Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Cloud Crushing and Dissipation of Uniformly-Driven Adiabatic Turbulence in Circumgalactic Media

Alex Lv Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, People’s Republic of China Department of Astronomy, School of Physics, Peking University, Beijing 100871, People’s Republic of China alexanderlu@stu.pku.edu.cn Lile Wang Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, People’s Republic of China Department of Astronomy, School of Physics, Peking University, Beijing 100871, People’s Republic of China Lile Wang lilew@pku.edu.cn Renyue Cen Institute for Advanced Study in Physics, Zhejiang University, Hangzhou 310027, People’s Republic of China Institute of Astronomy, School of Physics, Zhejiang University, Hangzhou 310027, People’s Republic of China renyuecen@zju.edu.cn Luis C. Ho Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, People’s Republic of China Department of Astronomy, School of Physics, Peking University, Beijing 100871, People’s Republic of China lho.pku@gmail.com
Abstract

The circumgalactic medium (CGM) is responsive to kinetic disruptions generated by nearby astrophysical events. In this work, we study the saturation and dissipation of turbulent hydrodynamics within the CGM through an extensive array of 252 numerical simulations with a large parameter space. These simulations are endowed with proper cooling mechanisms to consistently explore the parameter space spanned by the average gas density, metallicity, and turbulence driving strength. A dichotomy emerges in the dynamics dissipation behaviors. Disturbances that are hot and subsonic are characterized by weak compression and slow dissipation, resulting in density fluctuations typically 102less-than-or-similar-toabsentsuperscript102\lesssim 10^{-2}≲ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. Conversely, warm supersonic turbulence, marked by significant compression shocks and subsequent rapid cooling, is associated with substantial clumping factors 100101similar-toabsentsuperscript100superscript101\sim 10^{0}-10^{1}∼ 10 start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT. In the supersonic cases, the kinetic energy decay is divided into a rate-limiting phase of shock dissipation and a comparatively swift phase of thermal dissipation, predominantly occurring within the overdense regions. Upon turbulence driving turnoff, the strong density contrasts decay within a relatively brief timescale of 30300Myrsimilar-toabsent30300Myr\sim 30-300~{}{\rm Myr}∼ 30 - 300 roman_Myr, depending on the average gas density. Dense clouds are crushed on similar timescales of 30100Myrsimilar-toabsent30100Myr\sim 30-100~{}{\rm Myr}∼ 30 - 100 roman_Myr, depending on turbulence driving strength but independent from average gas density. Results of this work also contribute a novel dataset of dissipation timescales that incorporates an understanding of kinematics and thermodynamics in addition to the traditional cooling rate tables, which may serve as a valuable asset for forthcoming simulations that aim to explore gas dynamics on galactic and cosmological scales.

Circumgalactic medium (1879) — Collapsing clouds (267) — Hydrodynamical simulations (767) — Extragalactic astronomy (506)
\declare@file@substitution

revtex4-1.clsrevtex4-2.cls

1 Introduction

The circumgalactic medium (CGM) represents a rich, multiphased environment surrounding a galaxy, extending out up to the virial radius of the halo of 101102superscript101superscript10210^{1}-10^{2}10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT kpc. It plays a crucial role in galaxy evolution, serving as an important source of baryons (Werk et al., 2014) for gas accretion onto the galactic disc. The physical state of the CGM can be split into distinct phases (Cen, 2013): The cool phase of the CGM typically hovers around 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT to 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT K, while the hot phase 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT to 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT K (Rudie et al., 2012; Werk et al., 2013), meaning the bulk of the CGM consists of ionized gas. The mode of accretion depends heavily on the cooling timescale of the gas in relation to the freefall time: cold mode accretion characterized by filamentary streams and clumps when tcool<tffsubscript𝑡coolsubscript𝑡fft_{\mathrm{cool}}<t_{\mathrm{ff}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT < italic_t start_POSTSUBSCRIPT roman_ff end_POSTSUBSCRIPT and hot mode accretion characterized by smooth cooling flows when tcool>tffsubscript𝑡coolsubscript𝑡fft_{\mathrm{cool}}>t_{\mathrm{ff}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT > italic_t start_POSTSUBSCRIPT roman_ff end_POSTSUBSCRIPT (White & Rees, 1978; Shen et al., 2013; Crighton et al., 2013; Stern et al., 2020)). However, given the multiphased nature of the CGM, with varying metallicities, ionization states and temperatures, calculating and characterizing cooling timecales is a highly complex and nonlinear problem. Generally speaking, cooling streams only occur in the non-virialized regions of the circumgalactic medium where the accretion rate exceeds a critical threshold (Faucher-Giguère & Oh, 2023). Thermal instabilities due to rapid cooling shocks the gas, leading to the formation of cold supersonic filaments (Balbus & Soker, 1989; Dekel et al., 2009; Stern et al., 2021).

Unlike typical giant molecular cloud (GMC) conditions in the interstellar medium (ISM), in addition to being much hotter and mostly ionized, the CGM is also more diffuse. Observations reveal typical hydrogen column densities ranging from 1020cm2superscript1020superscriptcm210^{20}\;\mathrm{cm}^{-2}10 start_POSTSUPERSCRIPT 20 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT near the disc down to 1014cm2superscript1014superscriptcm210^{14}\;\mathrm{cm}^{-2}10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT out to the virial radius (Werk et al., 2014), corresponding to number densities of around 102superscript10210^{-2}10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT to 103cm3superscript103superscriptcm310^{-3}\;\mathrm{cm}^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT (Zahedy et al., 2019; Crighton et al., 2015). In such an environment, Jeans’ instabilities alone are insufficient in forming the cool phase of the CGM. Turbulence, evidenced by observations of broad line-widths (Werk et al., 2016; Rudie et al., 2012), plays a pivotal role in the gas dynamics of the CGM. A key driver in nonvirialized halos with high accretion rates (>10Myr1absent10subscript𝑀direct-productsuperscriptyr1>10\;M_{\odot}\;\mathrm{yr}^{-1}> 10 italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT roman_yr start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT for 1012Msuperscript1012subscript𝑀direct-product10^{12}M_{\odot}10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT haloes), as mentioned previously, are supersonic cool accretion flows which can stir turbulence via localized thermodynamics such as turbulent mixing between hot and cold phases and gas entrainment onto supersonic cold streams (Ji et al., 2019; Yang & Ji, 2023). Internal feedback processes from supernovae and/or active galactic nuclei (AGN) within the galaxy, are also capable of driving turbulence in the CGM via large-scale outflows (Fielding et al., 2017). These feedback-induced turbulence motions in turn significantly impact the accretion of gas back onto the ISM or into the AGN, rendering it chaotic and asymmetric (Gaspari et al., 2013). They can also “stimulate” the clumping and eventual precipitation/accretion of CGM gas back onto the ISM in regimes where tcoolsubscript𝑡coolt_{\mathrm{cool}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT exceeds tffsubscript𝑡fft_{\mathrm{ff}}italic_t start_POSTSUBSCRIPT roman_ff end_POSTSUBSCRIPT (Voit et al., 2017; Voit, 2018). Within a cold cloud, turbulence may fragment the cloud into smaller ”droplets”, exponentially increasing the surface area and growth rate of cool clumps (Gronke et al., 2022), which may explain the observational evidence for cold streams in the CGM. Additionally, turbulent winds can damp fluid instabilities in cool clouds in the presence of hot winds, stabilizing the cloud against destruction (Sparre et al., 2020), although compared to radiative cooling processes this only has a significant impact in highly supersonic flow (Li et al., 2020).

However, turbulence in such diffuse astrophysical media is not necessarily a continuous phenomenon, particularly in hot virialized haloes where accretion is subsonic and quasi-spherically symmetric (Faucher-Giguère & Oh, 2023). Internal sources such as feedback are tied to episodic starburst and AGN activity from the disc (Rubin et al., 2014; Nielsen et al., 2015; Borthakur et al., 2013), and environmental sources are tied to episodic mergers or ram pressure on infall into a group or cluster. Hence, this necessitates the study of not merely steady-state turbulence driving, but also the dissipation of turbulence during quiescent periods. Using magnetohydrodynamic simulations of turbulence driving on molecular cloud scales with an isothermal equation of state, Stone et al. (1998) found saturation timescales to be independent of magnetic field strength, and energy dissipation timescales to be inversely correlated with magnetic field strength once turbulence was turned off. Magnetic fields provide magnetic pressure, which plays an important role in the gravitational stability of GMCs (Ostriker et al., 2001), although the role it plays in CGM cloud stability is uncertain and somewhat of an open question (Faucher-Giguère & Oh, 2023). While Stone et al. (1998) considered variable magnetic field strength (damping strength) across multiple runs, they fixed their turbulence driving strength.

In this study, using hydrodynamic simulations, we investigate the effects of a generalized source of turbulence in a circumgalactic environment, though it can be extended to broader environments such as the intracluster medium or diffuse phases of the ISM. This work applies and builds on the methodology of Stone et al. (1998) to such environments, with variable turbulence driving strengths, an adiabatic equation of state and a standard collisional ionization equilibrium (CIE) cooling curve (Gnedin & Hollon, 2012). We will elaborate on the methodological differences from Stone et al. (1998) in Section 2.

2 Methods

2.1 Hydrodynamics and Thermodynamics

Studying turbulence numerically requires solving a consistent set of hydrodynamic equations. We employ the hetrogeneous hydrodynamic code KRATOS (Wang et al. in prep), integrating the following Euler equations for an ideal gas:

ρt+(ρ𝐯)=0,(ρ𝐯)t+(ρ𝐯𝐯+p𝐈)=ρΦ,εt+[𝐯(ε+p)]=ρ𝐯Φ+S.formulae-sequence𝜌𝑡𝜌𝐯0formulae-sequence𝜌𝐯𝑡𝜌𝐯𝐯𝑝𝐈𝜌Φ𝜀𝑡delimited-[]𝐯𝜀𝑝𝜌𝐯Φ𝑆\begin{split}\dfrac{\partial\rho}{\partial t}+\nabla\cdot(\rho\mathbf{v})&=0\ % ,\\ \dfrac{\partial(\rho\mathbf{v})}{\partial t}+\nabla\cdot(\rho\mathbf{v}\mathbf% {v}+p\mathbf{I})&=-\rho\nabla\Phi,\\ \dfrac{\partial\varepsilon}{\partial t}+\nabla\cdot[\mathbf{v}(\varepsilon+p)]% &=-\rho\mathbf{v}\cdot\nabla\Phi+S\ .\end{split}start_ROW start_CELL divide start_ARG ∂ italic_ρ end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ ( italic_ρ bold_v ) end_CELL start_CELL = 0 , end_CELL end_ROW start_ROW start_CELL divide start_ARG ∂ ( italic_ρ bold_v ) end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ ( italic_ρ bold_vv + italic_p bold_I ) end_CELL start_CELL = - italic_ρ ∇ roman_Φ , end_CELL end_ROW start_ROW start_CELL divide start_ARG ∂ italic_ε end_ARG start_ARG ∂ italic_t end_ARG + ∇ ⋅ [ bold_v ( italic_ε + italic_p ) ] end_CELL start_CELL = - italic_ρ bold_v ⋅ ∇ roman_Φ + italic_S . end_CELL end_ROW (1)

Here, ρ𝜌\rhoitalic_ρ, 𝐯𝐯\mathbf{v}bold_v, p𝑝pitalic_p and ϵitalic-ϵ\epsilonitalic_ϵ are the mass density, velocity, gas pressure, and total energy density, respectively. ΦΦ\Phiroman_Φ is the gravitational potential, 𝐈𝐈\mathbf{I}bold_I is the identity tensor, and S𝑆Sitalic_S is the source term which captures additional thermodynamics beyond adiabatic compression and expansion. The hydrodynamic solver employs the slope-limited piecewise linear method (PLM) reconstruction scheme, the HLLC approximate Riemann solver, and the second order Runge-Kutta (RK2) time integrator. The Jeans length under typical CGM physical parameters approximately reads,

lJ104pc×(T104K)1/2(ρ102mpcm3)1/2.similar-tosubscript𝑙𝐽superscript104pcsuperscript𝑇superscript104K12superscript𝜌superscript102subscript𝑚𝑝superscriptcm312l_{J}\sim 10^{4}~{}\mathrm{pc}\times\left(\dfrac{T}{10^{4}~{}\mathrm{K}}\right% )^{1/2}\left(\dfrac{\rho}{10^{-2}~{}m_{p}~{}\mathrm{cm}^{-3}}\right)^{-1/2}\ .italic_l start_POSTSUBSCRIPT italic_J end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_pc × ( divide start_ARG italic_T end_ARG start_ARG 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_K end_ARG ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( divide start_ARG italic_ρ end_ARG start_ARG 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT . (2)

Such scales are considerably greater than the box size, let alone the typical cloud sizes, within our turbulent CGM simulations, making Φ0Φ0\Phi\rightarrow 0roman_Φ → 0 a safe approximation in this study.

Within each sub-step in the RK2 cycles, we integrate the impact of S𝑆Sitalic_S in every zone via a semi-implicit method. The cooling rate, represented by S𝑆Sitalic_S in eq. (1), is based on the standard table from Gnedin & Hollon (2012). It covers the temperature range 103108Ksuperscript103superscript108K10^{3}-10^{8}~{}{\rm K}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT roman_K and varying metallicities (see Figure 1). For typical CGM gases with density ρ102mpcm3less-than-or-similar-to𝜌superscript102subscript𝑚𝑝superscriptcm3\rho\lesssim 10^{-2}~{}m_{p}~{}\mathrm{cm}^{-3}italic_ρ ≲ 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, the column density required for the cooling photons to escape in typical CGM scenarios is estimated by N1019cm2less-than-or-similar-to𝑁superscript1019superscriptcm2N\lesssim 10^{19}~{}\mathrm{cm}^{-2}italic_N ≲ 10 start_POSTSUPERSCRIPT 19 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT per kpc along the escape path. For dense clouds with 102×\sim 10^{2}\times∼ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT × the mean density, the typical sizes (101pcsimilar-toabsentsuperscript101pc\sim 10^{1}~{}\mathrm{pc}∼ 10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_pc) are sufficiently small to allow cooling photons to escape. These are sufficient to guarantee the optical-thin condition for the standard cooling table.

