SIMULATION OF THE COSMIC EVOLUTION OF ATOMIC AND MOLECULAR HYDROGEN IN GALAXIES

, , , , and

Published 2009 June 1 © 2009. The American Astronomical Society. All rights reserved.
, , Citation D. Obreschkow et al 2009 ApJ 698 1467 DOI 10.1088/0004-637X/698/2/1467

0004-637X/698/2/1467

ABSTRACT

We present a simulation of the cosmic evolution of the atomic and molecular phases of the cold hydrogen gas in about 3 × 107 galaxies, obtained by postprocessing the virtual galaxy catalog produced by De Lucia & Blaizot on the Millennium Simulation of cosmic structure. Our method uses a set of physical prescriptions to assign neutral atomic hydrogen (H i) and molecular hydrogen (H2) to galaxies, based on their total cold gas masses and a few additional galaxy properties. These prescriptions are specially designed for large cosmological simulations, where, given current computational limitations, individual galaxies can only be represented by simplistic model objects with a few global properties. Our recipes allow us to (1) split total cold gas masses between H i, H2, and helium, (2) assign realistic sizes to both the H i and H2 disks, and (3) evaluate the corresponding velocity profiles and shapes of the characteristic radio emission lines. The results presented in this paper include the local H i and H2 mass functions, the CO luminosity function, the cold gas mass–diameter relation, and the Tully–Fisher relation (TFR), which all match recent observational data from the local universe. We also present high-redshift predictions of cold gas diameters and the TFR, both of which appear to evolve markedly with redshift.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Observations of gas in galaxies play a vital role in many fields of astrophysics and cosmology. Detailed studies of atomic and molecular material now possible in the local universe with radio and millimeter telescopes will, over the coming decades, be extended to high redshifts as new facilities come on line.

First, hydrogen is the prime fuel for galaxies, when it condenses from the hot ionized halo onto the galactic disks. The fresh interstellar medium (ISM) thus acquired mainly consists of atomic hydrogen (H i), but in particularly dense regions, called molecular clouds, it can further combine to molecular hydrogen (H2). Only inside these clouds can new stars form. Mapping H i and H2 in individual galaxies therefore represents a key tool for understanding their growth and evolution. Second, the characteristic H i radio line permits the measurement of the radial velocity and velocity dispersion of the ISM with great accuracy, thereby leading to solid conclusions about galaxy dynamics and matter density profiles. Third, and particularly with regard to next-generation radio facilities, surveys of H i are also discussed as a powerful tool for investigating the large-scale structure of the universe out to high redshifts. While such large-scale surveys are currently dominated by the optical and higher frequency bands (e.g., Spitzer (Fang et al. 2005), Sloan Digital Sky Survey (SDSS; Eisenstein et al. 2005), DEEP2 (Davis et al. 2003), 2dFGRS (Cole et al. 2005), Galaxy Evolution Explorer (GALEX; Milliard et al. 2007), and Chandra (Gilli et al. 2003)), they may well be overtaken by future radio arrays, such as the Square Kilometre Array (SKA; Carilli & Rawlings 2004). The latter features unprecedented sensitivity and survey speed characteristics regarding the H i line and could allow the construction of a three-dimensional map of ∼109 H i galaxies in just a few years survey time. The cosmic structure hence revealed, specifically the baryon acoustic oscillations (BAOs) manifest in the power spectrum, will, for example, constrain the equation of state of dark energy an order of magnitude better than possible nowadays (Abdalla & Rawlings 2005; Abdalla et al. 2009). Fourth, deep low-frequency detections will presumably reveal H i in the intergalactic space of the dark ages (Carilli et al. 2004)—one of the ultimate jigsaw pieces concatenating the radiation-dominated early universe with the matter-dominated star-forming universe.

Typically, H i and H2 observations are considered part of radio and millimeter astronomy, as they rely on the characteristic radio line of H i at a rest-frame frequency of ν = 1.42 GHz and several carbon monoxide (CO) radio lines, indirectly tracing H2 regions, in the ν = 102–103 GHz band. Such line detections will soon undergo a revolution with the advent of new radio facilities such as the SKA and the Atacama Large Millimeter/submillimeter Array (ALMA). These observational advances regarding H i and H2 premise equally powerful theoretical predictions for both the optimal design of the planned facilities and for the unbiased analysis of future detections.

This is the first paper in a series of papers aiming at predicting basic H i and H2 properties in a large sample of evolving galaxies. Here, we introduce a suite of tools to assign H i and H2 properties, such as masses, disk sizes, and velocity profiles, to simulated galaxies. These tools are subsequently applied to the ∼3 × 107 simulated evolving galaxies in the galaxy catalog produced by De Lucia & Blaizot (2007, hereafter the "DeLucia catalog") for the Millennium Simulation of cosmic structure (Springel et al. 2005). In forthcoming publications, we will specifically investigate the cosmic evolution of H i and H2 masses and surface densities predicted by this simulation, and we will produce mock-observing cones, from which predictions for the SKA and the ALMA will be derived.

Section 2 provides background information about the DeLucia catalog. In particular, we highlight the hybrid simulation scheme that separates structure formation from galaxy evolution, and discuss the accuracy of the cold gas masses of the DeLucia catalog. In Section 3, we derive an analytic model for the H2/H i ratio in galaxies in order to split the hydrogen of the cold gas of the DeLucia catalog between H i and H2. We compare the resulting mass functions (MFs) and the CO luminosity function (LF) with recent observations. Sections 4 and 5 explain our model to assign diameters and velocity profiles to H i and H2 disks. The simulation results are compared to observations from the local universe and high-redshift predictions are presented. In Section 6, we discuss some consistency aspects and limitations of the approaches taken in this paper. Section 7 concludes the paper with a brief summary and outlook.

2. BACKGROUND: SIMULATED GALAXY CATALOG

N-body simulations of cold dark matter (CDM) on supra-galactic scales proved to be a powerful tool to analyze the nonlinear evolution of cosmic structure (e.g., Springel et al. 2006). Starting with small primordial perturbations in an otherwise homogeneous part of a model universe, such simulations can quantitatively reproduce the large-scale structures observed in the real universe, such as galaxy clusters, filaments, and voids. These simulations further demonstrate that most dark matter aggregations, especially the self-bound halos, grow hierarchically, that is, through successive mergers of smaller progenitors. Hence, each halo at a given cosmic time can be ascribed a "merger tree" containing all its progenitors. One of the most prominent simulations is the "Millennium run" (Springel et al. 2005), which followed the evolution of 21603 ≈ 1010 particles of mass 8.6 × 108Mh−1 over a redshift range z = 127 → 0 in a cubic volume of (500 Mpc h−1)3 with periodic boundary conditions. The dimensionless Hubble parameter h, defined as $H_0=100\,h\rm \;km\;s^{-1}\;Mpc^{-1}$, was set equal to h = 0.73, and the other cosmological parameters were chosen as Ωmatter = 0.25, Ωbaryon = 0.045, ΩΛ = 0.75, and σ8 = 0.9.

Despite impressive results, modern N-body simulations of CDM in comoving volumes of order (500 Mpc h−1)3 cannot simultaneously evolve the detailed substructure of individual galaxies. The reasons are computational limitations, which restrict both the mass resolution and the degree to which baryonic and radiative physics can be implemented. Nevertheless, an efficient approximate solution for the cosmic evolution of galaxies can be achieved by using a hybrid model that separates CDM-dominated structure growth from more complex baryonic physics (Kauffmann et al. 1999). The idea is to first perform a purely gravitational large-scale N-body simulation of CDM and to reduce the evolving data cube to a set of halo merger trees. These dark matter merger trees are assumed independent of the baryonic and radiative physics taking place on smaller scales, but they constitute the mass skeleton for the formation and evolution of galaxies. As a second step, each merger tree is populated with a list of galaxies, which are represented by simplistic model objects with a few global properties (stellar mass, gas mass, Hubble type, star formation rate (SFR), etc.). The galaxies are formed and evolved according to a set of physical prescriptions, often of a "semianalytic" nature, meaning that galaxy properties evolve analytically unless a merger occurs. This hybrid approach tremendously reduces the computational requirements compared to hydro-gravitational N-body simulations of each galaxy.

Croton et al. (2006) were the first to apply this hybrid scheme to the Millennium Simulation, thus producing a catalog with ∼1.1 × 109 galaxies in 64 time steps, corresponding to about ∼3 × 107 evolving galaxies. This catalog was further improved by De Lucia & Blaizot (2007), giving rise to the DeLucia catalog used in this paper. The underlying semianalytic prescriptions to form and evolve galaxies account for the most important mechanisms known today. In brief, the hot gas associated with the parent halo is converted into galactic cold gas according to a cooling rate that scales with redshift and depth of the halo potential. Stars form at a rate proportional to the excess of the cold gas density above a critical density, below which star formation is suppressed. In return, supernovae reheat some fraction of the cold gas, and, if the energy injected by supernovae is large enough, their material can escape from the galaxy and later be reincorporated into the hot gas. In addition, when galaxies become massive enough, cooling gas can be reheated via feedback from active galaxy nuclei (AGNs) associated with continuous or merger-based black hole mass accretion. These basic mechanisms are completed with additional prescriptions regarding merger-related starbursts, morphology changes, metal enrichment, dust evolution, and change of photometric properties (see Croton et al. 2006; De Lucia & Blaizot 2007). The free parameters in this model were adjusted such that the simulated galaxies at redshift z = 0 fit the joint luminosity/color/morphology distribution of observed low-redshift galaxies (Cole et al. 2001; Huang et al. 2003; Norberg et al. 2002). A good first-order accuracy of the model is suggested by its ability to reproduce the observed bulge-to-black hole mass relation (Häring & Rix 2004), the Tully–Fisher relation (TFR; Giovanelli et al. 1997), and the cold gas metallicity as a function of stellar mass (Tremonti et al. 2004, see also Figures 4 and 6 in Croton et al. 2006).

Some galaxies in the DeLucia catalog have no corresponding halo in the Millennium Simulation. Such objects can form during a halo merger, where the resulting halo is entirely ascribed to the most massive progenitor galaxy. In the model, the other galaxies continue to exist as "satellite galaxies" without halos. These galaxies are identified as "type 2" objects in the DeLucia catalog. If the halo properties of a satellite galaxy are required, they must be extrapolated from the original halo of the galaxy or estimated from the baryonic properties of the galaxy.

We emphasize that the semianalytic recipes of the DeLucia catalog are simplistic and may require an extension or readjustment, when new observational data become available. In particular, recent observations (Bigiel et al. 2008; Leroy et al. 2008) suggest that star formation laws based on a surface density threshold are suspect, especially in low surface density systems. Moreover, galaxies in the DeLucia catalog with stellar masses Mstars below 4 × 109M typically sit at the centers of halos with less than 100 particles, whose merging history could only be followed over a few discrete cosmic time steps. It is likely that the physical properties of these galaxies are not yet converged. Croton et al. (2006) noted that especially the morphology (colors and bulge mass) of galaxies with Mstars ≲ 4 × 109M is poorly resolved, since, according to the model, the bulge formation directly relies on the galaxies' merging history and disk instabilities. Nevertheless, the simulated cosmic space densities of stars in early- and late-type galaxies at redshift z = 0 and the cosmic star formation history are consistent with observations (see figures and references in Croton et al. 2006). It is therefore probable that at least the more massive galaxies in the simulation are not significantly affected by the mass resolution and the simplistic law for star formation.

