A publishing partnership

The following article is Open access

Metallicity Dependence of Molecular Cloud Hierarchical Structure at Early Evolutionary Stages

, , , , , and

Published 2023 August 22 © 2023. The Author(s). Published by the American Astronomical Society.
, , Citation Masato I. N. Kobayashi et al 2023 ApJ 954 38 DOI 10.3847/1538-4357/ace34e

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/954/1/38

Abstract

The formation of molecular clouds out of H i gas is the first step toward star formation. Its metallicity dependence plays a key role in determining star formation throughout cosmic history. Previous theoretical studies with detailed chemical networks calculate thermal equilibrium states and/or thermal evolution under one-zone collapsing background. The molecular cloud formation in reality, however, involves supersonic flows, and thus resolving the cloud internal turbulence/density structure in three dimensions is still essential. We here perform magnetohydrodynamics simulations of 20 km s−1 converging flows of warm neutral medium (WNM) with 1 μG mean magnetic field in the metallicity range from the solar (1.0 Z) to 0.2 Z environment. The cold neutral medium (CNM) clumps form faster with higher metallicity due to more efficient cooling. Meanwhile, their mass functions commonly follow ${dn}/{dm}\propto {m}^{-1.7}$ at three cooling times regardless of the metallicity. Their total turbulence power also commonly shows the Kolmogorov spectrum with its 80% in the solenoidal mode, while the CNM volume alone indicates the transition toward Larson's law. These similarities measured at the same time in units of the cooling time suggest that the molecular cloud formation directly from the WNM alone requires a longer physical time in a lower-metallicity environment in the 1.0–0.2 Z range. To explain the rapid formation of molecular clouds and subsequent massive star formation possibly within ≲10 Myr as observed in the Large/Small Magellanic Clouds, the H i gas already contains CNM volume instead of pure WNM.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Molecular clouds host star formation, and thus their formation and evolution is an essential step for star formation in galaxies and galaxy evolution (Kennicutt & Evans 2012). The mass and volume of galactic disks are dominated by H i gas (Kalberla & Kerp 2009), but the spatial distribution of star formation rate is correlated more with molecular clouds (traced by CO lines) rather than H i gas (Schruba et al. 2011; Izumi et al. 2022). Therefore, the formation efficiency of molecular clouds out of H i gas and the resultant molecular cloud properties set the initial conditions of galactic star formation. In particular, the metallicity controls the heating/cooling rate and the thermal state of the interstellar medium (ISM; Field et al. 1969; Wolfire et al. 1995; Liszt 2002; Wolfire et al. 2003; Glover & Clark 2014; Bialy & Sternberg 2019) as well as it changes the formation/destruction rate of molecules and the resultant chemical state of the ISM (see, e.g., Glover & Clark 2012 for the CO-to-H2 conversion factor, and Glover & Clark 2014 and Bialy & Sternberg 2019 for the conditions of the metallicity and interstellar radiation field that makes the H2 cooling and heating important). Observations show that the metallicity increases with the cosmic time by the metal production from massive stars (e.g., Lehner et al. 2016). The metallicity also has a galactocentric gradient even within individual galaxies (e.g., the Milky Way; Fernández-Martín et al. 2017). Therefore, it is crucial to investigate metallicity dependence of the ISM evolution below the solar value (Z) to understand galactic star formation and cosmic star formation history.

In 1.0 Z environments, theoretical studies show that the thermal instability in the H i phase initiates the phase transition from the warm neutral medium (WNM) to the cold neutral medium (CNM; Field et al. 1969). Observational studies of the emission and absorption of H i, C ii, and CO lines show the existence of such multiphase ISM, and they now try to constrain the geometrical structure of the WNM and the CNM, for example, the studies with the Arecibo Telescope (e.g., Heiles & Troland 2003), Hubble Space Telescope (e.g., Jenkins & Tripp 2011), Giant Metrewave Radio Telescope (e.g., Roy et al. 2013), and Herschel, Very Large Array (VLA), and IRAM 30 m telescopes (e.g., Herrera-Camus et al. 2017; Murray et al. 2018). Such measurements are also performed toward lower-metallicity environments such as the Large Magellanic Cloud (LMC) with the Australian Telescope Compact Array (e.g., Marx-Zimmer et al. 2000). Previous one-zone theoretical studies have developed detailed chemical networks to comprehensively study the metallicity dependence of ISM evolution from the present day to primordial gas, such as thermal evolution of collapsing protostellar clouds (Omukai 2000; Omukai et al. 2005) and the thermal/chemical steady states (Bialy & Sternberg 2019). Theses studies show that the thermal instability still plays an important role even in low-metallcity environments with ≳10−2 Z where fine structure lines of [O i] (63.2 μm) and [C ii] (157.7 μm) are the dominant coolant. 9

The WNM and the CNM are in pressure equilibrium under a typical Galactic pressure and metallicity (Field et al. 1969; Wolfire et al. 1995; Liszt 2002; Wolfire et al. 2003; Bialy & Sternberg 2019). Supersonic flows are believed to be an important first step to initiate the thermal instability by compressing/destabilizing the previously stable WNM to the thermally unstable neutral medium (UNM), which subsequently evolves to the CNM due to cooling. Such supersonic flows originate from the passage of galactic spiral arms, the expansion of supernova remnants, H ii regions, etc. (Inutsuka et al. 2015). Many authors investigated this condition by performing numerical simulations of WNM converging flows, initially in one dimension (e.g., Hennebelle & Pérault 1999; Koyama & Inutsuka 2000; Vázquez-Semadeni et al. 2006), and later in multiple dimensions (e.g., Koyama & Inutsuka 2002; Audit & Hennebelle 2005; Heitsch et al. 2006b). They showed that the dynamically condensing motion of the UNM due to the thermal instability results in the formation of turbulent clumpy CNM structures, which are important progenitors of the filamentary structures observed in molecular clouds.

This multiphase nature of the ISM seems ubiquitous also in a wide range of low-metallcity environments as suggested by large-scale simulations on a galactic/cosmological scale (e.g., supersonic flows by supernovae, galaxy mergers, Arata et al. 2018; gas accretion from the host dark matter halo, Dekel & Birnboim 2006). Therefore, multidimensional numerical studies on low-metallicity molecular clouds are still essential to understand metallicity dependence of their formation and subsequent star formation, where the thermal instability and turbulence operate simultaneously. Inoue & Omukai (2015) performed three-dimensional simulations of converging flows as well as a linear stability analysis. They confirmed the development of the thermal instability as long as metal lines dominate in the cooling process under modest far-UV (FUV) background with $\mathrm{log}({G}_{0})\gt -3$ (see, Hu et al. 2021, for the validity of the chemical equilibrium assumption in low-metallicity environments). These previous approaches, however, employ a coarser spatial resolution in a lower metallicity, aimed at resolving the thermal instability with the same number of numerical cells between different metallicities, because the maximum growth scale of the thermal instability is larger at lower metallicities. The CNM clumps on subparsec scales are not fully resolved yet, and their properties and statistics in low-metallicity environments remain unclear.

Recent observations with the Atacama Large Millimeter/submillimeter Array (ALMA) reveal the existence of filamentary structures whose width is ∼0.1 pc in the outer disk of the Milky Way and the Magellanic Clouds (Tokuda et al. 2019; Fukui et al. 2019; Indebetouw et al. 2020; Wong et al. 2022; N. Izumi et al. 2023, in preparation), which is similar to those in the solar neighborhood (André et al. 2010; Arzoumanian et al. 2011). These structures are likely inherited from clumpy/filamentary structures in the H i phase (see Fujii et al. 2021, for a comparison of H i and molecular gas structures). The number of such high-resolution observations from radio to optical/near-infrared bands toward extragalactic sources will increase significantly in upcoming years with ALMA, James Webb Space Telescope (JWST), the next generation VLA, the Five-hundred-meter Aperture Spherical radio Telescope, the Square Kilometre Array, etc. Therefore, understanding subparsec-scale structures during molecular cloud formation from H i gas is critical to understanding the possible universality of star formation processes in different metallicity environments, e.g., the common existence of filamentary molecular clouds as observations suggest.

In this article, we perform magnetohydrodynamics (MHD) simulations of WNM converging flows to investigate the metallicity dependence of molecular cloud formation, especially focusing on thermal instability development and the resultant CNM properties. We investigate three cases with the metallicities of 1.0, 0.5, and 0.2 Z, which correspond to the typical values of the Milky Way, LMC, and Small Magellanic Cloud (SMC), respectively. By aiming at coherently resolving the turbulence/density structures comparable to the scale of molecular filaments/cores, we employ the 0.02 pc spatial resolution at all metallicities, which is enough to resolve the cooling length of the UNM evolving to the CNM (see Kobayashi et al. 2020). The typical cooling length of the WNM and UNM is 1 pc (1 Z)–3 pc (0.2 Z) and that of the CNM is 0.1 pc (1 Z)–0.3 pc (0.2 Z) in our simulation. This spatial resolution motivated by recent observations is higher than previous studies (e.g., compared with Inoue & Omukai 2015, at 0.2 Z), which enables us to coherently compare the statistics of CNM structures and discuss their possible universality/diversity between different metallicities.

