{2},\theta_{2}\right)\,, (2.4)
Γ(R90;ki,θi)\displaystyle\Gamma\left(R_{90\star};k_{i},\theta_{i}\right) =\displaystyle= 1f(ki)θikiR90ki1exp(R90θi),fori{1,2}\displaystyle\frac{1}{f(k_{i})\theta^{k_{i}}_{i}}R_{90\star}^{k_{i}-1}\exp\left(-\frac{R_{90\star}}{\theta_{i}}\right)\,,\,\,{\rm for}\,\,i\in\{1,2\} (2.5)
f(ki)\displaystyle f(k_{i}) =\displaystyle= 0tki1et𝑑t.\displaystyle\int_{0}^{\infty}t^{k_{i}-1}e^{-t}\,dt\,. (2.6)

Here, ξ\xi and 1ξ1-\xi represent the fractions of the first and second Γ\Gamma distributions with characteristic parameters (k1,θ1)(k_{1},\theta_{1}) and (k2,θ2)(k_{2},\theta_{2}), respectively.

Employing the χ2\chi^{2}-statistics333Technically, the python package, ``scipy.optimize.curve_fit"{\rm``scipy.optimize.curve\_fit"} is utilized for the computation of the best-fit parameters and associated errors for the two Γ\Gamma distributions., we fit the numerically obtained p(R90)p(R_{90\star}) to eqs. (2.4)-(2.6) by adjusting the values of (k1,θ1)(k_{1},\ \theta_{1}) and (k2,θ2)(k_{2},\ \theta_{2}) as well as ξ\xi. Figure 5 compares the best-fit bimodal Γ\Gamma distributions (thick solid lines) with the numerical results (red filled circles), confirming its validity in all of the three mm_{\star}-ranges. Table 2 lists the best-fit parameter values for the three cases of mm_{\star} ranges. To physically understand this bimodal nature of p(R90)p(R_{90\star}), recall the previous result that the probability density function of τ\tau was also found to follow a single Γ\Gamma-distribution whose best-fit parameters (k,θ)(k,\ \theta) turned out to be scale dependent [21]. In other words, if τ\tau is measured from the initial tidal fields smoothed on a different scale, then p(τ)p(\tau) is described by a single Γ\Gamma distribution with a different set of two parameters. Given this previous finding, we interpret the bimodal nature of p(R90)p(R_{90\star}) as an evidence supporting that the galaxy stellar sizes R90R_{90\star} are affected by multi-scale initial conditions.

3 Summary and conclusion

In light of the recent numerical results that the spin parameters and formation epochs of galactic subhalos exhibited strong dependences on the single parameter initial condition for the first order generation of protogalaxy angular momentum [20], we have numerically investigated if and how strongly the single parameter initial condition also contributes to the establishment of six observable galaxy properties: stellar sizes (R90R_{90\star} and R50R_{50\star}), ages (aa_{\star}), specific star formation rates (sSFR), optical grg-r colors (grg-r), and metallicities (ZZ_{\star}). As in our prior work [20], the current analysis has been focused on the sample of central galaxies in the stellar mass range of 9mM/(h1M)119\leq m_{\star}\equiv M_{\star}/(h^{-1}M_{\odot})\leq 11 at z=0z=0 from the TNG 300-1 hydrodynamic simulations, adopting the tidal torque picture according to which the single parameter initial condition is quantified by the degree of misalignments between the initial tidal field (Tij)(T_{ij}) and protogalaxy inertia tensors (Iij)(I_{ij}) [29, 30, 31].

Unlike in ref. [20], however, we have replaced (Iij)(I_{ij}) by the density hessian tensor, (Hij)(H_{ij}), defined as the second derivative of the initial density field on two grounds: (Hij)(H_{ij}) was found to approximate (Iij)(I_{ij}) quite well on the scale RfR_{f} comparable to a protogalaxy Lagrangian radius [10], and the former is much more robust than the latter against the variation of subhalo finding algorithm and definition of subhalo virial radius. In addition, (Hij)(H_{ij}) is observationally reconstructable at initial redshifts unlike the protogalaxy inertia tensors [27]. Defining the single parameter initial condition as the degree of misalignment between (Hij)(H_{ij}) and (Tij)(T_{ij}), and calling it the primordial spin factor, τ\tau, we have measured the values of τ\tau at the protogalactic sites of the TNG central galaxies at z=127z=127. Given that the τ\tau values vary with the smoothing scale, rfRf/(h1Mpc)r_{f}\equiv R_{f}/(h^{-1}{\rm Mpc}), we have considered three different cases of rf=0.5, 1r_{f}=0.5,\ 1 and 22.

