Sei sulla pagina 1di 8

Draft version August 5, 2019

Typeset using LATEX twocolumn style in AASTeX62

A Markov model for non-lognormal density distributions in compressive isothermal turbulence


Philip Mocz∗1 and Blakesley Burkhart2, 3
1 Department of Astrophysical Sciences, Princeton University, 4 Ivy Lane, Princeton, NJ, 08544, USA∗
2 Centerfor Computational Astrophysics, Flatiron Institute, 162 Fifth Avenue, New York, NY 10010, USA
3 Department of Physics and Astronomy, Rutgers University, 136 Frelinghuysen Rd, Piscataway, NJ 08854, USA
arXiv:1908.00544v1 [astro-ph.GA] 1 Aug 2019

(Received xxx xx, 2019; Revised xxx xx, 2019; Accepted xxx xx, 2019)

Submitted to ApJL

ABSTRACT
Compressive isothermal turbulence is known to have a near lognormal density probability distribution
function (PDF) with a width that scales with the sonic Mach number and nature of the turbulent
driving (solenoidal vs compressive). However, the physical processes that mold the extreme high
and low density structures in a turbulent medium can be different, with the densest structures being
composed of strong shocks that evolve on shorter timescales than the low density fluid. The density
PDF in a turbulent medium exhibits intermittency (deviations from lognormal due to shocks), that
can increase with the sonic Mach number, which is often ignored in analytic models for turbulence and
star formation. We develop a simple model for intermittent turbulence by treating it as a continuous
Markov process, which explains both the density PDF and the transient time-scales of structures in
turbulence as a function of density. Our analytic model depends on only a single parameter, the
effective compressive sonic Mach number, and successfully describes the intermittency seen in both
1D and 3D simulations of supersonic and subsonic compressive isothermal turbulence. The model
quantifies the role of intermittency on the distribution of density structures in turbulent environments,
and has application to star forming molecular clouds and star formation efficiencies.

Keywords: methods: statistical — stars: formation — turbulence

1. INTRODUCTION experience collapse under self-gravity, and form stars.


The interstellar medium (ISM) is a supersonic, highly It is thus important to understand the distribution and
turbulent, magnetized medium (Larson 1981; Shu et al. timescales of high density structures from a theoretical
1987). The turbulent density distribution plays a crucial perspective, to understand the initial conditions of star
role in the process of star formation, including setting formation and predict star formation efficiencies.
the initial mass function, star formation rates and star Despite the wide usage of the lognormal model
formation efficiency. There is a well established numer- (Vazquez-Semadeni 1994; Federrath et al. 2008; Price
ical and analytic result that the density distribution of et al. 2011; Hopkins 2013), supersonic turbulence is
a supersonic, isothermal medium is approximately log- known to also exhibit intermittency (deviations from a
normal, with a variance that grows with the sonic Mach lognormal density PDF) due to the presence of strong
number M (Passot & Vázquez-Semadeni 1998; Padoan shocks (Kowal et al. 2007; Federrath 2013; Squire &
& Nordlund 2011a; Price et al. 2011; Molina et al. 2012; Hopkins 2017; Pan et al. 2019). The intermittency
Burkhart & Lazarian 2012). Past a critical density becomes stronger at higher Mach numbers (i.e., for
threshold, the high density tail of the distribution may stronger density compressions and larger degree of den-
sity contrast). The intermittent density PDF can be
described well with a simple spatial model that consid-
Corresponding author: Philip Mocz ers the density to be arranged as a collection of strong
pmocz@astro.princeton.edu shocks of width M−2 (Squire & Hopkins 2017).
∗ Einstein Fellow Recently, Robertson & Goldreich (2018) put forth the
theoretical picture (confirmed by simulations) that a su-
2 Mocz & Burkhart

