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

Saturation level of turbulence in collapsing gas clouds

Sho Higashi Department of Physics, Konan University, Okamoto, Kobe, Japan Hajime Susa Department of Physics, Konan University, Okamoto, Kobe, Japan Gen Chiaki Astronomical Institute, Tohoku University, Aoba, Sendai, Japan National Astronomical Observatory of Japan, Mitaka, Tokyo, Japan E-mail:[email protected]
(Received ????; Revised ????; Accepted ????)
Abstract

We investigate the physical mechanism that decides the saturation level of turbulence in collapsing gas clouds. We perform a suite of high-resolution numerical simulations following the collapse of turbulent gas clouds with various effective polytropic exponents γeff\gamma_{\rm eff}, initial Mach numbers 0\mathcal{M}_{0}, and initial turbulent seeds. Equating the energy injection rate by gravitational contraction and the dissipation rate of turbulence, we obtain an analytic expression of the saturation level of turbulence, and compare it with the numerical results. Consequently, the numerical results are well described by the analytic model, given that the turbulent driving scale in collapsing gas clouds is one-third of Jeans length of collapsing core. These results indicate that the strength of turbulence at the first core formation in the early universe/present-day star-formation process can be estimated solely by γeff\gamma_{\rm eff}.

journal: ApJsoftware: Enzo (Bryan et al., 2014; Brummel-Smith et al., 2019), Grackle (Smith et al., 2017), yt (Turk et al., 2011)

1 Introduction

Astrophysical fluids are well known to be turbulent and magnetized over a wide range of spatial scales, from planetary systems to galaxies. Turbulent motion and magnetic fields play important roles for the formation of astronomical objects. In particular, many earlier studies have shown that both of them crucially contribute to star formation. In fact, turbulence and magnetic fields affect the fragmentation process of parent gas clouds or accretion disks around protostars, in various environments, from ancient to present-day star formation sites.

The first generation of stars, termed Population III (Pop III) stars, are believed to form in host dark matter minihalos with masses M106MM\sim 10^{6}M_{\odot} at redshifts z10z\gtrsim 10 in the Λ\LambdaCDM paradigm (Haiman et al., 1996; Tegmark et al., 1997; Nishi & Susa, 1999; Fuller & Couchman, 2000; Abel et al., 2002; Bromm et al., 2002; Yoshida et al., 2003). Due to the lack of heavy elements, which are efficient coolants, Pop III forming clouds are more massive and hotter than present-day clouds. As a result, the typical mass of Pop III stars (101000M)(\sim 10-1000M_{\odot}) should be higher than present-day stars because of the larger fragmentation scale and higher mass accretion rate (e.g. Susa et al., 1996; Omukai & Nishi, 1998; Yoshida et al., 2008). However, recent numerical studies for the mass accretion phase have shown that less massive stars (1M)(\sim 1M_{\odot}) form through the fragmentation of accretion disks (so-called disk fragmentation) (Clark et al., 2011b; Greif et al., 2011, 2012; Susa, 2013, 2019; Susa et al., 2014; Stacy et al., 2016; Hirano & Bromm, 2017; Inoue & Yoshida, 2020; Chiaki & Yoshida, 2022). Besides, other previous studies have demonstrated that disk fragmentation is promoted by turbulence (Clark et al., 2011a; Riaz et al., 2018; Wollenberg et al., 2020) or suppressed by magnetic fields (Machida & Doi, 2013; Sharda et al., 2020; Sadanari et al., 2021). Thus, the strength of turbulence and magnetic fields is crucial for determining the initial mass function (IMF) of Pop III stars.

On the other hand, in present-day galaxies, stars form from rotating molecular cloud cores with typical temperatures of 10\sim 10 K and magnetic fields of a few tens of μG\mu\mathrm{G}, composed of molecular hydrogen (H2\mathrm{H_{2}}), carbon oxide (CO), ammonia (NH3\mathrm{NH_{3}}), dust grains, etc. Depending on the properties of their environments, the cores have various masses from low (<10M<10M_{\odot}) to high (>100M>100M_{\odot}) and turbulent velocities from sub- to supersonic. Accordingly, the masses of present-day stars lie over a wide range with a power-law distribution (e.g., Salpeter, 1955; Kroupa et al., 2001; Chabrier, 2005). Such a cloud core distribution could be modeled by magneto-turbulence in parent molecular clouds (e.g., Padoan & Nordlund, 2002) which is reflected in IMF, while disk fragmentation after the formation of hydrostatic cores may also modify the IMF. Thus, the turbulence and magnetic fields are also crucial for the star formation process in the present-day universe.

As for the amplification of turbulence, previous studies have shown that gravitational contraction can take on that task (Vázquez-Semadeni et al., 1998; Robertson & Goldreich, 2012; Murray & Chang, 2015; Birnboim et al., 2018; Guerrero-Gamboa & Vázquez-Semadeni, 2020; Mandal et al., 2020; Hennebelle, 2021). In our latest study, Higashi et al. (2021) (hereafter HSC21), we numerically followed the evolution of turbulence in collapsing polytropic/barotropic clouds during the collapse phase. We also obtained an analytic formula that properly describes the growth of turbulence in collapsing clouds. In these simulations, we also found that the root-mean-square of the turbulent velocity can reach a few times the sound speed and eventually saturates, even when the initial seed field is feeble. Although we found the saturation level seems to depend on the effective polytropic exponent, we did not step into detailed mechanisms.

In this paper, we investigate the saturation mechanisms of amplified turbulence by gravitational collapse in detail. For this purpose, we first perform numerical simulations following the cloud collapse with various initial conditions. We describe the setup of the collapse simulations in Section 2, and we present our results in Section 3. We then theoretically estimate the saturation level of turbulence in Section 4 and compare our numerical results and theoretical estimates in Section 5. Section 6 is devoted to discussion. We summarize the main points in this paper in Section 7.

2 Collapse Simulations

Here, we perform a suite of high-resolution three-dimensional collapse simulations by using the N-body/adaptive mesh refinement (AMR) cosmological hydrodynamics code enzo (Bryan et al., 2014; Brummel-Smith et al., 2019)111http://enzo-project.org/. This code solves the compressive hydrodynamic equations with the piecewise parabolic method (PPM) in an Eulerian frame using an HLLC Riemann solver. The basic equations are

ρt+(ρ𝒗)\displaystyle\frac{\partial\rho}{\partial t}+\nabla\cdot(\rho\mbox{\boldmath$v$}) =\displaystyle= 0,\displaystyle 0, (1)
ρ𝒗t+ρ(𝒗)𝒗\displaystyle\frac{\partial\rho\mbox{\boldmath$v$}}{\partial t}+\rho(\mbox{\boldmath$v$}\cdot\nabla)\mbox{\boldmath$v$} =\displaystyle= Pρϕ,\displaystyle-\nabla P-\rho\nabla\phi, (2)
Et+(E+P)𝒗\displaystyle\frac{\partial E}{\partial t}+\nabla\cdot\left(E+P\right)\mbox{\boldmath$v$} =\displaystyle= ρ𝒗ϕΛ+Γ.\displaystyle-\rho\mbox{\boldmath$v$}\cdot\nabla\phi-\Lambda+\Gamma. (3)

In these equations, E,ρ,𝒗,ϕ,ΛE,~{}\rho,~{}\mbox{\boldmath$v$},~{}\phi,~{}\Lambda, and Γ\Gamma are the total energy density, density, velocity, gravitational potential, radiative/chemical cooling rate, and radiative heating rate, respectively. The total energy density is given by E=e+ρv2/2E=e+\rho v^{2}/2, where ee is the thermal energy density. Remark that we solve Equation (3) only for a model with primordial chemistry (§2.2) but not for barotropic runs (§2.1). The gravitational potential ϕ\phi is obtained from the Poisson’s equation below with a Fast Fourier Transform (FFT) technique (Hockney & Eastwood, 1988)

2ϕ=4πGρ.\nabla^{2}\phi=4\pi G\rho. (4)

We generate a gravitationally unstable Bonner-Ebert sphere with a central density ρpeak,0=4.65×1020gcm3\rho_{\rm peak,0}=4.65\times 10^{-20}~{}\mathrm{g~{}cm^{-3}}, uniform temperature T0=200KT_{0}=200~{}\mathrm{K}, and cloud radius rcr_{\rm c} = 1.5 pc as an initial condition. This is the same initial condition as HSC21. The computational domain has a size of 4 rcr_{\rm c} with periodic boundary conditions.