The net τ\tau-dependences of the aforementioned six stellar properties have been evaluated by the amounts of their mutual information (MI), for which the controlled sample of the central galaxies have been created, where the effects of mass and environment differences are all minimized. It has been found that the amounts and statistical significances of the MI signals from the six stellar properties as well as the stellar property containing the largest amount of MI depend on mm_{\star} and rfr_{f}. In the following, we provide a summary of the key results of our analysis.

  • In the stellar mass range of 9m<9.59\leq m_{\star}<9.5, only the three stellar properties, R90R_{90\star}, afa_{f\star} and ZZ_{\star}, among the six are found to contain significant amounts of MI(τ){\rm MI}(\tau) for the case of rf=0.5r_{f}=0.5. The largest amount of MI about the primordial spin factor τ\tau is exhibited by the stellar metallicity, ZZ_{\star}. Although the amount of MI between each of these three properties and τ\tau shows a trend to diminish as rfr_{f} increases, the 90 percent-mass radius still exhibits substantial amounts of MI(τ){\rm MI}(\tau) even at rf=1r_{f}=1 and 22.

  • In the stellar mass range of 9.5m<109.5\leq m_{\star}<10, the 50 percent-mass radius, R50R_{50\star}, as well as R90R_{90\star}, afa_{f\star} and ZZ_{\star}, are shown to contain significant amounts of MI(τ){\rm MI}(\tau) for the case of rf=0.5r_{f}=0.5. Among these four, it is the 90 percent-mass radius R90R_{90\star} that contains the largest amount of MI with τ\tau.

  • In the stellar mass range of 10m1110\leq m_{\star}\leq 11, significant MI signals are yielded for the case of rf=0.5r_{f}=0.5 by all of the six stellar properties, among which afa_{f\star}, and grg-r colors yield the most significant amounts of MI(τ){\rm MI}(\tau).

  • The higher stellar mass a galaxy has, the larger amount of MI(τ){\rm MI}(\tau) its stellar property has with τ\tau. The only exception is the stellar metallicity, ZZ_{\star}, which yields a larger amount of MI(τ){\rm MI}(\tau) in the lower stellar mass range of m10m_{\star}\leq 10 than in the higher counterpart.

  • The amounts of MI(τ){\rm MI}(\tau) are found to decrease as rfr_{f} increases for all of the stellar properties except for the 50 percent-mass radius, R50R_{50\star}, which exhibits a larger amount of MI(τ){\rm MI}(\tau) at the larger smoothing scales in the lowest stellar mass range 9m9.59\leq m_{\star}\leq 9.5.

Noting that the most robust signal of MI(τ){\rm MI}(\tau) is produced by R90R_{90\star} in all of the three mm_{\star}-ranges, we speculate that the probability density function of R90R_{90\star} may also be similar to that of τ\tau, which was found in our prior work [21] to follow a Γ\Gamma-distribution. Upon this speculation, we have numerically determined p(R90)p(R_{90\star}) and performed a χ2\chi^{2}-fitting of it to the Γ\Gamma-distribution in each mm_{\star}-range, which has eventually led us to discover two different modes of p(R90)p(R_{90\star}), being best described by a bimodal Γ\Gamma distribution. This bimodal nature of p(R90)p(R_{90\star}) has been interpreted as an evidence of multi-scale influences of the initial conditions on the establishment of galaxy stellar properties.

The final conclusion is that the six stellar properties of present galaxies still retain significant and net dependences on the initial condition quantified by the primordial spin factor and that among the six, the stellar sizes reflect well the multi-scale effects of the initial density and tidal fields, being well modeled by an analytic bimodal Γ\Gamma distribution. Our results should provide a whole new insight into the long-standing debate on the roles of nature vs. nurture for the establishments of present galaxy properties. We intend to pursue a follow-up work in the direction of applying the current methodology to real data for observational detections of MI signals between the six stellar properties and primordial spin factors. We hope to report the result in the near future.

Acknowledgments

The IllustrisTNG simulations were undertaken with compute time awarded by the Gauss Centre for Supercomputing (GCS) under GCS Large-Scale Projects GCS-ILLU and GCS-DWAR on the GCS share of the supercomputer Hazel Hen at the High Performance Computing Center Stuttgart (HLRS), as well as on the machines of the Max Planck Computing and Data Facility (MPCDF) in Garching, Germany. JSM acknowledges the support by the National Research Foundation (NRF) of Korea grant funded by the Korean government (MEST) (No. 2019R1A6A1A10073437). JL acknowledges the support by Basic Science Research Program through the NRF of Korea funded by the Ministry of Education (No.2019R1A2C1083855).