2.2 Turbulence Driving

This study focuses on the local dissipative behaviors of various phases of the CGM, hence the turbulence energy cascade is emulated via kinetic energy injection. We follow an approach similar to the one outlined in Mac Low (1999). During each timestep, the simulation domain is subject to a uniform perturbation via an acceleration vector aligned with a randomly selected direction vector a^^𝑎\hat{a}over^ start_ARG italic_a end_ARG. Denoting the amplitude of perturbation as A𝐴Aitalic_A, the turbulence energy injection rate per by unit mass ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT (which will be referred to as turbulence driving strength) can be evaluated as,

ϵ˙i=A{[ρL3]1Δt𝐢δV𝐢ρ𝐢𝐯𝐢a^},subscript˙italic-ϵ𝑖𝐴superscriptdelimited-[]delimited-⟨⟩𝜌superscript𝐿31Δ𝑡subscript𝐢𝛿subscript𝑉𝐢subscript𝜌𝐢subscript𝐯𝐢^𝑎\dot{\epsilon}_{i}=A\left\{[\langle\rho\rangle L^{3}]^{-1}\Delta t\ \sum_{% \mathbf{i}}\delta V_{\mathbf{i}}\;\rho_{\mathbf{i}}\mathbf{v}_{\mathbf{i}}% \cdot\hat{a}\right\}\ ,over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_A { [ ⟨ italic_ρ ⟩ italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Δ italic_t ∑ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT italic_δ italic_V start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT bold_v start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT ⋅ over^ start_ARG italic_a end_ARG } , (3)

where L𝐿Litalic_L is the box length, ρdelimited-⟨⟩𝜌\langle\rho\rangle⟨ italic_ρ ⟩ is the mean mass density throughout the domain, and ΔtΔ𝑡\Delta troman_Δ italic_t is the current timestep. The summation goes through all cells, where 𝐢𝐢\mathbf{i}bold_i is the cell index, and δV𝐢𝛿subscript𝑉𝐢\delta V_{\mathbf{i}}italic_δ italic_V start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT, ρ𝐢subscript𝜌𝐢\rho_{\mathbf{i}}italic_ρ start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT and 𝐯𝐢subscript𝐯𝐢\mathbf{v}_{\mathbf{i}}bold_v start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT are the cell volume, mass density, and velocity vector for the 𝐢𝐢\mathbf{i}bold_i-th cell. In practice, we establish a constant ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as a key model parameter, which remains fixed throughout the simulation’s progression up until turbulence turnoff. For each timestep, we utilize the enclosed summation in eq. (3) to normalize the amplitude A𝐴Aitalic_A. Subsequently, we inject kinetic energy into the system by updating the velocity vector to 𝐯𝐢=𝐯𝐢+Aa^Δtsuperscriptsubscript𝐯𝐢subscript𝐯𝐢𝐴^𝑎Δ𝑡\mathbf{v}_{\mathbf{i}}^{\prime}=\mathbf{v}_{\mathbf{i}}+A\hat{a}\Delta tbold_v start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT = bold_v start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT + italic_A over^ start_ARG italic_a end_ARG roman_Δ italic_t, and update the energy density accordingly.

Refer to caption
Figure 1: The cooling curves used in the simulations, adopting the data provided by Gnedin & Hollon (2012). The interpolated and extrapolated curves are computed as Λ(Z/Z)/n2=Λ(0)/n2+(Z/Z)(Λ(1)/n2Λ(0)/n2)Λ𝑍subscript𝑍direct-productsuperscript𝑛2Λ0superscript𝑛2𝑍subscript𝑍direct-productΛ1superscript𝑛2Λ0superscript𝑛2\Lambda(Z/Z_{\odot})/n^{2}=\Lambda(0)/n^{2}+(Z/Z_{\odot})(\Lambda(1)/n^{2}-% \Lambda(0)/n^{2})roman_Λ ( italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) / italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = roman_Λ ( 0 ) / italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ) ( roman_Λ ( 1 ) / italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - roman_Λ ( 0 ) / italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ).

2.3 Simulation Setup

Focusing on the local structures and characteristics of turbulence compression and subsequent dissipation processes, these simulations are conducted on L=256pc𝐿256pcL=256~{}\mathrm{pc}italic_L = 256 roman_pc cubic domains with periodic boundary conditions. Various simulations are carried out to cover the parameter space, with 6 different metallicity values Z/Z{0,0.01,0.1,0.3,1,3}𝑍subscript𝑍direct-product00.010.10.313Z/Z_{\odot}\in\{0,0.01,0.1,0.3,1,3\}italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ∈ { 0 , 0.01 , 0.1 , 0.3 , 1 , 3 }, 6 different average gas density values ρ/(mpcm3){104,103,102,101,100,101}delimited-⟨⟩𝜌subscript𝑚𝑝superscriptcm3superscript104superscript103superscript102superscript101superscript100superscript101\langle\rho\rangle/(m_{p}~{}\mathrm{cm}^{-3})\in\{10^{-4},10^{-3},10^{-2},10^{% -1},10^{0},10^{1}\}⟨ italic_ρ ⟩ / ( italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) ∈ { 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT }, and 7 different turbulence energy injection rate values ϵ˙i/(ergs1g1){0.01,0.1,0.3,1,3,10,100}subscript˙italic-ϵ𝑖ergsuperscripts1superscriptg10.010.10.31310100\dot{\epsilon}_{i}/(\mathrm{erg}~{}\mathrm{s}^{-1}~{}\mathrm{g}^{-1})\in\{0.01% ,0.1,0.3,1,3,10,100\}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT / ( roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) ∈ { 0.01 , 0.1 , 0.3 , 1 , 3 , 10 , 100 }. In total, there are 252 simulations. Selection of these parameters extend beyond the density and metallicity range typical of the CGM to account for denser, metal enriched phases captured in cool clouds, though we note that the Jeans length remains larger than our box even for our highest density runs.

Methodologically, setting aside the different astrophysical medium of interest and our exclusion of magnetic fields, a key difference between this work and Stone et al. (1998) is our usage of an adiabatic equation of state with the inclusion of thermal energy, adiabatic heating and a radiative cooling function. This is necessary given the differences between CGM and GMC environments: the CGM is orders of magnitude hotter and reside in efficient cooling temperature regimes, and is subject to much more violent turbulence driving events. Without artificial viscosity, energy dissipation happens entirely via the radiative cooling function, with kinetic energy dissipating via the energy cascade into thermal energy.

The wide coverage over the multi-dimensional parameter space limits the viable resolution of each run due to computational and storage limits. We adopt a relatively low 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resolution for each run, resulting in a spatial resolution of 2pc2pc2~{}\mathrm{pc}2 roman_pc. Such resolution is sufficient to resolve compressive shocks that form relatively dense gas clumps with sizes 10pcsimilar-toabsent10pc\sim 10~{}\mathrm{pc}∼ 10 roman_pc (see Section 3.6). For conditions matching more closely to those of the dense phase of the CGM, namely Z/Z{0.1,0.3}𝑍subscript𝑍direct-product0.10.3Z/Z_{\odot}\in\{0.1,0.3\}italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT ∈ { 0.1 , 0.3 } and ρ/(mpcm3){102,101}𝜌subscript𝑚𝑝superscriptcm3superscript102superscript101\rho/(m_{p}~{}\mathrm{cm}^{-3})\in\{10^{-2},10^{-1}\}italic_ρ / ( italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) ∈ { 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT , 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT }, we run an additional set of higher resolution 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs with the exact same setup otherwise. The denser and higher metallicity phases are chosen given their efficient cooling regimes, which allows the turbulence to remain supersonic for higher values of ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and thus more analyzable high resolution runs.

The initial conditions of each run are set so that the initial gas temperature is uniformly T=3×104K𝑇3superscript104KT=3\times 10^{4}~{}\mathrm{K}italic_T = 3 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT roman_K.

The initial velocity field is randomized, with each component of each cell’s velocity having a random value between ±1kms1plus-or-minus1kmsuperscripts1\pm 1~{}\mathrm{km}~{}\mathrm{s}^{-1}± 1 roman_km roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, representing a homogenous gas with initial subsonic motions. The choice of initial velocity field is arbitrary as long as the velocities are subsonic, since all runs will undergo turbulence driving until the gas reaches saturated steady-state turbulence. An initial non-zero velocity field is necessary, since given a periodic box and entirely uniform perturbations, theoretically no substructures would form, and any resulting substructures that do form are the result of numerical errors and approximations in the flux advection and reconstruction schemes. The randomized subsonic initial motions guarantee a physical basis for any resulting turbulence driven substructures.

Each run is fixed to 210000 cycles. Turbulence driving is turned on for the first 200000 cycles to guarantee turbulence saturation and a turbulent quasi-steady states. Turbulence is turned off for the last 10000 cycles to study the dissipation of turbulences. This scheme of fixing the number of cycles is related to the Courant-Friedrichs-Lewy (CFL) conditions, which limits the timestep of each cycle ΔtΔ𝑡\Delta troman_Δ italic_t by

Δt=C×min𝐢{Δx|𝐯𝐢|+cs,𝐢}.Δ𝑡𝐶subscript𝐢Δ𝑥subscript𝐯𝐢subscript𝑐𝑠𝐢\Delta t=C\times\min_{\mathbf{i}}\bigg{\{}\dfrac{\Delta x}{|\mathbf{v}_{% \mathbf{i}}|+c_{s,\mathbf{i}}}\bigg{\}}\ .roman_Δ italic_t = italic_C × roman_min start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT { divide start_ARG roman_Δ italic_x end_ARG start_ARG | bold_v start_POSTSUBSCRIPT bold_i end_POSTSUBSCRIPT | + italic_c start_POSTSUBSCRIPT italic_s , bold_i end_POSTSUBSCRIPT end_ARG } . (4)

In this work we choose C=0.3𝐶0.3C=0.3italic_C = 0.3. As one can easily deduce, each cycle with timestep constrained by eq. (4) represents a fraction of the fluid crossing time (in case of high Mach numbers) or sound crossing time (in case of low Mach numbers). Therefore, a fixed number of cycles naturally covers a similar number of fluid or sound crossing timescales, providing sufficient temporal resolution for analysis while minimizing data storage requirements.

3 Results

Unless specified otherwise, plots showing a sample of data use the 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs, while plots showing single representative samples use the 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs. Plots showing 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs will have this fact relayed in figure captions, and we will discuss resolution convergence in Appendix A.

3.1 Subsonic and Supersonic Dissipation

Refer to caption
Figure 2: Two scatter plots showing the mass averaged mach number ()\mathcal{M})caligraphic_M ) vs the fractional energy decay at three saturation times (Δϵ/ϵ0Δitalic-ϵsubscriptitalic-ϵ0\Delta\epsilon/\epsilon_{0}roman_Δ italic_ϵ / italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) on top and the kinetic energy saturation time (tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) on the bottom, with each point representing a run. \mathcal{M}caligraphic_M is defined as vavg/cs,0subscript𝑣avgsubscript𝑐𝑠0v_{\mathrm{avg}}/c_{s,0}italic_v start_POSTSUBSCRIPT roman_avg end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT, where vavgsubscript𝑣avgv_{\mathrm{avg}}italic_v start_POSTSUBSCRIPT roman_avg end_POSTSUBSCRIPT is the mass averaged velocity across all cells and the initial sound speed cs,0subscript𝑐𝑠0c_{s,0}italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT is computed from the mass-averaged temperature of the entire box. tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT represents the kinetic energy saturation time defined as the ratio of the kinetic specific energy to the turbulence driving strength ts=ϵk/ϵ˙isubscript𝑡𝑠subscriptitalic-ϵ𝑘subscript˙italic-ϵ𝑖t_{s}=\epsilon_{k}/\dot{\epsilon}_{i}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The temperatures on the top plot represent initial temperatures at turbulence turnoff.
Refer to caption
Figure 3: A comparison between two 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resolution runs where Z/Z=0.3𝑍subscript𝑍direct-product0.3Z/Z_{\odot}=0.3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.3 and n=102cm3𝑛superscript102superscriptcm3n=10^{-2}\;\mathrm{cm}^{-3}italic_n = 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, with the top and bottom rows representing ϵ˙i=1ergs1g1subscript˙italic-ϵi1ergsuperscripts1superscriptg1\mathrm{\dot{\epsilon}_{i}}=1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT = 1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and ϵ˙i=3ergs1g1subscript˙italic-ϵi3ergsuperscripts1superscriptg1\mathrm{\dot{\epsilon}_{i}}=3\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT = 3 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT respectively. The left column shows density slice plots at z=0𝑧0z=0italic_z = 0, the middle column the mass-weighted density-temperature phase plots (contoured 2D histogram), and the right column the specific total and thermal energies and the mass-averaged temperature over time. The dotted blue lines in the top middle phase plot represent lines of constant pressure.

A bimodality in the initial turbulent state and the energy dissipation is seen in Figure 2, with a distinct supersonic and subsonic turbulence regime. In the subsonic case, the initial averaged mach numbers (\mathcal{M}caligraphic_M) are less than 1111, the initial temperatures are above 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT K, and little dissipation is present within 3ts3subscript𝑡𝑠3t_{s}3 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Meanwhile in the supersonic case, the \mathcal{M}caligraphic_M values are of order unity and above, the initial temperatures are below 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT K, and significant energy dissipation of 1080%10percent8010-80\%10 - 80 % within three kinetic saturation timescales is present within 3ts3subscript𝑡𝑠3t_{s}3 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. There is a temperature gradient across the supersonic runs, where hotter steady-state turbulence leads to increased energy dissipation. Among the subsonic runs, the energy loss hovers around 1%percent11\%1 %, but runs with ϵ˙i=100ergs1g1subscript˙italic-ϵ𝑖100ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=100\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 100 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and initial temperatures of around 108superscript10810^{8}10 start_POSTSUPERSCRIPT 8 end_POSTSUPERSCRIPT K lose around 36%3percent63-6\%3 - 6 % of their initial energy within 3ts3subscript𝑡𝑠3t_{s}3 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, with initial \mathcal{M}caligraphic_M higher than the other subsonic runs. The distribution of fractional energy losses almost seems multimodal, with two distinct subsonic regimes and one distinct supersonic regime. Across runs with fixed initial conditions (same n𝑛nitalic_n and Z/Z𝑍subscript𝑍direct-productZ/Z_{\odot}italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT), increasing ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT leads to an eventual transition point from supersonic to subsonic turbulence (see Figure 3, though this offset in both \mathcal{M}caligraphic_M and Δϵ/ϵ0Δitalic-ϵsubscriptitalic-ϵ0\Delta\epsilon/\epsilon_{0}roman_Δ italic_ϵ / italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT between the subsonic ϵ˙i=100ergs1g1subscript˙italic-ϵ𝑖100ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=100\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 100 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and other subsonic runs may imply that with even stronger turbulence driving there is another supersonic regime. The transition occurs when thermal dissipation via radiative cooling is unable to reach an equilibrium with energy injection via turbulence driving and energy cascade until the gas temperature exceeds 107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT K and enters the bremsstrahlung regime. Regions along the cooling curve (see Figure 1) where Λ/T<0Λ𝑇0\partial\Lambda/\partial T<0∂ roman_Λ / ∂ italic_T < 0 are unstable during turbulence driving, hence giving rise to a temperature (and by proxy the subsonic/supersonic turbulence) bimodality.