In this paper, we postprocess the DeLucia catalog to estimate realistic H i and H2 properties for each galaxy. Our prescriptions will make use of the cold gas masses given for each galaxy in the DeLucia catalog, and hence it is crucial to verify these cold gas masses against current observations. Based on a new estimation of the H2 MF, we have recently calculated the normalized density of cold neutral gas in the local universe as Ωobsgas = (4.4 ± 0.8) × 10−4h−1, hence Ωobsgas ≈ 6.0 × 10−4 for h = 0.73 (Obreschkow & Rawlings 2009). This value was obtained by integrating the best-fitting Schechter functions of the local H i MF and H2 MF and therefore it includes an extrapolation toward masses below the respective detection limits of H i and H2. The simulated local cold gas density Ωsimgas of the DeLucia catalog, obtained from the sum of the cold gas masses of all galaxies at redshift z = 0, exceeds the observed value by a factor ζ = Ωsimgasobsgas = 1.45, as shown in Table 1. For comparison, this table also lists the cold gas densities of other galaxy catalogs produced for the Millennium Simulation, using different semianalytic recipes (Bertone et al. 2007) and different schemes for the construction of dark matter merger trees (Bower et al. 2006).

Table 1. Normalized Cold Gas Densities at z = 0 of Three Different Semianalytic Galaxy Simulations Applied to the Millennium Simulation of Cosmic Structure

Catalog Ωsimgas Ωsimgasobsgas
De Lucia & Blaizot (2007) 8.7 × 10−4 1.45
Bower et al. (2006) 14.8 × 10−4 2.45
Bertone et al. (2007) 9.0 × 10−4 1.50

Note. The rightmost column shows the multiplicative offset from the observed value as determined by Obreschkow & Rawlings (2009).

Download table as:  ASCIITypeset image

There are plausible reasons for the excess of cold gas in the DeLucia catalog compared to observations. Most importantly, the semianalytic recipes only distinguish between two gas phases: the hot (T ≈ 106–107 K) and ionized material located in the halo of the galaxy or group of galaxies, and the cold (T ≈ 102–103 K) gas in galactic disks. However, recent observations have clearly revealed that some hydrogen in the disk of the Milky Way is warm (T ≈ 104 K) and ionized, too. For example, Reynolds (2004) analyzed faint optical emission lines from hydrogen, helium, and trace atoms, leading to the conclusion that about 1/3 of all the hydrogen gas in the local interstellar cloud (LIC) is ionized. If this were true for all the gas in disk galaxies, one would expect a correction factor around 1.5 between simulated disk gas and cold neutral gas. Justified by this considerations, we decided to divide all the cold gas masses in the DeLucia catalog MDeLuciagas by the constant ζ = 1.45 in order to obtain more realistic estimates,

Equation (1)

3. GAS MASSES AND MASS FUNCTIONS

In this section, we establish a physical prescription to subdivide the cold and neutral hydrogen mass $M_{\rm H}=M_{\rm H\,\hbox{\scriptsize {\sc i}}}+M_{{\rm H}_2}$ of a galaxy into its atomic (H i) and molecular (H2) component based on the observed and theoretically confirmed relation between local gas pressure and local molecular fraction (Elmegreen 1993; Blitz & Rosolowsky 2006; Leroy et al. 2008; Krumholz et al. 2009). This prescription shall be applied to the DeLucia catalog. The resulting simulated H i and H2 MFs and the related CO LF will be compared to observations in the local universe.

3.1. Prescription for Subdividing Cold Gas

Before addressing the subcomposition of cold hydrogen, we note that the total cold hydrogen mass MH can be inferred from the total cold gas mass Mgas by a constant factor MH = 0.74 Mgas, which corresponds to the universal abundance of hydrogen (e.g., Arnett 1996) that changes insignificantly with cosmic time. The remaining gas is composed of helium (He) and a minor fraction of heavier elements, collectively referred to as metals (Z). The DeLucia catalog gives an estimate for the metal mass in cold gas MZ, and hence we shall compute the masses of cold hydrogen and He as

Equation (2)

where the hydrogen fraction β = 0.75 is chosen slightly above 0.74 to account for the subtraction of the 1%–2% metals in Equation (2).

The subdivision of the cold hydrogen mass $M_{\rm H}=M_{\rm H\,\hbox{\scriptsize {\sc i}}}+M_{{\rm H}_2}$ depends on the galaxy and evolves with cosmic time. We shall tackle this complexity using the variable H2/H i ratio $R_{\rm mol}^{\rm galaxy}\equiv M_{{\rm H}_2}/M_{\rm H\,\hbox{\scriptsize {\sc i}}}$, hence

Equation (3)

Detailed observations of H i and CO in nearby regular spiral galaxies revealed that virtually all cold gas of these galaxies resides in flat, often approximately axially symmetric, disks (e.g., Walter et al. 2008; Leroy et al. 2008). CO maps recently obtained for five nearby elliptical galaxies (Young 2002) show that even these galaxies, who carry most of their stars are in a spheroid, have most of their cold gas in a disk. There is also empirical evidence that most cold gas in high-redshift galaxies resides in disks (e.g., Tacconi et al. 2006). Based on these findings, we assume that galaxies generally carry their cold atomic and molecular cold gas in flat disks with axially symmetric surface density profiles $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$ and $\Sigma _{\rm H_2}(r)$, where r denotes the galactocentric radius in the plane of the disk. Using these functions, Rgalaxymol can be expressed as

Equation (4)

To solve Equation (4), we shall now derive an analytic model for $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$ and $\Sigma _{\rm H_2}(r)$. To this end, we analyzed the observed density profiles $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$ and $\Sigma _{\rm H_2}(r)$ presented by Leroy et al. (2008) for 12 nearby spiral galaxies of The H i Nearby Galaxy Survey (THINGS).5 In general, the surface density of the total hydrogen component (H i + H2) is well fitted by a single exponential profile,

Equation (5)

where rdisk is a scale length and $\tilde{\Sigma }_{\rm H}\equiv M_{\rm H}/(2\pi r_{\rm disk}^2)$ is a normalization factor, which can be interpreted the maximal surface density of the cold hydrogen disk.

Equation (5) can be solved for $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$ and $\Sigma _{\rm H_2}(r)$, if we know the local H2/H i ratio in the disk, i.e., the radial function $R_{\rm mol}(r)\equiv \Sigma _{\rm H_2}(r)/\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$. Following the theoretical prediction that Rmol(r) scales as some power of the gas pressure (Elmegreen 1993), Blitz & Rosolowsky (2006) presented compelling observational evidence for this power law based on 14 nearby spiral galaxies of various types. Perhaps the most complete empirical study of Rmol(r) today has recently been published by Leroy et al. (2008), who analyzed the correlations between Rmol(r) and various disk properties in 23 galaxies of the THINGS catalog. This study confirmed the power-law relation between Rmol(r) and pressure. On theoretical grounds, Krumholz et al. (2009) argued that Rmol(r) is most fundamentally driven by density rather than pressure. However, by virtue of the thermodynamic relation between pressure and density, it is, in the context of this paper, irrelevant which quantity is considered, and the density law for Rmol(r) by Krumholz et al. (2009) is indeed consistent with the pressure laws by Blitz & Rosolowsky (2006) and Leroy et al. (2008). Here, we shall apply the pressure law

Equation (6)

where P(r) is the kinematic midplane pressure outside molecular clouds, and P* = 2.35 × 10−13 Pa and α = 0.8 are empirical values adopted from Leroy et al. (2008).

Elmegreen (1989) showed that the equations of hydrostatic equilibrium for an infinite thin disk with gas and stars exhibit a simple approximate solution for the macroscopic kinematic midplane pressure P(r) of the ISM,

Equation (7)

where G is the gravitational constant, Σgas(r) is the surface density of the total cold gas component (H i + H2 + He+ metals), Σdiskstars(r) is the surface density of stars in the disk (thus excluding the bulge stars of early-type spiral galaxies and elliptical galaxies), and fσ(r) ≡ σgasstars,z is the ratio between the vertical velocity dispersions of gas and stars. The impact of supernovae and other small-scale effects on the gas pressure are implicitly included in Equation (7) via the velocity dispersion σgas. For Σdiskstars = 0, Equation (7) reduces to P(r) = 0.5 π G Σgas(r)2, which is sometimes used as an approximation for the ISM pressure in gas-rich galaxies (e.g., Crosthwaite & Turner 2007).

To simplify Equation (7), we note that Σgas(r) can be expressed as Σgas(r) = Mgas/(2π r2disk)exp(−r/rdisk), which is identical to Equation (5) up to the constant factor correcting for helium and metals. To find a similar expression for Σdiskstars(r), we analyzed the stellar surface densities Σstars(r) of the 12 THINGS galaxies mentioned before. In agreement with many other studies (e.g., Courteau et al. 1996), we found that Σstars(r) is generally well approximated by a double-exponential profile, i.e., the sum of an exponential profile Σbulgestars for the bulge and an exponential profile Σdiskstars for the disk. On average, the scale length of the stellar disk $\tilde{r}_{\rm disk}$ is 30%–50% smaller than the gas scale length rdisk, which traces the fact that stars form in the more central H2-dominated parts of galaxies. Indeed, several observational studies revealed that the stellar scale length is nearly identical to that of molecular gas (e.g., Young et al. 1995; Regan et al. 2001; Leroy et al. 2008), and hence smaller than the scale length of H i or the scale length rdisk of the total cold gas component. For simplicity, we shall here assume $r_{\rm disk}=2\,\tilde{r}_{\rm disk}$ for all galaxies, such that Σdiskstars(r) = 4 Mstars/(2π r2disk)exp(−2 r/rdisk). Finally, we approximate the dispersion ratio fσ(r) as fσ(r) = f0σ exp(r/rdisk), where f0σ is a constant. This approximation is motivated by empirical evidence that the gas dispersion σgas remains approximately constant across galactic discs (e.g., Dickey et al. 1990; Boulanger & Viallefond 1992; Leroy et al. 2008), combined with theoretical and observational studies showing that the stellar velocity dispersion σstars,z decreases approximately exponentially with a scale length twice that of the stellar surface density (e.g., Bottema 1993). Within those approximations, Equation (7) reduces to

Equation (8)

where 〈fσ〉 ≡ 4f0σ is a constant, which can be interpreted as the average value of fσ(r) weighted by the stellar surface density, since ∫2π r Σdiskstars(r)fσ(r)/Mdiskstars = 4f0σ.

Substituting P(r) in Equation (6) for Equation (8), we obtain

Equation (9)

with

Equation (10)

where $K\equiv G/(8\pi \,P_\ast)=11.3\rm \;m^4\;kg^{-2}$. Equation (9) reveals that the H2/H i ratio Rmol(r) is described by an exponential profile with scale length rdisk/1.6. It should be emphasized that the central value Rcmol does not necessarily correspond to the H2/H i ratio measured at the center of real galaxies due to an additional H2 enrichment caused by the central stellar bulge. However, Rcmol represents the extrapolated cental H2/H i ratio of the exponential profile, which approximates Rmol(r) in the outer, disk-dominated galaxy parts.