References

  • [1] P. E. Seiden, L. S. Schulman, and H. Gerola, Stochastic star formation and the evolution of galaxies, Astrophys. J. 232 (1979) 702, doi:10.1086/157329
  • [2] T. Statler, B. F. Smith and N. Comins, Stochastic self-propagating star formation in three-dimensional disk galaxy simulations., Astrophys. J. 270 (1983) 79, doi:10.1086/161099
  • [3] N. Caplar and S. Tacchella, Stochastic modelling of star-formation histories I: the scatter of the star-forming main sequence., Mon. Not. Roy. Astron. Soc. 487 (2019), 3845, , https://doi.org/10.1093/mnras/stz1449
  • [4] S. Tacchella, J.  C.  Forbes and N. Caplar, Stochastic modelling of star-formation histories II: star-formation variability from molecular clouds and gas inflow, Mon. Not. Roy. Astron. Soc. 497 (2020), 698, https://doi.org/10.1093/mnras/staa1838
  • [5] Y. Wang, et al., A stochastic model to reproduce the star formation history of individual galaxies in hydrodynamic simulations, Mon. Not. Roy. Astron. Soc, 515 (2022), 3249, https://doi.org/10.1093/mnras/stac1956
  • [6] A. Pallottini and A. Ferrara, Stochastic star formation in early galaxies: Implications for the James Webb Space Telescope, Astron. Astrophys. 677 (2023) L4, doi:10.1051/0004-6361/202347384[arXiv:2307.03219 [astro-ph.GA]].
  • [7] L. V. Sales, J. F. Navarro, T. Theuns, J. Schaye, S. D. M. White, C. S. Frenk, R. A. Crain and C. Dalla Vecchia, The origin of discs and spheroids in simulated galaxies, Mon. Not. Roy. Astron. Soc. 423 (2012), 1544. doi:10.1111/j.1365-2966.2012.20975.x, [arXiv:astro-ph/1112.2220 [astro-ph.CO]].
  • [8] J. Km and J. Lee, How does the surface density and size of disc galaxies measured in hydrodynamic simulations correlate with the halo spin parameter?, Mon. Not. Roy. Astron. Soc. 432 (2013), 1701, doi:10.1093/mnras/stt632, [arXiv:1210.8321].
  • [9] J. Shi, A. Lapi, C. Mancuso, H.  Wang, and L. Danese, Angular Momentum of Early- and Late-type Galaxies: Nature or Nurture?, Astrophys. J. 843 (2017) 105, doi:10.3847/1538-4357/aa7893
  • [10] H. R. Yu, P. Motloch, U. L. Pen, Y. Yu, H. Wang, H. Mo, X. Yang and Y. Jing, Probing primordial chirality with galaxy spins, Phys. Rev. Lett. 124 (2020) 101302, doi:10.1103/PhysRevLett.124.101302 [arXiv:1904.01029 [astro-ph.CO]].
  • [11] C. Cadiou, A. Pontzen and H. V. Peiris, Stellar angular momentum can be controlled from cosmological initial conditions, Mon. Not. Roy. Astron. Soc. 517 (2022), 3459, doi:10.1093/mnras/stac2858, [arXiv:2206.11913 [astro-ph.GA]].
  • [12] L. E. Pérez-Montaño, V. Rodriguez-Gomez, B. C. Sodi, Q. Zhu, A. Pillepich, M. Vogelsberger and L. Hernquist, The formation of low surface brightness galaxies in the IllustrisTNG simulation, Mon. Not. Roy. Astron. Soc. 514 (2022), 5840, doi:10.1093/mnras/stac1716, [arXiv:2206.08942 [astro-ph.GA]].
  • [13] J. S. Moon and J. Lee, Reoriented Memory of Galaxy Spins for the Early Universe, Astrophys. J. 952 (2023) 101, doi:10.3847/1538-4357/acd9ac, [arXiv:2303.03058 [astro-ph.CO]].
  • [14] M. J. Sheng, L. Zhu, H. R. Yu, H. Ma, H. Li, P. Wang and X. Kang, Spin speed correlations and the evolution of galaxy-halo systems, Phys. Rev. D 109 (2024) 123548, doi:10.1103/PhysRevD.109.123548 [arXiv:2311.07969 [astro-ph.CO]].
  • [15] M. J. Disney, J. D. Romano, D. A. Garcia-Appadoo, A. A. West, J. J. Dalcanton and L. Cortese, Galaxies appear simpler than expected, Nature 455 (2008), 1082, doi:10.1038/nature07366, [arXiv:0811.1554 [astro-ph]].