We start the calculations with a base grid with 2563256^{3} cells and progressively refine cells as the cloud collapses in order to resolve the Jeans length at least by 128 cells (we refer to this minimum refinement criterion as a ‘Jeans parameter’). Except for the model with γeff=1.3\gamma_{\rm eff}=1.3 (explained later), we terminate simulations when the maximum cloud density reaches 104gcm310^{-4}~{}\mathrm{g~{}cm^{-3}}. By the end of the simulations, the maximum refinement level reaches 27 for γeff=1.0\gamma_{\rm eff}=1.0 and 20 for γeff=1.25\gamma_{\rm eff}=1.25. The corresponding minimum spatial resolutions are 3.0×1053.0\times 10^{-5} AU and 0.038 AU, respectively. The maximum levels in the other runs are in the range of 20-27. We use the yt toolkit (Turk et al., 2011)222https://yt-project.org/ to analyze the simulation data.

Table 1: Properties of simulations
Name Mach number Initial rotation γeff\gamma_{\rm eff} turbulence seed Remark
LowA 0.1 No All A fiducial
LowB 0.1 No All except 1.25 and 1.3 B
LowC 0.1 No All except 1.25 and 1.3 C
LowA w/c 0.1 No With chemistry A
Middle 0.5 No 1.09 A
Middle w/r 0.5 Yes 1.09 A
High 1.0 No 1.09 A
“Low”, “Middle”, and “High” correspond to the strength of initial Mach numbers. ‘w/c’ and ‘w/r’ represent the
model with chemical reactions and initial rotation, respectively. Note: we run simulations for γeff=1.25\gamma_{\rm eff}=1.25
and 1.31.3 only with seed A because of high computational cost.

2.1 Barotropic EoS

We calculate the temperature evolution of the collapsing gas by using a simplified polytropic model as HSC21. The relation between the pressure PP and the density ρ\rho of gas is expressed as

PργeffP\propto\rho^{\gamma_{\rm eff}} (5)

by using an effective polytropic exponent γeff\gamma_{\rm eff}. In this model, we use this relation instead of solving Equation (3). We perform simulations for eight γeff\gamma_{\rm eff}: 1.0, 1.05, 1.09, 1.1, 1.15, 1.2, 1.25 and 1.3. The parameters γeff=\gamma_{\rm eff}=1.0, 1.09 and 1.2 were also employed in HSC21, which enables us to compare the results directly. Note that we terminate the simulation when the mean density within the core reaches 1010gcm310^{-10}~{}\mathrm{g~{}cm^{-3}} only for γeff=1.3\gamma_{\rm eff}=1.3 because of high computational cost. Here, the models for γeff=1.09\gamma_{\rm eff}=1.09 and 1.11.1 mimic the primordial gas in minihalos (Omukai & Nishi, 1998). In these models, we set the gaseous mean molecular weight μ=1.22\mu=1.22.

2.2 The model with primordial chemistry

Although not explicitly calculated in the barotropic EoS, solving the entropy production/reduction by hydrodynamic shock or radiative cooling, may change the amplification/saturation level of turbulence through the baroclinic term of the vorticity equation (e.g., Wise & Abel, 2007). In order to investigate the effects of baroclinic term on the evolution of turbulent velocities, we explicitly take into account non-equilibrium chemistry and radiative cooling of primordial gas. For this, we solve 49 chemical reactions of 15 primordial species (e,H,H+,H,H2,H2+,He,He+,He2+,HeH+,D,D+,D,HD\mathrm{e,~{}H,~{}H^{+},~{}H^{-},~{}H_{2},~{}H_{2}^{+},~{}He,~{}He^{+},~{}He^{2+},~{}HeH^{+},~{}D,}\\ \mathrm{D^{+},~{}D^{-},~{}HD}, and HD+\mathrm{HD^{+}}) by using grackle (Smith et al., 2017)333https://grackle.readthedocs.io/ modified/extended in Chiaki & Wise (2019) (see Chiaki & Wise 2019 in detail). In this model, μ\mu is directly calculated from the abundance of the chemical species. In the high-density region with a number density above 1017cm310^{17}~{}\mathrm{cm^{-3}}, the time step of non-equilibrium chemical solver becomes prohibitively smaller than the dynamical time, resulting in a huge computational cost. Hence, in this region, we calculate the abundances of hydrogen and helium by equilibrium chemical solver using the Saha-Boltzman equation. As the initial condition, we assume the mass fractions of H2,H+,H\mathrm{H_{2},~{}H^{+},~{}H}, and He\mathrm{He} as xH2=7.60×104x_{\rm H_{2}}=7.60\times 10^{-4}, xH+=7.60×108x_{\rm H^{+}}=7.60\times 10^{-8}, xH=0.759x_{\rm H}=0.759, and xHe=0.240x_{\rm He}=0.240, respectively.

2.3 Initial turbulent flow

It is known that collapsing gas cores have some degree of turbulent velocities at the onset of run-away collapse (see the reviews of Greif 2015 for primordial clouds and McKee & Ostriker 2007 for present-day clouds).

Additionally, the power spectrum of the turbulent velocity follows the Larson’s law (P(k)|𝒗k|2k4P(k)\equiv\langle|\mbox{\boldmath$v$}_{\rm k}|^{2}\rangle\propto k^{-4}) in both present-day (Solomon et al., 1987) and primordial gas clouds (Prieto et al., 2011), where 𝒗k\mbox{\boldmath$v$}_{\rm k} and kk are the velocity in the wavenumber space and a wavenumber, respectively. \langle\cdots\rangle denotes the average of a quantity. In order to mimic the turbulent flow of the initial cloud, we give the initial velocity field following the Larson’s law, by varying the initial root-mean-square Mach number.

Thus, the 1D kinetic energy spectrum E(k)k2P(k)E(k)\propto k^{2}P(k) is proportional to k2k^{-2}, which has the same power-law index of Burgers (compressible) turbulence (Burgers, 1948) 444Strictly speaking, this initial velocity field is not self-consistent “real” turbulence since there is no correlation between density and velocity. See section 6.3..

This turbulent velocity field is composed of the natural mixture of solenoidal (divergence-free) and compressive (curl-free) modes with a 2:1 ratio. It is the reasonable initial condition, given that transverse modes account for 2/3 of the spatial directions in a 3D system, and longitudinal modes account for the remaining 1/3. Earlier studies showed that star-forming cores have a few percent of rotation energy against gravitational potential (e.g., Goodman et al., 1993; Yoshida et al., 2006). Therefore, we also perform the simulation for the more realistic case, where the rotational energy is 5% of the gravitational potential, to investigate the effects of initial rotation on the temporal evolution of the turbulent velocity.

To investigate the effect of the initial random seed of turbulence, we perform simulations with three different initial turbulent seeds (tagged “A”, “B” and “C”) for γeff=\gamma_{\rm eff}= 1.0, 1.05, 1.1, 1.15 and 1.2. We consider a single seed (A) for γeff=\gamma_{\rm eff}=1.25 and 1.3 because of the high computational cost. The properties of each simulation are summarized in Table 1.