The remainder of this article is organized as follows. In Section 2, we explain our simulation setups, and we show the main results in Section 3. In Section 4, we discuss the implications of low-metallicity cloud formation based on our results. We summarize our results in Section 5 and follow with a description of future prospects.

2. Method

2.1. Basic Equations and Setups

To investigate the development of thermal instability and the formation of multiphase ISM from the WNM, we calculate supersonic WNM converging flows by solving the following basic equations:

Equation (1)

Equation (2)

Equation (3)

Equation (4)

Equation (5)

Equation (6)

where ρ is the mass density, v represents the velocity, P represents the thermal pressure, T without any subscript is the temperature, Φ is the gravitational potential, G is the gravitational constant, and ∇i = ∂/∂xi , where xi spans x, y, and z. δij is the identity matrix. We calculate the total energy density, e, as e = P/(γ − 1) + ρ v2/2 + B2/8π where the ratio of the specific heat is γ = 5/3. We introduce the thermal conductivity, κ, as κ(T) = 2.5 × 103 T0.5 erg cm−1 s−1 K−1, which considers collision between hydrogen atoms (Parker 1953).

${ \mathcal L }(T,Z)$ is the net cooling rate. We employ the functional form of ${ \mathcal L }(T,1.0\,{Z}_{\odot })$ from Kobayashi et al. (2020) in the case of 1.0 Z condition. This functional form combines the results from Koyama & Inutsuka (2002) in T ≤ 14,577 K and Cox & Tucker (1969) and Dalgarno & McCray (1972) in T > 14, 577 K by considering the cooling rates due to Lyα, CII, He, C, O, N, Ne, Si, Fe, and Mg lines with the photoelectric heating. The shock heating and compression destabilize the WNM to join the UNM (defined as ${(\partial ({ \mathcal L }/T)/\partial T)}_{P}\lt 0;$ see, e.g., Balbus 1986, 1995), leading to the formation of the CNM clumps.

To investigate the lower-metallicity cases in this article, we apply three modifications to ${ \mathcal L }(T,1.0\,{Z}_{\odot })$ to prepare ${ \mathcal L }(T,Z)$. First, the cooling rate due to the metal lines is set to be linearly scaled with the metallicity, which is a good approximation in >10−4 Z as long as the metal lines dominate the cooling (Inoue & Omukai 2015). Second, we set the photoelectric heating rate to be linearly scaled with the metallicity by assuming that the dust abundance is also scaled with the metallicity. We, therefore, use ${{\rm{\Gamma }}}_{\mathrm{pe}}=2.0\times {10}^{-26}\left(Z/{Z}_{\odot }\right)$ erg s−1. Third, we implement the X-ray and cosmic-ray heating rates as ΓX = 2.0 × 10−27 erg s−1 and ΓCR = 8.0 × 10−28 erg s−1, respectively (Koyama & Inutsuka 2000). The X-ray and cosmic-ray heating processes are subdominant in the metallicity range of this study; for example, X-rays (cosmic rays) contribute 10% (30%) of the total heating rates in the 0.2 Z environment. Their relative importance decreases even more when the metallicity increases because the photoelectric heating rate increases with metallicity. We show the detailed functional form of this revised ${ \mathcal L }(T)$ in Appendix A. Figure 1 shows the thermal equilibrium state, i.e., ${ \mathcal L }(T,Z)=0$. This shows that, in lower-metallicity environments, the combination of the inefficient cooling and metallicity-independent X-ray and cosmic-ray heatings allows for the existence of the WNM phase until higher density in the range of 1–10 cm−3. In each calculation, we assume a uniform metallicity distribution in space, as representation of low-metallicity environments.

Figure 1.

Figure 1. The thermal equilibrium states as a function of pressure (P), density (n), and the metallicity (from 1.0–0.2 Z, shown with differently colored lines). The black circle at the bottom left shows the initial condition at 1.0 Z.

Standard image High-resolution image

We use the publicly available MHD simulation code Athena++ (Stone et al. 2020) to solve the basic equations, where we employ the HLLD MHD Riemann solver (Miyoshi & Kusano 2005) and the constrained transport method to integrate the magnetic fields (Evans & Hawley 1988; Gardiner & Stone 2005, 2008). The self-gravity is calculated with the full multigrid method (Tomida et al. 2023).

Our simulation domain has sizes of Lx,y,z = 20, 10, and 10 pc on each side in Cartesian coordinates. We continuously inject supersonic WNM flows from the two x-boundaries so that the two flows collide head-on. We employ the periodic boundary condition on the y- and z-boundaries. The collision forms a shock-compressed layer sandwiched by two shock fronts, in which the thermal instability converts the WNM to CNM. We employ Vinflow = 20 km s−1 as the flow velocity; this choice corresponds to a representation of the late phase of supernova remnant expansion or normal component of galactic spiral shocks. The initial velocity field is set as ${v}_{x}(x)\,={V}_{\mathrm{inflow}}\tanh (-x/0.78\,\mathrm{pc})$ and vy,z = 0 km s−1.

The initial WNM flow, and also the injected WNM flow, are thermally stable, with a mean number density n0 = 0.57 cm−3. The corresponding pressure, temperature, and sound speed at each metallicity are listed on Table 1, where we use ρ0 = n0 μM mp with the mean molecular weight μM of 1.27 (Inoue & Inutsuka 2012). The ram pressure of the converging flow is ${P}_{\mathrm{ram}}/{k}_{{\rm{B}}}^{}=3.5\times {10}^{4}\,{\rm{K}}\,{\mathrm{cm}}^{-3}$. The WNM flows have density fluctuation following the Kolmogorov spectrum Pρ (k) ∝ k−11/3 (Kolmogorov 1941; Armstrong et al. 1995). We impose a random phase in each k up to k/2π = 3.2 pc−1, which corresponds to a wavelength of 0.32 pc. The mean dispersion of the density fluctuation is chosen as $\sqrt{\langle \delta {n}_{0}^{2}\rangle }/{n}_{0}=0.5$. The injected flows have periodic distributions smoothly connected to the initial condition, so that we impose ρ(t, x = − 10 pc, y, z) = ρ(t = 0, x = 10 pc − Vinflow t, y, z) and ρ(t, x = 10 pc, y, z) = ρ(t = 0, x = − 10 pc + Vinflow t, y, z) on the x-boundaries where t represents the time. The interaction between this density inhomogeneity and shock fronts generates turbulence (e.g., Koyama & Inutsuka 2002; Inoue & Inutsuka 2012; Carroll-Nellenback et al. 2014; Kobayashi et al. 2022).

Table 1. Parameters at Different Metallicities

  ${P}_{0}/{k}_{{\rm{B}}}^{}$ T0 Cs,0 tcool tfinal
 [K cm−3][K][km s−1][Myr][Myr]
1 Z 3.6 × 103 6.4 × 103 6.41.03.0
0.5 Z 3.5 × 103 6.2 × 103 6.32.06.0
0.2 Z 3.4 × 103 6.1 × 103 6.25.115.0

Note. n0 = 0.57 cm−3 and B0 = 1 μG Are Common at All Metallicities.

Download table as:  ASCIITypeset image

The magnetic field is initially threaded in the x-direction with 1 μG strength. Previous studies show that successful molecular cloud formation occurs in such a configuration where the flow and the mean magnetic field are close to parallel (Hennebelle & Pérault 2000; Inoue & Inutsuka 2008; Iwasaki et al. 2019). The typical mean field strength in H i gas varies from 1–10 μG in the Milky Way (Crutcher 2012) and 0.5–5 μG in the Magellanic Clouds (Gaensler et al. 2005; Beck 2013; Livingston et al. 2022). We, therefore, opt to choose 1 μG as representative strength in investigating the metallicity dependence. The detailed dependence on the field strength in low-metallicity environments is left for future studies at this moment.

We employ a uniform spatial resolution of 0.02 pc to resolve the typical cooling length of the UNM. This resolution is required to have convergence in the CNM mass fraction after 1.0 tcool (where tcool is the cooling time defined as $P/(\gamma -1)/\rho { \mathcal L }$; Kobayashi et al. 2020). We identify the shock front position by P > 1.3P0 to define the volume of the shock-compressed layer.

Figure 2 shows the three-dimensional view of the initial density field with a uniform 1 μG magnetic field.

Figure 2.

Figure 2. The initial condition. The color shows n [cm−3]. The white lines represent the initial uniform magnetic fields with 1 μG strength. The two supersonic flows are injected through the two x-boundaries at x = − 10 and 10 pc, where the origin of the coordinate is at the box center.

Standard image High-resolution image

3. Results

3.1. Expectations

Figure 1 shows that the CNM thermal state at n > 102 cm−3 is similar within the 1.0–0.2 Z range. In this metallicity range, the cooling rate is proportional to the metallicity (see Inoue & Omukai 2015). These suggest that the tcool is longer in low-metallicity environments as ∝ Z−1 but CNM properties may become similar between 1.0–0.2 Z at the same time measured in units of the cooling time.