personic, turbulent, isothermal medium is a network of sound-speed is c = 1. The equation of state of such
quasi-1D strong shock sheet fragments with exponential a fluid is P = ρc2 = ρ. We define the log-density as:
post-shock profiles1 . These strong shocks make up most s ≡ log(ρ/ρ). Then, the isothermal fluid equations are:
of the mass fraction (and high density tail of the density
PDF), but fill just a tiny fraction of the volume. The Dt s = −∇ · v (1)
high density shocks are transient structures, that evolve
Dt v = −∇s (2)
and dissipate on faster timescales than the cloud cross-
ing time (but see also Falceta-Gonçalves et al. 2008). where Dt = ∂t + v · ∇ is the convective derivative.
A complimentary picture to describing the density Density changes are entirely due to the divergence of
PDF as a spatial composition as was done in Robertson the velocity field (compressibility of the fluid).
& Goldreich (2018), is to consider the statistical time-
evolution of the densities of fluid elements. The volume 3. LOG-DENSITY PDF
elements that make up the density PDF, for example Our object of study is the density PDF. Consider the
may be considered a Markov process. Indeed, a lognor- volume-weighted PDF of the log-density, PV (s), which
mal distribution is suggestive of an Ornstein-Uhlenbeck is the probability of a given unit volume V having an
stochastic process, i.e., a mean-reverting random walk average density ρ = exp(s), per unit s.
(Uhlenbeck & Ornstein 1930). The density PDF in su- The PDF must obey 2 constraints, the conservation
personic turbulence has been modelled as such recently of probability,
in Konstandin et al. (2016); Scannapieco & Safarzadeh Z ∞
(2018). These models, however, did not focus on the PV (s) ds = 1, (3)
non-lognormality of the density PDF and predict near −∞
lognormal distributions.
and the conservation of mass,
Thus our focus is to construct a simple, predictive
Z ∞
Markov model to describe the intermittent distribution
of density in compressive turbulence. We do so by ap- ρPV (s) ds = ρ = 1. (4)
−∞
plying simple assumptions for the timescales of evolving
structures as a function of density. We then measure the We can also define the related mass-weighted density
timescales in 1D simulations of isothermal ‘turbulence’ PDF, as PM (s) = ρPV (s).
– such simulations are entirely shock-dominated, or ‘in- A common model for the density PDF is the lognor-
termittent’, as there is no true turbulence in 1D. Thus, mal:
(s − s0 )2
 
LN 1
the simple set up we work with here is akin to Burgers PV (s) = √ exp − (5)
turbulence. These 1D simulations allow us to fit the free 2πσ 2 2σ 2
parameters of our model and study the dependence on where s0 = −σ 2 /2 is the center of the normal distri-
the sonic Mach number. We then compare our model bution, and σ 2 is its variance. The variance has been
to 3D simulations, which exhibit an interplay between empirically determined from numerical simulations of
strong shocks and a turbulent cascade. driven isothermal turbulence to be a function of the
We lay out our notation and a brief overview of the sonic Mach number and the turbulent driving (b):
isothermal fluid equations and the lognormal model in
§ 2 and 3. We describe out analytic Markov model for in- σ 2 = log(1 + b2 M2 ) (6)
termittent turbulence in § 4, as well as fitting the model
where b ' 1 for compressive driving and b ' 1/3 for
parameters with 1D simulations. We compare our model
solenoidal (Federrath et al. 2008; Price et al. 2011).
to previous models in the literature for the PDF, and to
If the volume-averaged PDF is lognormal, the mass-
results from 3D supersonic isothermal turbulence simu-
weighted PDF is too, just with shifted mean: s0 → σ 2 /2.
lations in § 5. We summarize our main findings in § 6.
A well-known motivation for the origin of a lognor-
2. ISOTHERMAL FLUID EQUATIONS mal PDF comes from considering the statistics of the
divergences in the velocity field in the right-hand-side
The equations for an isothermal, inviscid fluid can
of Eqn. 1 responsible for evolving s and applying the
be written compactly as follows. Let ρ be the aver-
Central Limit Theorem (e.g. Vazquez-Semadeni 1994).
age density in the medium, and use units where the
If a Lagrangian fluid element of log-density s encounters
instances of diverging velocity fields in the fluid which
1 see also Mocz & Burkhart 2018 for the magnetohydrodynamic behave like independent random variables, then by the
version of the picture Central Limit Theorem, the mass-weighted distribution
Non-lognormal Markov model 3

of the log-density s is normal, and thus so is the volume- is a drift term that returns s towards the mean s∗ of the
weighted distribution. Any deviation from the assump- distribution on a timescale of τA (s), and
tion of independent random variables will lead to devi-
ations in the lognormal distribution. 2σs2
D(s) = (9)
τD (s)

is is a diffusion term, with timescale τD (s) and N (0, 1)