In our analysis of the evolution of turbulence (Section 3), we calculate the cell-volume weighted root-mean-square turbulent velocity within a sphere with a radius of half the Jeans length LJL_{\rm J} (hereafter called the “Jeans volume”), as well as HSC21. To define a cloud core, it is necessary to evaluate the Jeans length because the length scale of a collapsing gas core is typically the Jeans length. However, the Jeans length depends on the mean density and sound speed within the core of a certain radius. Thus, we need to assess the core radius accurately at first. We calculate the core radius iteratively as follows

  1. 1.

    First, we cut out the overdense region of ρρth\rho\geq\rho_{\rm th}, where ρth=ρpeak/16\rho_{\rm th}=\rho_{\rm peak}/16 in this paper. ρpeak\rho_{\rm peak} is the maximum density in the computational domain.

  2. 2.

    Calculate the center of mass 𝒓c\mbox{\boldmath$r$}_{\rm c} in the region and ‘tentative’ Jeans length (=πcs,t2/Gρmean,t)\left(=\sqrt{\pi c_{\rm s,t}^{2}/G\rho_{\rm mean,t}}\right) using cell volume-weighted average density ρmean,t\rho_{\rm mean,t}, and sound speed cs,tc_{\rm s,t}. And then, we assume this Jeans length as tentative radius r0r_{0}.

  3. 3.

    Considering a sphere with a radius of r0r_{0} from 𝒓c\mbox{\boldmath$r$}_{\rm c}, calculate tentative Jeans length and radius r1r_{1} as well as 2.

  4. 4.

    If |r1r0||r_{1}-r_{0}| is smaller than the smallest cell width, we terminate this calculation. If not, we substitute r1r_{1} to r0r_{0} and go back to 3 to repeat until the above condition is fulfilled.

  5. 5.

    We obtain r1r_{1} as the core radius r1(=LJ/2)r_{1}(=L_{\rm J}/2) after convergence.

Within the Jeans volume VJV_{\rm J}, the turbulent velocity is defined as follows:

vturb2LJ/2ViVJ(𝒗i𝒗rad,i)2,v_{\rm turb}^{2}\equiv\sum_{\leq L_{\rm J}/2}\frac{V_{i}}{V_{\rm J}}(\mbox{\boldmath$v$}_{i}-\mbox{\boldmath$v$}_{{\rm rad},i})^{2}, (6)

where Vi,𝒗i,𝒗rad,iV_{i},~{}\mbox{\boldmath$v$}_{i},~{}\mbox{\boldmath$v$}_{\rm{rad},i} are the cell volume, the velocity, and smoothed radial velocity of the iith cell, respectively. To calculate the turbulent velocity, we estimate the smoothed background radial velocity through the following procedure:

  1. 1.

    Calculate the radial velocity profile with NradN_{\rm rad} radial bins. NradN_{\rm rad} should be smaller than the Jeans parameter so that we can remove the radial velocity fluctuations at the scale of the cell size. Here, we set Nrad=16N_{\rm rad}=16.

  2. 2.

    Linearly interpolate this profile to estimate the smoothed radial velocity 𝒗rad,i\mbox{\boldmath$v$}_{{\rm rad},i} at the position of each cell.

Refer to caption
Figure 1: Evolution of the turbulent velocity vturbv_{\rm turb} as a function of mean density ρmean\rho_{\rm mean} in the Jeans volume. The solid curves with different colors denote the numerical results for different γeff\gamma_{\rm eff}. The dotted lines denote the mean sound speed in the Jeans volume.

3 Results

In this section, we initially present the overall turbulence evolution in the models with the barotropic EoS focusing on the effects of the polytropic index, initial turbulent seed, and initial turbulent velocity in Section 3.1. We then compare the results with/without the baroclinic term and show its effects on the evolution of turbulence in Section 3.2.

3.1 Overall turbulence evolution in barotropic EoS

Figure 1 shows the evolution of the turbulent velocity as a function of the mean density ρmean\rho_{\rm mean} within the Jeans volume of each snapshot in the fiducial model (LowA). Note that we divide the results into two panels for visibility. As shown in HSC21, the turbulent velocities (solid curves) increase at the rate that depends on γeff\gamma_{\rm eff} initially. Eventually they saturate at ρmean1013gcm3\rho_{\rm mean}\simeq 10^{-13}~{}\mathrm{g~{}cm^{-3}} and then increase with the growth rate similar to the sound speed (dashed lines). Additionally, the ratio of the saturation velocity to the sound speed depends on γeff\gamma_{\rm eff} and decreases as γeff\gamma_{\rm eff} increases. The physical mechanism that determines these saturation levels is explained in sections 4 and 5 in detail.

Refer to caption
Figure 2: Same as Figure 1 but for the different initial seeds of turbulence (LowA, LowB, and LowC). Difference in line style corresponds to the difference of initial seeds and difference in colors corresponds to the difference of γeff\gamma_{\rm eff}. For ease of viewing, we multiply the results by factors for each γeff\gamma_{\rm eff} model.
Refer to caption
Figure 3: Same as Figure 1 but for the different 0\mathcal{M}_{0} and rotation energies (LowA, Middle, Middle w/r, and High). Solid curves are numerical results for each model. The magenta dotted line denotes the averaged sound speed in the Jeans volume for γeff=1.09\gamma_{\rm eff}=1.09. The black dashed line is obtained by an analytic estimate in HSC21 Equation (14).

Next, we plot the evolution of turbulent velocities with each initial turbulent seed in Figure 2. For ease of viewing, we multiply the results by factors for each γeff\gamma_{\rm eff}. We can see the growth rates of turbulent velocities before the saturation (ρmean1013gcm3)\left(\rho_{\rm mean}\lesssim 10^{-13}~{}\mathrm{g~{}cm^{-3}}\right) do not depend on the initial turbulent seed. We can also see that the final turbulent velocities are almost converged in all γeff\gamma_{\rm eff} and the difference in the initial seeds is a factor of two at most. The difference in Mach numbers among different seeds and its details are explained in Section 5.

Figure 3 shows the evolution of turbulent velocities for each model with different initial turbulent velocities. We also plot the results of the model with rotation (Middle w/r). In all the models, the turbulent velocities saturate at ρmean1013gcm3\rho_{\rm mean}\gtrsim 10^{-13}~{}\mathrm{g~{}cm^{-3}}. The saturation levels in these models almost converge and the difference among different initial strengths of turbulence or the effect of rotation are about 20% at most. This implies the saturation level of turbulent velocity does not depend on the initial strength of turbulence or the presence of cloud rotation.

We find that the rotational velocity is subdominant, compared to the turbulent component in our particular models in the entire course of collapse. Thus the resultant Mach number is not affected much by the initial rotation. However, we have to keep in mind that we must explicitly consider the rotation velocity if it is the dominant component. In such cases the definition of turbulent velocity (Equation 6) should be modified as Equation (18) in Chiaki & Yoshida (2022).

3.2 Comparison with/without baroclinic term

As mentioned in Section 2.1, γeff=1.09\gamma_{\rm eff}=1.09 and 1.11.1 well reproduce the temperature evolution of collapsing primordial gas clouds. Here we compare the temporal evolution of the turbulent velocities and the saturation levels between the models with the barotropic EoS and the model explicitly treating radiative cooling as shown in Figure 4.

Refer to caption
Figure 4: Same as Figure 3 but for barotropic models and the model with primordial cooling. The solid curves denote the numerical results for γeff=1.09\gamma_{\rm eff}=1.09 (red), γeff=1.1\gamma_{\rm eff}=1.1 (cyan), and the model with radiative cooling (green). The black dashed line is obtained by an analytic estimate in HSC21 Equation (14).