Examining the distribution of kinetic saturation times, which is defined as ts=ϵk/ϵ˙isubscript𝑡𝑠subscriptitalic-ϵ𝑘subscript˙italic-ϵ𝑖t_{s}=\epsilon_{k}/\dot{\epsilon}_{i}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT during steady state turbulence, we observe it to be smooth and continuous among the supersonic runs, with timescales ranging from 3×102101Myr3superscript102superscript101Myr3\times 10^{-2}-10^{1}\;\mathrm{Myr}3 × 10 start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT roman_Myr. As with temperature, we see a ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT gradient where stronger turbulence driving leads to lower tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. \mathcal{M}caligraphic_M also shows a peculiar trend where it scales positively with ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for ts>3×101Myrsubscript𝑡𝑠3superscript101Myrt_{s}>3\times 10^{-1}\;\mathrm{Myr}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT > 3 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Myr, but negatively with ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for ts<3×101Myrsubscript𝑡𝑠3superscript101Myrt_{s}<3\times 10^{-1}\;\mathrm{Myr}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT < 3 × 10 start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_Myr The subsonic runs appear clustered, with distinct vertical offsets in tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT congruent to distinct values of ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, which is far different from the continuous distribution of tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT among the supersonic runs. This would suggest in the subsonic regime, tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT depends primarily on ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, with weaker dependancies (if any) on n𝑛nitalic_n or Z𝑍Zitalic_Z.

The physical distinction between subsonic and supersonic dissipation becomes clearer in Figure 3, where we examine the steady state of two particular 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resolution runs in more detail. On the top where ϵ˙i=1ergs1g1subscript˙italic-ϵi1ergsuperscripts1superscriptg1\mathrm{\dot{\epsilon}_{i}}=1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT = 1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, we observe large variations in ρ𝜌\rhoitalic_ρ and T𝑇Titalic_T across many orders of magnitude, and distinct overdense clumps in the gas. Meanwhile on the bottom, where ϵ˙i=3ergs1g1subscript˙italic-ϵi3ergsuperscripts1superscriptg1\mathrm{\dot{\epsilon}_{i}}=3\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT = 3 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, we observe a near uniform gas with perturbations of order unity. Subsonic motions in the gas are unable to produce shocks, which is a prerequisite for substructure formation in diffuse, non self-gravitating regimes. The right column paints a clearer picture of the different regimes of dissipation seen in Figure 2. Supersonic turbulence sees two distinct epochs of dissipation, characterized by an initial rapid dissipation epoch followed by a slow dissipation epoch. Subsonic turbulence on the other hand, sees only a slow dissipation epoch, cotemporal with an increase in thermal energy. This increase can also be seen in thermal energy of the supersonic run, and represents the cascading rate of kinetic energy into thermal energy via numerical hydrodynamic processes exceeding the radiative cooling rate. The physical nature behind the two dissipation epochs and the individual dissipative behaviours of thermal and kinetic energy will be examined more closely in Section 3.2.

Refer to caption
Figure 4: The dissipation rate of specific total energy, normalized by initial sound speed cs,02superscriptsubscript𝑐𝑠02c_{s,0}^{2}italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, over time for supersonic (green) and subsonic runs (red). \mathcal{M}caligraphic_M follows the same definition as the one in Figure 2, which is vavg/cs,0subscript𝑣avgsubscript𝑐𝑠0v_{\mathrm{avg}}/c_{s,0}italic_v start_POSTSUBSCRIPT roman_avg end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT, and tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT represent the kinetic energy saturation time. The shaded error regions represent 1σ𝜎\sigmaitalic_σ.

This distinction between the dissipation rates of supersonic and subsonic turbulence can be seen in Figure 4. Here the energy is normalized by the initial sound speed at turbulence turnoff cs,02superscriptsubscript𝑐𝑠02c_{s,0}^{2}italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT consistent with Stone et al. (1998). The normalized dissipation rates of the supersonic runs as a whole are a full dex higher than those of the subsonic runs up to a few tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, beyond which the rates begin to converge. The convergence is indicative of the dissipation of supersonic motions and dispersal of substructures as the box becomes homogenized, and is discussed more in depth in the following sections. Figure 4 encapsulates most of the entire range in tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT spanned by the subsonic runs, with the supersonic runs spanning a larger range up to hundreds and thousands of tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. As mentioned earlier, this is the result of the shorter timesteps of the subsonic runs.

3.2 Thermal vs Kinetic Supersonic Dissipation

Refer to caption
Figure 5: Normalized (top two) and unnormalized (bottom two) energy dissipation curves over time. Specific energy is normalized by initial sound speed cs,02superscriptsubscript𝑐𝑠02c_{s,0}^{2}italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, and time is normalized by kinetic saturation time tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The shaded regions represent 1σ𝜎\sigmaitalic_σ error regions and each coloured curve represents different values of ϵ˙isubscript˙italic-ϵi\mathrm{\dot{\epsilon}_{i}}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT roman_i end_POSTSUBSCRIPT. The first and third plots show thermal energy, and the second and fourth plots show kinetic energy.

Fixing our attention towards the supersonic runs, Figure 5 shows the time evolution of specific thermal (ϵgsubscriptitalic-ϵ𝑔\epsilon_{g}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT) and kinetic (ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT) energy binned in ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, with the top two plots normalized by initial sound speed cs,02superscriptsubscript𝑐𝑠02c_{s,0}^{2}italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at turbulence turnoff and tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. It is immediately clear that ϵgsubscriptitalic-ϵ𝑔\epsilon_{g}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT and ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT dissipate very differently on very different timescales. In the normalized plots, there are significant variations in thermal dissipative behaviour with respect to ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Beyond 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, 1 dex increases in ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT result in an approximate 0.5 dex drop in normalized thermal energy. Meanwhile, kinetic dissipation curves are more or less within each others’ error regions, although there is some weak dependancy on ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, with stronger turbulence driving leading to slightly shorter dissipation timescales relative to tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT.

Thermal dissipation is characterized by the initial rapid dissipation epoch as previously seen in Figure 3, on timescales of order 0.1ts0.1subscript𝑡𝑠0.1t_{s}0.1 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The separation between different binned values of ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is somewhat orderly in the normalized plot, manifesting themselves as a ”branch-off” from their universal initial values of cs,02/(γ1)superscriptsubscript𝑐𝑠02𝛾1c_{s,0}^{2}/(\gamma-1)italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_γ - 1 ) starting at 0.01ts0.01subscript𝑡𝑠0.01t_{s}0.01 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Higher ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT sees higher amounts of normalized thermal dissipation during the rapid epoch, before all runs eventually see dissipation rates decreasing significantly and their curves flattening. The bottom unnormalized thermal dissipation plot reveals a degree of non-linearity in the starting times of the rapid dissipation epoch with the ϵ˙i=10ergs1g1subscript˙italic-ϵ𝑖10ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=10\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 10 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT beginning at later times compared to the other runs. The unnormalized plot also reveals a range of energy plateaus shared across all runs afterwards, representing an inefficient cooling regime once the gas has reached 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K, and will be examined in more detail in Figures 6 and 7. Kinetic dissipation is characterized by a steady exponential drop over time. Unlike thermal dissipation, kinetic dissipation is far more weakly coupled to ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, despite the kinetic energy during steady state turbulence having a strong dependency on ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as seen in the initial values of the kinetic dissipation plots, both normalized and unnormalized. In the unnormalized plot, despite the gradient of initial ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, the dissipation curves all converge after 100100100100 Myr.

Refer to caption
Figure 6: Phase plots showing the time tracks of thermal versus kinetic energy for 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resolution runs corresponding to Z/Z=0.3𝑍subscript𝑍direct-product0.3Z/Z_{\odot}=0.3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.3 and n=0.1cm3𝑛0.1superscriptcm3n=0.1\;\mathrm{cm}^{-3}italic_n = 0.1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. Stars denote the turbulence driving turn off time where each curve begins, circular dots denote tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as one moves from the star and along the curve, and the squares denote t=100Myr𝑡100Myrt=100\;\mathrm{Myr}italic_t = 100 roman_Myr. The black line represents the x=y𝑥𝑦x=yitalic_x = italic_y line.
Refer to caption
Figure 7: A similar plot as Figure 5 but with curves binned in n𝑛nitalic_n as opposed to ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The shaded regions represent 1σ𝜎\sigmaitalic_σ. The thermal energy (top) plot shows only Z/Z=1𝑍subscript𝑍direct-product1Z/Z_{\odot}=1italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 1 runs with the shaded regions only representing variations in ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, while the kinetic energy (bottom) plot shows all metallicity runs with the shaded regions representing variations in both Z𝑍Zitalic_Z and ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. The black dotted line represents the specific thermal energy floor corresponding to the 1000100010001000 K temperature floor.
Refer to caption
Figure 8: The turbulence power spectra of the four high resolution 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs shown in Figure 6, where Z/Z=0.3𝑍subscript𝑍direct-product0.3Z/Z_{\odot}=0.3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.3 and n=0.1cm3𝑛0.1superscriptcm3n=0.1\;\mathrm{cm}^{-3}italic_n = 0.1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, and ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT being 0.01, 0.1, 1.0 and 10.0 ergs1g1ergsuperscripts1superscriptg1\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT from left to right. The bottom row plots the turbulence power spectra during steady state turbulence, after tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, as well as k5/3superscript𝑘53k^{-5/3}italic_k start_POSTSUPERSCRIPT - 5 / 3 end_POSTSUPERSCRIPT reference lines representing the inertial regime of the kolmogorov spectrum. The cyan coloured region represents 1σ1𝜎1\sigma1 italic_σ across the last twenty outputs of the steady state turbulence epoch. The top row plots the power spectra from turbulence turn off up to 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, with the colour denoting ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. The value of tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is also shown across each of the four runs. Both the top and bottom row plots are smoothed using degree-5 polynomial splines for visualization purposes. The leftmost ϵ˙i=0.01ergs1g1subscript˙italic-ϵ𝑖0.01ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=0.01\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0.01 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT column does not show a 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT power spectrum since the simulation run ended before 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT.

We show the time tracks of a few higher resolution runs comparing thermal and kinetic energy directly in Figure 6. We observe rapid dissipation in thermal energy within the first 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for higher values of ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and relatively consistent exponential kinetic dissipation rate with no plateaus, consistent with our observations of Figure 5. The latter can be further evidenced by the approximately even horizontal spacing between circles (most pronounced for ϵ˙i=10ergs1g1subscript˙italic-ϵ𝑖10ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=10\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 10 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT), with ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT being a viable proxy for time. We also confirm the convergence of kinetic and thermal energy as seen in the unnormalized bottom plots of Figure 5 - regardless of initial energy or turbulence driving strength, for t>30ts𝑡30subscript𝑡𝑠t>30t_{s}italic_t > 30 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT the dissipation curves overlap. Additionally, the squares which mark t=100Myr𝑡100Myrt=100\;\mathrm{Myr}italic_t = 100 roman_Myr are more or less at the same location across all four runs where the kinetic and thermal dissipation curves have overlapped. This would suggest 100100100100 Myr to be a ”universal” turbulence dissipation timescale, depending on n𝑛nitalic_n and Z𝑍Zitalic_Z but fully indepedent from ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. This independence will also be explored further in Section 3.5 on the dissipation of overdensities. The overlap also coincides with a plateau in thermal energy, formed by a combination of ineffficient cooling as the gas reaches 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K and the continuous cascade from kinetic to thermal energy.

We observe variations in the end state with respect to initial conditions in Figure 7, with all quantities unnormalized and curves separated by number density. The top thermal energy plot affirms the convergence of thermal energy curves observed in Figure 6 across a larger, lower resolution sample size, with varied turbulence driving across fixed initial conditions. Since the metallicity is fixed, the spread within each n𝑛nitalic_n bin represents only variations in ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. We denote the temperature floor with the black dotted line as an unphysical asymptote the thermal energy approaches across all runs, but as seen in Figure 1, the cooling rates drop by many orders of magnitude below 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K, so the actual physical evolution of thermal energy would not be a significant downwards deviation from a horizontal asymptote. We see that this convergence is independant of turbulence driving as the error regions decrease in size, and the value of ϵgsubscriptitalic-ϵ𝑔\epsilon_{g}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT at which the convergence begins, depends on n𝑛nitalic_n and Z𝑍Zitalic_Z (since inefficient cooling can also arise from low n𝑛nitalic_n and/or low Z𝑍Zitalic_Z). We also affirm that these convergences occur at thermal energy plateaus after the initial rapid dissipation epoch. Kinetic energy, as with Figure 6 and Figure 5 also shows convergence and overlap. This convergence and overlap via binning both in n𝑛nitalic_n and in ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, suggests a universal kinetic energy dissipation timescale. We note that this convergence point is at 100100100100 Myr in both Figures 5 and 7, the same location marked by the squares in Figure 6.

The spatial scaling of the turbulence can be analyzed via the turbulence power spectrum (Bavassano et al., 1982; Burkhart et al., 2015; Bustard & Oh, 2023), shown in Figure 8. We compute the turbulence power spectra as follows:

Pk(k)=12k2𝑑Ωk|v^(k)|2subscript𝑃𝑘𝑘12superscript𝑘2differential-dsubscriptΩ𝑘superscript^𝑣𝑘2P_{k}(k)=\dfrac{1}{2}k^{2}\int d\Omega_{k}|\hat{v}(\vec{k})|^{2}italic_P start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_k ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∫ italic_d roman_Ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | over^ start_ARG italic_v end_ARG ( over→ start_ARG italic_k end_ARG ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (5)

where 𝑑Ωkdifferential-dsubscriptΩ𝑘\int d\Omega_{k}∫ italic_d roman_Ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT represents a solid angle integral over k𝑘kitalic_k-space, and k𝑘kitalic_k represents the radial component of k𝑘\vec{k}over→ start_ARG italic_k end_ARG. v^(k)^𝑣𝑘\hat{v}(\vec{k})over^ start_ARG italic_v end_ARG ( over→ start_ARG italic_k end_ARG ) represents the Fourier Transform (FT) of the velocity field:

v^(k)=1(2π)3v(r)eikrd3r^𝑣𝑘1superscript2𝜋3𝑣𝑟superscript𝑒𝑖𝑘𝑟superscript𝑑3𝑟\hat{v}(\vec{k})=\dfrac{1}{(2\pi)^{3}}\int\vec{v}(\vec{r})e^{-i\vec{k}\cdot% \vec{r}}d^{3}rover^ start_ARG italic_v end_ARG ( over→ start_ARG italic_k end_ARG ) = divide start_ARG 1 end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ∫ over→ start_ARG italic_v end_ARG ( over→ start_ARG italic_r end_ARG ) italic_e start_POSTSUPERSCRIPT - italic_i over→ start_ARG italic_k end_ARG ⋅ over→ start_ARG italic_r end_ARG end_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_r (6)

The factor of 1/2121/21 / 2 in eq. (5) represents the conversion from velocity-squared to specific kinetic energy.