10-2 f=0 is a random number drawn from a normal distribution
of mean zero and variance one.
5
0.1 In the case that τA (s) = τD (s) = const, the Markov
10-5 f= 
S =16 process is the well-known Ornstein-Uhlenbeck process
S =8
S =4
(mean-reverting random walk), and results in a lognor-
10-8 S =2 mal distribution for s (Eqn. 5). Having constant drift
S=1 and diffusion timescales is the unique solution that pre-
-20 -15 -10 -5 0 5 10 dicts a lognormal distribution, although modifications
to the timescales can yield distributions that are very
close to lognormal (Scannapieco & Safarzadeh 2018).
4
0.1τeddy
Despite the turbulent density PDF being near lognor-
3 0.2τeddy mal, the types of structures at high and low densities
0.4τeddy
2 0.8τeddy are quite different. High density gas is composed of
1 strong shocks arranged in quasi-1D sheet fragments that
0 can interact. They have a large mass fraction but small
-1
volume-filling fraction. Low density regions are created
by the interaction of 3-D expansion waves, structured
-2
as large volume-filling irregular voids. The high density
-3
-4 -2 0 2 4 part of the PDF establishes itself first very quickly when
the medium is turbulently stirred, in just a fraction of
the dynamical time (Nordlund & Padoan 1999). These
shocks are known to operate on faster timescales than
Figure 1. Top panel : A plot of our model for the in-
termittent density PDF (Eqn. 13). f = 0 (dotted) is a the rest of the fluid (Scannapieco & Safarzadeh 2018;
simple Gaussian, while f > 0 introduces negative skewness Robertson & Goldreich 2018).
(shown in solid lines is f = 0.15). Bottom panel : Measure Thus we consider a drift timescale that is constant
of the density-dependent transient timescales in 1D isother- at low densities s < s∗ , but decreases at high densities
mal shock simulations used to tune our model. ∆s reverts s > s∗ :
back to the straight gray line after ∼ τeddy , meaning that the τA,0
τA (s) = (10)
structure at a given location is now uncorrelated with what 1 + H(s − s∗ )3f /2
used be there. Shown is our model fit (very thick, transpar-
where H(s) is the Heaviside step function, τA,0 is a char-
ent line) and simulation data (M = 2 (thin), 4, 8, 16 (thick))
at 4 different times. High density structures tend to revert acteristic timescale, and f ≥ 0 is a free-parameter set-
back to mean distribution (gray line) faster than structure ting how strongly the timescale decreases at high densi-
at low densities. ties. This is a phenomenological model meant to capture
the fact that strong shocks can form, expand, and revert
back to the mean on faster time-scales than the typical
dynamical time of the system, but it is rather quite gen-
4. A MARKOV MODEL FOR INTERMITTENT eral to describe a two-state behaviour which ends up
TURBULENCE yielding tractable analytic solutions. We also consider a
We model the volume-weighted density PDF as a constant dispersion term:
Markov process: 2σs2
D(s) = (11)
p τD,0
s(t + dt) = s(t) + A(s)dt + N (0, 1) D(s)dt (7)
where σs is the size of the dispersion and τD,0 is the
where associated dispersion timescale.
This Markov process has a steady-state analytic solu-
s∗ − s
A(s) = (8) tion for the density PDF, which can be obtained from
τA (s)
4 Mocz & Burkhart

solving the time-steady Fokker-Planck equation:


6
∂ 2 D(s)
 

0 = − [A(s)P (s)] + 2 P (s) (12) 4
∂s ∂s 2 2
The solution is: 0

(s − s∗ )2 (1 + f (s − s∗ )H(s − s∗ ))
 
-2
PV (s) ∝ exp − .
2S̃ -4
(13) -6
where we have defined -8
τA,0
S̃ ≡ σs2 (14) -10
τD,0 -12