As well as the results of barotropic EoS runs (Figure 4) in HSC21, we can see that all of the results for LowA with γeff=1.09\gamma_{\rm eff}=1.09, 1.11.1, and w/c are in good agreement with the analytic estimate (dashed line) at ρmean1013gcm3\rho_{\rm mean}\lesssim 10^{-13}\mathrm{g~{}cm^{-3}}. We can also see that the turbulent velocities in all models saturate at ρmean1013gcm3\rho_{\rm mean}\gtrsim 10^{-13}~{}\mathrm{g~{}cm^{-3}} and converge. This shows that the growth and saturation of turbulence in the gravitational collapse is not affected greatly by the baroclinic term induced by shock waves and chemical reactions/radiative cooling but depends only on γeff\gamma_{\rm eff}.

4 Analytic estimate of turbulence saturation

We find the growth and saturation of turbulence depend solely on γeff\gamma_{\rm eff} in Section 3. In this section, extending the analytic model proposed by Robertson & Goldreich (2012), we analytically estimate the maximum saturation level of turbulence induced by gravitational collapse from the balance between the amplification and the decay due to dissipation of turbulence.

In the “adiabatic heating” model in Robertson & Goldreich (2012), the temporal evolution of rms turbulent velocity on the uniformly contracting background is expressed as

dvdt=Hvηv2a.\frac{dv}{dt}=-Hv-\eta\frac{v^{2}}{a\ell}. (7)

Here, aa, H(a˙/a)H(\equiv\dot{a}/a), η\eta, and \ell are the scale factor, “Hubble parameter”, dissipation coefficient of turbulence, and the typical driving scale of turbulence in comoving coordinates, respectively. In the uniformly changing background, the mean density field ρ\rho and aa are given by the relation ρa3\rho\propto a^{-3}.

The first-term on the right-hand side in Equation (7) denotes the amplification of turbulence due to gravitational collapse, while the second term represents the decay term due to dissipation, respectively. We can rewrite this equation by using the turbulent eddy turnover frequency ω=va\omega=\frac{v}{a\ell} as

dvda=(1+ηωH)va.\frac{dv}{da}=-\left(1+\eta\frac{\omega}{H}\right)\frac{v}{a}. (8)

Robertson & Goldreich (2012) also derived the relation between ω\omega and HH, from Equations (7) and (8),

dlog(ω/H)dlog(1/a)=(2+ηωH)dlogHdlog(1/a).\frac{d\log(\omega/H)}{d\log(1/a)}=\left(2+\eta\frac{\omega}{H}\right)-\frac{d\log H}{d\log(1/a)}. (9)

Here \ell is assumed to be a constant. If the collapse time tcollapse|H|1t_{\rm collapse}\sim|H|^{-1} is proportional to the free-fall time (1/ρ\propto 1/\sqrt{\rho}); the second term in Equation (9) becomes

dlog(H)dlog(1/a)=32.\frac{d\log(H)}{d\log(1/a)}=\frac{3}{2}. (10)

The asymptotic relation is approached as dlog(ω/H)/dlog(1/a)0d\log(\omega/H)/d\log(1/a)~{}\rightarrow~{}0. We have

2+ηωH320(forlog(1/a)).2+\eta\frac{\omega}{H}-\frac{3}{2}\rightarrow 0\quad(\mathrm{for}~{}\log(1/a)\rightarrow\infty). (11)

We apply this model to the collapsing gas core in a self-similar fashion. Assuming the driving scale in Equations (7) is a factor ff times Jeans length LJL_{\rm J}, and re-deriving the relation corresponding to Equation (9) and (10), we obtain

dlog(ω/H)dlog(1/a)=53γeff2+ηωH(forafLJ).\frac{d\log(\omega/H)}{d\log(1/a)}=\frac{5-3\gamma_{\rm eff}}{2}+\eta\frac{\omega}{H}~{}~{}~{}({\rm for}~{}a\ell\sim fL_{\rm J}). (12)

Here, the Jeans length is

LJ=πcs2Gρa3(2γeff)/2,L_{\rm J}=\sqrt{\frac{\pi c_{\rm s}^{2}}{G\rho}}\propto a^{3(2-\gamma_{\rm eff})/2}, (13)

where csc_{\rm s} is the sound speed of a collapsing cloud. The asymptotic value in the limit of ρ\rho\rightarrow\infty (or a0a\rightarrow 0)is

ω|H|=53γeff2η.\frac{\omega}{|H|}=\frac{5-3\gamma_{\rm eff}}{2\eta}. (14)

Next, we estimate |H|(=a˙/a=(1/3)(ρ˙/ρ))|H|(=\dot{a}/a=(1/3)(\dot{\rho}/\rho)) in the collapsing gas core. Since collapse proceeds in a self-similar fashion (see Appendix A), the growth of the central density is obtained by integrating a set of ordinary differential equations (Yahil, 1983; Suto & Silk, 1988). From the equations, the density ρ(r,t)\rho(r,t) as a function of radius rr from the cloud center and time tt is given as ρ(r,t)=α(x)/(4πGt2)\rho(r,t)=\alpha(x)/(4\pi Gt^{2}). α(x)\alpha(x) is a dimensionless quantity as a function of dimensionless coordinate xr/κtnx\equiv r/\sqrt{\kappa}t^{n}, where κ\kappa and nn are constants. See Suto & Silk 1988 in detail. At the center of the core (x=0)(x=0), the central density ρc\rho_{\rm c} is given as ρc=α(0)/(4πGt2)\rho_{\rm c}=\alpha(0)/(4\pi Gt^{2}), where α(0)\alpha(0) is a constant depending on γeff\gamma_{\rm eff} (see Table 2). In a self-similarly collapsing gas core, we assume the gas density to be uniform (ρmeanρc\rho_{\rm mean}\simeq\rho_{\rm c}), then we have

|H|=|13ρ˙ρ|=|23t1|=234πGρ1α(0).|H|=\left|\frac{1}{3}\frac{\dot{\rho}}{\rho}\right|=\left|\frac{2}{3}t^{-1}\right|=\frac{2}{3}\sqrt{4\pi G\rho}\frac{1}{\sqrt{\alpha(0)}}. (15)

Substituting this relation to Equation (14), we obtain

ω=53γeff3η4πGρmeanα(0).\omega=\frac{5-3\gamma_{\rm eff}}{3\eta}\sqrt{\frac{4\pi G\rho_{\rm mean}}{\alpha(0)}}. (16)

Then, with ω=v/fLJ\omega=v/fL_{\rm J}, we can derive the saturated turbulent velocity as,

vsat=53γeff3η2πcsα(0)f.v_{\rm sat}=\frac{5-3\gamma_{\rm eff}}{3\eta}\frac{2\pi c_{\rm s}}{\sqrt{\alpha(0)}}f. (17)

Finally, we obtain the maximum saturation Mach number of turbulence

sat53γeff3η2πα(0)f.\mathcal{M}_{\rm sat}\simeq\frac{5-3\gamma_{\rm eff}}{3\eta}\frac{2\pi}{\sqrt{\alpha(0)}}f. (18)

This expression depends on γeff\gamma_{\rm eff} but not on ρmean\rho_{\rm mean}. This feature nicely agrees with the numerical results obtained in the previous section as we see in Section 5.

Table 2: Calculated α(0)\alpha(0) for each γeff\gamma_{\rm eff}
γeff\gamma_{\rm eff} 1.0 1.05 1.09 1.10 1.15 1.2 1.25 1.3
α(0)\alpha(0) 1.67 2.15 3.03 3.26 4.73 7.19 11.6 22.0

5 Driving scale of turbulence

In this section, we compare the numerical results in Section 3 with the analytic estimate in Section 4. First, we solve the ordinary differential equations (Equations 15a and b) of Suto & Silk (1988) by using the fourth-order Runge-Kutta method to obtain α(0)\alpha(0), which is necessary to calculate sat\mathcal{M}_{\rm sat}. We summarize obtained α(0)\alpha(0) corresponding to each γeff\gamma_{\rm eff} in Table 2.