The dissipation is visible in the downward shift in the power spectra in the bottom plot, with relatively even spacing between the power spectra at tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The upwards bump at higher k𝑘kitalic_k along the inertial regimes shows the bottleneck effect (Dobler et al., 2003; Donzis & Sreenivasan, 2010), representing a buildup of energy below the dissipation range. The inertial regimes of our runs match the kolmogorov k5/3superscript𝑘53k^{-5/3}italic_k start_POSTSUPERSCRIPT - 5 / 3 end_POSTSUPERSCRIPT power law (Kolmogorov, 1941)fairly well, both during and after steady state turbulence. This is indicative of the weakly supersonic nature of our runs with \mathcal{M}caligraphic_M only being on order unity (Ossenkopf & Mac Low, 2002; Larson, 1981). Notably, stronger turbulence driving does not lead to an increasingly shocked gas with a k2superscript𝑘2k^{-2}italic_k start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT power law (Burgers, 1948), but rather a near-congruent upwards shift in the entire spectrum. The shape of the spectrum remains preserved during dissipation, consistent with (Vazza et al., 2009).

3.3 Comparisons with Isothermal, Compressible GMC Dissipation

Refer to caption
Figure 9: A comparison between our thermal and kinetic dissipation curves and the kinetic dissipation curves of Stone et al. (1998) (black), Ostriker et al. (1999) (cyan) and Ostriker et al. (2001) (magenta). The latter three curves show dissipation under artificial viscosity, an isothermal equation of state and the presence of magnetic fields where B=β1/2×1.4(T/10K)1/2(n/102cm3)μG𝐵superscript𝛽121.4superscript𝑇10K12𝑛superscript102superscriptcm3𝜇GB=\beta^{-1/2}\times 1.4\;(T/10\;\mathrm{K})^{1/2}(n/10^{2}\;\mathrm{cm}^{-3})% \;\mu\mathrm{G}italic_B = italic_β start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT × 1.4 ( italic_T / 10 roman_K ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ( italic_n / 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT ) italic_μ roman_G (so β=𝛽\beta=\inftyitalic_β = ∞ represents pure hydrodynamics). The cyan and magenta dissipation curves include self-gravity. tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT represents the kinetic energy saturation time defined as ϵk/ϵ˙isubscriptitalic-ϵ𝑘subscript˙italic-ϵ𝑖\epsilon_{k}/\dot{\epsilon}_{i}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT during steady state turbulence, and the energies are normalized by sound speed squared cs,02superscriptsubscript𝑐𝑠02c_{s,0}^{2}italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT right at turbulence driving turnoff.

We compare our dissipation curves with various kinetic dissipation curves in GMCs in Figure 9. The black curve from Stone et al. (1998) serves as a fiducial comparison representing dissipation with pure hydrodynamics without magnetic fields or self gravity. Two key distinctions emerge - as also seen in Figure 2 our supersonic runs are only weakly supersonic with M𝑀Mitalic_M on order unity, while Stone et al. (1998) sees stronger supersonic motions with M𝑀Mitalic_M being an order of magnitude higher. The dissipation timescales are also significantly shorter relative to tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, where kinetic energy dissipates well within tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT while for our runs both thermal and kinetic energy dissipate on timescales greater than 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Both distinctions are reflective of the different astrophysical medium of interest (GMCs are much cooler than CGMs, which allows significantly weaker velocity perturbations to drive stronger supersonic turbulent motions) and of the different physical processes through which dissipation occurs (energy cascade and radiative cooling for us, artificial viscosity for Stone et al. (1998)). The cyan and magenta curves from Ostriker et al. (1999) and Ostriker et al. (2001) illustrate the effects of self gravity (both) and magnetic fields (latter). While they play a minor role in stabilization against dissipation, their dissipation timescales remain well within tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT.

We note that the artificial viscosity employed by the aforementioned authors via the code ZEUS (Stone & Norman, 1992a) is necessary to capture and thermalize shocks, while KRATOS adopts higher order Godunov solver. Physically, in our work, kinetic dissipation occurs only via shock thermalization through adiabatic compression. Numerical viscosity also plays a role in shock thermalization (Nelson et al., 2013), though its effects diminish with more accurate solvers and higher resolution grids (Zhang et al., 2003). Hearkening back to Figure 8, it’s likely that the bottleneck (Dobler et al., 2003) is the result of both a lack of subgrid viscous forces accounting for dissipation at kolmogorov microscales, and insufficient numerical viscosity given our numerical scheme and resolution.

3.4 Energy Dissipation Timescales

Refer to caption
Figure 10: 2D contour maps of the half energy dissipation timescales thsubscript𝑡t_{h}italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, defined as ϵ(th)=0.5ϵ0italic-ϵsubscript𝑡0.5subscriptitalic-ϵ0\epsilon(t_{h})=0.5\epsilon_{0}italic_ϵ ( italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) = 0.5 italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, across only supersonic runs. The blank regions on the top left above the diagonal represent subsonic runs, and the blank regions on the bottom represent supersonic runs which did not dissipate half the component of energy in question. The first and second rows shows total energy, the middle third and fourth rows show thermal energy, and the bottom fifth and sixth rows show kinetic energy

We quantify the raw dissipation timescales thsubscript𝑡t_{h}italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT in Figure 10, defined as the time it takes for the total, thermal or kinetic energy to drop by half. The primary trend on the top plot shows decreasing dissipation timescales with increasing ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT across all metallicities, with some n𝑛nitalic_n dependence for higher metallicities particularly at Z/Z=1𝑍subscript𝑍direct-product1Z/Z_{\odot}=1italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 1 and Z/Z=3𝑍subscript𝑍direct-product3Z/Z_{\odot}=3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 3. There is little discernible correlation between the dissipation timescale and metallicity - rather, increased metallicity leads to enhanced cooling, allowing for some runs to reach the supersonic regime which would have been subsonic at lower metallicities, as seen with the increased number of supersonic runs above the diagonal for higher metallicity contour plots. On the middle and bottom plots, we can affirm our earlier observations from Section 3.2. Noting that blank bins below the diagonal in the thermal dissipation timescale represent supersonic runs that could not dissipate half their thermal energy over the simulation and interpreting those bins has having longer dissipation timescales, we see a steep drop in thsubscript𝑡t_{h}italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT with increased ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and n𝑛nitalic_n of several orders of magnitude. Meanwhile in the bottom plot, while a ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT dependence in the kinetic dissipation timescales can be observed, it is far less steep and spans fewer orders of magnitude compared to the thermal dissipation timescales. No clear n𝑛nitalic_n dependence can be observed either. The subsonic runs occupy the top left regions of all plots, above the diagonal, representing initial parameters where cooling is insufficient in preventing the gas from reaching a hot subsonic turbulent state.

Refer to caption
Figure 11: Dimensionless dissipation timescales thsubscript𝑡t_{h}italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT, defined as ϵ(th)=0.5ϵ0italic-ϵsubscript𝑡0.5subscriptitalic-ϵ0\epsilon(t_{h})=0.5\epsilon_{0}italic_ϵ ( italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT ) = 0.5 italic_ϵ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, for both kinetic (top) and thermal (bottom) energy as a function of ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Kinetic energy is normalized by tcross,0subscript𝑡cross0t_{\mathrm{cross,0}}italic_t start_POSTSUBSCRIPT roman_cross , 0 end_POSTSUBSCRIPT while thermal energy is normalized by tcool,0subscript𝑡cool0t_{\mathrm{cool,0}}italic_t start_POSTSUBSCRIPT roman_cool , 0 end_POSTSUBSCRIPT. Kinetic dissipation timescales are additionally separated between supersonic (blue) and subsonic (black) runs.

We also present a set of dimensionless dissipation timescales in Figure 11. We define a kinetic timescale in tcross,0=/vrmssubscript𝑡cross0delimited-⟨⟩subscript𝑣rmst_{\mathrm{cross,0}}=\ell/\langle v_{\mathrm{rms}}\rangleitalic_t start_POSTSUBSCRIPT roman_cross , 0 end_POSTSUBSCRIPT = roman_ℓ / ⟨ italic_v start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT ⟩ and a thermal timescale in tcool,0=Eg,0/ini2Λ(Ti)Δxi3subscript𝑡cool0subscript𝐸𝑔0superscript𝑖superscriptsubscript𝑛𝑖2Λsubscript𝑇𝑖Δsuperscriptsubscript𝑥𝑖3t_{\mathrm{cool,0}}=E_{g,0}/\sum^{i}n_{i}^{2}\Lambda(T_{i})\Delta x_{i}^{3}italic_t start_POSTSUBSCRIPT roman_cool , 0 end_POSTSUBSCRIPT = italic_E start_POSTSUBSCRIPT italic_g , 0 end_POSTSUBSCRIPT / ∑ start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT italic_n start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_Λ ( italic_T start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) roman_Δ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. The former represents the initial crossing time at turbulence turnoff, where vrmsdelimited-⟨⟩subscript𝑣rms\langle v_{\mathrm{rms}}\rangle⟨ italic_v start_POSTSUBSCRIPT roman_rms end_POSTSUBSCRIPT ⟩ represents the mass averaged rms velocity across every cell, while the latter represents the initial cooling time at turbulence turnoff, where Eg,0subscript𝐸𝑔0E_{g,0}italic_E start_POSTSUBSCRIPT italic_g , 0 end_POSTSUBSCRIPT represents the total initial thermal energy and the denominator represents the total cooling rate integrated across every cell. Subsonic kinetic dissipation shows a universal dimensionless timescale of approximately 0.6tcross,00.6subscript𝑡cross00.6t_{\mathrm{cross,0}}0.6 italic_t start_POSTSUBSCRIPT roman_cross , 0 end_POSTSUBSCRIPT, while supersonic thermal dissipation shows a convergence towards a universal dimensionless timescale of roughly 8.5tcool,08.5subscript𝑡cool08.5t_{\mathrm{cool,0}}8.5 italic_t start_POSTSUBSCRIPT roman_cool , 0 end_POSTSUBSCRIPT.

3.5 Density Homogenization

Refer to caption
Figure 12: The dissipation of the clumping factor ρ2/ρ21delimited-⟨⟩superscript𝜌2superscriptdelimited-⟨⟩𝜌21\langle\rho^{2}\rangle/\langle\rho\rangle^{2}-1⟨ italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_ρ ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 over time. Consistent with previous plots, the coloured lines show various binnings in initial simulation parameters across supersonic runs. The black line and shaded region shows the average and 1σ1𝜎1\sigma1 italic_σ across all subsonic runs, and the dotted line shows 5σ5𝜎5\sigma5 italic_σ across all timesteps for all subsonic runs. Top: Clumping factor dissipation binned in ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Bottom: Clumping factor dissipation binned in n𝑛nitalic_n

.

As seen in Figure 3 and as discussed in previous sections, there are density contrasts spanning multiple orders of magnitude during steady state turbulence. We define overall scale of such density contrasts via the clumping factor, which is defined as C=ρ2/ρ2𝐶delimited-⟨⟩superscript𝜌2superscriptdelimited-⟨⟩𝜌2C=\langle\rho^{2}\rangle/\langle\rho\rangle^{2}italic_C = ⟨ italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_ρ ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where ρdelimited-⟨⟩𝜌\langle\rho\rangle⟨ italic_ρ ⟩ represents a spatial average across all cells. We can examine the broad density homogenization of the medium via dissipation in the clumping factor. We label the full definition of C𝐶Citalic_C in subsequent plots and figures for clarity, and plot C1𝐶1C-1italic_C - 1 given C=1𝐶1C=1italic_C = 1 represents a completely uniform medium.

Refer to caption
Figure 13: Phase plots tracking the temporal evolution of specific thermal (Top), specific kinetic energy (Bottom) relative to clumping factor. Four 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resolution runs are shown, corresponding to Z/Z=0.3𝑍subscript𝑍direct-product0.3Z/Z_{\odot}=0.3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.3, ϵ˙i=1ergs1g1subscript˙italic-ϵ𝑖1ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and n𝑛nitalic_n ranging from 0.01cm30.01superscriptcm30.01\;\mathrm{cm}^{-3}0.01 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT to n=10cm3𝑛10superscriptcm3n=10\;\mathrm{cm}^{-3}italic_n = 10 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. Similar to Figure 6, stars denote the turbulence driving turn off time where each curve begins, circular dots denote tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as one moves from the star and along the curve, and squares denote the 100 Myr mark. The black dotted line in the top plot marks the T=1000K𝑇1000KT=1000\;\mathrm{K}italic_T = 1000 roman_K point.

We show the time evolution of C𝐶Citalic_C in Figure 12 with respect to both ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and n𝑛nitalic_n. An expected observation of the top plot shows a positive correlation between ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and steady state clumping, as well as a significant valley between the clumping factors of subsonic and supersonic runs. Subsonic clumping factors do show a decrease of around a single dex, although this is insignificant compared to the several-dex decreases in the C𝐶Citalic_C of the supersonic runs, especially when considering the 11-1- 1 shift in the y axis. We define a supersonic tubulent medium to be homogenized if its clumping factor falls below a density homogenization limit of Csub1.0132subscript𝐶sub1.0132C_{\mathrm{sub}}\approx 1.0132italic_C start_POSTSUBSCRIPT roman_sub end_POSTSUBSCRIPT ≈ 1.0132 representing 5σ5𝜎5\sigma5 italic_σ from the mean inital subsonic clumping factor. Functionally, this definition translates to a density homogenization timescale tDissCsuperscriptsubscript𝑡DissCt_{\mathrm{Diss}}^{\mathrm{C}}italic_t start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_C end_POSTSUPERSCRIPT, beyond which the density contrasts of a dissipated supersonic turbulent gas becomes indistinguishable from a subsonic turbulent gas.