We can now solve Equations (5) and (9) for the atomic and molecular surface density profiles, i.e.,

Equation (11)

Equation (12)

These model profiles can be checked against the observed H i - and H2-density profiles of the nearby galaxies analyzed by Leroy et al. (2008). In particular, we can test the limitations implied by our assumption that $r_{\rm disk}=2\,\tilde{r}_{\rm disk}$. To this end, we selected two observed regular spiral galaxies with $r_{\rm disk}\approx 2\,\tilde{r}_{\rm disk}$ (NGC 3184) and $r_{\rm disk}\approx \tilde{r}_{\rm disk}$ (NGC 5505). To evaluate Equations (11) and (12) for those galaxies, we require the quantities Mdiskstars, Mgas, rdisk, and 〈fσ〉. Mdiskstars, Mgas, and rdisk were determined by fitting a single exponential profile to the total cold gas component and a double-exponential profile (bulge and disk) to the stellar component, and 〈fσ〉 was chosen as 〈fσ〉 = 0.4, i.e., the value given by Elmegreen (1993) for nearby galaxies. As shown in Figure 1, the resulting model profiles $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$ and $\Sigma _{\rm H_2}(r)$ approximately match the empirical data. The fact that the fit is rather good for both galaxies demonstrates that the quality of the model predictions does not sensibly depend on the goodness of the model assumption $r_{\rm disk}\approx 2\,\tilde{r}_{\rm disk}$. Similarly good fits are indeed found for most of the 12 THINGS galaxies, for which Leroy et al. (2008) published radial H i - and H2-density profiles.

Figure 1.

Figure 1. Column density profiles of two nearby spiral galaxies. The filled triangles, circles, and squares, respectively, represent the measured column density profiles of stars, H i, and H2 (Leroy et al. 2008). The solid lines show the best-fitting double-exponential functions for the stellar densities. The dashed lines and dash-dotted lines represent the predictions of our pressure-based model given in Equations (11) and (12).

Standard image High-resolution image

Equations (11) and (12) can be solved for the maximal surface densities of H i and H2. $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$ exhibits its maximum at the radius $r_{\rm H\,\hbox{\scriptsize {\sc i}}}^{\max }=0.625\,r_{\rm disk}\,\ln (3/5\times R_{\rm mol}^{c})$, as long as Rcmol>5/3. Galaxies in this category show an H i drop toward their center, such as observed in most galaxies in the THINGS catalog (Walter et al. 2008). By contrast, disk galaxies with Rcmol ⩽ 5/3 have H i-density profiles peaking at the center, $r_{\rm H\,\hbox{\scriptsize {\sc i}}}^{\max }=0$. Galaxies with such small values of Rcmol have low gas densities by virtue of Equation (10), such as the irregular galaxies NGC 4214 and NGC 3077 (see profiles in Leroy et al. 2008). $\Sigma _{\rm H_2}(r)$ given in Equation (12) always peaks at the disk center, $r_{\rm H_2}^{\max }=0$.

The maximal values of $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$ and $\Sigma _{\rm H_2}(r)$, called $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}^{\rm max}\equiv \Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r_{\rm H\,\hbox{\scriptsize {\sc i}}}^{\max })$ and $\Sigma _{\rm H_2}^{\rm max}\equiv \Sigma _{\rm H_2}(r_{\rm H_2}^{\max })$, can be computed as

Equation (13)

Equation (14)

The density profiles of Equations (11) and (12) can be substituted into Equation (4). The exact solution of Equation (4) is quite unhandy, but Rgalaxymol only depends on Rcmol and an excellent approximation, accurate to better than 5% over the 9 orders of magnitude Rcmol = 10−3to106 (covering the most extreme values at all redshifts), is given by

Equation (15)

Equations (10) and (15) constitute a physical prescription to estimate the H2/H i ratio of any regular galaxy based on four global quantities: the disk stellar mass Mdiskstars, the cold gas mass Mgas, the scale radius of the cold gas disk rdisk, and the dispersion parameter 〈fσ〉. In Obreschkow & Rawlings (2009), we showed that the H2/H i ratios inferred from this model are consistent with observations of nearby galaxies. Moreover, this model presumably extends to high redshifts, since it essentially relies on the fundamental relation between pressure and molecular fraction and on a few other physical assumptions with weak or absent dependence on cosmic epoch. However, a critical discussion of the limitations of this model is presented Sections 6.2 and 6.3.

3.2. Application to the DeLucia Catalog

We applied the model given in Equations (10) and (15) together with Equations (2) and (3) to the DeLucia catalog in order to assign H i, H2, and He masses to the simulated galaxies. The quantities Mdiskstars and MZ used in these Equations are directly contained in the DeLucia catalog, and Mgas was inferred from the given cold gas masses via the correction of Equation (1). The dispersion parameter 〈fσ〉 is approximated by the constant 〈fσ〉 = 0.4, consistent with the local observational data used by (Elmegreen 1989, but see discussion in Section 6.3).

The remaining and most subtle ingredient for our prescription of Equations (10) and (15) is the scale radius rdisk. This radius can be estimated from the virial radius rvir of the parent halo, but their relation is intricate. Even modern N-body plus smoothed particle hydrodynamics (SPH) simulations of galaxy formation still struggle at reproducing observed disk diameters (Kaufmann et al. 2007), and hence the more simplistic semianalytic approaches are likely to require some empirical adjustment. Mo et al. (1998) studied the case of a flat exponential disk in an isothermal singular halo. When assuming that the disk's mass can be neglected for its rotation curve, they find

Equation (16)

where λ is the spin parameter of the halo and ξ is the ratio between the specific angular momentum of the disk (angular momentum per unit mass) and the specific angular momentum of the halo. The model behind Equation (16) does not distinguish between different scale radii for stars and cold gas, but it assumes a single exponential disk in hydrostatic equilibrium without including the effects of star formation. It is therefore natural to identify rdisk in Equation (16) with the cold gas scale radius rdisk of Equations (10)–(12).

In the Millennium Simulation, rvir was calculated from the virial mass Mvir using the relation

Equation (17)

where ρc(z) is the critical density for closure ρc(z) = 3H2(z)/(8πG) and Mvir is the virial mass of the halo. For central halos, Mvir was approximated as M200, i.e., the mass in the region with an average density equal to 200ρc(z); and for subhalos, Mvir was approximated as the total mass of the gravitationally bound simulation particles.

The spin parameter λ was calculated directly from the N-body Millennium Simulation according to the definition λ ≡ JhaloEhalo1/2G−1Mhalo−5/2, where Jhalo denotes the angular momentum of the halo, Ehalo its energy, and Mhalo its total mass. For the satellite galaxies in the DeLucia catalog, i.e., the ones without halo (see Section 2), the value of λ × rvir was approximated as the respective value of the original galaxy halo before its disappearance.

The only missing parameter for the calculation of rdisk via Equation (16) is the angular momentum ratio ξ. It is of order unity for evolved disk galaxies (e.g., Fall & Efstathiou 1980; Zavala et al. 2008), but its exact value is uncertain because of the difficulty of measuring the spin of dark matter halos and convergence issues in numerical simulations. For example, Kaufmann et al. (2007) showed that N-body plus SPH simulations with as many as 106 particles per galaxy do not reach convergence in angular momentum, because of the difficulties to model the transport of angular momentum.

Here, we shall chose ξ, such that our simulation reproduces the empirical relation between the galaxy baryon mass Mbary and the stellar scale radius $\tilde{r}_{\rm disk}\approx r_{\rm disk}/2$, measured in the local universe. To ensure consistency with Section 3.1, we use again the data from the THINGS galaxies analyzed by Leroy et al. (2008). Of the 23 galaxies in this sample, we reject the six irregular objects, since our model for H i and H2 assumes regular galaxies. The remaining 17 galaxies cover all spiral types and include H i-rich and H2-rich galaxies. For each galaxy, we adopted the stellar scale radii $\tilde{r}_{\rm disk}$ and baryon masses $M_{\rm bary}=M_{\rm stars}+M_{\rm H\,\hbox{\scriptsize {\sc i}}}+M_{{\rm H}_2}$ directly from the data presented in Leroy et al. (2008).6 The Hubble types T, i.e., the numerical stage indexes along the revised Hubble sequence of the RC2 system (de Vaucouleurs et al. 1976), were drawn from the HyperLeda database (Paturel et al. 2003).

The resulting empirical relation between Mbary and $\tilde{r}_{\rm disk}$ is displayed in Figure 2. The scatter of these data probably underestimates the true scatter caused by all galaxies, since we exclusively considered non-interacting regular spiral galaxies. The zero point of the mean relation between Mbary and $\tilde{r}_{\rm disk}$ depends on the morphological galaxy type, as is revealed by the distinction of early- and late-type spiral galaxies in Figure 2. Given identical baryon masses, the stellar scale radii of early-type galaxies tend to be smaller than the radii of late-type galaxies. This trend was also detected in other data samples (e.g., data from Kregel et al. 2002 shown in Obreschkow & Rawlings 2009). Several reasons could explain this finding: (1) early-type galaxies have more massive stellar bulges, which present an additional central potential that contracts the disk; (2) bulges often form from disk instabilities, occurring preferably in systems with relatively low angular momentum, and hence early-type galaxies are biased toward smaller angular momenta and smaller scale radii; (3) larger bulges like those of lenticular and elliptical galaxies, often arise from galaxy mergers, which tend to reduce the specific angular momenta and scale radii.

Figure 2.

Figure 2. Relation between the baryon mass (stars + cold gas) and the stellar scale radius of disk galaxies. The filled and empty squares, respectively, represent observed early- and late-type spiral galaxies (Leroy et al. 2008). Typical 1σ error bars are shown in the bottom right corner. The empirical fit given in Equation (19) is represented by a solid line for early-type spiral galaxies (T = 2) and by a dashed line for late-type spiral galaxies (T = 10). The dotted and dash-dotted lines represent the respective relations of the simulated galaxies in the DeLucia catalog, if ξ = 1. In order for the simulation to maximally align with the observations, ξ must be chosen according to Equation (20). For the latter case, 102 random early- and late-type spiral galaxies of the DeLucia catalog at z = 0 are represented by filled and empty dots, respectively.

Standard image High-resolution image

To parameterize the dependence of the scale radius on the morphological galaxy type, the latter shall be quantified using the stellar mass fraction of the bulge, $\mathcal {\bm B}\equiv M_{\rm stars}^{\rm bulge}/M_{\rm stars}$. In the observed sample, the values of $\mathcal {\bm B}$ can be approximately inferred from the Hubble type T. Here, we shall use the relation

Equation (18)

which approximately parameterizes the mean behavior of 146 moderately inclined barred and unbarred local spiral galaxies analyzed by Weinzirl et al. (2009). Equation (18) satisfies the boundary conditions $\mathcal {\bm B}=1$ for T = −6 (i.e., pure spheriods) and $\mathcal {\bm B}=0$ for T = 10 (pure disks).

We shall approximate the relation between Mbary and $\tilde{r}_{\rm disk}$ as a power law with an additional term for the observed secondary dependence on morphological type,

Equation (19)

where a0, a1, and a2 are free parameters. The best fit to the empirical data in terms of a maximum likelihood approach is given by the choice a0 = 0.3, a1 = 0.4, and a2 = −0.6.

The fit of Equation (19) is displayed in Figure 2 for early-type spiral galaxies ($T=2\leftrightarrow \mathcal {\bm B}=0.25$, solid line) and for late-type spiral galaxies ($T=10\leftrightarrow \mathcal {\bm B}=0$, dashed line). For comparison, the mean power-law relations for the early- and late-type spiral galaxies in the DeLucia catalog at z = 0 are displayed as dotted and dash-dotted lines for the choice ξ = 1. The simulated scale radii using ξ = 1 are consistent with the observed ones for very massive galaxies (Mbary ≈ 1011M), but less massive galaxies in the simulation turn out slightly too large if ξ = 1. Furthermore, the morphological dependence of the simulation is too small compared to the observations. This can be corrected ad hoc by introducing a variable ξ that depends on both Mbary and $\mathcal {\bm B}$, i.e.,

Equation (20)

where b0, b1, and b2 are free parameters. The parameters minimizing the rms deviation between the simulated galaxies and the empirical model of Equation (19) are b0 = −0.1, b1 = 0.3, and b2 = −0.6. In addition, we chose a lower limit for ξ equal to 0.5, in order to prevent unrealistically small-scale radii.