Besides, we also need the dissipation coefficient η\eta of turbulence. Mac Low (1999) performed the simulations of uniform, randomly driven turbulence and found that the driven turbulence saturates, resulting in the balance between energy injection and dissipation. He also found that the energy injection rate and the dissipation rate of turbulence can be well approximated by the linear relation, and derived the value of the dissipation coefficient. However, his simulations were performed only with the isothermal equation of state, so that the γeff\gamma_{\rm eff} dependence is still unclear. Therefore, we perform another set of simulations, based on the simulations of Mac Low (1999), to confirm the value and γeff\gamma_{\rm eff} dependence of η\eta. Consequently, we obtain η=0.42\eta=0.42 for any γeff\gamma_{\rm eff}, which is the same as Mac Low (1999), i.e. the dissipation coefficient has no γeff\gamma_{\rm eff} dependence. Thus, we use this value to obtain the saturation Mach numbers from Equation (18). The detailed setup and results of simulations are summarized in Appendix B.

Refer to caption
Figure 5: Saturation Mach number sat\mathcal{M}_{\rm sat} v.s. γeff\gamma_{\rm eff} (upper panel) and ratio of eddy turnover frequency ω\omega and the absolute value of Hubble parameter |H||H| v.s. γeff\gamma_{\rm eff} (lower panel). The red square, blue cross, and inverse triangle symbols denote the numerical results for 0=0.1\mathcal{M}_{0}=0.1 with seeds A, B, and C, respectively. The blue, orange, and green solid lines in the upper panel are estimated saturation Mach numbers using Equation (18) which assumes turbulence driving scale aLJ/2a\ell\sim L_{\rm J}/2, LJ/3L_{\rm J}/3, and LJ/4L_{\rm J}/4, respectively. The black dashed line in the lower panel is obtained from Equation (14).

The upper panel of Figure 5 shows the saturation Mach numbers for each γeff\gamma_{\rm eff} calculated from Equation (18). The red square, blue cross, and green inverse triangle symbols are obtained from our numerical results for 0=0.1\mathcal{M}_{0}=0.1 with the initial seeds A, B, and C, respectively. The black circle symbols are the average of the three models. As shown in Section 3, the turbulent velocities saturate at ρmean1013gcm3\rho_{\rm mean}\gtrsim 10^{-13}~{}\mathrm{g~{}cm^{-3}} in all models. Therefore, we regard the time-averaged Mach number in ρmean1011gcm3\rho_{\rm mean}\geq 10^{-11}~{}\mathrm{g~{}cm^{-3}} as the saturation Mach number sat\mathcal{M}_{\rm sat}. The blue, orange, and green solid curves denote sat\mathcal{M}_{\rm sat} estimated from Equation (18) with turbulence driving scales of aLJ/2a\ell\sim L_{\rm J}/2, LJ/3L_{\rm J}/3, and LJ/4L_{\rm J}/4, respectively.

We can see that, if the turbulent driving scale is one-third of Jeans length, the average of the three initial seeds (black circle symbols) are in good agreement with the theoretical estimate of the saturation Mach number. This is also almost consistent with the result of Guerrero-Gamboa & Vázquez-Semadeni (2020) for an isothermal core, who obtained the driving scale of 0.285 LJL_{\rm J} with the same dissipation coefficient as Mac Low (1999) but with a different theoretical model based on virial equilibrium of a core.

The lower panel of Figure 5 shows the ratio of eddy turnover frequency ω\omega and the absolute value of Hubble parameter |H||H| as a function of γeff\gamma_{\rm eff} in the saturation regime. The symbols indicate the results obtained with the same method as the upper panel, and the black dashed line is obtained from Equation (14). Unlike the solid curves in the upper panel, the dashed line in the lower panel has a weaker dependency against γeff\gamma_{\rm eff} because Equation (14) does not contain α(0)\alpha(0).

As well as the saturation Mach number, we can see that the average of the results with the three initial seeds (black circle symbols) are in good agreement with the theoretical estimate. These results provide clear evidence that the saturation mechanism of turbulence is a balance between the amplification and the decay of turbulence.

Important point of the present study is that our theory can well describe the numerical results of various γeff\gamma_{\rm eff}. This versatility of the theory supports its significance.

6 Discussion

6.1 On the PopIII formation

In this paper, we have focused on the turbulence in the runaway collapse phase. As a result, given an effective polytropic index γeff\gamma_{\rm eff}, we find an analytic formula to assess the saturation level of turbulence in collapsing cores. In case we consider the primordial star formation which is well approximated by γeff1.09\gamma_{\rm eff}\simeq 1.09, the turbulence will be supersonic sat1.52\mathcal{M}_{\rm sat}\sim 1.5-2 (Figure 5 and Equation 18), which is consistent with the results of earlier works (e.g., Greif et al., 2012). Therefore, supersonic turbulence should be a general feature of the collapsing star-forming clouds in primordial environments.

Not only turbulence but also magnetic fields can have substantial impacts on the IMF of the first stars if they are sufficiently strong. It has been considered that magnetic fields are very weak in the early universe, but recent high-resolution simulations have revealed that they are amplified to certain levels by the small-scale dynamo effect due to turbulence (e.g., Sur et al., 2010, 2012; Federrath et al., 2011; Turk et al., 2012; Sharda et al., 2021). Thus, the magnetic energy can be potentially amplified to the level of the turbulent energy. We can estimate the maximum magnetic field strength as Bmax1.6×104GB_{\rm max}\sim 1.6\times 10^{4}~{}{\rm G} for nH=1019cm3n_{\rm H}=10^{19}~{}{\rm cm}^{-3}, assuming that the magnetic energy can reach the turbulent energy estimated from the results in Section 3, which in turn will affect the dynamics of the infalling gas.

The accretion phase, which is not investigated in this paper, is crucial for the Pop III formation. Recent numerical studies have indicated that both the turbulence and magnetic field play important roles in the accretion phase, for instance, a fragmentation of the accretion disk is promoted/suppressed by their effects, whereas the impact on the fragmentation depends on their strength (e.g., Wollenberg et al., 2020; Sharda et al., 2020, 2021; Prole et al., 2022). In this paper and HSC21, we have found that the amplification of turbulence always occurs in the collapsing primordial gas clouds and the supersonic turbulence is developed. Given that the small-scale dynamo is extremely efficient at generating strong magnetic fields from turbulence, both strong turbulence and magnetic fields likely always exist at the onset of the accretion phase and should be included in the simulations of accretion phase for the first star formation.

6.2 On the present-day star formation

In the primordial case, the relation between the gas density and pressure is described by a single power law index γeff\gamma_{\rm eff}. On the other hand, the temperature evolution in the present-day star formation is more complicated. The temperature evolution from cloud collapse to star formation tracks 4 phases: (i) isothermal collapse (γeff=1.0\gamma_{\rm eff}=1.0, ρ1014gcm3\rho\lesssim 10^{-14}~{}\mathrm{g~{}cm^{-3}}), (ii) first-core formation (γeff=1.4\gamma_{\rm eff}=1.4, 1014gcm3ρ108gcm310^{-14}~{}\mathrm{g~{}cm^{-3}}\lesssim\rho\lesssim 10^{-8}~{}\mathrm{g~{}cm^{-3}}), (iii) the second collapse (γeff=1.1\gamma_{\rm eff}=1.1 , 108gcm3ρ104gcm310^{-8}~{}\mathrm{g~{}cm^{-3}}\lesssim\rho\lesssim 10^{-4}~{}\mathrm{g~{}cm^{-3}}), (iv) protostar formation (γeff=1.67\gamma_{\rm eff}=1.67, 104gcm3ρ10^{-4}~{}\mathrm{g~{}cm^{-3}}\lesssim\rho) (e.g., Masunaga & Inutsuka, 2000; Tomida et al., 2013). The gas is almost isothermal until the gas density reaches 1014gcm310^{-14}~{}\mathrm{g~{}cm^{-3}}, and the turbulent velocity is already transonic in the molecular clouds at much lower densities. Therefore, the saturation level of turbulence can be estimated as sat23\mathcal{M}_{\rm sat}\sim 2-3 just before the first core formation according to Fig 5 and Equation (12). From the first core phase, the equation of state becomes much stiffer, which may result in lower sat\mathcal{M}_{\rm sat}. It is difficult to predict the evolution of turbulence when γeff\gamma_{\rm eff} changes during the collapse, because we have investigated just for the single power index γeff\gamma_{\rm eff}. We will investigate the effects of such a complex equation of state in future works.