The typical tcool of the injected WNM is ∼1 Myr at 1 Z, ∼2 Myr at 0.5 Z, and ∼5 Myr at 0.2 Z (see Appendix B.1). Kobayashi et al. (2020) investigated the converging flow at 1 Z and suggested that the shock-compressed layer achieves a quasi-steady state at ∼3tcool. Therefore, to make a comparison between different metallicities both at the same physical time and at the same time measured in units of the cooling time, we integrate until 3 tcool in each metallicity. Table 1 lists these parameters.

3.2. The Thermal States and Magnetic Field Evolution

Figure 3 shows examples of the three-dimensional view of our simulation results. Panels (a), (b), and (c) compare the results at 3 Myr from the 1.0 Z, 0.5 Z, and 0.2 Z runs. Panels (d) and (e) compare the results at 3tcool from the 0.5 Z and 0.2 Z runs. These panels show that CNM clumpy/filamentary structures form more slowly in lower-metallicity environments. The development of such CNM structures similar to those in 1 Z environment requires similar times measured in units of the cooling time, for example, at 3tcool. Compared to the same physical time of 3 Myr, the geometry of the shock-compressed layer in a lower-metallicity environment is less disturbed (e.g., panel (c)), close to a plane-parallel configuration because inefficient cooling keeps the layer more adiabatic.

Figure 3.

Figure 3. Three-dimensional view of the density and magnetic fields. Panels (a), (b), and (c) are at t = 3 Myr from 1.0 Z, 0.5 Z, and 0.2 Z. Panels (d) and (e) are at t = 3tcool from 0.5 Z and 0.2 Z, which therefore correspond to t = 6.0 Myr and 15.0 Myr, respectively. We show the slices of the density field on the three boundaries at x = 10 pc, y = − 5 pc, and z = − 5 pc. The density is colored with n [cm−3] from blue (0.1 cm−3) to yellow (103 cm−3). The magnetic field lines are integrated from the two x-boundaries, whose strength $\left|B\right|$ is colored in orange.

Standard image High-resolution image

Figure 4 compares the thermal states of the different metallicities at 3 Myr and at 3 tcool. This shows that, at 3 Myr, the shock-compressed layer is still dominated more by the shock-heated WNM/UNM in the lower-metallicity environment. Their thermal pressure is still close to the flow ram pressure. Therefore, the temperature just starts to decrease in the shock-compressed layer of 0.2 Z at 3 Myr (0.6 tcool), so that the plane-parallel geometry of the shock-compressed layer seen in Figure 3 is close to a simple one-dimensional shock compression (see also Section 3.3 and Appendix C for its impact on the turbulent velocity).

Figure 4.

Figure 4. The phase diagram of the shock-compressed layer. The upper (lower) panels compare the thermal states at t = 3 Myr (at t = 3tcool). The color represents the number of cells. The red circle at the lower left shows the initial state of the injected WNM, and the gray solid line emanating from that red point shows the isothermal line (i.e., Pn). The black horizontal line shows the flow ram pressure.

Standard image High-resolution image

The net magnetic flux in our simulation does not increase in time because the mean magnetic field is completely parallel to the WNM flow. Nevertheless as we see in Figure 3, the pre-shock density fluctuation induces a number of oblique shocks at the shock front, which locally fold the field lines. The magnetic fields in the shock-compressed layer are further twisted and stretched due to the turbulence, which introduces a larger scatter in the local field strength. We investigate the relation between the field strength and the number density at 3 tcool. Figure 5 shows the phase histogram on the plane of the magnetic field strength and the number density in the shock-compressed layer. This figure shows that the field strength is initially amplified toward the shock-heated WNM volume through the shock compression almost following Bn1 (as indicated with the translucent gray line). The strength further varies due to the turbulence at n < 10 cm−3. Note that the initial mean magnetic field is completely parallel to the flow velocity, and not all of the volume experiences a perfect one-dimensional shock compression. Therefore, the most frequent field strength is slightly weaker than that the relation B ∝ n1 (see the upper panels of Figure 5).

Figure 5.

Figure 5. The phase histogram on the plane of the magnetic field strength and the number density. The color represents the number of the numerical cells. The black small circle shows the initial condition of the injected WNM flow, and the translucent dashed gray line emanating from this black circle shows the relation of Bn1. The horizontal black dashed line shows the typical maximum field strength by assuming the balance between the magnetic pressure and the inflow ram pressure is that of Equation (7). The black solid curve shows the volume-weighted mean field strength 〈B〉 as a function of density in the range of n ≥ 1 cm−3. The gray straight line at the top shows the relation of 〈B〉 ∝ n1/5.

Standard image High-resolution image

Accompanying the density enhancement toward n ∼ 103 cm−3 by the thermal instability, the field strength gradually increases but with a limited level. The black solid curve in Figure 5 shows the volume-weighted average of the magnetic field strength as a function of the number density 〈B〉(n), where $B=\left|{\bf{B}}.\right|$. 10 This shows that the amplification roughly scales with 〈B〉 ∝ n1/5 in the n > 1 cm−3 range. This slow evolution occurs because condensing motion by the thermal instability is confined along the magnetic field orientations. Similar results have also been obtained by previous numerical studies of the solar metallicity environment (e.g., Hennebelle et al. 2008; Heitsch et al. 2009; Inoue & Inutsuka 2012). Our results show that this gradual amplification of the magnetic field occurs also in lower-metallicity environments.

Figure 5 indicates that there is a maximum magnetic field strength attained. We can estimate this maximum strength by assuming there is balance between the postshock magnetic pressure and the ram pressure of the injected WNM as ${B}_{\mathrm{eq}}^{2}/8\pi ={\rho }_{0}{V}_{\mathrm{inflow}}^{2}$. This gives a typical value of

Equation (7)

Although this is an extreme case where the magnetic energy dominates the shock-compressed layer, Equation (7) should give a good approximation even when we consider local enhancement by the turbulence and the condensing motion by the thermal instability, because the inflow ram pressure determines the typical maximum pressure of the shock-compressed layer. Beq, shown in Figure 5 (the horizontal black dashed line), indeed outlines the typical maximum strength of the magnetic fields.

Note that the field strength can increase beyond Beq at the densest volume where the self-gravity plays a role. This occurs in n ≳ 2.6 × 103 cm−3 in our simulation setup. Such a critical density of n ∼ 2.6 × 103 cm−3 can be estimated as the density at which the CNM plasma beta with B = Beq becomes unity (Iwasaki & Tomida 2022). It is difficult to clearly confirm this amplification in our current simulations due to our limited volume and due to our diffuse WNM-only initial condition. Nevertheless, we expect that the field strength also increases in low-metallicity environments because the CNM thermal states are similar between 1.0–0.2 Z, as are the critical densities at which the self-gravity starts to dominate. Previous simulations of a converging flow at 1.0 Z show 〈B〉 ∝ n1/2 at n > 103 cm−3 either when they integrated much longer times to accumulate mass or when they started with two-phase atomic flows with their mean number density already n ≳ 5 cm−3 (e.g., Hennebelle et al. 2008; Inoue & Inutsuka 2012; Iwasaki & Tomida 2022).

Our results show that the development of CNM clumpy/filamentary structure and the magnetic field strength is similar between metallicities if they are measured at the same time in units of the cooling time in each metallicity (i.e., the evolution slows down linearly with the metallicity in terms of the physical time). The field strength is relatively constant from a few microgauss to 10 μG up to n < 103 cm−3 as 〈B〉 ∝ n1/5, and possibly starts to amplify further by self-gravity. Note that this field strength in n < 103 cm−3 is consistent with Zeeman measurements of low (column) density regions of molecular clouds in the Milky Way (e.g., Crutcher 2012) and Faraday rotation measurements toward the ISM in the LMC and SMC (Gaensler et al. 2005; Beck 2013; Livingston et al. 2022).

3.3. The Overall Turbulent Structure

In Kobayashi et al. (2022), we showed that the coevolution of the turbulence and the thermal evolution by the thermal instability plays a significant role to determine the multiphase density structure and its density probability distribution function at the molecular cloud formation stage. Kobayashi et al. (2022), however, neglected the magnetic fields and studied simple hydrodynamic converging flows. In this section, we will confirm those findings even in the current MHD simulations and investigate how the turbulent structure differs between/resembles the cases with different metallicities.

Panels (a), (b), and (c) of Figure 6 show the evolution of the velocity dispersion, the mean density, and the turbulent pressure, respectively, as a function of the time measured in units of the cooling time (i.e., t/tcool). In the early stage at t < 0.5 tcool, the shock-compressed layer has plane-parallel geometry in the low-metallicity environment (see Section 3.2). Due to the resultant efficient deceleration of the inflow and the inefficient cooling, the mass is dominated by the shock-heated WNM state with slow turbulence, with negligible volume/mass of the CNM. Therefore, the volume-weighted velocity dispersion $\sqrt{\langle \delta {v}^{2}\rangle }$ is close to the density-weighted velocity dispersion $\sqrt{\langle \delta {v}^{2}{\rangle }_{\rho }}$ of ∼2 km s−1 in 0.2 Z (panel (a) of Figure 6; see also Appendix C).