We emphasize that Equation (20) is merely an empirical correction; this choice of ξ should not be considered as an estimate of the true ratio between the specific angular momenta of the disk and the halo, but it also accounts for the imperfection of the simplistic halo model by Mo et al. (1998), for missing physics in the semianalytic modeling, and for possible systematic errors in the spin parameters λ of the Millennium Simulation. The average value of Equation (20) over all galaxies in the DeLucia catalog is 〈ξ〉 = 0.7 (with σ = 0.2), which is approximately consistent with ξ ≈ 1 of modern high-resolution simulations of galaxy formation (e.g., Zavala et al. 2008), even though the latter still suffer from issues with the transport of angular momentum as mentioned above.

Using Equations (16) and (20), we estimated a scale radius rdisk for each galaxy in the DeLucia catalog. A sample of 102 simulated early- and late-type spiral galaxies at z = 0 is shown in Figure 2. Given rdisk as well as Mstars, Mgas, and 〈fσ〉 = 0.4, we then applied Equations (2), (3), (10), and (15) in order to subdivide the non-metallic cold gas mass (MgasMZ) of each galaxy into H i, H2, and He.

3.3. Atomic and Molecular Mass Functions

We shall now compare the H i and H2 masses predicted by our model of Sections 3.1 and 3.2 to recent observations in the local universe. From the viewpoint of the simulation, a fundamental output are the MFs of H i and H2, while the available observational counterparts are the LFs of the H i-emission line (Zwaan et al. 2005) and the CO(1–0)-emission line (Keres et al. 2003). Therefore, either the simulated data or the observed data need a luminosity-to-mass (or vise versa) conversion to compare the two. Section 3.3 focuses on the MFs, adopting the standard luminosity-to-mass conversion for H i used by Zwaan et al. (2005) and the CO-luminosity-to-H2-mass conversion of Obreschkow & Rawlings (2009). As a complementary approach, Section 3.4 will focus on the LFs, which will require a model for the conversion of simulated H2 masses into CO luminosities.

We define the MFs as ϕx(Mx) ≡ dρx/dlog Mx, where ρx(Mx) is the space density (number per comoving volume) of galaxies containing a mass Mx of the constituent x (H i, H2, He, etc.). Given a mass, such as $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$, for each galaxy in the DeLucia catalog, the derivation of the corresponding MF only requires the counting of the number of sources per mass interval. We chose 60 mass intervals, logarithmically spaced between 108M and 1011M, giving about ∼106 galaxies per mass interval in the central mass range, while keeping the mass error relatively small (Δlog(M) < 0.05). Since MFs combine units of mass and length, they generally depend on the Hubble constant H0, or the dimensionless Hubble parameter h, defined as H0 = 100 h km s−1 Mpc−1. Although MFs are often plotted in units making no assumption on h (e.g., Mh−2 for the mass scale), this is impossible when observations are compared to cosmological simulations. The reason is that simulated masses in the Millennium Simulation scale to first order as h−1, whereas empirical masses, when determined from electromagnetic fluxes, are proportional to the square of the distance and hence scale as h−2. For all plots in this paper, we shall therefore use h = 0.73, which corresponds to the value adopted by the Millennium Simulation (see Section 2).

Figure 3 displays the H i and H2 MF of our simulation (solid lines), as well as the corresponding empirical MFs for the local universe (points with error bars). The empirical H i MF was obtained by Zwaan et al. (2005) based on 4315 galaxies of the HI-Parkes All Sky Survey (HIPASS) and the empirical H2 MF was derived in Obreschkow & Rawlings (2009) from the CO LF presented by Keres et al. (2003). Both empirical MFs approximately match the simulated data. We note, however, that the consistency between observation and simulation decreases if we skip the overall correction of the cold gas masses in the DeLucia catalog by the constant factor ζ (dotted lines), which, as argued in Section 2 can be justified by a fraction of the disk gas being electronically excited or ionized.

Figure 3.

Figure 3. Simulated galaxy MFs for H i and H2 with (solid lines) and without (dotted lines) the constant correction for all cold gas masses given in Equation (1). The filled and open squares with error bars represent the corresponding empirical MFs from Zwaan et al. (2005) and Obreschkow & Rawlings (2009).

Standard image High-resolution image

Our simulation slightly overpredicts the observed number of the largest H i and H2 masses, i.e., the ones in the exponential tail of the MFs in Figure 3. These tails contain the most massive systems, whose emergent luminosities are most likely to be biased by opacity and thermal effects. Including these effects would probably correct the space density of massive systems toward the simulated MFs. Additionally, we note that the presented empirical MFs neglect mass measurement errors, which might have an important effect on the slope of the exponential tails. Another difference between observations and simulation are the spurious bumps in the low-mass range of the simulated MFs, i.e., $\log (M_{\rm H\,\hbox{\scriptsize {\sc i}}})\approx 8.5$ and $\log (M_{{\rm H}_2})\approx 8.0$, where the number density is about doubled compared to observations. This feature can also be seen in the optical bJ-band LF shown by Croton et al. (2006) and stems from an imprecision in the number density of the smallest galaxies in the DeLucia catalog, where the mass resolution of the Millennium Simulation implies a poorly resolved merger history. The overdensity of sources around this resolution limit roughly balances the mass of even smaller galaxies, i.e., $\log (M_{\rm H\,\hbox{\scriptsize {\sc i}}}/{M}_{\odot })\ll 8.0$, that are missing in the simulation.

The universal gas densities of the simulation, expressed relative to the critical density for closure, are $\Omega _{\rm H\,\hbox{\scriptsize {\sc i}}}^{\rm sim}=3.4\times 10^{-4}$ and $\Omega _{{\rm H}_2}^{\rm sim}=1.1\times 10^{-4}$, in good agreement with the observations $\Omega _{\rm H\,\hbox{\scriptsize {\sc i}}}^{\rm obs}=(3.6\pm 0.4)\times 10^{-4}$ (Zwaan et al. 2005) and $\Omega _{{\rm H}_2}^{\rm obs}=(0.95\pm 0.37)\times 10^{-4}$ (Obreschkow & Rawlings 2009).

Figure 4 shows our simulated H i MF and H2 MF together with the MF for the cold gas metals given in the original DeLucia catalog and the MF for He as trivially derived using Equation (2). This picture reveals that in the cold gas of the local universe He is probably more abundant than H2, but less abundant than H i.

Figure 4.

Figure 4. Simulated galaxy MFs for H i (solid line), H2 (dashed line), cold He (dotted line), and cold gas metals (dash-dotted line). The H i MF and H2 MF are identical to the solid lines in Figure 3.

Standard image High-resolution image

We shall now consider the H i masses in elliptical and spiral galaxies separately. This division is based on the Hubble type T, where we consider galaxies with T < 0 as "ellipticals" and galaxies with T ⩾ 0 as "spirals"—a rough separation that neglects other types like irregular galaxies as well as various subclassifications. In the simulation, T is computed from the bulge mass fraction according to Equation (18).

The simulated H i MFs of both elliptical and spiral galaxies are shown in Figure 5. In order to determine the observational counterparts, we split the HIPASS galaxy sample into elliptical and spiral galaxies according to the Hubble types provided in the HyperLeda reference database (Paturel et al. 2003). For both subsamples, the H i MF was evaluated using the 1/Vmax  method (Schmidt 1968), where Vmax  was estimated from the analytic completeness function for HIPASS, which characterizes the completeness of each source given its H i-peak flux density Sp and integrated H i-line flux Sint (Zwaan et al. 2004). In order to estimate the uncertainties of the MFs, we derived them for 104 random half-sized subsets of the HIPASS sample—a bootstrapping approach. The standard deviation of the 104 values of $\log (\phi _{\rm H\,\hbox{\scriptsize {\sc i}}})$ for each mass bin was divided by $\sqrt{2}$ to estimate the 1σ errors of $\log (\phi _{\rm H\,\hbox{\scriptsize {\sc i}}})$ for the full sample.

Figure 5.

Figure 5. H i MFs for elliptical and spiral galaxies. The solid line and dashed line, respectively, represent the simulated result, where the galaxies have been divided in ellipticals and spirals according to their Hubble type estimated using Equation (18). The filled and open dots with error bars represent the corresponding empirical H i MFs, which we derived from the HIPASS sample. The shaded zone represents the H i-mass range $M_{\rm H\,\hbox{\scriptsize {\sc i}}}\lesssim 10^9\;{M}_{\odot }$, approximately corresponding to the simulated galaxies with poorly resolved morphologies.

Standard image High-resolution image

Figure 5 demonstrates that our simulation successfully reproduces the H i masses of both spiral and elliptical galaxies for H i masses greater than ∼109M, although the nearly perfect match between simulation and observation may be somewhat coincidental due to the uncertainties of the Hubble types T calculated via Equation (18). For H i masses smaller than 109M, the morphological separation seems to breakdown (shaded zone in Figure 5). Indeed the H i-mass range $M_{\rm H\,\hbox{\scriptsize {\sc i}}}\lesssim 10^9\;{M}_{\odot }$ approximately corresponds to the stellar mass range Mstars ≲ 4 × 109M, for which morphology properties are poorly resolved (see Section 2).

3.4. Observable H i and CO Luminosities

The characteristic radio line of H i stems from the hyperfine energy level splitting of the hydrogen atom and lies at 1.42 GHz rest-frame frequency. The velocity-integrated luminosity of this line $L_{\rm H\,\hbox{\scriptsize {\sc i}}}$ can be calculated from the H i mass via

Equation (21)

Equation (21) neglects H i-self-absorption effects, but this is likely to be a problem only for the largest disk galaxies observed edge-on (Rao et al. 1995). The strict proportionality between $L_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ assumed in Equation (21) means that the H i LF is geometrically identical to the H i MF.

By contrast, the H2 masses used for the empirical H2 MFs in Figures 3 and 4 rely on measurements of the CO(1–0) line, i.e., the 115 GHz radio line stemming from the fundamental rotational relaxation of the most abundant CO isotopomer $\rm ^{12}C^{16}O$. Here, we only consider this line, but luminosities of other CO lines can be estimated using approximate empirical line ratios (e.g., Braine et al. 1993; Righi et al. 2008).

The CO(1–0)-to-H2 conversion generally depends on the galaxy and the cosmic epoch, and it is often represented by the dimensionless factor

Equation (22)

where $N_{{\rm H}_2}$ is the column density of H2 molecules and ICO is the integrated CO(1–0)-line intensity per unit surface area defined via the surface brightness temperature in the Rayleigh–Jeans approximation. The definition of Equation (22) implies the mass–luminosity relation (e.g., review by Young & Scoville 1991)

Equation (23)

where LCO is the velocity-integrated luminosity of the CO(1–0) line.