  • [16] D. A. Garcia-Appadoo, A. A. West, J. J. Dalcanton, L. Cortese and M. J. Disney, Correlations among the properties of galaxies found in a blind HI survey, which also have SDSS optical data, Mon. Not. Roy. Astron. Soc. 394 (2009), 340, doi:10.1111/j.1365-2966.2008.14292.x, [arXiv:0809.1434 [astro-ph]].
  • [17] Y. Y. Chang, R. Chao, W. H. Wang and P. Chen, A Principle Component Analysis of Galaxy Properties from a Large, Gas-Selected Sample., Advances in Astronomy 2012 (2012), 208901, doi:10.1155/2012/208901
  • [18] Z. Sharbaf, I. Ferreras and O. Lahav, What drives the variance of galaxy spectra?., Mon. Not. Roy. Astron. Soc. 526 (2023), 585, doi:10.1093/mnras/stad2668
  • [19] A. J. Benson, Galaxy formation theory, Physics Reports 495 (2010) 33, doi:10.1016/j.physrep.2010.06.001
  • [20] J. S. Moon and J. Lee, Mutual information between galaxy properties and the initial predisposition, JCAP 05 (2024) 111, doi:10.1088/1475-7516/2024/05/111, [arXiv:2311.03632 [astro-ph.CO]].
  • [21] J. S. Moon and J. Lee, Galaxy Spin Transition Driven by the Misalignments between the Protogalaxy Inertia and Initial Tidal Tensors, Astrophys. J. 966 (2024) 100, doi:10.3847/1538-4357/ad3825 [arXiv:2401.11707 [astro-ph.CO]].
  • [22] Z. Berta, R. Jimenez, A. F. Heavens and B. Panter, The role of spin in the formation and evolution of galaxies, Mon. Not. Roy. Astron. Soc. 391 (2008) 197, doi:10.1111/j.1365-2966.2008.13742.x [arXiv:0802.1934 [astro-ph]].
  • [23] A. F. Teklu et al., Connecting Angular Momentum and Galactic Dynamics: The Complex Interplay between Spin, Mass, and Morphology., Astrophys. J. 812 (2015), 29, doi:10.1088/0004-637X/812/1/29
  • [24] V. Rodriguez-Gomez and et al., Galactic angular momentum in the IllustrisTNG simulation – I. Connection to morphology, halo spin, and black hole mass, Mon. Not. Roy. Astron. Soc. 512 (2022) 5978, doi:10.1093/mnras/stac806, [arXiv:2203.10098 [astro-ph.GA]].
  • [25] M. Du, L. C. Ho, H. R.Yu, and V. P. Debattista, The Origin of the Relation Between Stellar Angular Momentum and Stellar Mass in Nearby Disk-dominated Galaxies., Astrophys. J. Lett. 937 (2022) L18, doi:10.3847/2041-8213/ac911e
  • [26] A. B. Romeo, O. Agertz and F. Renaud, The specific angular momentum of disc galaxies and its connection with galaxy morphology, bar structure and disc gravitational instability, Mon. Not. Roy. Astron. Soc. 518 (2023) 1002, doi:10.1093/mnras/stac3074, [arXiv:2204.02695 [astro-ph.GA]].
  • [27] P. Motloch, H. R. Yu, U. L. Pen and Y. Xie, An observed correlation between galaxy spins and initial conditions, Nature Astron. 5 (2021) 283, doi:10.1038/s41550-020-01262-3 [arXiv:2003.04800 [astro-ph.CO]].
  • [28] P. Motloch, U. L. Pen and H. R. Yu, Correlating galaxy shapes and initial conditions: An observational study, Phys. Rev. D 105 (2022) 083504, doi:10.1103/PhysRevD.105.083504[arXiv:2111.12578 [astro-ph.CO]].
  • [29] A. G. Doroshkevich, The space structure of perturbations and the origin of rotation of galaxies in the theory of fluctuation., Astrofizika 6 (1970) 581.
  • [30] S. D. M. White, Angular momentum growth in protogalaxies, Astrophys. J. 286 (1984) 38, doi:10.1086/162573
  • [31] J. Lee and U. L. Pen, Cosmic shear from galaxy spins, Astrophys. J. Lett. 532 (2000) L5, doi:10.1086/312556 [arXiv:astro-ph/9911328].