Despite the clear gradient in inital clumping factor positively correlated with ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, the curves and error regions converge and reach Csubsubscript𝐶subC_{\mathrm{sub}}italic_C start_POSTSUBSCRIPT roman_sub end_POSTSUBSCRIPT on roughly the same timescales of around 100 Myr, albeit with some spread between 30303030 and 300300300300 Myr. When binned in n𝑛nitalic_n on the bottom plot, a positive n𝑛nitalic_n dependence in tDissCsuperscriptsubscript𝑡DissCt_{\mathrm{Diss}}^{\mathrm{C}}italic_t start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_C end_POSTSUPERSCRIPT emerges. The shrinkage of the error regions encapsulates the same variations in ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT converging in the top plot. Hearkening back to Figures 5 and 7, dissipation in C𝐶Citalic_C is similar to but not congruent to dissipation in ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. While both quantities dissipates exponentially, kinetic energy dissipation does not depend on n𝑛nitalic_n.

We examine the intercorrelation between clumping factor dissipation and energy dissipation more closely in Figure 13. Since the locations of the squares denote a fixed time t=100𝑡100t=100italic_t = 100 Myr, the lower the position of the square, the shorter tDissCsuperscriptsubscript𝑡DissCt_{\mathrm{Diss}}^{\mathrm{C}}italic_t start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT start_POSTSUPERSCRIPT roman_C end_POSTSUPERSCRIPT is. We note that only the two lowest values of n𝑛nitalic_n are remotely representative of conditions in the CGM, and our purpose of showing n=1cm3𝑛1superscriptcm3n=1\;\mathrm{cm}^{-3}italic_n = 1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and n=10cm3𝑛10superscriptcm3n=10\;\mathrm{cm}^{-3}italic_n = 10 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT is to illustrate a trend. Given their similar dissipative behaviours in Figures 5 and 12, the middle plot unsurpsingly shows a steady power law relation between C𝐶Citalic_C and ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. The correlation with thermal dissipation in the top plot is much less smooth, and mirrors the the time tracks in Figure 6. C𝐶Citalic_C only shows a power law relation with ϵgsubscriptitalic-ϵ𝑔\epsilon_{g}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT during the rapid thermal dissipation epoch, continues to dissipate during the ϵgsubscriptitalic-ϵ𝑔\epsilon_{g}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT plateau. There is a nonlinear relation between initial C𝐶Citalic_C and n𝑛nitalic_n, with n=0.01cm3𝑛0.01superscriptcm3n=0.01\;\mathrm{cm}^{-3}italic_n = 0.01 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and n=10cm3𝑛10superscriptcm3n=10\;\mathrm{cm}^{-3}italic_n = 10 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT exhibiting higher clumping factors than the other runs. We again observe and affirm the negative correlation in the top plot between n𝑛nitalic_n and C𝐶Citalic_C at 100 Myr as seen in Figure 12, but we also note the dissipated steady state behaviour of ϵgsubscriptitalic-ϵ𝑔\epsilon_{g}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT. The more diffuse runs with lower clumping factors at 100 Myr have hotter thermal energy plateaus, with the n=0.01cm3𝑛0.01superscriptcm3n=0.01\;\mathrm{cm}^{-3}italic_n = 0.01 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and n=0.1cm3𝑛0.1superscriptcm3n=0.1\;\mathrm{cm}^{-3}italic_n = 0.1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT not even reaching the 1000 K temperature floor.

Refer to caption
Figure 14: The ratio between tcoolsubscript𝑡coolt_{\mathrm{cool}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT and tcrosssubscript𝑡crosst_{\mathrm{cross}}italic_t start_POSTSUBSCRIPT roman_cross end_POSTSUBSCRIPT at turbulence turnoff versus the clumping factor ρ2/ρ21delimited-⟨⟩superscript𝜌2superscriptdelimited-⟨⟩𝜌21\langle\rho^{2}\rangle/\langle\rho\rangle^{2}-1⟨ italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_ρ ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1 at 100 Myr. Each point represents a supersonic turbulent run, with the colour bar and marker colours denoting the normalized thermal energy ϵg/cs,02subscriptitalic-ϵ𝑔superscriptsubscript𝑐𝑠02\epsilon_{g}/c_{s,0}^{2}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT at 100 Myr. As with previous figures, cs,0subscript𝑐𝑠0c_{s,0}italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT represents the sound speed at turbulence turnoff.

We affirm this trend holds statistically via an examination across all supersonic turbulent runs in Figure 14, and it becomes evident the n𝑛nitalic_n dependence of tDissCsuperscriptsubscript𝑡Diss𝐶t_{\mathrm{Diss}}^{C}italic_t start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT represents a more fundamental positive dependance on tcool/tcrosssubscript𝑡coolsubscript𝑡crosst_{\mathrm{cool}}/t_{\mathrm{cross}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT roman_cross end_POSTSUBSCRIPT. The more inefficient the cooling, the more uniform the medium becomes. There are no discernable trends in the colouring of the data points, showing that neither tcool/tcrosssubscript𝑡coolsubscript𝑡crosst_{\mathrm{cool}}/t_{\mathrm{cross}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT roman_cross end_POSTSUBSCRIPT nor tDissCsuperscriptsubscript𝑡Diss𝐶t_{\mathrm{Diss}}^{C}italic_t start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT correlate significantly with the process of thermal dissipation. Rather, the dependence of tDissCsuperscriptsubscript𝑡Diss𝐶t_{\mathrm{Diss}}^{C}italic_t start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT on tcool/tcrosssubscript𝑡coolsubscript𝑡crosst_{\mathrm{cool}}/t_{\mathrm{cross}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT roman_cross end_POSTSUBSCRIPT reflects a dependence on the current dynamical state of the gas. Going back to Figure 13, despite having similar kinetic energies at 100 Myr, the n=0.01cm3𝑛0.01superscriptcm3n=0.01\;\mathrm{cm}^{-3}italic_n = 0.01 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT run has distinctly more thermal energy than the n=10cm3𝑛10superscriptcm3n=10\;\mathrm{cm}^{-3}italic_n = 10 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT run at that time, representing a smaller tcool/tcrosssubscript𝑡coolsubscript𝑡crosst_{\mathrm{cool}}/t_{\mathrm{cross}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT / italic_t start_POSTSUBSCRIPT roman_cross end_POSTSUBSCRIPT and hence a shorter density homogenization timescale, as the gas cannot cool fast enough to prevent the bulk motions from diffusing and smoothing out overdensities.

3.6 Fourier Analysis of Turbulent Clouds

In this section we characterize the turbulence driven overdensities and clouds in more detail. While C𝐶Citalic_C can broadly describe the overall ”clumpiness” of the medium, it is insufficient in characterizing properties such as the spatial scales or \mathcal{M}caligraphic_M of the substructures. We extend our power spectrum analysis from Figure 8 to more gas properties, following similar methods to those of Federrath & Klessen (2013). For some field q(r)𝑞𝑟q(\vec{r})italic_q ( over→ start_ARG italic_r end_ARG ) over 3superscript3\mathbb{R}^{3}blackboard_R start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, its power spectrum is defined as

Pq(k)=k2𝑑Ωk|q^(k)|2subscript𝑃𝑞𝑘superscript𝑘2differential-dsubscriptΩ𝑘superscript^𝑞𝑘2P_{q}(k)=k^{2}\int d\Omega_{k}|\hat{q}(\vec{k})|^{2}italic_P start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( italic_k ) = italic_k start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∫ italic_d roman_Ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT | over^ start_ARG italic_q end_ARG ( over→ start_ARG italic_k end_ARG ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (7)

where 𝑑Ωkdifferential-dsubscriptΩ𝑘\int d\Omega_{k}∫ italic_d roman_Ω start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT represents a solid angle integral over k𝑘kitalic_k-space, and k𝑘kitalic_k represents the radial component of k𝑘\vec{k}over→ start_ARG italic_k end_ARG. q^(k)^𝑞𝑘\hat{q}(\vec{k})over^ start_ARG italic_q end_ARG ( over→ start_ARG italic_k end_ARG ) represents the Fourier Transform (FT) of the field q𝑞qitalic_q:

q^(k)=1(2π)3q(r)eikrd3r^𝑞𝑘1superscript2𝜋3𝑞𝑟superscript𝑒𝑖𝑘𝑟superscript𝑑3𝑟\hat{q}(\vec{k})=\dfrac{1}{(2\pi)^{3}}\int q(\vec{r})e^{-i\vec{k}\cdot\vec{r}}% d^{3}rover^ start_ARG italic_q end_ARG ( over→ start_ARG italic_k end_ARG ) = divide start_ARG 1 end_ARG start_ARG ( 2 italic_π ) start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG ∫ italic_q ( over→ start_ARG italic_r end_ARG ) italic_e start_POSTSUPERSCRIPT - italic_i over→ start_ARG italic_k end_ARG ⋅ over→ start_ARG italic_r end_ARG end_POSTSUPERSCRIPT italic_d start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_r (8)
Refer to caption
Figure 15: Density (Left) and temperature (Right) slice plots of four 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resolution runs corresponding to Z/Z=0.3𝑍subscript𝑍direct-product0.3Z/Z_{\odot}=0.3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.3, n=0.1cm3𝑛0.1superscriptcm3n=0.1\;\mathrm{cm}^{-3}italic_n = 0.1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. Each row corresponds to a different value of ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - from top to bottom they are: 0.1ergs1g10.1ergsuperscripts1superscriptg10.1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}0.1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, 1ergs1g11ergsuperscripts1superscriptg11\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, 3ergs1g13ergsuperscripts1superscriptg13\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}3 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and 10ergs1g110ergsuperscripts1superscriptg110\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}10 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. The densities and colour bar are normalized by the initial density n=0.1cm3𝑛0.1superscriptcm3n=0.1\;\mathrm{cm}^{-3}italic_n = 0.1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT. The columns show the gas at different times, the first and third columns right at turbulence turnoff, the second and fourth columns at 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. On the density slice plots, dashed white lines show =11\mathcal{M}=1caligraphic_M = 1 contours and solid magenta lines show =33\mathcal{M}=3caligraphic_M = 3 contours. On the temperature slice plots, dashed lime lines show ρ/mp/n=3𝜌subscript𝑚𝑝𝑛3\rho/m_{p}/n=3italic_ρ / italic_m start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / italic_n = 3 contours .

Figure 15 paints a physical picture of the gas during steady state turbulence and during dissipation. While it’s evident from Figure 12 there should be higher density contrasts with higher ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, another effect of stronger driving is a dramatic shrinking in the typical clump sizes. The ϵ˙i=1ergs1g1subscript˙italic-ϵ𝑖1ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT run sees clumps of order 102superscript10210^{2}10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT pc, while the ϵ˙i=10ergs1g1subscript˙italic-ϵ𝑖10ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=10\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 10 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT run sees clumps of order 100superscript10010^{0}10 start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT to 101superscript10110^{1}10 start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT pc in size. There is a nonlinear though clear inverse relation between driving strength and clump size. Weak turbulence driving, as seen in the ϵ˙i=0.1ergs1g1subscript˙italic-ϵ𝑖0.1ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=0.1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0.1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT run, sees diffuse ”bubbles” rather than dense clumps. The =11\mathcal{M}=1caligraphic_M = 1 contours weakly trace the boundaries between diffuse and dense regions, and whose ubiquity shows both regions to be broadly supersonic. The =33\mathcal{M}=3caligraphic_M = 3 contours show even more limited overlap with overdense regions in the top two plots, but trace dense regions very well in the bottom two plots. Only under strong turbulence driving where ϵ˙i3ergs1g1subscript˙italic-ϵ𝑖3ergsuperscripts1superscriptg1\dot{\epsilon}_{i}\geq 3\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≥ 3 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT do we observe strongly supersonic clumps. In the temperature plots on the two left columns, the cool regions correlate very well with the dense regions on the two right columns, and share similar temperatures of 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT to 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K across all four runs. This would suggest that faster bulk motions within the clumps is a significant contributor towards their increased compressibility.

Refer to caption
Refer to caption
Figure 16: The density (Top) and \mathcal{M}caligraphic_M (Bottom) power spectra of the four high resolution 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs shown in Figure 6, where Z/Z=0.3𝑍subscript𝑍direct-product0.3Z/Z_{\odot}=0.3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.3 and n=0.1cm3𝑛0.1superscriptcm3n=0.1\;\mathrm{cm}^{-3}italic_n = 0.1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, and ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT being 0.1, 1, 3.0 and 10.0 ergs1g1ergsuperscripts1superscriptg1\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT from left to right. The bottom row plots the turbulence power spectra during steady state turbulence, after tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The cyan coloured region represents 1σ1𝜎1\sigma1 italic_σ across the last twenty outputs of the steady state turbulence epoch. The top row plots the power spectra from turbulence turn off up to 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, with the colour denoting ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT. The value of tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is also shown across each of the four runs. Both the top and bottom row plots are smoothed using degree-5 polynomial splines for visualization purposes.

We charactize the spatial scalings of \mathcal{M}caligraphic_M via a power spectrum of the mach number. Following equations (7) and (8) we define Pq(k)=P(k)subscript𝑃𝑞𝑘subscript𝑃𝑘P_{q}(k)=P_{\mathcal{M}}(k)italic_P start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ( italic_k ) = italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT ( italic_k ) and q^(k)=^(k)^𝑞𝑘^𝑘\hat{q}(\vec{k})=\hat{\mathcal{M}}(\vec{k})over^ start_ARG italic_q end_ARG ( over→ start_ARG italic_k end_ARG ) = over^ start_ARG caligraphic_M end_ARG ( over→ start_ARG italic_k end_ARG ) from q(r)=(r)𝑞𝑟𝑟q(\vec{r})=\mathcal{M}(\vec{r})italic_q ( over→ start_ARG italic_r end_ARG ) = caligraphic_M ( over→ start_ARG italic_r end_ARG ), where (r)𝑟\mathcal{M}(\vec{r})caligraphic_M ( over→ start_ARG italic_r end_ARG ) represents the spatial mach number define as:

(r)=|v(r)|cs(r)𝑟𝑣𝑟subscript𝑐𝑠𝑟\mathcal{M}(\vec{r})=\dfrac{|\vec{v}(\vec{r})|}{c_{s}(\vec{r})}caligraphic_M ( over→ start_ARG italic_r end_ARG ) = divide start_ARG | over→ start_ARG italic_v end_ARG ( over→ start_ARG italic_r end_ARG ) | end_ARG start_ARG italic_c start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( over→ start_ARG italic_r end_ARG ) end_ARG (9)

Similarly we an characterize the spatial scalings of the overdensities using the matter power spectrum Pδ(k)subscript𝑃𝛿𝑘P_{\delta}(k)italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT ( italic_k ), for which q(r)=δ(r)𝑞𝑟𝛿𝑟q(\vec{r})=\delta(\vec{r})italic_q ( over→ start_ARG italic_r end_ARG ) = italic_δ ( over→ start_ARG italic_r end_ARG ) where δ𝛿\deltaitalic_δ is a dimensionless overdensity parameter defined as

δ(r)=ρ(r)ρ¯ρ¯𝛿𝑟𝜌𝑟¯𝜌¯𝜌\delta(\vec{r})=\dfrac{\rho(\vec{r})-\bar{\rho}}{\bar{\rho}}italic_δ ( over→ start_ARG italic_r end_ARG ) = divide start_ARG italic_ρ ( over→ start_ARG italic_r end_ARG ) - over¯ start_ARG italic_ρ end_ARG end_ARG start_ARG over¯ start_ARG italic_ρ end_ARG end_ARG (10)

where ρ¯¯𝜌\bar{\rho}over¯ start_ARG italic_ρ end_ARG represents the spatially averaged density across the box, although functionally this equates to the initial density of the run.