As discussed in Obreschkow & Rawlings (2009), the theoretical and observational determination of the X-factor is a subtle task with a long history. Most present-day studies assume a constant X-factor Xc, such as

Equation (24)

which is typical for spiral galaxies in the local universe (Leroy et al. 2008). By contrast, Arimoto et al. (1996) and Boselli et al. (2002) suggested that X is variable, Xv, and approximately inversely proportional to the metallicity O/H, i.e., the ratio between the number of oxygen ions and hydrogen ions in the hot ISM. Using their data, we found that (Obreschkow & Rawlings 2009)

Equation (25)

At first sight, the empirical negative dependence of X on the metallicity seems to contradict the fact that $\rm ^{12}C^{16}O$ is optically thick for 115 GHz radiation. Indeed, the radiated luminosity should not depend on the density of metals, as long as the latter is high enough for the radiation to remain optically thick (Kutner & Leung 1985). However, detailed theoretical investigations (e.g., Maloney & Black 1988) of the sizes and temperatures of molecular clumps were indeed able to explain, and in fact predict, the negative dependence of X on metallicity.

Equation (25) links the metallicity of the hot ISM to the X-factor of cold molecular clouds, and it is likely a consequence of a more fundamental relation between cold gas metallicity and X. To uncover such a relation, we assume that the O/H metallicity of the cold ISM in local galaxies is approximated by O/H of the hot ISM. Given an atomic mass of 16 for oxygen, the fact that hydrogen makes up a fraction 0.74 of the total baryon mass, and assuming that oxygen accounts for a fraction of 0.4 of the mass of all metals (based on Arnett 1996; Kobulnicky & Zaritsky 1999), Equation (25) translates to

Equation (26)

where Mgas is the total cold gas mass and MZ is the mass of metals in cold gas. Equation (26) only relates cold gas properties to each other and therefore is more fundamental than Equation (25).

To evaluate Equation (26) for each galaxy in the simulation, we used the cold gas metal masses MZ given in the DeLucia catalog. Those masses are reasonably accurate as demonstrated by De Lucia et al. (2004) and Croton et al. (2006) through a comparison of the simulated stellar mass–metallicity relation to the empirical mass–metallicity relation obtained from 53,000 star-forming galaxies in the SDSS (Tremonti et al. 2004). For most galaxies at z = 0, the simulation yields metal fractions MZ/Mgas ≈ 0.01–0.04 in the local universe, thus implying Xv ≈ 1–4 in agreement with observed values (e.g., Boselli et al. 2002).

Figure 6 displays the simulated CO LF for the variable X-factor Xv (solid line) and the constant X-factor Xc (dashed line) together with the empirical CO LF (Keres et al. 2003), adjusted to h = 0.73. The comparison supports the variable X-factor of Equation (26) against Xc = 2 (and the same conclusion is found for other constant values of Xc). Using Equation (26) also has the advantage that the cosmic evolution of the X-factor due to the evolution of metallicity is implicitly accounted for. Nevertheless, Equation (26) may not be appropriate at high redshift as discussed in Section 6.3.

Figure 6.

Figure 6. LF of CO(1–0)-emission (CO LF) in the local universe. The solid line represents the simulated CO LF, obtained using the variable conversion factor Xv of Equation (26), and the dashed line represents the CO LF, obtained using the constant conversion factor Xc of Equation (24). The square dots and error bars represent the empirical CO LF determined by Keres et al. (2003).

Standard image High-resolution image

4. COLD GAS DISK SIZES

Using the axially symmetric surface density profiles for H i and H2 given in Equations (11) and (12), we can define the H i radius $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and H2 radius $r_{\rm H_2}$ of an axially symmetric galaxy as the radii corresponding to a detection limit Σ0, i.e.,

Equation (27)

Equation (28)

In this paper, we chose Σ0 = 1 M pc−2, corresponding to the deep survey of the Ursa Major group by, e.g., Verheijen (2001), but any other value could be adopted. In general, Equations (27) and (28) do not have explicit closed-form solutions and must be solved numerically for each galaxy.

Results for $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $r_{\rm H_2}$ at three epochs are displayed in Figure 7. Each graph shows 103 simulated galaxies, drawn randomly from the catalog with a probability proportional to their cold gas mass. This selection rule ensures that rare objects at the high end of the MF are included. The arithmetic average of the points in each graph can be interpreted as the cold gas mass-weighted average of the displayed quantities. This average is marked in each graph to emphasize changes with redshift. The data in Figure 7(a) are shown again in Figure 8 together with measurements of 39 spiral galaxies in the Ursa Major group (Verheijen 2001).

Figure 7.

Figure 7. Simulated mass–radius relations for H i and H2 at redshifts z = 0,  4.89,  10.07, corresponding to the simulation snapshots 63, 21, 12. The black dots represent 103 simulated galaxies and the solid lines show the power-law regression for the data in Figure 7(a) (i.e., H i at z = 0). The red crosses represent the cold gas mass-weighted averages of ($M_{\rm H\,\hbox{\scriptsize {\sc i}}}$, $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$) and ($M_{{\rm H}_2}$, $r_{\rm H_2}$) in the simulation at each of the three redshifts.

Standard image High-resolution image
Figure 8.

Figure 8. Relation between H i mass $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and H i radius $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ for galaxies at redshift z = 0. The black dots represent 103 simulated galaxies and the solid line their linear regression. The simulated data are identical to those plotted in Figure 7(a). The red squares show measurements in the Ursa Major group by Verheijen (2001), who used the same definition of $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ as this paper.

Standard image High-resolution image

Figures 7 and 8 reveal several features, which we shall discuss hereafter: (1) the mass–radius relation for H i is a nearly perfect power law with surprisingly small scatter; (2) in general, radii become smaller with increasing redshift; (3) the evolution of the mass–radius relation is completely different for H i and H2.

The first result, i.e., the strict power-law relation between $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$, is strongly supported by measurements in the Ursa Major group (Verheijen 2001; see Figure 8). The best power-law fit to the simulation is

Equation (29)

The rms scatter of the simulated data around Equation (29) is σ = 0.03 in log space, while the rms scatter of the observations by Verheijen (2001) is σ = 0.06. This small scatter is particularly surprising as the more fundamental relation between Mbary and $\tilde{r}_{\rm disk}$ shown in Figure 2 exhibits a much larger scatter of σ = 0.26. The square-law form of the power law in Equation (29) implies that the average H i surface density inside the radius $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ is nearly identical for all galaxies, which have most of their H i mass inside the radius $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$,

Equation (30)

The existence of such a constant average density of H i can to first order be interpreted as a consequence of the fact that H i transforms into H2 and stars as soon as its density and pressure are raised. In fact, observations show that $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}$ saturates at about $6\hbox{--}10\,{M}_{\odot }\;\rm pc^{-2}$ (Blitz & Rosolowsky 2006; Leroy et al. 2008) and that higher cold gas densities are generally dominated by $\Sigma _{\rm H_2}$. Therefore, H i maintains a constant surface density during the evolution of any isolated galaxy as long as enough H i is supplied from an external source, e.g., by cooling from a hot medium as assumed in the recipes of the DeLucia catalog. This also explains why the power-law relation between $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ remains nearly constant toward higher redshift in the simulation (Figures 7(a)–(c)). About 1% of the simulated galaxies at redshift z = 0 lie far off the power-law relation (i.e., are outside 5σ of the best fit), typically toward smaller radii (see Figure 8). One might first expect that these objects have a higher H i surface density, while, in fact, the contrary applies. These galaxies have very flat H i profiles with most of the H i mass lying outside the radius $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$, and therefore they would require a lower sensitivity limit than $1\,{M}_{\odot }\;\rm pc^{-2}$ for a useful definition of $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$. Such galaxies are indeed very difficult to map due to observational surface brightness limitations.

The radii $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $r_{\rm H_2}$ become smaller toward higher redshift. This is a direct consequence of the cosmic evolution of the virial radii rvir of the halos in the Millennium Simulation, which affects the disk scale radius $\tilde{r}_{\rm disk}$ in Equation (16). As shown by Mo et al. (1998), rvir scales as (1 + z)−1.5 for a fixed circular velocity or as (1 + z)−1 for a fixed halo mass, consistent with high-redshift observations (z = 2.5–6) in the Hubble Ultra Deep Field (UDF) by Bouwens et al. (2004). Their selection criteria include all but the reddest starburst galaxies in the UDF and some evolved galaxies.

In analogy to the mass–radius power-law relation for H i, our simulation predicts a similar relation, again nearly a square law, for H2 at redshift z = 0 (see Figure 7(d)). This power law is consistent with observations of the two face-on spiral galaxies M 51 (Schuster et al. 2007) and NGC 6946 (Crosthwaite & Turner 2007). For the small H2/H i ratios found in the local universe the $M_{{\rm H}_2}$– $r_{\rm H_2}$ relation is linked to the $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$– $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ relation, because both $M_{{\rm H}_2}$ and $r_{\rm H_2}$ can be regarded as a fraction (<1) of, respectively, $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$. However, there is no fundamental reason for a constant surface density of H2 and the smaller sizes of high-redshift galaxies implies a higher pressure of the ISM and thus a much higher molecular fraction by virtue of Equations (10) and (15). Therefore, H2 masses become uncorrelated to H i and tend to increase with redshift out to z ≈ 5, while $r_{\rm H_2}$ decreases. Hence, the $M_{{\rm H}_2}$– $r_{\rm H_2}$ relation must move away from the power law $M_{{\rm H}_2}\sim r_{\rm H_2}^2$ found at z = 0 (see Figures 7(d)–(f)).

5. REALISTIC VELOCITY PROFILES

In this section, we derive circular velocity profiles and atomic and molecular radio line profiles for the simulated galaxies in the DeLucia catalog. Circular velocity profiles Vc(r) for various galaxies are derived over the Sections 5.15.3 and transcribed to radio line profiles for edge-on galaxies in Section 5.4. Results for the local and high-redshift universe are presented in Section 5.5.

5.1. Velocity Profile of a Spherical Halo

To account for the narrowness of the emission lines observed in the central gas regions of many galaxies (e.g., Sauty et al. 2003), we require a halo model with vanishing velocity at the center, as opposed to, for example, the commonly adopted singular isothermal sphere with a density ρhalo(r) ∼ r−2 and a constant velocity profile. We chose the Navarro–Frenk–White (NFW; Navarro et al. 1995, 1996) model, which relies on high-resolution numerical simulations of dark matter halos in equilibrium. These simulations revealed that halos of all masses in a variety of dissipationless hierarchical clustering models are well described by the spherical density profile

Equation (31)

where ρ0 is a normalization factor and rs is the characteristic scale radius of the halo. This profile is also supported by the Hubble Space Telescope analysis of the weak lensing induced by the galaxy cluster MS 2053-04 at redshift z = 0.58 (Hoekstra et al. 2002). ρhalo(r) varies as r−1 at the halo center and continuously steepens to r−3 for r. It passes through the equilibrium profile of the self-gravitating isothermal sphere, i.e., ρhalo(r) ∼ r−2, at r = rs.

The definition of rvir in the Millennium Simulation given in Equation (17) implies that

Equation (32)

where chalorvir/rs is referred to as the halo concentration parameter. Most numerical models predict that chalo scales with the virial mass Mvir, defined as the mass inside the radius rvir, according to a power law (e.g., Navarro et al. 1997; Bullock et al. 2001; Dolag et al. 2004; Hennawi et al. 2007). Here, we shall use the result of Hennawi et al. (2007),

Equation (33)

which is consistent with recent empirical values of the matter concentration in galaxy clusters derived from X-ray measurements and strong lensing data (Comerford & Natarajan 2007).