  • [32] J. Lee and U. L. Pen, Galaxy spin statistics and spin-density correlation, Astrophys. J. 555 (2001) 106, doi:10.1086/321472 [arXiv:astro-ph/0008135].
  • [33] M. Vitvitska, A. A. Klypin, A. V. Kravtsov, R. H. Wechsler, J. R. Primack and J. S. Bullock, The origin of angular momentum in dark matter halos, Astrophys. J. 581 (2002) 809, doi:10.1086/344361 [arXiv:astro-ph/0105349 [astro-ph]].
  • [34] C. E. Shannon, A Mathematical Theory of Communicationt, Bell System Technical J. 27 (1948) 379, doi:10.1002/j.1538-7305.1948.tb01338.
  • [35] M. Postman and M. J. Geller, The morphology-density relation - The group connection., Astrophys. J. 281 (1984), 95, doi:10.1086/162078
  • [36] F. C. van den Bosch, X. Yang and H. J. Mo, Linking early and late type galaxies to their dark matter halos, Mon. Not. Roy. Astron. Soc. 340 (2003) 771, doi:10.1046/j.1365-8711.2003.06335.x, [arXiv:astro-ph/0210495 [astro-ph]].
  • [37] P. L. Gomez et al. [SDSS], Galaxy star - formation as a function of environment in the early data release of the Sloan Digital Sky Survey, Astrophys. J. 584 (2003), 210, doi:10.1086/345593, [arXiv:astro-ph/0210193 [astro-ph]].
  • [38] M. Tanaka, T. Goto, S. Okamura, K. Shimasaku and J. Brinkman, The Environmental dependence of galaxy properties in the local universe: Dependence on luminosity, local density, and system richness, Astron. J. 128 (2004) 2677, doi:10.1086/425529, [arXiv:astro-ph/0411132 [astro-ph]].
  • [39] F. Shankar, A. Lapi, P. Salucci, G. De Zotti and L. Danese, New relationships between galaxy properties and host halo mass, and the role of feedbacks in galaxy formation, Astrophys. J. 643 (2006) 14, doi:10.1086/502794, [arXiv:astro-ph/0601577 [astro-ph]].
  • [40] J. Lee and B. Lee, The Variation of Galaxy Morphological Type with the Shear of Environment, Astrophys. J. 688 (2008), 78, doi:10.1086/592181, [arXiv:0801.1558 [astro-ph]].
  • [41] J. Lee, A New Perspective on the Large-Scale Tidal Effect on the Galaxy Luminosity and Morphology, Astrophys. J. 867 (2018) 36, doi:10.3847/1538-4357/aae376 [arXiv:1805.07492 [astro-ph.GA]].
  • [42] S. Sarkar and B. Pandey, A study on the statistical significance of mutual information between morphology of a galaxy and its large-scale environment, Mon. Not. Roy. Astron. Soc. 497 (2020) 4077, doi:10.1093/mnras/staa2236, [arXiv:2003.13974 [astro-ph.GA]].
  • [43] B. Pandey and S. Sarkar, Exploring galaxy colour in different environments of the cosmic web with SDSS, Mon. Not. Roy. Astron. Soc. 498 (2020) 6069, doi:10.1093/mnras/staa2772 [arXiv:2002.08400 [astro-ph.GA]].
  • [44] A. Nandi, B. Pandey and P. Sarkar, The correlations between galaxy properties in different environments of the cosmic web, JCAP 02 (2024) 012, doi:10.1088/1475-7516/2024/02/012 [arXiv:2306.05354 [astro-ph.CO]].
  • [45] F. Marinacci, M. Vogelsberger, R. Pakmor, P. Torrey, V. Springel, L. Hernquist, D. Nelson, R. Weinberger, A. Pillepich and J. Naiman, et al. First results from the IllustrisTNG simulations: radio haloes and magnetic fields, Mon. Not. Roy. Astron. Soc. 480 (2018) 5113, doi:10.1093/mnras/sty2206 [arXiv:1707.03396].
  • [46] J. P. Naiman, et al. First results from the IllustrisTNG simulations: a tale of two elements - chemical evolution of magnesium and europium., Mon. Not. Roy. Astron. Soc. 477 (2018) 1206, doi:10.1093/mnras/sty618 [arXiv:1707.03401]