Figure 6.

Figure 6. The tcool-normalized time evolution of the volume-weighted velocity dispersion $\sqrt{\langle \delta {v}^{2}\rangle }$ and the density-weighted velocity dispersion $\sqrt{\langle \delta {v}^{2}{\rangle }_{\rho }}$ (panel (a)), the volume-weighted mean density 〈n〉 (panel (b)), and the averaged turbulent pressure 〈Pturb〉 (panel (c)), where 〈Pturb〉 = 〈ρ〉〈δ v2ρ . The black, red, and blue colors represent 1.0, 0.5, and 0.2 Z, respectively.

Standard image High-resolution image

At t ≳ 0.5 tcool in the lower-metallicity environment, the phase transition from the shock-heated WNM to the CNM tends to occur at a pressure closer to the inflow ram pressure (see panel (c) Figure 4). The mean number density is higher in lower-metallicity environments accordingly (panel (b) of Figure 6). During this evolution, the turbulent pressure gradually grows until it balances against the inflow ram pressure. As a result, the turbulent pressure is almost the same between the three metallicities at 3 tcool (panel (c) of Figure 6).

These results suggest that the turbulent pressure is almost the same at 3 tcool. The difference exists due to efficient compression in the lower-metallicity environments, because the inflow and the shock front incidents with an almost 90° angle, which results in a denser postshock WNM with slower velocity. However, this difference is limited by a factor of 2 (4) in the velocity dispersion (in the mean density, respectively). The value of $\sqrt{\langle \delta {v}^{2}\rangle }\sim 6$–9 km s−1, close to the sound speed of the WNM, suggests that turbulence on the molecular cloud scale is powered by the WNM super-Alfvénic turbulence not only in the Milky Way galaxy but also in the LMC and SMC. 11

To understand the turbulence structure, we perform a Fourier analysis of the turbulence by using a fast Fourier transform in the west (3.3; Frigo & Johnson 2005), and decompose the turbulence into the solenoidal and compressive modes, which are defined, respectively, as

Equation (8)

Equation (9)

Here, $\hat{k}$ represents a unit wavevector and $\tilde{{\boldsymbol{v}}}$ represents the Fourier component of the velocity field. We here denote $k=\left|{\boldsymbol{k}}\right|$. Figure 7 shows the power spectrum at 3 tcool. The solenoidal mode dominates the turbulence power on all scales. Table 2 summarizes the total fraction of each turbulent mode at 3 tcool. The solenoidal mode fraction amounts to 80%–90%. In calculating this, we employ powers between k/2π = 0.1 and 2.56 pc−1 to avoid numerical diffusion effects on a small scale (we refer the reader to the caption of Figure 7). 12

Figure 7.

Figure 7. The one-dimensional averaged turbulent power spectrum of the shock-compressed layer at t = 3tcool. The cyan, blue, and red curves correspond to the total power, the solenoidal mode power, and the compressive mode power, respectively. Their width represents the Poisson noise at each frequency k. It is difficult to visually recognize the cyan curve because the solenoidal mode dominates the total power so that they almost overlap in this presentation. The two vertical gray lines show the lowest frequency of the shock-compressed layer and the Nyquist frequency. The thin gray dotted line shows one-tenth of the Nyquist frequency, above which the numerical diffusion impacts the Fourier power. We employ the powers between the left gray line (k/2π = 0.1 pc−1) and the thin dotted line (k/2π = 2.56 pc−1) to measure the total fraction of the solenoidal and compressive modes in the subsequent analyses and figures. The black dashed line shows the one-dimensional Kolmogorov spectrum of k2 Pv (k) ∝ k−5/3.

Standard image High-resolution image

Table 2. Mode Fraction at 3 tcool, where the Total Power Is Normalized to Unity

Metallicity1.0 Z 0.5 Z 0.2 Z
Solenoidal0.860.910.93
Compressive0.140.090.07

Download table as:  ASCIITypeset image

Figure 8 shows the time evolution of this mode fraction as a function of t/tcool. The solenoidal (compressive) mode fraction evolves quasi-steadily with ∼80% (20%, respectively) after 1 tcool.

Figure 8.

Figure 8. The tcool-normalized time evolution of the turbulence mode fraction. The upper (lower) three lines correspond to the solenoidal (compressive, respectively) mode.

Standard image High-resolution image

Such solenoidal-mode-dominated turbulence is a natural consequence of the accreting system followed with the thermal instability, as we discuss in our previous converging flow simulations at 1.0 Z (Kobayashi et al. 2022). At early stages, the shock fronts deform by interaction with the inflow density inhomogeneity. These curved shock fronts introduce inhomogeneity in the postshock entropy distribution. Such entropy inhomogeneity generates turbulent vorticities accounting for a small fraction of the solenoidal mode (see Kida & Orszag 1990). At later stages after 1.0 tcool, the interaction between the formed CNM clumps and the shock fronts significantly deforms the shock fronts. This deformation creates a number of oblique shocks, whose maximum size is comparable to the size of the shock-compressed layer. This induces strong shear motion into the shock-compressed layer, so that the solenoidal mode dominates the turbulence. Our results show that the solenoidal-mode-dominated turbulence also emerges in the low-metallicity environment.

The results presented in this Section suggest that in the range of 1.0–0.2 Z, molecular clouds forming in the shock-compressed layer have a turbulent pressure close to the inflow ram pressure. Note that ≳80% of the turbulence power is in the solenoidal mode at all of the metallicities. This indicates that even in a galactic-scale converging region, forming molecular clouds are always solenoidal-mode dominated. Therefore, galactic-scale compressive motion is important in the formation of molecular clouds, but it does not immediately imply enhancement of star formation efficiency by enhancing compressive motion in molecular clouds.

3.4. The Properties/Statistics of the CNM Clumps

We identify the CNM structures to further investigate their properties. In this section, we define the CNM as the volume with T < 200 K and n > 20 cm−3.

First, we perform the Friends-of-Friends algorithm to identify CNM clumps as groups of connected CNM cells. Each clump has more than 64 member cells to avoid numerical noise on small scales. Figure 9 compares the size and mass functions of the identified CNM clumps at 3 tcool. We define the size l of each CNM clump as $l=\sqrt{{I}_{\max }/M}$ where ${I}_{\max }$ and M are the maximum eigenvalue of its inertia matrix and the mass of each clump. The size distribution peaks at ∼0.1 pc at all of the metallicities. 13 The mass functions follow a power-law distribution of ${dn}/{dm}\propto {m}^{-1.7}$ up to ∼102 M. Some CNM mass tends to stagnate in the central region of the shock-compressed layer due to the converging flow configuration (see Figure 3 and Appendix C). Large CNM clumps obtain more mass or coagulate with other CNM gas so that the most massive clumps deviate from the power-law distribution.

Figure 9.

Figure 9. The size and mass functions of CNM clumps. The gray filled histogram shows 1.0 Z, and the other step histograms show 0.5 Z (red) and 0.2 Z (blue). The mass function shows a reference power-law distribution of ${dn}/{dm}\propto {m}^{-1.7}$ (black dashed line).

Standard image High-resolution image

An index of −1.7 has often been reported in the 1.0 Z environment by previous converging WNM flow simulations in two dimensions (e.g., Heitsch et al. 2005; Hennebelle & Audit 2007; Hennebelle et al. 2008) and in three dimensions (e.g., Inoue & Inutsuka 2012). This power-law function is explained via the statistical growth of the thermal instability with a given density fluctuation spectrum. A functional form of  $dn/dm\propto {m}^{(\alpha -3)/3-2}$ is expected when the seed density fluctuation has a three-dimensional averaged spectral index of α (we refer the readers to Hennebelle & Audit (2007), for a detailed explanation and the derivation of the index (α -3)/3-2)). When α = 11/3, i.e., the Kolmogorov fluctuation, this gives ${dn}/{dm}\propto {m}^{-1.7}$, which is indeed consistent with the numerical results of previous papers in 1.0 Z environments. Our results suggest that, at the same t/tcool, the same CNM mass spectrum can also appear in lower-metallicity environments due to the thermal instability with the Kolmogorov turbulence background.

Second, we investigate the turbulence structure of the CNM volume alone by measuring the two-point correlation of the CNM velocity field. We measure the second-order velocity structure function

Equation (10)

where $r=\left|{\boldsymbol{r}}\right|$, and x is the three-dimensional position of the CNM cells. We select 27 subvolumes in the shock-compressed layer, where each subvolume is a (3.3 pc)3 cube, and their volume-centered position is (x, y, z) = (0 ± 3.3 pc, 0 ± 3.3 pc, 0 ± 3.3 pc). Figure 10 shows $\sqrt{S(r)}$ from each metallicity at 3 tcool. The CNM volume overall shows the transition toward the Larson-type scale-dependence as $\sqrt{S(r)}\propto {r}^{1/2}$. This relation extends from small scales within individual CNM clumps to large scales beyond those clumps. This indicates that the commonly observed Larson's law in molecular clouds is inherited from the CNM phase and is also ubiquitous in low-metallicity environments down to 0.2 Z.