For a spherical halo, the circular velocity profile is given by Vhaloc2(r) = GMhalo(r)/r with $M_{\rm halo}(r)=4\pi \int _0^rd\tilde{r}\,\tilde{r}^2\,\rho _{\rm halo}(\tilde{r})$. Using Equations (31) and (32), this implies that

Equation (34)

where xr/rvir (thus chalox = r/rs). This velocity vanishes at the halo center, then climbs to a maximal value $V_{\max }=1.65\,r_{s}\sqrt{G\rho _0}$ at r = 2.16 rs, from where it decreases monotonically with r, typically reaching 0.65–0.95Vmax  at r = rvir (x = 1) with the extremes corresponding, respectively, to chalo = 25 and chalo = 5. For larger radii, the velocity asymptotically approaches the point-mass velocity profile Vhaloc2(r) = GMvir/r.

5.2. Velocity Profile of a Flat Disk

For simplicity, we assume in this section that the galactic disk is described by a single exponential surface density for stars and cold gas,

Equation (35)

where Mdisk is the total disk mass, taken as the sum of the cold gas mass and the stellar mass in the disk. In most real galaxies, the stellar surface densities are slightly more compact (see Section 3.1), but we found that including this effect does not significantly modify the shape of the atomic and molecular emission lines. In fact, the radius, which maximally contributes to the disk mass, i.e., the maximum of r Σdisk(r), is r = rdisk. Therefore, we expect the gravitational potential to differ significantly from the point-mass potential only for r of order rdisk or smaller. Applying Poisson's equation to the surface density of Equation (35), the gravitational potential in the plane of the disk becomes

Equation (36)

where the integration surface D is given by $\tilde{r}\in [0,\infty)$, θ = [0, 2π).

The velocity profile for circular orbits in the plane of the disk can be calculated as Vdiskc2 = rdφdisk/dr. The integral in Equation (36) is elliptic, and hence there are no exact closed-from expressions for φdisk and Vdiskc. However, in this study we numerically found that an excellent approximation is given by

Equation (37)

where cdiskrvir/rdisk is the disk concentration parameter in analogy to the halo concentration parameter chalo of Section 5.1. Equation (37) is accurate to less than 1% over the whole range r = 0–10 rdisk and it correctly converges toward the circular velocity of a point-mass potential, Vhaloc2(r) = GMdisk/r, for r.

Like in Section 3.2, we shall use Equation (16) to compute rdisk and cdisk in the DeLucia catalog. This approach is slightly inconsistent because Equation (16) was derived by Mo et al. (1998) under the assumption that the disk is supported by an isothermal halo with ρhalo(r) ∼ r−2, while in Section 5.1 we have assumed the more complex NFW profile. We argue, however, that Equation (16) with the empirical correction of Equation (20) is sufficiently accurate, as it successfully reproduces the observed relation between Mbary and rdisk (see Figure 2) as well as the relation between $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ (see Figure 8). It can also be shown that, for realistic values of the halo concentration chalo (10–20 for one-galaxy systems) and the spin parameter λ (0.05–0.1, e.g., Mo et al. 1998), the scale radius of the halo rs and the disk radius rdisk are similar. Hence, the main mass contribution of the disk comes indeed from galactocentric radii, where the halo profile is approximately isothermal, thus justifying the assumption made by Mo et al. (1998) to derive Equation (16).

5.3. Velocity Profile of the Bulge

Many models for the surface brightness or surface density profiles of bulges have been proposed (e.g., overview by Balcells et al. 2001). A rough consensus seems established that no single surface density profile can describe a majority of the observed bulges, but that they are generally well matched by the class of Sersic functions (Sersic 1968), Σbulge(r) ∼ exp [ − (r/rbulge)1/n], where the exponent n depends on the morphological type (Andredakis et al. 1995), such that n ≈ 4 for lenticular/early-type galaxies (de Vaucouleurs profile) and n ≈ 1 for the bulges of late-type galaxies (exponential profile). Courteau et al. (1996) find slightly steeper profiles with n = 1–2 for nearly all spirals in a sample of 326 spiral galaxies using deep optical and IR photometry, and they show that by imposing n = 1 for all late-type galaxies, the ratio between the exponential scale radius of the bulge rbulge and the scale radius of the disk rdisk is roughly constant, $r_{\rm bulge}\approx 0.1\,\tilde{r}_{\rm disk}\approx 0.05\,r_{\rm disk}$. We shall therefore assume that all bulges have an exponential projected surface density,

Equation (38)

with rbulge = 0.05 rdisk.

For simplicity, we assume that the bulges of all galaxies are spherical and thus described by a radial space density function ρbulge(r). This function is linked to the projected surface density via Σbulge(r) = ∫dz ρbulge[(r2 + z2)1/2]. Numerically, we find that this model for ρbulge(r) is closely approximated by the Plummer model (Plummer 1911), more often used in the context of clusters,

Equation (39)

with a characteristic Plummer radius rPlummer ≈ 1.7 rbulge. The circular velocity profile Vbulgec corresponding to Equation (39) is given by Vbulgec2(r) = GMbulge(r)/r with $M^{\rm bulge}(r)=4\pi \int _0^r{ d}\tilde{r}\,\tilde{r}^2\,\rho _{\rm bulge}(\tilde{r})$. This solves to

Equation (40)

where cbulgervir/rPlummer ≈ 12 cdisk is the bulge concentration parameter.

5.4. Line Shapes from Circular Velocities

The addition rule for gravitational potentials implies that the circular velocity profile of the combined halo–disk–bulge system in the plane of the disk is given by

Equation (41)

where xr/rvir as in Sections 5.15.3. According to Equations (34), (37), and (40), this profile is determined by six parameters: the three form parameters chalo, cdisk, and cbulge and the three mass scales Mvir/rvir, Mdisk/rvir, and Mbulge/rvir. The form parameters were calculated as explained in Sections 5.15.3, while the mass scales were directly adopted from the DeLucia catalog. For the satellite galaxies with no resolved halo (see Section 2), Mvir and rvir were approximated as the corresponding quantities of the original galaxy halo just before its disappearance. An exemplar circular velocity profile for a galaxy in the DeLucia catalog at redshift z = 0 is shown in Figure 9.

Figure 9.

Figure 9. Circular velocity profile of a typical simulated galaxy with a small bulge at redshift z = 0. The total circular velocity (solid line) is given by the circular velocity of the halo (dashed line), the disk (dash-dotted line), and the bulge (dotted line) via Equation (41).

Standard image High-resolution image

In order to evaluate the profile of a radio emission line associated with any velocity profile Vc(r), we shall first consider the line profile of a homogeneous flat ring with constant circular velocity Vc and a total luminosity of unity. If a point of the ring is labeled by the angle γ it forms with the line of sight (see Figure 10), the apparent projected velocity of that point is given by Vobs = Vc sin γ. The ensemble of all angles γ ∈ [0, 2π) therefore spans a continuum of apparent velocities Vobs ∈ (−Vc, Vc) with a luminosity density distribution $\tilde{\psi }(V_{\rm obs},V_{c})\sim { d}\gamma /{ d}V_{\rm obs}$. Imposing the normalization condition $\int { d}V_{\rm obs}\tilde{\psi }(V_{\rm obs})=1$, we find that the edge-on line profile of the ring is given by

Equation (42)

This profile exhibits spurious divergent singularities at |Vobs| → Vc, which, in reality, are smoothed by the random, e.g., turbulent, motion of the gas. We assume that this velocity dispersion is given by the constant $\sigma _{\rm gas}=8\rm \;km\;s^{-1}$, which is consistent with the velocity dispersions observed across the disks of several nearby galaxies (e.g., Shostak & van der Kruit 1984; Dickey et al. 1990; Burton 1971). The smoothed velocity profile is then given by

Equation (43)

which conserves the normalization ∫dVobsψ(Vobs) = 1. Some examples of the functions $\tilde{\psi }(V_{\rm obs},V_{c})$ and ψ(Vobs, Vc) are plotted in Figure 11.

Figure 10.

Figure 10. Apparent velocity Vobs induced by the infinitesimal ring element at the angle γ.

Standard image High-resolution image
Figure 11.

Figure 11. Illustration of the functions ψ (Equation (43)) and $\tilde{\psi }$ (Equation (42)), which represent the normalized emission line of a homogeneous edge-on disk or ring with constant circular velocity.

Standard image High-resolution image

From the edge-on line profile ψ(Vobs, Vc) of a single ring and the face-on surface densities of atomic and molecular gas, $\Sigma _{\rm H\,\hbox{\scriptsize {\sc i}}}(r)$ and $\Sigma _{\rm H_2}(r)$, we can now evaluate the edge-on profiles of emission lines associated with the entire H i and H2 disks, respectively. Since H2 densities are most commonly inferred from CO detections, we shall hereafter refer to all molecular emission lines as "the CO line." The edge-on line profiles (or "normalized luminosity densities") $\Psi _{\rm H\,{\mathsc i}}(V_{\rm obs})$ and ΨCO(Vobs) are given by

Equation (44)

Equation (45)

These two functions satisfy the normalization conditions $\int { d}V_{\rm obs}\Psi _{\rm H\,{\mathsc i}}(V_{\rm obs})=1$ and ∫dVobsΨCO(Vobs) = 1. To obtain intrinsic luminosity densities, $\Psi _{\rm H\,{\mathsc i}}(V_{\rm obs})$ must be multiplied by the integrated luminosity of the H i line (see Equation (21)) and ΨCO(Vobs) must be multiplied by the integrated luminosity of the considered molecular emission line, e.g., the integrated luminosity of the CO(1–0) line given in Equation (23).

Figure 12 displays the line profiles $\Psi _{\rm H\,{\mathsc i}}(V_{\rm obs})$ and ΨCO(Vobs) for the exemplar galaxy with the velocity profile shown in Figure 9. All line profiles produced by our model are mirror symmetric, but they can, in principle, differ significantly from the basic double-horned function ψ(Vobs). Especially for CO, where the emission from the bulge can play an important role, several local maxima can sometimes be found in the line profile, in qualitative agreement with various observations (e.g., Lavezzi & Dickey 1998).

Figure 12.

Figure 12. Simulated edge-on H i- and CO-emission lines for the exemplar galaxy, for which the circular velocity profile is shown in Figure 9. The line profiles have been computed using Equation (44).

Standard image High-resolution image

5.5. Results and Discussion

For every galaxy in the DeLucia catalog, we computed the edge-on line profiles $\Psi _{\rm H\,\hbox{\scriptsize {\sc i}}}(V_{\rm obs})$ and ΨCO(Vobs), from which we extracted the line parameters indicated in Figure 12. $\Psi _{\rm H\,\hbox{\scriptsize {\sc i}}}^0\equiv \Psi _{\rm H\,\hbox{\scriptsize {\sc i}}}(0)$ and Ψ0CO ≡ ΨCO(0) are the luminosity densities at the line center, and $\Psi _{\rm H\,\hbox{\scriptsize {\sc i}}}^{\rm max}$ and ΨmaxCO are the peak luminosity densities, i.e., the absolute maxima of $\Psi _{\rm H\,\hbox{\scriptsize {\sc i}}}(V_{\rm obs})$ and ΨCO(Vobs). $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{\rm peak}$ and wpeakCO are the line widths measured between the left and the right maximum. These values vanish if the line maxima are at the line center, such as found, for example, in slowly rotating systems. $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{50}$, w50CO, $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{20}$, and w20CO are the line widths measured at, respectively, the 50 percentile level or the 20 percentile level of the peak luminosity densities—the two most common definitions of observed line widths.