From the observational side, high-resolution instruments, such as ALMA, have begun to provide insights on the effects of turbulence on star formation, targeting dense cores in molecular clouds. For example, several authors reported multiple systems with large separation and highly misaligned rotational axes, indicating the fragmentation due to turbulence during the protostar formation (Williams et al., 2014; Fernández-López et al., 2017; Lee et al., 2017). Besides, observations of Taurus by Tokuda et al. (2018) indicate that intense turbulent shock heating possibly occurs in star-forming cores. Our results may indicate that strong turbulence at such small scales is driven by gravitational collapse and in turn, it causes the fragmentation and the shock waves. It will be clarified whether such turbulence is actually driven by gravitational collapse or not by further high-resolution observations of gravitationally unstable “starless” dense cores in the future.

6.3 A caveat on the initial condition

In this study and HSC21, we initially give the velocity fluctuation which has a characteristic energy spectrum (E(k)k2E(k)\propto k^{-2}). Strictly speaking, this initial condition mimics only the velocity field of turbulence in realistic gas clouds, but not the density field that should be correlated with the velocity. In order to develop this “real” turbulence, twice the large eddy turnover times are necessary at least. In contrast, the eddy turnover timescale is initially longer than the collapse timescale (vturb<csv_{\rm turb}<c_{\rm s}) as shown in Section 4.6 in HSC21. As a result, the cloud collapse proceeds before the initial field well develops into a turbulent flow. Thus we have to keep in mind this incompleteness of the initial field. However, considering the cosmological simulations that naturally generate the initial gas distribution also obtain similar growth/saturation of the velocity field (e.g., Greif et al., 2012), we guess that the numerical results from the “real” initial turbulence will be similar to the results in the present paper.

7 Summary

In this paper, we have studied the saturation mechanism of amplified turbulence in collapsing gas clouds.

First of all, assuming the polytropic equation of state, we perform high-resolution collapse simulations with various polytropic exponents γeff\gamma_{\rm eff} to investigate the evolution of the turbulent velocity. We confirm that turbulence can be amplified through gravitational collapse and its saturation level depends on γeff\gamma_{\rm eff}, as found in HSC21. We also perform collapse simulations with various initial velocities, such as high Mach number and turbulence with initial cloud rotation, and find that there is no significant difference in the saturation level. To investigate the effects of baroclinic term leading to the time variation of turbulent vorticity, we follow cloud collapse until protostar formation, self-consistently solving chemical reactions in primordial gas to introduce radiative cooling explicitly. We find that the baroclinic term due to shock waves/radiative cooling has a quite small effect on the evolution of turbulence. Finally, we analytically estimate the saturation level of turbulence by using similarity solutions and the turbulence dissipation coefficient η\eta.

We find that the numerical results are well described by the analytic formula with the turbulence driving scale of one-third the Jeans scale.

We emphasize that the present results show the saturation level of turbulence in contracting prestellar cores is given by the balance between the energy input rate from gravitational collapse and the dissipation rate of turbulence. The former depends only on γeff\gamma_{\rm eff} (Equation 12) while the latter is a constant (see Appendix B). Therefore, the strength of turbulence at the epoch of first-core formation can be estimated solely by γeff\gamma_{\rm eff} both in the early and present-day universe.

We are grateful to anonymous referee for careful reading of the manuscript and constructive comments. We thank T. Inoue, K. Tomida, K. Omukai, T. Hosokawa, M. Machida, K. Sugimura, and M. I. N. Kobayashi for fruitful discussions and useful comments. We are thankful for the support by Ministry of Education, Science, Sports and Culture, Grants-in-Aid for Scientific Research No. 22K03689, 17H02869, and 17H06360. This work was supported by JST SPRING, Grant Number JPMJSP2117. A part of numerical calculations in this work was carried out on Yukawa-21 at the Yukawa Institute Computer Facility. Computations and analysis described in this work were performed using the publicly-available Enzo , Grackle, and yt codes, which is the product of a collaborative effort of many independent scientists from numerous institutions around the world.

Appendix A Comparison with similarity solution

Assuming that the evolution of the density in a gas cloud follows the similarity solution , we have analytically estimated the saturation Mach number in Section 4. Although we have shown that this assumption is reasonable in three-dimensional collapse simulations for γeff=1.09\gamma_{\rm eff}=1.09 (HSC21), we did not confirm its validity for other γeff\gamma_{\rm eff}. Therefore, we compare our simulation results with the similarity solution for all the γeff\gamma_{\rm eff} that we consider.

Figure 6 shows that the evolution of the mean core density as a function of negative time, where the origin t=0t=0 is the time of the final snapshot. All the curves except for ‘with cooling’, γeff=1.09\gamma_{\rm eff}=1.09, and γeff=1.1\gamma_{\rm eff}=1.1 are multiplied by factors for ease of viewing. Although the density evolution becomes faster due to the cloud deformation by the turbulence for γeff=1.0\gamma_{\rm eff}=1.0 and 1.15, it can be clearly seen that the density evolution in the three-dimensional simulations (solid curves) and the similarity solution (dashed curves) is almost in good agreement for all γeff\gamma_{\rm eff}. This result supports that our assumption in Section 4 is reasonable.

Refer to caption
Figure 6: Evolution of the core mean densities as a function of time t-t. The solid lines are numerical results. The dashed lines which have corresponding colors denote ρmean=(α(0)/4πG)t2\rho_{\rm mean}=(\alpha(0)/4\pi G)t^{-2}, where α(0)\alpha(0) are given in Table 2. Note: These values except for ‘with cooling’, γeff=1.09\gamma_{\rm eff}=1.09, and γeff=1.1\gamma_{\rm eff}=1.1 are multiplied by factors for ease of viewing.

Appendix B Simulations of turbulence dissipation

To estimate sat\mathcal{M}_{\rm sat} from Equation (18), we need the dissipation coefficient η\eta of turbulence. In this appendix, we extend the simulations of Mac Low (1999), who investigated turbulent dissipation for isothermal gas. We perform simulations for various γeff\gamma_{\rm eff}, based on the method of Mac Low (1999). Here we remark that the dissipation coefficient ηv\eta_{\rm v} defined by Mac Low (1999) is slightly different from η\eta in Equation (18). They can be converted to each other with the relationship of η=2πηv\eta=2\pi\eta_{\rm v} (equation 14 of Guerrero-Gamboa & Vázquez-Semadeni, 2020).

Refer to caption
Figure 7: Temporal evolution of the turbulent velocity in the case with a kinetic energy input rate E˙in=1\dot{E}_{\rm in}=1. The color of the curves corresponds to each γeff\gamma_{\rm eff} indicated by the box in bottom right. We can see no γeff\gamma_{\rm eff} dependence.

First, we note that each physical quantity in this simulation is described using computational code units. All runs are performed on a three-dimensional uniform grid with 1283128^{3} cells.555Mac Low (1999) showed that, if enough resolution (1283128^{3} at most) is provided, the growth of velocity in this kind of turbulence simulations does not depend on resolution. Hence we choose 1283128^{3} uniform grids. Also, the resolution is comparable to the effective resolution of the collapse simulations in Section 2. We set the computational domain length L=1L=1 with a periodic boundary condition. The gas is uniformly distributed in the domain with the initial density ρ0=1\rho_{0}=1 and the total mass mρL3=1m\equiv\rho L^{3}=1. For comparison with the results of Section 2.1, we set μ=1.22\mu=1.22 and γad=1.4\gamma_{\rm ad}=1.4 and perform simulations with the same set of γeff\gamma_{\rm eff}.