Figure 10.

Figure 10. The square root of the second-order velocity structure function S(r) as a function of the distance r between CNM cells at 3 tcool. The black shaded region shows the maximum and minimum of S(r) among the 27 sampled subvolumes. The red solid curve shows the volume-weighted average of all of the subvolumes. The black solid (dashed) line shows the Larson (Kolmogorov) power-law relation as $\sqrt{S(r)}\propto {r}^{1/2}$ ($\sqrt{S(r)}\propto {r}^{1/3}$). The vertical gray dotted line at k/2π = 0.4 pc−1 indicates one-tenth of the Nyquist frequency (i.e., the typical spatial scale below which the turbulent power is affected by the numerical diffusion).

Standard image High-resolution image

Note that the dynamic range is still limited if we consider only the scales without the numerical diffusion (i.e., the scales larger than the vertical dotted line). S(r) in this range is between the Kolmogorov and Larson relations. Further high-resolution simulations are required to finally conclude that the CNM velocity structure function follows the Larson-type relation on all scales. Nevertheless, the strongest turbulence power within the CNM clumps is in eddies whose scale is comparable to the typical CNM clump size ∼0.1 pc. Our results, therefore, suggest that the internal velocity dispersion within the CNM clumps remains ≲1 km s−1 while the clump-to-clump relative velocity is 3–5 km s−1.

Note that the amplitude of $\sqrt{S(r)}$ at 0.2 Z is smaller by a factor of 2 than that at 1.0 Z, as we see in $\sqrt{\langle \delta {v}^{2}{\rangle }_{\rho }}$ (see Figure 6). This is consistent with the comparison of the CO line width between clouds in the LMC/SMC and the Milky Way (Bolatto et al. 2008; Fukui & Kawamura 2010; Ohno et al. 2023), except for those in extreme star-forming systems, such as 30 Dor R136 cluster (Indebetouw et al. 2013). Also note that a variation at a given spatial displacement r is larger in the lower-metallicity environment because the shock-compressed layer is thicker, and thus the total number of cells in this analysis increases toward lower metallicity.

The results presented in this Section show that CNM mass spectrum is ${dn}/{dm}\propto {m}^{-1.7}$, common in the 1.0–0.2 Z range with a Kolmogorov turbulence background. The second-order structure function of the CNM volume alone indicates the transition toward the Larson-type turbulence scale-dependence ubiquitously at all metallicities with its amplitude smaller by a factor of 2 at 0.2 Z compared with 1.0 Z.

Combining all of the results in Section 3, Figure 11 schematically summarizes the hierarchical thermal/turbulent structure of the multiphase medium in this metallicity range.

Figure 11.

Figure 11. A schematic figure showing the relationship between the density structures and their thermal states in the metallicity range of 1.0–0.2 Z. The WNM/UNM turbulence on the entire cloud scale powers the relative velocity between CNM clumps with 3–5 km s−1, while the internal velocity dispersion within individual CNM clumps remains ≲1 km s−1. The velocity two-point correlation of the CNM volume follows Larson's law as $\sqrt{S(r)}\propto {r}^{1/2}$.

Standard image High-resolution image

4. Implications and Discussions

Our results show that the physical properties of a shock-compressed layer scales with the metallicity in the 1.0–0.2 Z range. This suggests that the properties of subsequently formed molecular clouds are likely similar between the Milky Way, LMC, and SMC if we somehow select and compare clouds at the same t/tcool.

4.1. Preexistence of CNM Structures in the WNM: CNM Determines the Formation of Molecular clouds?

On one hand in the 1.0 Z context, previous authors often investigated cloud formation by supersonic flows in supernova remnants, galactic spirals, etc. (Kim et al. 2020; Chevance et al. 2022). For example, a single supernova remnant expands to 50 pc size in 0.3 Myr (∼0.3 tcool) and accumulates 104 M H i mass (e.g., Kim & Ostriker 2015). The compilation of multiple supernovae events is able to create even more-massive molecular clouds (Inutsuka et al. 2015). On the other hand, in a low-metallicity environment, the molecular cloud formation out of the WNM alone requires longer physical time. This is the case even in the configuration where the mean magnetic field is completely parallel to the WNM inflow, as we have studied, which forms molecular clouds most efficiently compared with the magnetic fields inclined against the flow (Inoue & Inutsuka 2008; Iwasaki et al. 2019).

To estimate how this timescale impacts molecular cloud formation in low-metallicity environments, let us assume that all of the CNM volume eventually evolves to the molecular gas. Based on our simulation results, 14 the expected mass of the cloud, Mcloud, is approximately

Equation (11)

Equation (12)

Here L2 is the inflow cross section in our calculation domain, and thus the first three factors in Equation (11) come from the inflow mass flux. The exact dependence of tcool on n0 and Vinflow is still left for future studies, which is involved in the conversion from Equation (11) to Equation (12). We here employ ${t}_{\mathrm{cool}}\propto {n}_{0}^{-1}{Z}^{-1}$ as the fiducial dependence, which is consistent with our definition of tcool as calculated in Table 1 (see Appendix B.2 and Equation (B7)).

If we suppose the case of a single flow event with n0 = 0.57 cm−3 and Vinflow = 20 km s−1 creates a typical maximum cloud of a few 105 M in the LMC/SMC, Equation (12) indicates that such a WNM flow should continue coherently on scales of L ≃ 100 pc and 15 Myr. Note that 15 Myr is 1 order of magnitude longer compared with the typical expansion timescale of a single supernova remnant. A superbubble rather than a single supernova event is more likely to keep such a coherent one-directional flow over a 15 Myr timescale. However, the prevalent existence of molecular clouds that are not associated with superbubbles in the LMC/SMC indicates that faster flows and/or the preexistence of CNM structure in the WNM are important for cloud formation and evolution in a lower-metallicity environment (see Dawson et al. 2013). 15

One possibility is a fast H i gas flow, which is observed as the tidal interaction between the LMC and the SMC (Fukui et al. 2017; Tsuge et al. 2019). Its velocity is as high as 50–100 km s−1, comparable to the escape velocity of the LMC/SMC. We can straightforwardly expect the coherent continuation of such a flow over a few 10 Myr timescale because it travels on a 10 kpc scale between the LMC and the SMC. However, such a fast flow induces strong turbulence in molecular clouds, which can also destroy dense structures (see Klein et al. 1994; Heitsch et al. 2006a). In addition, even for such efficient mass accumulation, thermal evolution from the WNM to molecular clouds still requires cooling. For example, albeit on an L = 100 pc scale with a coarser spatial resolution of 0.2 pc, Maeda et al. (2021) performed convergence of WNM flow simulations with Vinflow = 100 km s−1 but without any CNM structure preexisting in the WNM flow. In the 0.2 Z environment with an initial WNM density of n0 = 0.75 cm−3, they showed that the molecular cloud formation takes 23 Myr. The cloud mass with n > 104 cm−3 is a few 105 M at 23 Myr, which is consistent with Equation (12). 16

Therefore, even in the fast flow environment, the preexistence of CNM structure in the WNM is important (see Pringle et al. 2001; Inoue & Inutsuka 2009). This leads to another question of, "how does the ISM form the CNM in low-metallicity environments?" As can be seen in our simulations, low-mass CNM clumps form if the mean magnetic field is parallel to compressive events, and this inflow can be as slow as Vinflow = 20 km s−1. Therefore, some previous generations of flows with a few tens of kilometers per second along with the mean magnetic field (such as a single supernova) are responsible for introducing CNM clumps in the WNM. Such CNM preexistence and subsequent compression due to shocks presumably determine the formation site, the mass, and the structure of molecular clouds. Indeed, based on the recent ALMA observations toward N159E/W regions in the LMC, Tokuda et al. (2022) indicated that a fast H i gas flow with a multiple-scale substructure potentially explains the formation of filamentary molecular clouds on various size scales. Such a hierarchical structure may also determine the fractal nature of the young stellar structures in the LMC, suggested by the spatial clustering analysis of star clusters (e.g., Miller et al. 2022). In addition, the similarity between the power-law spectrum of the molecular cloud mass function (e.g., ${dn}/{dm}\propto {m}^{-\alpha }$ where α = 1.7–1.9 in the LMC; Fukui et al. 2001; Ohno et al. 2023) and that of CNM clump mass function in our simulation also indicates that molecular cloud formation is predetermined by CNM structure (Figure 9). This implication requires further study to be confirmed in the future.

4.2. H2 Cooling

The H2 is an important coolant in the context of the primordial gas cooling. Chemical paths to H2 formation exist even in the primordial gas, especially that via the gas-phase interaction between H and H, where the electron fraction impacts the abundance of H. Susa et al. (1998) calculated the electron fraction in the postshock layer of supersonic ISM flows, and Susa & Umemura (2004) showed that, with Vinflow = 30 km s−1, H2 cooling is dominant over the metal line cooling in the postshock region even in 0.1 Z without any UV background that dissociates the H2. This is not the case in our calculation where we employ G0 = 1 for all of the metallicity by assuming ongoing active star formation as in the LMC/SMC. The metal line cooling is always dominant in this setup.