In Figure 16, we characterize the dissipation of the dense clumps as seen in Figure 15 via the time evolution Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT and Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT across the same four runs. Both power spectra exhibit broken kαsuperscript𝑘𝛼k^{-\alpha}italic_k start_POSTSUPERSCRIPT - italic_α end_POSTSUPERSCRIPT power laws, with higher ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT corresponding to shallower α𝛼\alphaitalic_α in the 0.07<k<kd0.07𝑘subscript𝑘𝑑0.07<k<k_{d}0.07 < italic_k < italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT range, where kd1pc1subscript𝑘𝑑1superscriptpc1k_{d}\approx 1\;\mathrm{pc}^{-1}italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ≈ 1 roman_pc start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT during turbulence driving. We will refer to this range in k𝑘kitalic_k as the linear range, the slope of the linear range as αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT and αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT for Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT and Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT repectively, and k>kd𝑘subscript𝑘𝑑k>k_{d}italic_k > italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT as the dissipation range. The maxima of Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT for ϵ˙i=3ergs1g1subscript˙italic-ϵ𝑖3ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=3\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 3 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT and ϵ˙i=10ergs1g1subscript˙italic-ϵ𝑖10ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=10\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 10 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT are around the same as if not lower than the maxima for ϵ˙i=1ergs1g1subscript˙italic-ϵ𝑖1ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, representing a larger proportion of the overdensities being captured at smaller scales. A shallow αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT represents not just the presence of distinct clumps at small scales, but hierarchically structured clumping, where dense clumps host even denser clouds within. For Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT, a shallower αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT represents those same clumps being increasingly supersonic due to underpressurization from enhanced cooling. The rise in the linear range is also noticeably sharper from ϵ˙i=3ergs1g1subscript˙italic-ϵ𝑖3ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=3\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 3 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT to ϵ˙i=10ergs1g1subscript˙italic-ϵ𝑖10ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=10\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 10 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, which is consistent with Figure 15.

During dissipation, Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT sees a steady increase in αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT and decrease in kdsubscript𝑘𝑑k_{d}italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT, which represents hierarchical dissipation of substructures beginning at the smallest scales. We see this in Figure 15 (primarily for ϵ˙i1ergs1g1subscript˙italic-ϵ𝑖1ergsuperscripts1superscriptg1\dot{\epsilon}_{i}\geq 1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≥ 1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT), where at 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, the gas is smoother on smaller scales compared to how it is at tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, while preserving its clumpiness on larger scales. Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT dissipates more irregularly, given its strong coupling with cooling rate ΛΛ\Lambdaroman_Λ. While kdsubscript𝑘𝑑k_{d}italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT also decreases, αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT evolves very differently when compared with αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT. For ϵ˙i=0.1ergs1g1subscript˙italic-ϵ𝑖0.1ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=0.1\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0.1 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, there is no significant change in αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT, while for the three runs with higher ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT increases between t=0𝑡0t=0italic_t = 0 and t=10ts𝑡10subscript𝑡𝑠t=10t_{s}italic_t = 10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and decreases between t=10ts𝑡10subscript𝑡𝑠t=10t_{s}italic_t = 10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and t=100ts𝑡100subscript𝑡𝑠t=100t_{s}italic_t = 100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The drop in amplitude is more significant than that of Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT and scales with increasing k𝑘kitalic_k, where cooling becomes increasingly dominant.

Refer to caption
Figure 17: The power law indices of Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT (Top) and Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT (Bottom) over time across supersonic 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs. The linear range power laws are separated in ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, and the dissipation range power laws are coagulated and shown in black. The shaded regions represent 1σ1𝜎1\sigma1 italic_σ.

The shortened cloud survival timescales at efficient cooling regimes does not contradict existing evidence that radiative cooling supports cloud (Gronke & Oh, 2018, 2020; Li et al., 2020). Rather, it speaks on the typical sizes of such clouds. The dissipation of clouds due to shockwaves is governed by cloud-crushing timescale (Klein et al., 1994) defined as

tcc=χ1/2a0vbsubscript𝑡ccsuperscript𝜒12subscript𝑎0subscript𝑣𝑏t_{\mathrm{cc}}=\dfrac{\chi^{1/2}a_{0}}{v_{b}}italic_t start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT = divide start_ARG italic_χ start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT end_ARG (11)

where χ𝜒\chiitalic_χ is the dimensionless cloud overdensity factor, a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the cloud radius and vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT the incoming shock velocity. Clouds can only survive shocks when tcool<tccsubscript𝑡coolsubscript𝑡cct_{\mathrm{cool}}<t_{\mathrm{cc}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT < italic_t start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT (Cooper et al., 2009; Gronke & Oh, 2018, 2020), when a significant pressure gradient can be maintained between the cloud and the intracloud medium. The destruction of clouds as seen in the steepening of αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT and the difference between the t=0𝑡0t=0italic_t = 0 and t=10ts𝑡10subscript𝑡𝑠t=10t_{s}italic_t = 10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT snapshots in Figure 15 show that tcool>tccsubscript𝑡coolsubscript𝑡cct_{\mathrm{cool}}>t_{\mathrm{cc}}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT > italic_t start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT on all scales. Assuming a uniform ”intracloud medium” shock speed vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT when integrated across all clouds in all directions given our uniform randomized turbulence driving scheme (See Section 2), the typical cloud size must then have a sharp inverse dependence on χ𝜒\chiitalic_χ, where a0χβproportional-tosubscript𝑎0superscript𝜒𝛽a_{0}\propto\chi^{-\beta}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∝ italic_χ start_POSTSUPERSCRIPT - italic_β end_POSTSUPERSCRIPT for β>3/2𝛽32\beta>3/2italic_β > 3 / 2 since tcoolχ2proportional-tosubscript𝑡coolsuperscript𝜒2t_{\mathrm{cool}}\propto\chi^{-2}italic_t start_POSTSUBSCRIPT roman_cool end_POSTSUBSCRIPT ∝ italic_χ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT. We note this as a generalized empirical result given the nonideal conditions of our simulations and our broader focus on global energy and substructure dissipation, whereas detailed studies on cloud survival and crushing employ idealized wind tunnel simulations (Klein et al., 1994; Xu & Stone, 1995; Cooper et al., 2009; Gronke & Oh, 2018, 2020; Li et al., 2020; Kanjilal et al., 2021).

We compute αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT and αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT via a curve fit using the Levenberg-Marquardt least-squares algorithm (Marquardt, 1963) for 0.07<k<kd0.07𝑘subscript𝑘𝑑0.07<k<k_{d}0.07 < italic_k < italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT. To account for the shrinking of kdsubscript𝑘𝑑k_{d}italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT over time, multiple curve fits are performed for each run at each timestep for 0.2<kdπ/40.2subscript𝑘𝑑𝜋40.2<k_{d}\pi/40.2 < italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT italic_π / 4, and the fit with the smallest error is chosen. The bounds are chosen from visual inspection of the 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resolution power spectra, encapsulating the full range of kdsubscript𝑘𝑑k_{d}italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT across all runs and all times.

Refer to caption
Figure 18: The full set of dissipation curves for the Z/Z=0.3𝑍subscript𝑍direct-product0.3Z/Z_{\odot}=0.3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.3, n=0.1cm3𝑛0.1superscriptcm3n=0.1\;\mathrm{cm}^{-3}italic_n = 0.1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, and ϵ˙i=10ergs1g1subscript˙italic-ϵ𝑖10ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=10\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 10 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. Across four distinct snapshots t=0,ts,10ts,100ts𝑡0subscript𝑡𝑠10subscript𝑡𝑠100subscript𝑡𝑠t=0,t_{s},10t_{s},100t_{s}italic_t = 0 , italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , 10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT , 100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, the top row plots density slices normalized by n𝑛nitalic_n while the middle row plots velocity magnitude. The bottom plot shows five dimensionless dissipation curves over time normalized by tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT: thermal energy (ϵg/cs,02subscriptitalic-ϵ𝑔superscriptsubscript𝑐𝑠02\epsilon_{g}/c_{s,0}^{2}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), kinetic energy (ϵk/cs,02subscriptitalic-ϵ𝑘superscriptsubscript𝑐𝑠02\epsilon_{k}/c_{s,0}^{2}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / italic_c start_POSTSUBSCRIPT italic_s , 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT), clumping factor (C=ρ2/ρ21𝐶delimited-⟨⟩superscript𝜌2superscriptdelimited-⟨⟩𝜌21C=\langle\rho^{2}\rangle/\langle\rho\rangle^{2}-1italic_C = ⟨ italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ / ⟨ italic_ρ ⟩ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 1), matter power spectrum linear regime power index (αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT) and mach power spectrum linear regime power index (αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT). The former three quantities are shown on the left y𝑦yitalic_y axis, while the latter two are shown on the right y𝑦yitalic_y axis. The red, green and cyan dotted lines on the bottom plot corresponding to the boxes of the same colour around the slice plots show the times slice plots represent.

We show the power indicies for Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT and Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT over time across all our supersonic 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs in Figure 17. Both power spectra show an initial gradient in α𝛼\alphaitalic_α, with higher ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT leading to shallower α𝛼\alphaitalic_α. The power law indices for Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT in the top plot initially converge up until 10101010 Myr, beyond which they begin to sharply steepen. Meanwhile for Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT, the power law indicies also initially converge up until the same time, but remain more or less unchanged as a whole beyond that time, with ϵ˙i=100ergs1g1subscript˙italic-ϵ𝑖100ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=100\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 100 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT serving as a mild outlier. Both behaviours are consistent with Figure 16. As with C𝐶Citalic_C in Figure 12, α𝛼\alphaitalic_α shows distinct and synchronous trends with respect to absolute time t𝑡titalic_t, but unlike Figure 12, this no distinct n𝑛nitalic_n dependence in the time evolution of α𝛼\alphaitalic_α. The steepening of α𝛼\alphaitalic_α in the top plot represents the crushing of small scale clouds, and from the convergent behaviour of the curves we see that similar to tDissCsubscriptsuperscript𝑡𝐶Disst^{C}_{\mathrm{Diss}}italic_t start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT, the cloud crushing timescales are weakly coupled to if not independent from ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Similar to Figure 12, we can roughly define a universal cloud crushing timescale tccsubscript𝑡𝑐𝑐t_{cc}italic_t start_POSTSUBSCRIPT italic_c italic_c end_POSTSUBSCRIPT to be apprxoimately 10 Myr, representing the convergence point in the power law indices across all supersonic runs.

In Figure 18 we compare the time evolution of α𝛼\alphaitalic_α in relation to the other quantities examined in this study, namely thermal energy, kinetic energy, and clumping factor. The initial state is broadly weakly supersonic where ϵg/ϵk1subscriptitalic-ϵ𝑔subscriptitalic-ϵ𝑘1\epsilon_{g}/\epsilon_{k}\approx 1italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT / italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ≈ 1 with the clumps being strongly supersonic as seen from Figure 15. C𝐶Citalic_C and ϵksubscriptitalic-ϵ𝑘\epsilon_{k}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT are more or less parallel, affirming our earlier observations and discussions. The bulk of the dissipation occurs between tssubscript𝑡𝑠t_{s}italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, where a sharp drop in ϵgsubscriptitalic-ϵ𝑔\epsilon_{g}italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT corresponds to a steepening in αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT and αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT. For a brief period of time between 3ts3subscript𝑡𝑠3t_{s}3 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and 20ts20subscript𝑡𝑠20t_{s}20 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, the gas is strongly supersonic due to rapid thermal cooling, and within this period we observe C𝐶Citalic_C, αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT and αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT to roughly flatten out at various points. This may be indicative of temporary stabilization against cloud collapse, where the clouds become significantly larger and colder resulting in longer tccsubscript𝑡cct_{\mathrm{cc}}italic_t start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT and stronger pressure gradients between the cloud and intracloud medium.