During the simulations, turbulence is continuously driven by the injection of velocity fluctuations δv\delta v at each time step δt\delta t. This δv\delta v corresponds to a constant kinetic energy input rate E˙in=ΔE/Δt\dot{E}_{\rm in}=\Delta E/\Delta t, which is set as an initial parameter. The turbulent kinetic energy spectrum at a wavenumber kk is set to follow the Larson’s law (i.e., E(k)k2E(k)\propto k^{-2}), which has a peak at kpeak=2k_{\rm peak}=2. Thus, the typical driving scale of turbulence is L/kpeakL/k_{\rm peak}. The driven velocity field is composed of fully solenoidal mode, which mimics the turbulent velocity field driven by gravitational collapse (Federrath et al., 2011; Higashi et al., 2021). We perform simulations with E˙in=0.1,0.3,1,3,10\dot{E}_{\rm in}=0.1,~{}0.3,~{}1,~{}3,~{}10 for every γeff\gamma_{\rm eff} and terminate them at twice the sound crossing time tsct_{\rm sc} calculated by the initial sound speed given by each γeff\gamma_{\rm eff}.

Refer to caption
Figure 8: Scatter plot for the energy dissipation rate (=E˙in=\dot{E}_{\rm in}) v.s. kvrms3kv_{\rm rms}^{3}. Difference of colors corresponds to the difference of γeff\gamma_{\rm eff}. The blue solid line is obtained by a least-squares method. Note: There is no difference in values among different γeff\gamma_{\rm eff}, so that, the points are almost degenerated to a single point at each energy dissipation rate.

First, as an example, we show the temporal evolution of the turbulent velocity for E˙in=1\dot{E}_{\rm in}=1 in Figure 7. The velocity increases quickly and saturates at tsc\simeq t_{\rm sc} for all γeff\gamma_{\rm eff}. We can see that the evolution of velocity is almost the same for every γeff\gamma_{\rm eff}, which indicates that there is no γeff\gamma_{\rm eff} dependence during the growth of turbulence.

Mac Low (1999) showed that, when the turbulence reaches an equilibrium state, the energy dissipation rate E˙kin\dot{E}_{\rm kin} of turbulence equals to the energy injection rate E˙in\dot{E}_{\rm in} and can be well approximated by the linear relation

E˙kinηvmk~vrms3,\dot{E}_{\rm kin}\simeq-\eta_{\rm v}m\tilde{k}v_{\rm rms}^{3}, (B1)

where k~(2π/L)kpeak\tilde{k}\equiv(2\pi/L)k_{\rm peak} is a “dimensionalized wavenumber”. Using this relation, we estimate the dissipation coefficient.

We plot kvrms3kv_{\rm rms}^{3} v.s. energy dissipation rate (=E˙in=\dot{E}_{\rm in}) after the sound crossing time from the beginning of the simulation in Figure 8. The circles in different colors denote the results for corresponding γeff\gamma_{\rm eff}. We can see that there is no difference among different γeff\gamma_{\rm eff}, and the points are almost overlapped at a single point for each energy dissipation rate. From the results and Equation (B1), we can estimate ηv=0.42/(2π)\eta_{\rm v}=0.42/(2\pi) with the least squares method. It is fully consistent with the result of the isothermal case in Mac Low (1999). Hence we use η=0.42\eta=0.42 in section 5.