The impact of H2 cooling on thermal instability is limited typically to G0 ≤ 10−3 environments (Inoue & Omukai 2015; see also Glover & Clark 2014). It is left for future studies to investigate the dynamical formation and evolution of the multiphase ISM driven by supersonic flows in lower-metallicity environments with a UV radiation field of G0 < 1.

4.3. Different Assumptions: Dust-to-gas Ratio, Electron Fraction, and Radiation Field

In this section, we introduce several previous studies to list how the adopted assumptions impact our results.

First, many previous studies also assume that the dust-to-gas ratio is linearly scaled with the gas-phase metallicity in their fiducial models. Meanwhile, based on the extragalactic observational indications (e.g., Rémy-Ruyer et al. 2014), there are also studies testing a broken power-law dependence. For example, ${ \mathcal D }\propto {Z}^{2}$ in Glover & Clark (2014) and Z3 in Bialy & Sternberg (2019), where ${ \mathcal D }$ is the dust-to-gas ratio. Such a superlinear decrease of the dust abundance at ≲0.2 Z reduces the photoelectric heating rate relatively to the C ii cooling rate. The thermally stable states of the UNM and the CNM are cooler in this superlinear case 17 . However, in the range of ≥0.2 Z, they show that this difference has a limited impact on the CNM thermal equilibrium state (see also Figure 5 of Kim et al. 2023b). In addition, this superlinear relation of the dust-to-gas ratio does not significantly impact the overall dynamics in our simulations because C ii cooling at the shock-heated WNM right after the shock compression provides the typical evolutionary timescale (as discussed in Appendix B.1).

Second, the gas-phase electron fraction changes the photoelectric heating efficiency. The photoelectric heating rate by FUV radiation can be estimated as Γphot(1.0 Z)= 1.3 × 10−24 epsilon G0 erg s−1 where epsilon is the photoelectric heating efficiency as

Equation (13)

(see Equation (43) in Bakes & Tielens 1994). Here ne is the electron number density so that the electron fraction can be defined as xe = ne /nH, and the polycyclic aromatic hydrocarbon (PAH) reaction rate is ϕPAH = 0.5 (Wolfire et al. 2003). The G0 T1/2/ne dependence in the denominator describes the relative importance of the ionization against the recombination on the dust surface. 18 Our fiducial value, Γphot(1.0 Z) = 2.0 × 10−26 erg s−1 in Equation (A2), corresponds to the typical condition of the thermally stable WNM at 1.0 cm−3 with xe ∼ 0.02 determined by cosmic-ray ionization (see Equation (12) of Bialy & Sternberg 2019). The photoelectric heating efficiency may deviate from this fiducial value right after the shock compression where the collisional ionization increases the electron fraction. Koyama & Inutsuka (2000) performed a high-resolution simulation (albeit one-dimensional) of shock propagation into the WNM. The postshock shock-heated WNM volume reaches n ∼ 5 cm−3, xe ∼ 0.1, and T ∼ 6400 K, so that Equation (13) predicts Γphot(1.0 Z) = 6 × 10−26 erg s−1. The factor of 3 difference from our fiducial value can expand the parameter space for the thermally stable WNM. However, note that a factor of a few uncertainty also exists in the assumption on the PAH size distribution and shape (Kim et al. 2023b). Also note that such high xe volume is limited to a 10−1 pc scale from the shock front by the recombination (i.e., Figure 3 of Koyama & Inutsuka 2000). Therefore, we opt to use the fiducial constant photoelectric heating efficiency (and its linear dependence on the metallicity) for our comparison between metallicities in this article.

Lastly, the interstellar radiation field varies across space. It is ideal to numerically investigate the 10 pc-scale local radiation field, e.g., zoomed-in approach consistently coupled with galactic-scale star formation (e.g., Peters et al. 2017; Kim et al. 2023a, 2023b), and this is left for future studies.

4.4. Applicability of Converging Flow Results to the ISM in Reality

Our results indicate that the solenoidal-mode-dominated turbulence is generally realized in molecular clouds undergoing supersonic compressions in the metallicity range 1.0–0.2 Z. The density inhomogeneity (and/or velocity fluctuation) in the converging flows can be larger (can exist) in reality, especially if the flows already have multiple phases over the one-phase WNM (Inoue & Inutsuka 2012; Iwasaki et al. 2019). Such a strong density/velocity inhomogeneity leads to stronger shock deformation and induces stronger shear motion into the shock-compressed layer. The stronger inhomogeneity in the inflow density/velocity induces stronger turbulence (e.g., the systematic survey by Kobayashi et al. 2020), but the turbulent speed is expected to be on the order of the WNM sound speed (see Appendix C).

Expansion of multiple supernova remnants also induces cloud compressions as often observed in galactic-scale numerical simulations (Girichidis et al. 2016; Kim et al. 2023a) and may explain the bubble features in nearby galaxies observed in JWST (Watkins et al. 2023). Each compression event occurs at diverse angles at various times, and the generation of solenoidal mode turbulence locally occurs due to the deformed shock fronts. It is likely that the compressive mode of turbulence does not become dominant unless multiple compressions from various angles occur simultaneously with relatively uniform density or unless the cloud itself becomes massive enough to gravitationally collapse. Previous analytic studies also show that the solenoidal mode of turbulence can grow faster than the compressive mode even in a gravitationally contracting background (e.g., Higashi et al. 2021).

4.5. Implications for Cosmic Star Formation History

Our simulations show that physical properties of molecular clouds resemble one another if compared at the same time in units of the cooling time. What about those in lower-metallicity ranges, which are important in understanding the overall cosmic star formation history beyond the Magellanic Clouds? The metallicity dependence that we have shown comes primarily from the line coolings of [O i] (63.2 μm) and [C ii] (157.7 μm). This metallicity dependence in cooling rate is expected to continue down to Z ∼ 10−3 Z, until which [O i] and [C ii] are the dominant coolants to induce thermal instability. In contrast, the main heating process changes from the photoelectric heating to the X-ray and cosmic-ray heating at ∼0.1 Z. The heating rate in <0.1 Z becomes independent of the metallicity if we employ the same constant X-ray and cosmic-ray heating rate. Under such conditions in <0.1 Z, Bialy & Sternberg (2019) showed that the density of the thermally stable CNM scales roughly as nCNMZ−1 (see their Equation (18)). This results in a roughly constant cooling time in <0.1 Z. In an even lower metallicity range of Z ≲ 10−3 Z, the UV field strength impacts the growth of thermal instability (due to H2 dissociation), and thus the comprehensive investigations remain by also changing the UV field strength for future studies. Overall, we assume that, in the range of ≥0.1 Z, physical properties of molecular clouds are similar between metallicities at the same time measured in units of the cooling time; we require further investigation for the range of <0.1 Z.

Meanwhile, our results also indicate that a difference in cloud formation conditions (e.g., different n0 and Vinflow) is important in the formation of molecular clouds with different properties, which is required to comprehensively understand cosmic star formation history. For example, recent ALMA observations started to spatially resolve the galactic disk of star-forming galaxies at Cosmic Noon at the redshifts of 2–4. They showed the existence of molecular clouds in gravitationally unstable galactic disks and that clouds tend to have higher column density/higher star formation rate density (e.g., Σgas > 100 M pc−2 and ΣSFR > 1 M yr−1 kpc−2) than those in the Milky Way (e.g., Tadaki et al. 2018). Such properties resemble those observed in nearby luminous infrared galaxies (Kennicutt & De Los Reyes 2021), who possibly experience galaxy mergers. These indicate a possibility that some drastic mechanisms, such as galaxy mergers with high velocity, form high column density molecular clouds that host active star formation at Cosmic Noon.

In addition, even starting with the same cloud formation condition to form a similar molecular cloud across the metallicity, the variation of the stellar initial mass function with the metallicity, if any, may arise from subsequent fragmentation in the collapse of clouds and circumstellar disks. We remark that some previous authors investigated this phenomena with numerical simulations using the same initial cloud properties for all metallicities (e.g., Bate 2019; Chon et al. 2021).

For the time being, the above discussions are merely speculations. Further simulations with various initial density, inflow velocity, etc., are needed and left for future studies.

5. Summary and Prospects

To understand the metallicity dependence of molecular cloud formation at its initial stages, we perform and compare the MHD simulations of the WNM supersonic converging flows with 20 km s−1 on 10 pc scales in 1.0, 0.5, and 0.2 Z environments. We impose the mean magnetic filed strength of 1 μG as representative strength in the metallicity range from the Milky Way to the Magellanic Clouds. The field orientation is parallel to the supersonic flows, which is a promising configuration for efficient molecular cloud formation. The flow forms a shock-compressed layer sandwiched by two shock fronts, within which thermal instability occurs to develop the multiphase ISM. We employ the 0.02 pc spatial resolution to resolve the thermal instability and turbulent structures.