4 Discussion

4.1 On the Nature of Supersonic Thermal and Kinetic Dissipation

Thermal and kinetic energy exhibit vastly different dissipative behaviours. The bulk of the thermal energy dissipates during the initial rapid dissipation epoch while kinetic energy maintains a consistent exponentially decaying rate. The physics behind the rapid thermal dissipation epoch results from enhanced cooling in dense clouds, evidenced by stronger turbulence driving leading to a clumpier medium, shallower Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT power law and increased thermal dissipation as seen in Figures 12, 16 and 5. Thermal dissipation is then strongly coupled to turbulence driving indirectly through increased clumping. Kinetic dissipation on the other hand occurs via energy cascade into thermal energy, per the following relation in the inertial regime as derived by Kolmogorov (1941):

Ek=Ckk5/3ϵ˙k2/3subscript𝐸𝑘subscript𝐶𝑘superscript𝑘53superscriptsubscript˙italic-ϵ𝑘23E_{k}=C_{k}k^{-5/3}\dot{\epsilon}_{k}^{2/3}italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = italic_C start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_k start_POSTSUPERSCRIPT - 5 / 3 end_POSTSUPERSCRIPT over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT (12)

As seen in Figure 8, our simulations follow the k5/3superscript𝑘53k^{-5/3}italic_k start_POSTSUPERSCRIPT - 5 / 3 end_POSTSUPERSCRIPT power law quite well, despite the bottleneck effect (Dobler et al., 2003). We note that since we cannot resolve the kolmogorov microscales and we did not introduce additional subgrid artificial viscosity, the cascade occurs via only shock heating and numerical dissipation, which may have represented an underestimation of up to 50% (Stone et al., 1998). In comparatively stronger supersonic regimes, the power law approaches k2superscript𝑘2k^{-2}italic_k start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT for a fully shocked gas (Burgers, 1948) and the bottleneck effect becomes less significant (Federrath et al., 2010; Federrath & Klessen, 2013). While we do achieve strongly supersonic regions within dense clumps, overall our supersonic runs are weakly supersonic where \mathcal{M}caligraphic_M and ϵk/ϵgsubscriptitalic-ϵ𝑘subscriptitalic-ϵ𝑔\epsilon_{k}/\epsilon_{g}italic_ϵ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT / italic_ϵ start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT are of order unity during steady state turbulence, due to shock heating of the gas and inefficient cooling regimes beyond the local maximum along ΛΛ\Lambdaroman_Λ beyond 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT to 105superscript10510^{5}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT K depending on Z/Z𝑍subscript𝑍direct-productZ/Z_{\odot}italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT. s This kinetic dissipation rate ϵ˙k2/3superscriptsubscript˙italic-ϵ𝑘23\dot{\epsilon}_{k}^{2/3}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT depends only on Eksubscript𝐸𝑘E_{k}italic_E start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT, which leads to an indirect coupling with ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT where strong turbulence driving leads to higher initial kinetic energy. However this correlation only results in exponential dissipation, which leads to convergence in the kinetic energy curves across all turbulence driving strengths. We observe similarities with Stone et al. (1998) and Ostriker et al. (1999) with the shapes of our kinetic dissipation curves, although in our case it takes up to 100ts100subscript𝑡𝑠100t_{s}100 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for kinetic energy to dissipate by 1 dex, while in their case 1 dex dissipation occurs within 0.1ts0.1subscript𝑡𝑠0.1t_{s}0.1 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Substructures, as the product of shock compression during turbulence driving, naturally couple with kinetic energy and dissipate similarly. Their lifetimes and dissipation rate are also coupled with thermal dissipation, since strong cooling and further underpressurization within the cloud leads to increased cloud stability against crushing (Cooper et al., 2009; Gronke & Oh, 2018, 2020).

4.2 On the Formation and Crushing of Turbulent Clouds

In circumgalatic environments stable against Jeans collapse, turbulence and radiative cooling become the primary drivers of turbulent cloud formation. Turbulence drives density fluctuations in the medium, which in turn overcool and trigger condensation (Mo & Miralda-Escude, 1996; Maller & Bullock, 2004; Armillotta et al., 2016; Chen et al., 2023), forming dense clouds. For strong turbulence driving we observe this to be hierarchical, with further condensation and denser clumps formining within exisiting overdensities. Such clouds remain stable via a strong pressure gradient between regions internal and external to the cloud, where the hot high pressure external region spatially confines the cool low pressure cloud (Li et al., 2020). Turbulence driving maintains the strong pressure gradient, where kinetic energy is continuously, uniformly and isotropically injected into our box, which cascades into thermal energy and heats the medum. Dense clouds remain cool at around 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K since they can efficiently thermally dissipate the additional energy, but the diffuse medium must climb to higher temperatures of 105106superscript105superscript10610^{5}-10^{6}10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT K, where ΛΛ\Lambdaroman_Λ reaches its local maxima depending on Z/Z𝑍subscript𝑍direct-productZ/Z_{\odot}italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT, in order to reach thermal equilibrium with turbulence driving (the run becomes subsonic if thermal equilibrium cannot be reached at those local maxima). Gronke & Oh (2018, 2020) showed that stable clouds can grow via entraining and mixing with hot gas on its boundary layers. However, those findings are based in idealized wind tunnel simulations, and we cannot confirm whether this is present in our simulations since our ”winds” are effectively isotropic.

Our clouds are eventually ”crushed” via cooling of the intracloud medium, which is consistent with Li et al. (2020). From Figure 15, the diffuse gas cools significantly after 10ts10subscript𝑡𝑠10t_{s}10 italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT for the ϵ˙i=10ergs1g1subscript˙italic-ϵ𝑖10ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=10\;\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 10 roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, representing the end of the rapid thermal dissipation epoch as seen in Figure 18. Figures 12 and 17 show trends only with respect to t𝑡titalic_t as opposed to dimensionless time such as t/ts𝑡subscript𝑡𝑠t/t_{s}italic_t / italic_t start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. The n𝑛nitalic_n dependence arises from enhanced cooling stabilizing clouds against dissipation, although it is only present for tDissCsuperscriptsubscript𝑡Diss𝐶t_{\mathrm{Diss}}^{C}italic_t start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT and absent for tccsubscript𝑡cct_{\mathrm{cc}}italic_t start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT. The independence of both timescales from ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT raises interesting questions on the interdependencies between χ𝜒\chiitalic_χ, a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT in eq, (11). From visual inspection, ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT shares an inverse relation with a0subscript𝑎0a_{0}italic_a start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, but a positive relation with vbsubscript𝑣𝑏v_{b}italic_v start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT and χ𝜒\chiitalic_χ, which may end up constraining tccsubscript𝑡cct_{\mathrm{cc}}italic_t start_POSTSUBSCRIPT roman_cc end_POSTSUBSCRIPT. A future study focused on the formation and dissipation of turbulence driven clouds, particularly in non-idealized environments such as ours, is highly warranted.

4.3 Physical Implications of the Density Homogenization and Cloud Crushing Timescales

The density homogenization timescales of 303003030030-30030 - 300 Myr and cloud crushing timescales of 10 Myr independent of turbulence driving strength may have multiple physical implications. On the brevity of such timescales, warm and cool clouds are ubiquitously observed in the CGM (Stocke et al., 2013), while the dynamical time on circumgalactic scales is several hundred Myr (Heckman et al., 1991) and freefall time up to a Gyr, several times longer than the structural dissipation timescales. While we neglected microphysics such as conduction which may play an important role in cloud stabilization ((Braginskii, 1965; Li et al., 2020), a more likely explanation would be periodic feedback-driven outflows from the galactic disc ((Rubin et al., 2014; Nielsen et al., 2015; Borthakur et al., 2013) with hot outflow gas accelerating and shocking the cool CGM gas (Thompson et al., 2016), though Muratov et al. (2015) found galactic outflows to be bursty with periods of up to hundreds of Myr in the FIRE simulations (Hopkins et al., 2014). On the other hand, feedback from galactic nuclear regions occur on much shorter timescales of 10s of Myr, particularly for barred galaxies (Krumholz & Kruijssen, 2015), from which the feedback of nuclear star clusters are capable of forming massive biconical outflows (Tenorio-Tagle & Muñoz-Tuñón, 1998; Tenorio-Tagle et al., 2005; Schneider et al., 2020). A possible alternate or compounding explanation is that haloes exhibiting such clouds are unvirialized and hence see turbulence driven by supersonic cold streams (Stern et al., 2021). On the independence of such timescales relative to ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT may also serve as a method of ascertaining when the last perturbation occured regardless of driving strength, from starburst/AGN driven outflows to accretion flows. However, the positive n𝑛nitalic_n dependance of the density homogenization timescales would require a proper spatial characterization of CGM densities, which is a challenging task given the complex multiphased nature of the CGM (Tumlinson et al., 2017).

4.4 Future Work

While the non self-gravitating and optically-thin approximations are valid for our diffuse runs, they become less accurate for our denser runs of n1cm3𝑛1superscriptcm3n\geq 1\;\mathrm{cm}^{-3}italic_n ≥ 1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT where our conditions begin to approach those of the ISM. More detailed exploration of these denser phases, with consistent thermochemistry and self-gravity, would serve as a bridge between our study and existing studies on turbulence dissipation in GMCs. Additionally, while magnetic fields are secondary to radiative cooling in stabilizing clouds against crushing, they play an important role in GMCs. We would recommend the inclusion of magnetic fields in subsequent studies of these denser phases, or even subsequent studies of the diffuse phases for the sake of completion.

The highly nonlinear thermal dissipative behaviours highlight the impact of kinematics, compression and overdense clouds on thermal dissipation. Our energy dissipation timescales and curves can be applied in larger scale cosmological simulations in lieu of conventional cooling curves, where those overdense clouds we observe in this study cannot be explicitly resolved.

5 Conclusion

This work explores the saturation and dissipation of hydrodynamic turbulence with a large array of 252 simulations, covering and extending beyond the typical range of physical parameters describing typical conditions within the CGM. The conclusions are summarized as follows.

  1. 1.

    Energy dissipation from uniformly driven turbulence can be characterized into subsonic and supersonic categories depending on the compressibility of the gas. Supersonic dissipation sees an initial rapid epoch of thermal-dominated dissipation followed by a slower epoch of energy dissipation, while subsonic dissipation only sees slow dissipation. Thermal dissipation occurs rapidly within a few kinetic saturation times before plateauing, while kinetic dissipation follows a consistent exponential curve.

  2. 2.

    Thermal dissipation occurs via enhanced cooling, with a highily nonlinear but strongly positive correlation with turbulence driving strength which creates overdense clouds via shocks. Kinetic dissipation occurs via the energy cascade, and positively but more weakly couples to turbulence driving strength which increases the initial kinetic energy.

  3. 3.

    Substructure formation is observed in supersonic runs, with clumping factors ranging from 2102102-102 - 10 depending on turbulent driving strength and the density field spanning a few orders of magnitude. Subsonic turbulence sees mostly uniform gas, with clumping factor within 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT of 1 and the density field spanning less than single order of magnitude. The density homogenization timescale tDissCsuperscriptsubscript𝑡Diss𝐶t_{\mathrm{Diss}}^{C}italic_t start_POSTSUBSCRIPT roman_Diss end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT, defined to be how long it takes for a supersonic run to become indistinguishable from a subsonic run, falls within the same order of magnitude across all runs at around 303003030030-30030 - 300 Myr depending on initial density but independent of turbulence driving strength.

  4. 4.

    Stronger turbulence driving yields denser, more concentrated and more compressible clouds, with flattened matter and \mathcal{M}caligraphic_M power spectra in the linear range. Cloud crushing timescales, defined using the power indices of the power spectra, are 10 Myr regardless of turbulence driving, and unlike density homogenization timescales, are independent from n𝑛nitalic_n.

This work was supported by the National Science Foundation of China (11991052, 12233001), the National Key R&D Program of China (2022YFF0503401), the China Manned Space Project (CMS-CSST-2021-A04, CMS-CSST-2021-A06), and the Zhejiang Laboratory (K2022PE0AB01). Renyue Cen is supported in part by the National Key Research and Development Program of China.

Appendix A Resolution Dependance

A.1 Energy Dissipation

Refer to caption
Figure 19: Thermal and kinetic dissipation curves comparisons between the fiducial high resolution 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs (solid line) and lower resolution 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT counterparts (dashed line). Z/Z=0.3𝑍subscript𝑍direct-product0.3Z/Z_{\odot}=0.3italic_Z / italic_Z start_POSTSUBSCRIPT ⊙ end_POSTSUBSCRIPT = 0.3 and n=0.1cm3𝑛0.1superscriptcm3n=0.1\;\mathrm{cm}^{-3}italic_n = 0.1 roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT for these runs.
ϵ˙i(ergs1g1)subscript˙italic-ϵ𝑖ergsuperscripts1superscriptg1\dot{\epsilon}_{i}\;(\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1})over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) 1283thsuperscript1283subscript𝑡128^{3}~{}t_{h}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT (Myr) 2563thsuperscript2563subscript𝑡256^{3}~{}t_{h}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT (Myr)
Total Energy
0.01 36.0 37.4
0.1 4.71 3.74
1 0.548 0.563
10 0.524 0.431
Thermal Energy
0.01 87.5 84.8
0.1 51.0 23.9
1 0.510 0.599
10 0.592 0.432
Kinetic Energy
0.01 8.25 7.64
0.1 2.49 3.24
1 0.623 0.502
10 0.434 0.425
Table 1: Comparisons between half energy dissipation timescales thsubscript𝑡t_{h}italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT between the fiducial high resolution 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs and their respective lower resolution 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT counterparts.

We compare the dissipation curves between our high resolution runs and their low resolution counterparts in Figure 19. The lower resolution runs show slightly different initial energies with variations on order unity between 0.80.80.80.8 and 1.51.51.51.5 times the initial thermal and kinetic energies of the high resolution runs. The resolution difference manifests itself primarily as slight differences in the turbulent steady state, with the dissipation curves themselves showing near congruent behaviours. Qualitatively there is good resolution convergence. However, quantitative differences in the initial states leads to noticeable (and in some cases significant) quantitative differences in the dissiption timescales as seen in Table 1. The largest disparity can be seen in the thermal energy dissipation timescales, where for ϵ˙i=0.1(ergs1g1)subscript˙italic-ϵ𝑖0.1ergsuperscripts1superscriptg1\dot{\epsilon}_{i}=0.1\;(\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1})over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 0.1 ( roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) thsubscript𝑡t_{h}italic_t start_POSTSUBSCRIPT italic_h end_POSTSUBSCRIPT varies by more than a factor of two. This is the result of the temperature dependence of the cooling curve, particularly at temperatures around 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT K where dΛdT𝑑Λ𝑑𝑇\dfrac{d\Lambda}{dT}divide start_ARG italic_d roman_Λ end_ARG start_ARG italic_d italic_T end_ARG is particularly steep. Disparities in the kinetic dissipation timescales are less significant, and may result from the finer resolution scales underestimating the energy cascading rate due to our lack of subgrid viscosity. Increased resolution leads both to an increase or decrease in initial energies depending on driving strength, which suggests that resolution has a highly nonlinear effect on the initial conditions.

As a comparison, Stone et al. (1998) saw variations in total (kinetic + magnetic) energy of around 6% between their 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs and their 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs, which is significantly less than our variations of up to 50% in kinetic energy. This highlights the additional resolution sensitivity that emerges from an adiabatic equation of state, where kinetic energy is dissipated via the energy cascade as opposed to artificial numerical viscosity as implemented in ZEUS (Stone & Norman, 1992b), which is the code used by Stone et al. (1998). We emphasize however that there is resolution convergence on the qualitative conclusions we make on the dissipation curves.

A.2 Power Spectra

Refer to caption
Figure 20: Comparisons between the power spectra of 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (blue) and 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (orange) resolution runs. From left to right we show the turbulence (Pksubscript𝑃𝑘P_{k}italic_P start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT), matter (Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT), and mach (Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT) power spectra. Pksubscript𝑃𝑘P_{k}italic_P start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT is multiplied by the k5/3superscript𝑘53k^{5/3}italic_k start_POSTSUPERSCRIPT 5 / 3 end_POSTSUPERSCRIPT such that kolmogorov power law (red dotted) is a horizontal line. From top to bottom we have ϵ˙i=1subscript˙italic-ϵ𝑖1\dot{\epsilon}_{i}=1over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = 1, 3 and 10 ergs1g1ergsuperscripts1superscriptg1\mathrm{erg}\;\mathrm{s}^{-1}\;\mathrm{g}^{-1}roman_erg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT.