References

  • Abel et al. (2002) Abel, T., Bryan, G. L., & Norman, M. L. 2002, Science, 295, 93, doi: 10.1126/science.295.5552.93
  • Birnboim et al. (2018) Birnboim, Y., Federrath, C., & Krumholz, M. 2018, MNRAS, 473, 2144, doi: 10.1093/mnras/stx2426
  • Bromm et al. (2002) Bromm, V., Coppi, P. S., & Larson, R. B. 2002, ApJ, 564, 23, doi: 10.1086/323947
  • Brummel-Smith et al. (2019) Brummel-Smith, C., Bryan, G., Butsky, I., et al. 2019, The Journal of Open Source Software, 4, 1636, doi: 10.21105/joss.01636
  • Bryan et al. (2014) Bryan, G. L., Norman, M. L., O’Shea, B. W., et al. 2014, ApJS, 211, 19, doi: 10.1088/0067-0049/211/2/19
  • Burgers (1948) Burgers, J. 1948, in Advances in Applied Mechanics, Vol. 1, Advances in Applied Mechanics, ed. R. Von Mises & T. Von Kármán (Elsevier), 171–199, doi: https://doi.org/10.1016/S0065-2156(08)70100-5
  • Chabrier (2005) Chabrier, G. 2005, in Astrophysics and Space Science Library, Vol. 327, The Initial Mass Function 50 Years Later, ed. E. Corbelli, F. Palla, & H. Zinnecker, 41, doi: 10.1007/978-1-4020-3407-7_5
  • Chiaki & Wise (2019) Chiaki, G., & Wise, J. H. 2019, MNRAS, 482, 3933, doi: 10.1093/mnras/sty2984
  • Chiaki & Yoshida (2022) Chiaki, G., & Yoshida, N. 2022, MNRAS, 510, 5199, doi: 10.1093/mnras/stab2799
  • Clark et al. (2011a) Clark, P. C., Glover, S. C. O., Klessen, R. S., & Bromm, V. 2011a, ApJ, 727, 110, doi: 10.1088/0004-637X/727/2/110
  • Clark et al. (2011b) Clark, P. C., Glover, S. C. O., Smith, R. J., et al. 2011b, Science, 331, 1040, doi: 10.1126/science.1198027
  • Federrath et al. (2011) Federrath, C., Sur, S., Schleicher, D. R. G., Banerjee, R., & Klessen, R. S. 2011, ApJ, 731, 62, doi: 10.1088/0004-637X/731/1/62
  • Fernández-López et al. (2017) Fernández-López, M., Zapata, L. A., & Gabbasov, R. 2017, ApJ, 845, 10, doi: 10.3847/1538-4357/aa7d51
  • Fuller & Couchman (2000) Fuller, T. M., & Couchman, H. M. P. 2000, ApJ, 544, 6, doi: 10.1086/317187
  • Goodman et al. (1993) Goodman, A. A., Benson, P. J., Fuller, G. A., & Myers, P. C. 1993, ApJ, 406, 528, doi: 10.1086/172465
  • Greif (2015) Greif, T. H. 2015, Computational Astrophysics and Cosmology, 2, 3, doi: 10.1186/s40668-014-0006-2
  • Greif et al. (2012) Greif, T. H., Bromm, V., Clark, P. C., et al. 2012, MNRAS, 424, 399, doi: 10.1111/j.1365-2966.2012.21212.x
  • Greif et al. (2011) Greif, T. H., Springel, V., White, S. D. M., et al. 2011, ApJ, 737, 75, doi: 10.1088/0004-637X/737/2/75
  • Guerrero-Gamboa & Vázquez-Semadeni (2020) Guerrero-Gamboa, R., & Vázquez-Semadeni, E. 2020, ApJ, 903, 136, doi: 10.3847/1538-4357/abba1f
  • Haiman et al. (1996) Haiman, Z., Thoul, A. A., & Loeb, A. 1996, ApJ, 464, 523, doi: 10.1086/177343
  • Hennebelle (2021) Hennebelle, P. 2021, A&A, 655, A3, doi: 10.1051/0004-6361/202141650
  • Higashi et al. (2021) Higashi, S., Susa, H., & Chiaki, G. 2021, ApJ, 915, 107, doi: 10.3847/1538-4357/ac01c7
  • Hirano & Bromm (2017) Hirano, S., & Bromm, V. 2017, MNRAS, 470, 898, doi: 10.1093/mnras/stx1220
  • Hockney & Eastwood (1988) Hockney, R. W., & Eastwood, J. W. 1988, Computer simulation using particles
  • Inoue & Yoshida (2020) Inoue, S., & Yoshida, N. 2020, MNRAS, 491, L24, doi: 10.1093/mnrasl/slz160
  • Kroupa et al. (2001) Kroupa, P., Aarseth, S., & Hurley, J. 2001, MNRAS, 321, 699, doi: 10.1046/j.1365-8711.2001.04050.x
  • Lee et al. (2017) Lee, J.-E., Lee, S., Dunham, M. M., et al. 2017, Nature Astronomy, 1, 0172, doi: 10.1038/s41550-017-0172
  • Mac Low (1999) Mac Low, M.-M. 1999, ApJ, 524, 169, doi: 10.1086/307784
  • Machida & Doi (2013) Machida, M. N., & Doi, K. 2013, MNRAS, 435, 3283, doi: 10.1093/mnras/stt1524
  • Mandal et al. (2020) Mandal, A., Federrath, C., & Körtgen, B. 2020, MNRAS, 493, 3098, doi: 10.1093/mnras/staa468
  • Masunaga & Inutsuka (2000) Masunaga, H., & Inutsuka, S.-i. 2000, ApJ, 531, 350, doi: 10.1086/308439
  • McKee & Ostriker (2007) McKee, C. F., & Ostriker, E. C. 2007, ARA&A, 45, 565, doi: 10.1146/annurev.astro.45.051806.110602
  • Murray & Chang (2015) Murray, N., & Chang, P. 2015, ApJ, 804, 44, doi: 10.1088/0004-637X/804/1/44
  • Nishi & Susa (1999) Nishi, R., & Susa, H. 1999, ApJ, 523, L103, doi: 10.1086/312277
  • Omukai & Nishi (1998) Omukai, K., & Nishi, R. 1998, ApJ, 508, 141, doi: 10.1086/306395
  • Padoan & Nordlund (2002) Padoan, P., & Nordlund, Å. 2002, ApJ, 576, 870, doi: 10.1086/341790
  • Prieto et al. (2011) Prieto, J., Padoan, P., Jimenez, R., & Infante, L. 2011, ApJ, 731, L38, doi: 10.1088/2041-8205/731/2/L38
  • Prole et al. (2022) Prole, L. R., Clark, P. C., Klessen, R. S., & Glover, S. C. O. 2022, MNRAS, 510, 4019, doi: 10.1093/mnras/stab3697
  • Riaz et al. (2018) Riaz, R., Bovino, S., Vanaverbeke, S., & Schleicher, D. R. G. 2018, MNRAS, 479, 667, doi: 10.1093/mnras/sty1635
  • Robertson & Goldreich (2012) Robertson, B., & Goldreich, P. 2012, ApJ, 750, L31, doi: 10.1088/2041-8205/750/2/L31
  • Sadanari et al. (2021) Sadanari, K. E., Omukai, K., Sugimura, K., Matsumoto, T., & Tomida, K. 2021, MNRAS, 505, 4197, doi: 10.1093/mnras/stab1330
  • Salpeter (1955) Salpeter, E. E. 1955, ApJ, 121, 161, doi: 10.1086/145971
  • Sharda et al. (2020) Sharda, P., Federrath, C., & Krumholz, M. R. 2020, MNRAS, 497, 336, doi: 10.1093/mnras/staa1926
  • Sharda et al. (2021) Sharda, P., Federrath, C., Krumholz, M. R., & Schleicher, D. R. G. 2021, MNRAS, 503, 2014, doi: 10.1093/mnras/stab531
  • Smith et al. (2017) Smith, B. D., Bryan, G. L., Glover, S. C. O., et al. 2017, MNRAS, 466, 2217, doi: 10.1093/mnras/stw3291
  • Solomon et al. (1987) Solomon, P. M., Rivolo, A. R., Barrett, J., & Yahil, A. 1987, ApJ, 319, 730, doi: 10.1086/165493
  • Stacy et al. (2016) Stacy, A., Bromm, V., & Lee, A. T. 2016, MNRAS, 462, 1307, doi: 10.1093/mnras/stw1728
  • Sur et al. (2012) Sur, S., Federrath, C., Schleicher, D. R. G., Banerjee, R., & Klessen, R. S. 2012, MNRAS, 423, 3148, doi: 10.1111/j.1365-2966.2012.21100.x
  • Sur et al. (2010) Sur, S., Schleicher, D. R. G., Banerjee, R., Federrath, C., & Klessen, R. S. 2010, ApJ, 721, L134, doi: 10.1088/2041-8205/721/2/L134
  • Susa (2013) Susa, H. 2013, ApJ, 773, 185, doi: 10.1088/0004-637X/773/2/185
  • Susa (2019) —. 2019, ApJ, 877, 99, doi: 10.3847/1538-4357/ab1b6f
  • Susa et al. (2014) Susa, H., Hasegawa, K., & Tominaga, N. 2014, ApJ, 792, 32, doi: 10.1088/0004-637X/792/1/32
  • Susa et al. (1996) Susa, H., Uehara, H., & Nishi, R. 1996, Progress of Theoretical Physics, 96, 1073, doi: 10.1143/PTP.96.1073
  • Suto & Silk (1988) Suto, Y., & Silk, J. 1988, ApJ, 326, 527, doi: 10.1086/166114
  • Tegmark et al. (1997) Tegmark, M., Silk, J., Rees, M. J., et al. 1997, ApJ, 474, 1, doi: 10.1086/303434
  • Tokuda et al. (2018) Tokuda, K., Onishi, T., Saigo, K., et al. 2018, ApJ, 862, 8, doi: 10.3847/1538-4357/aac898
  • Tomida et al. (2013) Tomida, K., Tomisaka, K., Matsumoto, T., et al. 2013, ApJ, 763, 6, doi: 10.1088/0004-637X/763/1/6
  • Turk et al. (2012) Turk, M. J., Oishi, J. S., Abel, T., & Bryan, G. L. 2012, ApJ, 745, 154, doi: 10.1088/0004-637X/745/2/154
  • Turk et al. (2011) Turk, M. J., Smith, B. D., Oishi, J. S., et al. 2011, ApJS, 192, 9, doi: 10.1088/0067-0049/192/1/9
  • Vázquez-Semadeni et al. (1998) Vázquez-Semadeni, E., Cantó, J., & Lizano, S. 1998, ApJ, 492, 596, doi: 10.1086/305064
  • Williams et al. (2014) Williams, J. P., Mann, R. K., Di Francesco, J., et al. 2014, ApJ, 796, 120, doi: 10.1088/0004-637X/796/2/120
  • Wise & Abel (2007) Wise, J. H., & Abel, T. 2007, ApJ, 671, 1559, doi: 10.1086/522876
  • Wollenberg et al. (2020) Wollenberg, K. M. J., Glover, S. C. O., Clark, P. C., & Klessen, R. S. 2020, MNRAS, doi: 10.1093/mnras/staa289
  • Yahil (1983) Yahil, A. 1983, ApJ, 265, 1047, doi: 10.1086/160746
  • Yoshida et al. (2003) Yoshida, N., Abel, T., Hernquist, L., & Sugiyama, N. 2003, ApJ, 592, 645, doi: 10.1086/375810
  • Yoshida et al. (2008) Yoshida, N., Omukai, K., & Hernquist, L. 2008, Science, 321, 669, doi: 10.1126/science.1160259
  • Yoshida et al. (2006) Yoshida, N., Omukai, K., Hernquist, L., & Abel, T. 2006, ApJ, 652, 6, doi: 10.1086/507978
\listofchanges