We summarize our findings as follows:

  • 1.  
    The development of the CNM structure in the shock-heated WNM requires a longer time in the lower-metallicity environment where the typical tcool is almost inversely proportional to the metallicity. The CNM thermal states at different metallicities are similar if compared at the same time measured in units of the cooling time, instead of the same physical time.
  • 2.  
    The typical field strength of magnetic fields evolves gradually with 〈B〉 ∝ n1/5 up to B ∼ 11 μG at n ∼ 103 cm−3. This is consistent with Zeeman measurements of low (column) density regions of molecular clouds in the Milky Way, and Faraday rotation measurements toward the ISM in the LMC and SMC.
  • 3.  
    The postshock turbulent pressure balances against the inflow ram pressure (panel (c) of Figure 6). The turbulent velocity is slower in a lower-metallicity environment (albeit the difference is within a factor of 2), which is consistent with the line width of molecular clouds observed in the LMC/SMC.
  • 4.  
    The velocity power spectrum follows Kolmogorov's law if averaged over the entire volume of the shock-compressed layer, while two-point velocity correlation of the CNM volume alone exhibits a transition toward Larson's law.
  • 5.  
    At all metallicities after 1.0 tcool, the solenoidal (compressive) mode of the turbulence in the shock-compressed layer accounts for >80% (<20%, respectively) of the total turbulence power. This indicates that, even in a galactic-scale converging region, forming molecular clouds are always solenoidal-mode dominated. Therefore, a galactic-scale compressive motion is important for molecular cloud formation, but it does not immediately insinuate an enhancement of star formation efficiency by enhancing compressive motion in molecular clouds.
  • 6.  
    The CNM clump mass function has a power-law distribution of ${dn}/{dm}\propto {m}^{-1.7}$, which can be explained by the thermal instability growth under the Kolmogorov turbulence background.
  • 7.  
    These results suggest the common existence of hierarchical thermal and turbulent structure in molecular cloud precursors in the 1.0–0.2 Z range. The WNM/UNM components occupy most of the volume with strong turbulence of 4–10 km s−1. Meanwhile, the CNM component has an inter-clump velocity of 3–5 km s−1 as well as an internal velocity dispersion of ≲1 km s−1 within individual clumps (Figure 11).
  • 8.  
    We expect that this similarity in molecular cloud properties across the metallicity at the same t/tcool continues to hold down to Z ∼ 10−3 Z, because the dominant coolants are [O i] (63.2 μm) and [C ii] (157.7 μm) until this metallicity. Meanwhile, this indicates that some different cloud formation condition (e.g., different n0 or Vinflow due to galaxy mergers) is required to form molecular clouds with a higher column density/higher star formation rate density, as observed in luminous infrared galaxies and star-forming galaxies at Cosmic Noon.
  • 9.  
    Our results show that, in lower-metallicity environments, a longer physical time is required for the development of CNM structures out of the pure WNM. This indicates that, at the formation stage of molecular clouds out of the WNM in low-metallicity environments, the preexistence of CNM structure in the WNM volume controls the formation site and the mass of molecular clouds.

Our calculation is still limited to the early phase of molecular cloud formation. Investigating further compression in low-metallicity environments is left for future studies. In a 1.0 Z environment, it is known that the efficiency of molecular cloud formation depends on the inclination of the mean magnetic field against the inflow (e.g., Inoue & Inutsuka 2009; Iwasaki et al. 2019). We are planning to investigate a similar dependence of the magnetic field geometry in our forthcoming article.

Collisions between flows with different metallicities are ubiquitous in the context of galaxy mergers, including the LMC–SMC tidal interaction. It is interesting to investigate the spatial and temporal variation of the metallicity in the shock-compressed layer created by WNM flows with different metallicities, but this is left for future studies.

Studying cloud formation in extremely low-metallicity environments down to 10−4 Z is also important to reveal the initial conditions of star formation in young galaxies. For example, the formation of close binaries of massive stars in such environments is interesting as an origin of massive binary black holes whose coalescence events are observed by gravitational waves (Abbott et al. 2016). Previous numerical studies in this context start with cosmological initial conditions (e.g., Safranek-Shrader et al. 2014a, 2014b), or with an idealized model of a star-forming core without its formation process (e.g., Chon et al. 2021; a supercritical Bonnor-Ebert sphere of 104 M with n ∼ 104 cm−3), or focus on the kiloparsec-scale thermal instability without resolving core scales ∼0.1 pc (e.g., Inoue & Omukai 2015). We are planning to reveal the formation and internal structure of such star-forming clouds as an extension of our studies in this article. The dependence on radiation field strength and cosmic-ray/X-ray intensities is another important parameter in such conditions (see Susa et al. 2015), but this is also left for future studies (see also Section 4.2).

Acknowledgments

We appreciate the reviewer for providing a careful reading and helpful comments, which improved this manuscript. Numerical computations were carried out on Cray XC50 at Center for Computational Astrophysics, National Astronomical Observatory of Japan. M.I.N.K. (JP18J00508, JP20H04739, JP22K14080), K.I. (JP19K03929, JP19H01938, P21H00056), K.T. (JP16H05998, JP16K13786, JP17KK0091, JP21H04487), T.I. (JP18H05436, JP20H01944), K.O. (JP17H01102, JP17H06360, JP22H00149), and K.T. (JP21H00049, JP21K13962) are supported by Grants-in-Aid from the Ministry of Education, Culture, Sports, Science, and Technology of Japan. We appreciate Atsushi J. Nishizawa and Chiaki Hikage for helping in our Fourier analysis, and appreciate Kohei Kurahara for discussions on Faraday rotation measurements toward the Magellanic Clouds. We are grateful to Tomoaki Matsumoto, Hajime Susa, Sho Higashi, Gen Chiaki, Hajime Fukushima, Shu-ichiro Inutsuka, Shinsuke Takasao, Tetsuo Hasegawa, Kengo Tachihara, and Jeong-Gyu Kim for fruitful comments. We are grateful to Hiroki Nakatsugawa, who contributed to the early phase of this study through his master thesis work.

Appendix A: Heating and Cooling Rate

We calculate the metallicity-dependent net cooling rate as

Equation (A1)

where mgas = μM mp. The heating part consists of the photoelectric heating, X-ray heating, and cosmic-ray heating as

Equation (A2)

Here, the photoelectric heating rate is proportional to the metallicity because it is dominated by dust grains (Bakes & Tielens 1994). On the other hand, X-ray and cosmic-ray heating rates do not depend on the metallicity because they mostly heat hydrogen directly.

The cooling part consists of cooling due to the Lyα, CII, He, C, O, N, Ne, Si, Fe, and Mg lines (see also Koyama & Inutsuka 2000; Kobayashi et al. 2020). The functional form of the total cooling rate normalized in units of Γphot(1.0 Z) is

Equation (A3)

Here, the cooling rate by Lyα and He does not scale with the metallicity, while the other coolings due to the metals' lines are proportional to the metallicity.

Appendix B: Cooling Time

B.1. The Typical Cooling Time

The cooling time, $P/(\gamma -1)/\rho { \mathcal L }$, varies locally with space and time, depending on the local density/temperature variation. Nevertheless, the typical cooling time of this converging flow system can be estimated based on the typical state of the shock-heated WNM as follows.

To determine the representative thermal state, let us start with a simple estimation in a perfect one-dimensional shock compression. Once the inflow WNM passes the shock front, the density in the downstream increases to n ≃ 4n0 = 2.3 cm−3. Meanwhile, the temperature also increases to TPram/kB/(4n0) = 1.5 × 104 K, but it rapidly decreases to T ≃ T0 because of the efficient coolings by Lyα and heavy metals' lines. For example, the temperature decreases on the timescale of 2.0 × 10−4 Myr, 4.1 × 10−4 Myr, and 1.0 × 10−3 Myr at 1.0, 0.5, and 0.2 Z based on the T > 14,577 K regime of Equation (A3) with n = 4n0 and T = Pram/kB/(4n0). These timescales are shorter than the typical sound crossing time on a single numerical cell of 0.02 pc size with T = Pram/kB/(4n0), which is 1.2 × 10−3 Myr. Therefore, we need to perform simulations with a higher spatial resolution to confirm whether this shock-heated WNM evolves isochorically or isobarically during the return to T≃T0. In the following, let us assume that this rapid cooling leads to rather isochorical evolution and employ n = 4n0 and T = T0 as the representative initial density and temperature to estimate the typical tcool (see also the discussions in Appendix B.2). The corresponding pressure, 4n0 kB T0, is smaller than the ram pressure by roughly a factor of 2 (P/kBPram/2kB = 1.75 × 104 erg cm−3). This is consistent with most of the volumes in our simulations as shown in Figure 4. Such a thermal pressure smaller than the ram pressure is also achieved because the turbulent pressure almost balances with the inflow ram pressure, as shown in Figure 6.