We present the resolution comparisons of the power spectra in Figure 20. The 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs are shifted down by approximately 2 dex and have smaller kdsubscript𝑘𝑑k_{d}italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT compared to the 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs. However, we did not utilize or consider the absolute amplitudes of the power spectra in our analysis, and kdsubscript𝑘𝑑k_{d}italic_k start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT was determined dynamically through a series of curve fits for each run at each snapshot when fitting αsubscript𝛼\alpha_{\mathcal{M}}italic_α start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT and αδsubscript𝛼𝛿\alpha_{\delta}italic_α start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT. Beyond these differences, the shapes and trends are broadly consistent, with both Pδsubscript𝑃𝛿P_{\delta}italic_P start_POSTSUBSCRIPT italic_δ end_POSTSUBSCRIPT and Psubscript𝑃P_{\mathcal{M}}italic_P start_POSTSUBSCRIPT caligraphic_M end_POSTSUBSCRIPT for both resolutions seeing a flatter linear range with higher ϵ˙isubscript˙italic-ϵ𝑖\dot{\epsilon}_{i}over˙ start_ARG italic_ϵ end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT.

The bottleneck effect, which decribes the accumulation of energy at high k𝑘kitalic_k in the inertial regime of the turbulence power spectrum, sees a dependence on both the compressibility of the gas (\mathcal{M}caligraphic_M) as well as the resolution, being particularly relevant for higher resolution and strongly supersonic simulations (Dobler et al., 2003; Haugen et al., 2004; Federrath et al., 2010). We find that the bottleneck effect remains visible for both 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and 2563superscript2563256^{3}256 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT runs, though strong compressibility does dampen it for the 1283superscript1283128^{3}128 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT resolution run in the bottom left plot.

References

  • Armillotta et al. (2016) Armillotta, L., Fraternali, F., & Marinacci, F. 2016, MNRAS, 462, 4157, doi: 10.1093/mnras/stw1930
  • Balbus & Soker (1989) Balbus, S. A., & Soker, N. 1989, ApJ, 341, 611, doi: 10.1086/167521
  • Bavassano et al. (1982) Bavassano, B., Dobrowolny, M., Mariani, F., & Ness, N. F. 1982, J. Geophys. Res., 87, 3617, doi: 10.1029/JA087iA05p03617
  • Borthakur et al. (2013) Borthakur, S., Heckman, T., Strickland, D., Wild, V., & Schiminovich, D. 2013, ApJ, 768, 18, doi: 10.1088/0004-637X/768/1/18
  • Braginskii (1965) Braginskii, S. I. 1965, Reviews of Plasma Physics, 1, 205
  • Burgers (1948) Burgers, J. M. 1948, Advances in Applied Mechanics, 1, 171. https://api.semanticscholar.org/CorpusID:117400324
  • Burkhart et al. (2015) Burkhart, B., Collins, D. C., & Lazarian, A. 2015, ApJ, 808, 48, doi: 10.1088/0004-637X/808/1/48
  • Bustard & Oh (2023) Bustard, C., & Oh, S. P. 2023, ApJ, 955, 64, doi: 10.3847/1538-4357/aceef9
  • Cen (2013) Cen, R. 2013, ApJ, 770, 139, doi: 10.1088/0004-637X/770/2/139
  • Chen et al. (2023) Chen, H.-W., Qu, Z., Rauch, M., et al. 2023, ApJ, 955, L25, doi: 10.3847/2041-8213/acf85b
  • Cooper et al. (2009) Cooper, J. L., Bicknell, G. V., Sutherland, R. S., & Bland-Hawthorn, J. 2009, ApJ, 703, 330, doi: 10.1088/0004-637X/703/1/330
  • Crighton et al. (2013) Crighton, N. H. M., Hennawi, J. F., & Prochaska, J. X. 2013, ApJ, 776, L18, doi: 10.1088/2041-8205/776/2/L18
  • Crighton et al. (2015) Crighton, N. H. M., Hennawi, J. F., Simcoe, R. A., et al. 2015, MNRAS, 446, 18, doi: 10.1093/mnras/stu2088
  • Dekel et al. (2009) Dekel, A., Sari, R., & Ceverino, D. 2009, ApJ, 703, 785, doi: 10.1088/0004-637X/703/1/785
  • Dobler et al. (2003) Dobler, W., Haugen, N. E. L., Yousef, T. A., & Brandenburg, A. 2003, Phys. Rev. E, 68, 026304, doi: 10.1103/PhysRevE.68.026304
  • Donzis & Sreenivasan (2010) Donzis, D. A., & Sreenivasan, K. R. 2010, Journal of Fluid Mechanics, 657, 171, doi: 10.1017/S0022112010001400
  • Faucher-Giguère & Oh (2023) Faucher-Giguère, C.-A., & Oh, S. P. 2023, ARA&A, 61, 131, doi: 10.1146/annurev-astro-052920-125203
  • Federrath & Klessen (2013) Federrath, C., & Klessen, R. S. 2013, ApJ, 763, 51, doi: 10.1088/0004-637X/763/1/51
  • Federrath et al. (2010) Federrath, C., Roman-Duval, J., Klessen, R. S., Schmidt, W., & Mac Low, M. M. 2010, A&A, 512, A81, doi: 10.1051/0004-6361/200912437
  • Fielding et al. (2017) Fielding, D., Quataert, E., McCourt, M., & Thompson, T. A. 2017, MNRAS, 466, 3810, doi: 10.1093/mnras/stw3326
  • Gaspari et al. (2013) Gaspari, M., Ruszkowski, M., & Oh, S. P. 2013, MNRAS, 432, 3401, doi: 10.1093/mnras/stt692
  • Gnedin & Hollon (2012) Gnedin, N. Y., & Hollon, N. 2012, ApJS, 202, 13, doi: 10.1088/0067-0049/202/2/13
  • Gronke & Oh (2018) Gronke, M., & Oh, S. P. 2018, MNRAS, 480, L111, doi: 10.1093/mnrasl/sly131
  • Gronke & Oh (2020) —. 2020, MNRAS, 492, 1970, doi: 10.1093/mnras/stz3332
  • Gronke et al. (2022) Gronke, M., Oh, S. P., Ji, S., & Norman, C. 2022, MNRAS, 511, 859, doi: 10.1093/mnras/stab3351
  • Haugen et al. (2004) Haugen, N. E. L., Brandenburg, A., & Dobler, W. 2004, Phys. Rev. E, 70, 016308, doi: 10.1103/PhysRevE.70.016308
  • Heckman et al. (1991) Heckman, T. M., Lehnert, M. D., van Breugel, W., & Miley, G. K. 1991, ApJ, 370, 78, doi: 10.1086/169794
  • Hopkins et al. (2014) Hopkins, P. F., Kereš, D., Oñorbe, J., et al. 2014, MNRAS, 445, 581, doi: 10.1093/mnras/stu1738
  • Ji et al. (2019) Ji, S., Oh, S. P., & Masterson, P. 2019, MNRAS, 487, 737, doi: 10.1093/mnras/stz1248
  • Kanjilal et al. (2021) Kanjilal, V., Dutta, A., & Sharma, P. 2021, MNRAS, 501, 1143, doi: 10.1093/mnras/staa3610
  • Klein et al. (1994) Klein, R. I., McKee, C. F., & Colella, P. 1994, ApJ, 420, 213, doi: 10.1086/173554
  • Kolmogorov (1941) Kolmogorov, A. 1941, Akademiia Nauk SSSR Doklady, 30, 301
  • Krumholz & Kruijssen (2015) Krumholz, M. R., & Kruijssen, J. M. D. 2015, MNRAS, 453, 739, doi: 10.1093/mnras/stv1670
  • Larson (1981) Larson, R. B. 1981, MNRAS, 194, 809, doi: 10.1093/mnras/194.4.809
  • Li et al. (2020) Li, Z., Hopkins, P. F., Squire, J., & Hummels, C. 2020, MNRAS, 492, 1841, doi: 10.1093/mnras/stz3567
  • Mac Low (1999) Mac Low, M.-M. 1999, ApJ, 524, 169, doi: 10.1086/307784
  • Maller & Bullock (2004) Maller, A. H., & Bullock, J. S. 2004, MNRAS, 355, 694, doi: 10.1111/j.1365-2966.2004.08349.x
  • Marquardt (1963) Marquardt, D. W. 1963, Journal of the Society for Industrial and Applied Mathematics, 11, 431, doi: 10.1137/0111030
  • Mo & Miralda-Escude (1996) Mo, H. J., & Miralda-Escude, J. 1996, ApJ, 469, 589, doi: 10.1086/177808
  • Muratov et al. (2015) Muratov, A. L., Kereš, D., Faucher-Giguère, C.-A., et al. 2015, MNRAS, 454, 2691, doi: 10.1093/mnras/stv2126
  • Nelson et al. (2013) Nelson, D., Vogelsberger, M., Genel, S., et al. 2013, MNRAS, 429, 3353, doi: 10.1093/mnras/sts595
  • Nielsen et al. (2015) Nielsen, N. M., Churchill, C. W., Kacprzak, G. G., Murphy, M. T., & Evans, J. L. 2015, ApJ, 812, 83, doi: 10.1088/0004-637X/812/1/83
  • Ossenkopf & Mac Low (2002) Ossenkopf, V., & Mac Low, M. M. 2002, A&A, 390, 307, doi: 10.1051/0004-6361:20020629
  • Ostriker et al. (1999) Ostriker, E. C., Gammie, C. F., & Stone, J. M. 1999, ApJ, 513, 259, doi: 10.1086/306842
  • Ostriker et al. (2001) Ostriker, E. C., Stone, J. M., & Gammie, C. F. 2001, ApJ, 546, 980, doi: 10.1086/318290
  • Rubin et al. (2014) Rubin, K. H. R., Prochaska, J. X., Koo, D. C., et al. 2014, ApJ, 794, 156, doi: 10.1088/0004-637X/794/2/156
  • Rudie et al. (2012) Rudie, G. C., Steidel, C. C., Trainor, R. F., et al. 2012, ApJ, 750, 67, doi: 10.1088/0004-637X/750/1/67
  • Schneider et al. (2020) Schneider, E. E., Ostriker, E. C., Robertson, B. E., & Thompson, T. A. 2020, ApJ, 895, 43, doi: 10.3847/1538-4357/ab8ae8
  • Shen et al. (2013) Shen, S., Madau, P., Guedes, J., et al. 2013, ApJ, 765, 89, doi: 10.1088/0004-637X/765/2/89
  • Sparre et al. (2020) Sparre, M., Pfrommer, C., & Ehlert, K. 2020, MNRAS, 499, 4261, doi: 10.1093/mnras/staa3177
  • Stern et al. (2020) Stern, J., Fielding, D., Faucher-Giguère, C.-A., & Quataert, E. 2020, MNRAS, 492, 6042, doi: 10.1093/mnras/staa198
  • Stern et al. (2021) Stern, J., Faucher-Giguère, C.-A., Fielding, D., et al. 2021, ApJ, 911, 88, doi: 10.3847/1538-4357/abd776
  • Stocke et al. (2013) Stocke, J. T., Keeney, B. A., Danforth, C. W., et al. 2013, ApJ, 763, 148, doi: 10.1088/0004-637X/763/2/148
  • Stone & Norman (1992a) Stone, J. M., & Norman, M. L. 1992a, ApJS, 80, 753, doi: 10.1086/191680
  • Stone & Norman (1992b) —. 1992b, ApJS, 80, 791, doi: 10.1086/191681
  • Stone et al. (1998) Stone, J. M., Ostriker, E. C., & Gammie, C. F. 1998, ApJ, 508, L99, doi: 10.1086/311718
  • Tenorio-Tagle & Muñoz-Tuñón (1998) Tenorio-Tagle, G., & Muñoz-Tuñón, C. 1998, MNRAS, 293, 299, doi: 10.1046/j.1365-8711.1998.01194.x
  • Tenorio-Tagle et al. (2005) Tenorio-Tagle, G., Silich, S., Rodríguez-González, A., & Muñoz-Tuñón, C. 2005, ApJ, 628, L13, doi: 10.1086/432665
  • Thompson et al. (2016) Thompson, T. A., Quataert, E., Zhang, D., & Weinberg, D. H. 2016, MNRAS, 455, 1830, doi: 10.1093/mnras/stv2428
  • Tumlinson et al. (2017) Tumlinson, J., Peeples, M. S., & Werk, J. K. 2017, ARA&A, 55, 389, doi: 10.1146/annurev-astro-091916-055240
  • Vazza et al. (2009) Vazza, F., Brunetti, G., Kritsuk, A., et al. 2009, A&A, 504, 33, doi: 10.1051/0004-6361/200912535
  • Voit (2018) Voit, G. M. 2018, ApJ, 868, 102, doi: 10.3847/1538-4357/aae8e2
  • Voit et al. (2017) Voit, G. M., Meece, G., Li, Y., et al. 2017, ApJ, 845, 80, doi: 10.3847/1538-4357/aa7d04
  • Werk et al. (2013) Werk, J. K., Prochaska, J. X., Thom, C., et al. 2013, ApJS, 204, 17, doi: 10.1088/0067-0049/204/2/17
  • Werk et al. (2014) Werk, J. K., Prochaska, J. X., Tumlinson, J., et al. 2014, ApJ, 792, 8, doi: 10.1088/0004-637X/792/1/8
  • Werk et al. (2016) Werk, J. K., Prochaska, J. X., Cantalupo, S., et al. 2016, ApJ, 833, 54, doi: 10.3847/1538-4357/833/1/54
  • White & Rees (1978) White, S. D. M., & Rees, M. J. 1978, MNRAS, 183, 341, doi: 10.1093/mnras/183.3.341
  • Xu & Stone (1995) Xu, J., & Stone, J. M. 1995, ApJ, 454, 172, doi: 10.1086/176475
  • Yang & Ji (2023) Yang, Y., & Ji, S. 2023, MNRAS, 520, 2148, doi: 10.1093/mnras/stad264
  • Zahedy et al. (2019) Zahedy, F. S., Chen, H.-W., Johnson, S. D., et al. 2019, MNRAS, 484, 2257, doi: 10.1093/mnras/sty3482
  • Zhang et al. (2003) Zhang, Y.-T., Shi, J., Shu, C.-W., & Zhou, Y. 2003, Phys. Rev. E, 68, 046709, doi: 10.1103/PhysRevE.68.046709