We shall now check the simulated line widths against observations by analyzing their relation to the mass of the galaxies. Here, we shall refer to all line width versus mass relations as TFRs, since they are generalized versions of the original relation between line widths and optical magnitudes of spiral galaxies (Tully & Fisher 1977). A variety of empirical TFRs have been published, such as the stellar mass TFR and the baryonic TFR (McGaugh et al. 2000). The latter relates the baryon mass (stars + gas) of spiral disks to their line widths (or circular velocities) and is probably the most fundamental TFR detected so far, obeying a single power law over 5 orders of magnitude in mass. We have also investigated the less fundamental empirical TFR between $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{50}$—hereafter the H i-TFR—using the spiral galaxies of the HIPASS catalog. Assuming no prior knowledge on the inclinations of the HIPASS galaxies, but taking an isotropic distribution of their axes as given, we found that the most-likely relation is

Equation (46)

for the Hubble parameter h = 0.73. Relative to Equation (46), the HIPASS data exhibit a Gaussian scatter with σ = 0.38 in $\log (M_{\rm H\,\hbox{\scriptsize {\sc i}}})$. Our method to find Equation (46) will be detailed in a forthcoming paper (D. O. Obreschkow et al. 2009, in preparation), especially dedicated to the H i-TFR.

Figures 13(a)–(d) show four TFRs at redshift z = 0. Each figure represents 103 simulated galaxies (black), randomly drawn from the simulation with a probability proportional to their cold gas mass in order to include the rare objects in the high end of the MF. Spiral and elliptical galaxies are distinguished as black dots and open circles.

Figure 13.

Figure 13. Relations between edge-on line widths and different mass tracers for the local universe. 103 simulated galaxies are represented by black dots (spiral galaxies) and black circles (elliptical galaxies). The solid lines represent power-law fits to the simulated spiral galaxies. In case of Figure 13(b), this fit only includes galaxies with Mstars>109M. Figure 13(d) does not include satellite galaxies without halos (see Section 2), for which Mvir is poorly defined. The dashed red line and rose-shaded zone in Figure 13(a) represent our observational determination and 1σ scatter of the H i-TFR from the HIPASS data (see Section 5.5). The rose dots and dashed lines in Figures 13(b) and (c) are the observational data and power-law regressions from McGaugh et al. (2000) and references therein.

Standard image High-resolution image

Figure 13(a) shows the simulated H i-TFR together with the empirical counterpart given in Equation (46). This comparison reveals good consistency between observation and simulation for spiral galaxies. However, the elliptical galaxies lie far off the H i-TFR. In fact, simulated elliptical galaxies generally have a significant fraction of their cold hydrogen in the molecular phase, consistent with the galaxy-type dependence of the H2/H i ratio first identified by Young & Knezek (1989). Therefore, H i is a poor mass tracer for elliptical galaxies, both in simulations and observations, leading to their offset from the TFR when only H i masses are considered. There seems to be no direct analog to the H i-TFR for elliptical galaxies.

Figures 13(b) and (c), respectively, display the simulated stellar mass TFR and the baryonic TFR, together with the observed data of McGaugh et al. (2000) corrected for h = 0.73. These data include various galaxies from dwarfs to giant spirals, whose edge-on line widths were estimated from the observed ones using the inclinations determined from the optical axis ratios. Figures 13(b) and (c) reveal a surprising consistency between simulation and observation. In Figure 13(b), both the simulated and observed data show a systematic offset from the power-law relation for all galaxies with $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{20}\lesssim 200\rm \;km\;s^{-1}$. Yet, the power-law relation is restored as soon as the cold gas mass is added to the stellar mass (Figure 13(c)), thus confirming that the TFR is indeed fundamentally a relation between mass and circular velocity.

It is interesting to consider the prediction of the simulation for the most fundamental TFR, i.e., the one between the total dynamical mass, taken as the virial mass Mvir, and the circular velocity, represented by the line width $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{20}$. This relation is shown in Figure 13(d) and reveals indeed a two to three times smaller scatter in $\log (\rm mass)$ than the baryonic TFR, hence confirming its fundamental character.

Although the simulated elliptical galaxies shown in Figures 13(b)–(d) roughly align with the respective TFRs for spiral galaxies, their scatter is larger. This is caused by the mass domination of the bulge, which leads to steep circular velocity profiles Vc(r) with a poorly defined terminal velocity. Therefore, line widths obtained by averaging over the whole elliptical galaxy are weak tracers of its spin. This picture seems to correspond to observed elliptical galaxies, where the central line widths, corresponding to the velocity dispersion in the bulge-dominated parts, are more correlated to the stellar mass than the line widths of the whole galaxy (see Faber–Jackson relation, e.g., Faber & Jackson 1976).

Kassin et al. (2007) noted that S0.5 ≡ (0.5 V2c + σ2gas)1/2 is a better kinematic estimator than the circular velocity Vc, in the sense that it markedly reduces the scatter in the stellar mass TFR. However, since our model assumes a constant gas velocity dispersion σgas for all galaxies, it is not possible to investigate this estimator.

The predicted evolution of the four TFRs in Figures 13(a)–(d) is shown in Figures 14(a)–(d). In all four cases, the simulation predicts two important features: (1) galaxies of identical mass have broader lines (and larger circular velocities) at higher redshift, and (2) the scatter of the TFRs generally increases with redshift. The first feature is mainly a consequence of the mass–radius–velocity relation of the dark matter halos assumed in the Millennium Simulation (see Croton et al. 2006; Mo et al. 1998). This relation predicts that, given a constant halo mass, Vc scales as (1 + z)1/2 for large z. Furthermore, the ratios $M_{\rm H\,\hbox{\scriptsize {\sc i}}}/M_{\rm vir}$ and Mstars/Mvir on average decrease with increasing redshift, explaining the stronger evolution found in Figures 14(a) and (b) relative to Figures 14(c) and (d).

Figure 14.

Figure 14. Simulated cosmic evolution of the different line width–mass relations shown in Figure 13. Spiral and elliptical galaxies are, respectively, represented by dots and circles. The black color corresponds to redshift z = 0 (identically to Figure 13), while blue and red color, respectively, represent z = 4.89 and z = 10.07. The solid black lines are power-law fits to the spiral galaxies at z = 0, where in case of Figure 14(b) only galaxies with Mstars>109M have been considered. The number of elliptical galaxies decreases with redshift—a consequence of the merger-driven and instability driven prescriptions for bulge formation in the DeLucia catalog.

Standard image High-resolution image

The increase of scatter in the TFRs with redshift is a consequence of the lower degree of virialization at higher redshifts, which, in the model, is accounted for via the spin parameter λ of the halos. λ is more scattered at high redshift due to the young age of the halos and the higher merger rates. More scatter in λ leads to more scatter in the radius rdisk via Equation (16) and thus to more scatter in the circular velocity Vc via Equations (37) and (40).

Current observational databases of resolved CO-line profiles are much smaller than H i databases and their signal/noise characteristics are inferior. Nevertheless efforts to check the use of CO-line widths for probing TFRs (e.g., Lavezzi & Dickey 1998) have led to the conclusion that in most spiral galaxies the CO-line widths are very similar to H i-line widths, even though the actual line profiles may radically differ. Figure 15 shows our simulated relation between $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{20}$ and w20CO, as well as the linear fit to observations of 44 galaxies in different clusters (Lavezzi & Dickey 1998). These observations are indeed consistent with the simulation. The simulated elliptical galaxies tend to have higher $w_{\rm CO}^{20}/w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{20}$ ratios than the spiral ones, due to fast circular velocity of the bulge component implied by its own mass.

Figure 15.

Figure 15. Relation between line widths of H i and CO. 103 simulated galaxies are represented by black dots (spiral galaxies) and black circles (elliptical galaxies). The red dashed line and rose-shaded zone represent the best fit and its 1σ confidence interval to observations of 44 galaxies in clusters presented by Lavezzi & Dickey (1998).

Standard image High-resolution image

The line profiles and widths studied in this section correspond to galaxies observed edge-on. First-order corrections for spiral galaxies seen at an inclination i ≠ 90° can be obtained by dividing the normalized luminosity densities $\Psi _{\rm H\,\hbox{\scriptsize {\sc i}}}^0$, $\Psi _{\rm H\,\hbox{\scriptsize {\sc i}}}^{\rm max}$, Ψ0CO, and ΨmaxCO by sin  i, while multiplying the line widths $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{50}$, $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{20}$, $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{\rm peak}$, w50CO, w20CO, and wpeakCO by sin  i. More elaborate corrections, accounting for the isotropy of the velocity dispersion σgas, were given by Makarov et al. (1997) and Kannappan et al. (2002).

6. DISCUSSION

We used a list of physical prescriptions to postprocess the DeLucia catalog and showed that many simulation results match the empirical findings from the local universe. However, this approach raises two major questions. (1) Are the applied prescriptions consistent with the DeLucia catalog in the sense that they represent a compatible extension of the semianalytic recipes used by De Lucia & Blaizot (2007) and Croton et al. (2006)? (2) What are the limitations of our prescriptions at low and high redshifts?

6.1. Consistency of the Model

The consistency question arises, because the DeLucia catalog relies on a simplified version of a Schmidt–Kennicutt law (Schmidt 1959; Kennicutt 1998), i.e., a prescription where the SFR scales as some power of the surface density of the ISM. However, in a smaller-scaled picture, new stars are bred inside molecular clouds, and hence it must be verified whether our prescription to assign H2 masses to galaxies is compatible with the macroscopic Schmidt–Kennicutt law. Our prescription exploited the empirical power law between the pressure of the ISM and its molecular content, as first presented by Blitz & Rosolowsky (2004, 2006). Based on this power-law relation, Blitz & Rosolowsky (2006) themselves formulated an alternative model for the computation of SFRs in galaxies, which seems more fundamental than the Schmidt–Kennicutt law. Applying both models for star formation to six molecule-rich galaxies in the local universe, they showed that their new pressure-based law predicts SFRs nearly identical to the ones predicted by the Schmidt–Kennicutt law. Therefore, our choice to divide cold hydrogen in H i and H2 according to pressure is indeed consistent with the prescription for SFRs used by De Lucia & Blaizot (2007) and Croton et al. (2006).

6.2. Accuracy and Limitations at z = 0

A first limitation of our simulation comes from the assumption that the surface densities of H i and H2 are axially symmetric (no spiral structures, no central bars, no warps, no satellite structures). In general, our models describe all galaxies as regular galaxies—as do all semianalytic models for the Millennium Simulation. Hence, the simulation results cannot be used to predict the H i and H2 properties of irregular galaxies.

While our models allowed us to reproduce the observed relation between $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ remarkably well for various spiral galaxies (e.g., Figure 8), it tends to underestimate the size of H i distributions in elliptical galaxies. For example, observations by Morganti et al. (1997) show that seven nearby E- and S0-type galaxies all have very complex H i distributions, often reaching far beyond the corresponding radius of a mass-equivalent disk galaxies. The patchy H i distributions found around elliptical galaxies are probably due to mergers and tidal interactions, which could not be modeled in any of the semianalytic schemes for the Millennium Simulation.