  • [47] D. Nelson, A. Pillepich, V. Springel, R. Weinberger, L. Hernquist, R. Pakmor, S. Genel, P. Torrey, M. Vogelsberger and G. Kauffmann, et al. First results from the IllustrisTNG simulations: the galaxy colour bimodality, Mon. Not. Roy. Astron. Soc. 475 (2018) 624, doi:10.1093/mnras/stx3040[arXiv:1707.03395].
  • [48] A. Pillepich, D. Nelson, L. Hernquist, V. Springel, R. Pakmor, P. Torrey, R. Weinberger, S. Genel, J. Naiman and F. Marinacci, et al., First results from the IllustrisTNG simulations: the stellar mass content of groups and clusters of galaxies, Mon. Not. Roy. Astron. Soc. 475 (2018) 648, doi:10.1093/mnras/stx3112 [arXiv:1707.03406].
  • [49] V. Springel, R. Pakmor, A. Pillepich, R. Weinberger, D. Nelson, L. Hernquist, M. Vogelsberger, S. Genel, P. Torrey and F. Marinacci, et al. First results from the IllustrisTNG simulations: matter and galaxy clustering,Mon. Not. Roy. Astron. Soc. 475 (2018) 676, doi:10.1093/mnras/stx3304 [arXiv:1707.03397].
  • [50] D. Nelson, V. Springel, A. Pillepich, V. Rodriguez-Gomez, P. Torrey, S. Genel, M. Vogelsberger, R. Pakmor, F. Marinacci and R. Weinberger, et al. The IllustrisTNG Simulations: Public Data Release, [arXiv:1812.05609].
  • [51] R. Adam et al. [Planck], Planck 2015 results. I. Overview of products and scientific results, Astron. Astrophys. 594 (2016) A1, doi:10.1051/0004-6361/201527101 [arXiv:1502.01582 [astro-ph.CO]].
  • [52] V. Springel, E pur si muove: Galilean-invariant cosmological hydrodynamical simulations on a moving mesh, Mon. Not. Roy. Astron. Soc. 401 (2010) 791, doi:10.1111/j.1365-2966.2009.15715.x
  • [53] V. Springel, S. D. M. White, G. Tormen and G. Kauffmann, Populating a cluster of galaxies. 1. Results at z = 0, Mon. Not. Roy. Astron. Soc. 328 (2001) 726, doi:10.1046/j.1365-8711.2001.04912.x [arXiv:astro-ph/0012055].
  • [54] P. Bett, V. Eke, C. S. Frenk, A. Jenkins, J. Helly and J. Navarro, The spin and shape of dark matter haloes in the Millennium simulation of a lambda-CDM universe, Mon. Not. Roy. Astron. Soc. 376 (2007) 215, doi:10.1111/j.1365-2966.2007.11432.x [arXiv:astro-ph/0608607 [astro-ph]].
  • [55] A. Gallazzi, S. Charlot, J. Brinchmann, S. D. M. White and C. A. Tremonti, The Ages and metallicities of galaxies in the local Universe, Mon. Not. Roy. Astron. Soc. 362 (2005) 41, doi:10.1111/j.1365-2966.2005.09321.x [arXiv:astro-ph/0506539 [astro-ph]].
  • [56] X. Z. Zheng, E. F. Bell, C. Papovich, C. Wolf, K. Meisenheimer, H. W. Rix, G. H. Rieke and R. Somerville, The dependence of Star Formation on Galaxy Stellar Mass, Astrophys. J. Lett. 661 (2007) L41, doi:10.1086/518690 [arXiv:astro-ph/0702208 [astro-ph]].
  • [57] M. S. Bothwell, R. Maiolino, C. Cicone, Y. Peng and J. Wagg, Galaxy metallicities depend primarily on stellar mass and molecular gas mass. Astronomy and Astrophysics 595 (2016) A48, doi:10.1051/0004-6361/201527918 [arXiv:astro-ph/1606.04102].
  • [58] A. Paranjape, O. Hahn and R. K. Sheth, Halo assembly bias and the tidal anisotropy of the local halo environment, Mon. Not. Roy. Astron. Soc. 476 (2018) no.3, 3631, doi:10.1093/mnras/sty496, [arXiv:1706.09906 [astro-ph.CO]].
  • [59] J. S. Bullock, A. Dekel, T. S. Kolatt, A. V. Kravtsov, A. A. Klypin, C. Porciani and J. R. Primack, A Universal angular momentum profile for galactic halos, Astrophys. J. 555 (2001), 240, doi:10.1086/321477 [arXiv:astro-ph/0011001 [astro-ph]].