Starting with n = 4n0 and T = T0, the net cooling rate during the subsequent thermal evolution is mostly determined by the CII cooling given by the first equation of Equation (A3):

Equation (B1)

Therefore, the cooling time can be estimated in general as

Equation (B2)

Equation (B3)

with n = 4n0, T = T0, and P/kB = Pram/2kB; the typical cooling time is

Equation (B4)

Equation (B5)

Equation (B4) gives the typical value of tcool as 1.01 Myr at 1.0 Z, 2.03 Myr at 0.5 Z, and 5.11 Myr at 0.2 Z, as summarized in Table 1. Equation (B5) gives a rough metallicity dependence.

B.2. Dependence of the Typical Cooling Time on the Initial Conditions

In the current article, we focus on the metallicity dependence starting with fixed initial conditions. Therefore, it is left for future studies to investigate the exact dependence of tcool on the initial conditions other than the metallicity. This requires parameter surveys by changing the mean density, the inflow velocity, the magnetic field strength, the inclination of the mean magnetic fields, the UV background, etc. The resultant impacts by all of these changes are coupled to each other because they modify the ram pressure, the turbulent pressure, and the following resultant shock-heated WNM state.

Nevertheless, as a first step, let us list several possible dependences of the two parameters, n0 and Vinflow. For simplicity, we here assume that the metallicity dependence comes only from the cooling/heating rate and is always independent from the choice of n0 and Vinflow. The other conditions are the same as our simulation setup in this article (e.g., the WNM inflow is parallel to the orientation of the mean magnetic field with B = 1 μG, the injected WNM is on the thermally stable state, etc.). These simplistic assumptions have to be investigated by further simulations, which is left for future studies.

Suppose that the postshock pressure, density, and temperature have dependences of $P\propto {n}_{0}^{\alpha }{V}_{\mathrm{inflow}}^{\beta }$, $n\propto {n}_{0}^{\gamma }{V}_{\mathrm{inflow}}^{\delta }$, and $T=P/{{nk}}_{{\rm{B}}}\propto {n}_{0}^{\alpha /\gamma }{V}_{\mathrm{inflow}}^{\beta /\delta }$, respectively, then

Equation (B6)

If we employ n = 4n0 and T = T0 as the shock-heated WNM state right behind the shock, as we did in Appendix B.1, α = γ = 1 and β = δ = 0. The cooling time is

Equation (B7)

We employ Equation (B7) as our fiducial dependence to derive Equation (12).

Alternatively, most of the shock-heated WNM volume evolves close to T = T0 even beyond n > 4n0 (Figure 4), and thus we may consider that an isothermal shock approximation describes the overall postshock state. In this case, α = γ = 1 and β = δ = 2. The cooling time is

Equation (B8)

accordingly.

We may also consider a simpler condition with P = Pram and T = T0, and n = Pram/kB/T. This gives α = β = δ = 2 and γ = 1, and the cooling time is

Equation (B9)

However as shown in our simulations, such a perfect isothermal compression is limited to the central volume during the very early stage when the shock compression is still close to a plane-parallel geometry without significant deformation (see panel (c) of Figure 3 and panel (c) of Figure 4).

Lastly, we would like to note that the above arguments assume the maximum pressure of the thermally stable state is independent of the metallicity. The flow ram pressure in the <0.1 Z environments must be higher (e.g., with a higher n0, a faster Vinflow) to successfully overcome this maximum pressure to enter the thermally unstable regime, because the maximum pressure depends on the metallicity roughly as ∝ 1/Z (see Figure 6 of Bialy & Sternberg 2019).

Appendix C: Turbulent Velocity

Panel (a) of Figure 6 shows that the turbulent velocity at earlier stages is much slower than the WNM sound speed, especially in lower-metallicity environments (e.g., ∼2 km s−1 in 0.2 Z at t ≲ 0.5tcool). The shock front configuration controls this turbulence strength. As we discussed in the first two paragraphs of Section 3.2, CNM forms more slowly in lower-metallicity environments and most of the volume resides in the shock-heated WNM state for longer physical times. Its thermal pressure, comparable to the inflow ram pressure, keeps the shock fronts in the plane-parallel configuration (see panels (a)–(c) of Figure 3 and 4). Such a configuration decelerates the inflow more efficiently than deformed shock fronts, and the postshock volume becomes denser and less turbulent. The physical state of the shock-heated WNM in this efficient deceleration regime can be estimated with a one-dimensional isothermal shock jump condition as

Equation (C1)

(see also panel (c) of Figure 4 and the discussion after Equation (B.2) for the validity of the one-dimensional isothermal approximation). Here v1 is the postshock flow velocity, n1 is the postshock density, P1 is the postshock thermal pressure, and ${ \mathcal M }$ is the inflow Mach number. The setup of the injected WNM expects v1 = 2.0 km s−1 and n1 = 5.8 cm−3, which can especially be seen in panel (c) of Figure 4. This is not prominent in higher-metallicity environments because the physical time of the plane-parallel shock configuration is shorter due to the efficient transition from the WNM to the CNM (i.e., the resultant interaction between CNM clumps and shock fronts induces more deformation of the shock fronts at earlier timing). In addition, in lower-metallicity environments, the shock-heated WNM travels a greater distance from the shock front when they become the CNM. Dense CNM clumps thus tend to form in the central region in lower-metallicity environments (panel (e) of Figure 3) and the turbulent development delays. This contributes to the difference between the metallicities in panels (a) and (b) of Figure 6 even after normalization with tcool until the turbulence is fully developed at t ∼ 3tcool.

The turbulent velocity increases in time once the shock fronts start to deform. Kobayashi et al. (2020) performed a systematic survey to show that the level of the shock deformation changes with the strength of the inflow density inhomogeneity. The turbulent velocity is, however, limited on the order of the WNM sound speed, and it does not exceed 10 km s−1. This is a natural consequence of the oblique shocks. As can be seen in our current simulation and in Kobayashi et al. (2020), the typical spatial scale of the shock deformation reaches the system size (see also Figure 8 of Kobayashi et al. 2020). Therefore, if we consider α ∼ 45° as the typical shock angle against the inflow, the oblique isothermal shock jump conditions provide the postshock physical states as

Equation (C2)

Here v1 is the postshock velocity component normal to the shock front, and this gives v1 ≃ 4.1 km s−1. The angle is locally smaller (α < 45°) to generate faster flow, but v1 > 10 km s−1 can be achieved when the inflow is parallel to the shock front with α ≤ 27°. Such a closely parallel configuration does not occur throughout the entire area of the shock front, so that the volume-averaged postshock velocity dispersion remains on the order of the WNM sound speed.

Footnotes

  • 9  

    See also Bialy & Sternberg (2019) and Inoue & Omukai (2015) for the effect of the H2 cooling and the H2 dissociation by UV radiation, as well as Omukai et al. (2005) and Chon et al. (2022) for the heating by the cosmic microwave background radiation at high redshifts and Omukai (2001) for atomic line cooling in the first star formation.

  • 10  

    Throughout this paper, 〈·〉 represents the volume-weighted average whereas 〈·〉ρ represents the density-weighted average.

  • 11  

    Note that the turbulence at n ∼ 103 cm−3 is typically super-Alfvénic. $\sqrt{\langle \delta {v}^{2}\rangle }$ ($\sqrt{\langle \delta {v}^{2}{\rangle }_{\rho }}$) roughly traces the turbulence of the WNM (CNM) because the volume is dominated by the WNM whereas 50% of the mass is locked in the CNM (Kobayashi et al. 2020). Therefore, combined with the slow evolution of 〈B〉 with n, the turbulent Alfvénic Mach number typically ranges from ${{ \mathcal M }}_{A}\simeq 1.6$–9.9.

  • 12  

    We hereafter denote the spatial frequency k/2π as the inverse of the wavelength so that the wave with k/2π = 0.1 pc−1 has a wavelength of 10 pc.

  • 13  

    The thermal instability also grows on scales smaller than 0.1 pc, but the sharp cutoff on <0.1 pc originates from our criteria of ≥ 64 member cells to avoid any numerical noise on that scale.

  • 14  

    The CNM mass in the shock-compressed layer at 15 Myr at 0.2 Z is 1.1 × 103 M in our simulation.

  • 15  

    Note that the CNM formation efficiency is higher in lower-metallicity environments, as we discussed in Section 3.2 This variation is, however, limited compared with the difference of the cooling time between metallicities (i.e., Z−1 in Equation (11)). In our simulations, the total mass of the CNM at 3tcool is 69 M at 1 Z, 174 M at 0.5 Z, and 694 M at 0.2 Z, respectively.

  • 16  

    In Equation (11), we count all of the CNM mass with n > 102 cm−3. This therefore gives a typical maximum mass of formed molecular clouds.

  • 17  

    See also Figure 2 of Bialy & Sternberg (2019) for the transition from the photoelectric-heating-dominated regime to the cosmic-ray-heating-dominated regime.

  • 18  

    See also Equations (19) and (20) in Wolfire et al. (1995) and Equations (32)–(34) in Bialy & Sternberg (2019) for the exact form; see also Bakes & Tielens (1994).

Please wait… references are loading.
10.3847/1538-4357/ace34e