Another limitation arises from neglecting stellar bulges as an additional source of disk pressure in Equation (7) (Elmegreen 1989). Especially, the heavier bulges of early-type spiral galaxies could introduce a positive correction of the central pressure and hence increase the molecular fraction, thus leading to very sharp H2 peaks in the galaxy centers, such as observed, for example, in the SBb-type spiral galaxy NGC 3351 (Leroy et al. 2008). Our model for the H2 surface density of Equation (12) fails at predicting such sharp peaks, although the predicted total H i and H2 mass and the corresponding radii and line profiles are not significantly affected by this effect.

6.3. Accuracy and Limitations at z>0

Additional limitations are likely to occur at higher redshifts, where our models make a number of assumptions based on low-redshift observations. Furthermore, the underlying DeLucia catalog itself may suffer from inaccuracies at high redshift, but we shall restrict this discussion to possible issues associated with the models in this paper.

Regarding the subdivision of hydrogen into atomic and molecular material (Section 3), our most critical assumption is the treatment of all galactic disks as regular exponential structures in hydro-gravitational equilibrium. This model is very likely to deviate more from the reality at high redshift, where galaxies were generally less virialized and mergers were much more frequent (de Ravel et al. 2008). Less virialized disks are thicker, which would decrease the average pressure and fraction of molecules compared to our model. Yet, galaxy mergers counteract this effect by creating complex shapes with locally increased pressures, where H2 forms more efficiently, giving rise to merger-driven starbursts. Therefore, it is unclear whether the assumption of regular disks tends to underestimate or overestimate the H2/H i ratios.

Another critical assumption is the high-redshift validity of the local relation between the H2/H i ratio Rmol and the external gas pressure P (Equation 6). This relation is not a fundamental thermodynamic relation, but represents the effective relation between the average H2/H i ratio and P, resulting from complex physical processes such as cloud formation, H2 formation on metallic grains, and H2 destruction by the photodissociative radiation field of stars and supernovae. Therefore, the RmolP relation could be subjected to a cosmic evolution resulting from the cosmic evolution of the cold gas metallicity or the cosmic evolution of the photodissociative radiation field. However, the metallicity evolution is likely to be problem only at the highest redshifts (z ≳ 10). Observations in the local universe show that spiral galaxies with metallicities differing by a factor 5 fall on the same RmolP relation (Blitz & Rosolowsky 2006). Yet, the average cold gas metallicity of the galaxies in the DeLucia catalog is only a factor 1.9 (3.6) smaller at z = 5 (z = 10) than in the local universe. These predictions are consistent with observational evidence from the SDSS that stellar metallicities were at most a factor 1.5–2 smaller at z ≈ 3 than today (Panter et al. 2008). The effect of the cosmic evolution of the photodissociative radiation field on the RmolP relation is difficult to assess. Blitz & Rosolowsky (2006) argued that the ISM pressure and the radiation field both correlate with the surface density of stars and gas, and therefore the radiation field is correlated to pressure. This is supported by observations in the local universe showing that the RmolP relation holds true for dwarf galaxies and spiral galaxies spanning almost 3 orders of magnitude in SFR. For those reasons, the RmolP relation could indeed extend surprisingly well to high redshifts.

In the expression for the disk pressure in Equation (7) (Elmegreen 1989), we assumed a constant average velocity dispersion ratio 〈fσ〉. Observations suggest that Vcgas decreases significantly with redshift (Förster Schreiber et al. 2006; Genzel et al. 2008), and therefore 〈fσ〉 perhaps increases. This would lead to even higher H2/H i ratios than predicted by our model. However, according to Equation (10) this is likely to be a problem only for galaxies with Mstars>Mgas, while most galaxies in the simulation at z>2 are indeed gas dominated.

Regarding cold gas geometries and velocity profiles, the most important limitation of our model again arises from the simplistic treatment of galactic disks as virialized exponential structures. Very young galaxies (≲$10^8\,\rm yr$) or galaxies undergoing a merger do not conform with this model, and therefore the predicted velocity profiles may be unreal and the disk radii may be meaningless. This is not just a limitation of the simulation, but it reveals a principal difficulty to describe galaxy populations dominated by very young or merging objects with quantities such as $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ or $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{50}$, which are common and useful for isolated systems in the local universe.

The radio line widths predicted by our model (Section 5.4) may be underestimated at high redshift, due to the assumption of a constant random velocity dispersion σgas. Förster Schreiber et al. (2006) found Vcgas ≈ 2–4 for 14 UV-selected galaxies at redshift z ≈ 2. This result suggests that radio lines at z ≈ 2 should be about 20%–30% broader than predicted by our model, and therefore the evolution of the TFRs could be slightly stronger than shown in Figure 14.

In summary, the H i and H2 properties predicted for galaxies at high redshift are generally uncertain, even though no significant, systematic trend of the model errors could be identified. Perhaps the largest deviations from the real universe occur for very young galaxies or merging objects, while isolated field galaxies, typically late-type spiral systems, might be well described by the model at all redshifts.

In Section 3.4, we ascribed CO(1–0) luminosities to the H2 masses using the metallicity-dependent X-factor of Equation (26). This model neglects several important aspects: (1) the projected overlap of molecular clouds, which is negligible in the local universe, may become significant at high redshifts, where galaxies are denser and richer in molecules; (2) the temperature of the cosmic microwave background (CMB) increases with redshift, hence changing the level population of the CO molecule (Silk & Spaans 1997; Combes et al. 1999); (3) the CMB presents a background against which CO sources are detected; (4) the molecular material in the very dense galaxies, such as ultraluminous infrared galaxies, may be distributed smoothly rather than in clouds and clumps (Downes et al. 1993); (5) the higher SFRs in early galaxies probably led to higher gas temperatures, hence changing the CO-level population.7 Combes et al. (1999) presented a simplistic simulation of the cosmic evolution of X, taking the cosmic evolution of metallicity and points (1) and (2) into account. They found that for an H2-rich disk galaxy 〈X〉 increases by a factor 1.8 from redshift z = 0 to z = 5. This value approximately matches the average increase of X by a factor 2 predicted by our simulation using the purely metallicity-based model of Equation (26). This indicates that the effects of (1) and (2) approximately balance each other. If a more elaborate model for X becomes available, the latter can be directly applied to correct our CO predictions. In fact, the X-factor only affects the CO(1–0) luminosity LCO calculated via Equation (23), but has no effect on the line properties considered in this paper, namely, the line widths wpeakCO, w50CO, and w20CO and the normalized luminosity densities Ψ0CO and ΨmaxCO.

7. CONCLUSION

In this paper, we presented the first attempt to incorporate detailed cold gas properties in a semianalytic simulation of galaxies in a large cosmological volume. To this end, we introduced a series of physical prescriptions to evaluate relevant properties of H i and H2 in simulated model galaxies.

When applied to the DeLucia-galaxy catalog for the Millennium Simulation, our recipes only introduce one additional free parameter. This parameter, i.e., the cold gas correction factor ζ (Section 2), was tuned to the cosmic space density of cold gas in the local universe. The additional parameter ξ, describing the transfer of angular momentum from the halo to the disk (Section 3.2), is not a free parameter for the hydrogen simulation, since it is fixed by the baryon mass–scale radius relation. In fact, we deliberately did not adjust ξ to match the observed H i mass–H i radius relation of Verheijen (2001), in order to check the reliability of our models against this observation.

Based on the DeLucia catalog, we produced a virtual catalog of ∼3 × 107 per redshift snapshot with various cold gas properties. This catalog represents an extension of the DeLucia catalog, and it can be used to investigate a broad variety of questions related to H i, H2, CO, and their cosmic evolution. The results presented in this paper have been restricted to some important examples, most of which could be compared directly to available observations and hence constitute key results for the verification of our simulation.

  • 1.  
    Based on a pressure model for the molecular content of cold gas, the simulation simultaneously reproduces the H i MF and the H2 MF (resp. the CO LF) observed in the local universe within the measurement uncertainties (Figure 3).
  • 2.  
    The simulated H i MFs for spiral and elliptical galaxies considered individually also match the observations for simulated galaxies with well defined galaxy types (Figure 5).
  • 3.  
    The simulated H i radii, imply a mass–radius relation for H i that matches the empirical counterpart (Figure 8), thus confirming that the relation between $M_{\rm H\,\hbox{\scriptsize {\sc i}}}$ and $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ is such that the average H i density inside $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$ is $3.8\;{M}_{\odot }\rm \;pc^{-2}$ for all galaxies in the local universe, although this value sensibly depends on the definition of $r_{\rm H\,\hbox{\scriptsize {\sc i}}}$.
  • 4.  
    The simulation predicts that the mass–radius relations for H i and H2 are similar in the local universe, but that their high-redshift evolution is completely different (Figure 7).
  • 5.  
    The simulated widths of the H i radio emission lines of spiral galaxies are consistent with the empirical H i-TFR derived from the HIPASS spiral galaxies (Figure 13(a)); and the simulation predicts that there is no analog H i-TFR for elliptical galaxies.
  • 6.  
    The simulated the stellar mass TFR and the baryonic TFR reveal good agreement with the empirical TFRs for both spiral and elliptical galaxies in the simulation (Figures 13(b) and (c)).
  • 7.  
    These TFRs are observable manifestations of a more fundamental relation between circular velocity and total dynamical mass, as suggested by the small scatter in the relation between $w_{\rm H\,\hbox{\scriptsize {\sc i}}}^{20}$ and Mvir (Figure 13(d)).
  • 8.  
    At higher redshift, the simulation predicts that the above TFRs remain valid (except for H i at z ≈ 10), but that their scatter increases and their zero point is shifted toward higher velocities at fixed mass—a fundamental prediction of hierarchical growth (Figure 14).

The good match between simulation and observation regarding gas masses, disk sizes, and velocity profiles supports the models and recipes established in this paper. It also validates the semianalytic recipes used by De Lucia & Blaizot (2007) and supports the Millennium Simulation (Springel et al. 2005) as a whole.

In forthcoming investigations, the presented extension of the DeLucia catalog toward cold gas properties could be used to investigate more elaborate questions. For example, what is the bias of the cosmic structure, for example, of the power spectrum, revealed in H i surveys or CO surveys compared to the underlying dark matter structure? How many H i sources can we expect to detect in future experiments performed by the SKA? Or how does the global H2/H i ratio evolve with redshift and how does it relate to the observed evolution of the SFR density?

This effort/activity is supported by the European Community Framework Programme 6, Square Kilometre Array Design Studies (SKADS), contract no. 011938. The Millennium Simulation databases and the web application providing online access to them were constructed as part of the activities of the German Astrophysical Virtual Observatory. D.O. thanks Gerard Lemson for his help in accessing the simulation data, as well as Erwin de Blok, Scott Kay, Raul Angulo, Carlton Baugh, and Carlos Frenk for fruitful discussions. Finally, we thank the anonymous referee for the helpful suggestions.

Footnotes

  • In total, Leroy et al. (2008) analyzed 23 galaxies. Here, we only use the 12 galaxies, for which radial density profiles are provided for both H i and H2 (based on CO(2–1) or CO(1–0) measurements), and we subtract the helium fraction included by Leroy et al. (2008).

  • Stellar masses Mstars rely on $3.6\;\mu \rm m$ maps (SINGS; Kennicutt et al. 2003), H i masses use the $21\rm \;cm$ maps from THINGS (Walter et al. 2008), H2 masses rely on CO maps (CO(2–1) from HERACLES, Leroy et al. 2008; CO(1–0) from BIMA SONG, Helfer et al. 2003).

  • This list is not exhaustive, see Maloney & Black (1988) and Wall (2007) for an overview of the physical complexity behind the X-factor.

Please wait… references are loading.
10.1088/0004-637X/698/2/1467