The solution depends only on the drift-to-dispersion


timescale ratio, τA,0 /τD,0 , not their absolute values.
15
The normalization of Eqn. 13 is computed from the
probability constraint (Eqn. 3), and s∗ from the mass 10
constraint (Eqn. 4). The analytic expressions are a bit
cumbersome but it is straightforward to compute these 5
values numerically. Note that the lognormal distribution
(Eqn. 5) is recovered when f = 0 and τA,0 = τD,0 . 0

The model, at this stage, has two free parameters that


-5
need to be fit to data: f , a knob for the degree of non-
lognormality, and S̃ which sets the width of the distri- -10
bution. In the top panel of Fig. 1 we plot the shapes
of the predicted PDFs for f = 0.15 and compare it to -15
f = 0 (lognormal), for various widths S̃. The f > 0 pa-
rameter introduces a degree of negative skewness. The
Figure 2. Space-time diagram for 1D isothermal supersonic
deviation from lognormal grows stronger as the PDF is
M = 4 ‘turbulence’ (top: density, bottom: velocity). The
made wider. We will fit the free parameters to 1D ‘tur- system quickly establishes a network of self-similar interact-
bulence’ simulations, described in sections 4.1 and 4.1.1. ing and expanding shocks.
4.1. 1D isothermal ‘turbulence’ simulations
We carry out 1D isothermal ‘turbulence’ simulations where A is the amplitude of the kicks, which sets the
to which we will tune the Markov model. This section resulting turbulent Mach number. We simulate turbu-
is dedicated to the numerical implementation details. lence with mach numbers M = 2, 4, 8, 16.
We solve Eqns. 1 and 2 using a 2nd order finite vol- We find the driving establishes a quasi-steady state
ume (Godunov) method, following Springel (2010) on a solution (approximately constant Mach number) where
Cartesian grid. We add a random driving term to the the energy injected is dissipated by the shocks. The re-
right-hand-side of Eqn. 2 to drive the turbulence. sult is ‘burgulence’: a network of 1D shocks (saw-tooth
Shocks are driven on a scale of Lstir = 1. The box size shaped as a function of s), with a velocity power spec-
is much larger, Lbox = 1000, to capture the statistics of trum of k −2 as in Burgers’ turbulence, as opposed to
shocks. The resolution is ∆x = 0.0625 to resolve shocks. the k −5/3 Kolmogorov result from 3D incompressible
The shocks are driven by a random driving term, turbulence. Fig. 2 shows a space-time diagram of the
which has to be compressive in 1D. The details of the full solution from t = 0 to t = 100.
driving do not matter significantly in 1D for the statis-
4.1.1. Fitting the Markov model timescales from 1D
tical properties of the shocks, unlike in 3D where they
simulations
can be a mix of compressive and solenoidal. We drive
the turbulence by adding 5 × Lbox Gaussian pulses and Our Markov model is best fit with f = 0.15 and S̃ =
uniform random locations xrand,Lbox in the box, in time 0.7/M with the 1D simulations (bottom panel Fig. 1).
intervals of ∆tkick = 0.1: We have carried out a suite of 1D driven isothermal
5×L ‘turbulence’ simulations, described in § 4.1, with sonic
Xbox
vsource = A × N (0, 1) exp −xrand,Lbox / 2L2stir
 Mach numbers of M = 2, 4, 8, 16. The simulations have
shock=1 a characteristic timescale τeddy = LMc
stir
associated with
(15) the stirring length scale.
Non-lognormal Markov model 5

We use the simulations to fit τA,0 and f , and found simulated: fully compressive, and fully solenoidal. The
τA,0 = τeddy and f = 0.15. The fit was performed as fol- compressive driving exhibits a much wider and more in-
lows, using an approach similar to that in Scannapieco termittent density distribution than does the solenoidal
& Safarzadeh (2018) (that work instead treated La- driving. In the compressive case, the PDF looks sim-
grangian tracer particles rather than volume elements). ilar to a 1D simulation with the a similar Mach num-
We considered the average change ∆s of a volume ele- ber (M ∼ 16). In the solenoidal case, the PDF looks
ment over a time interval, given a final value of s2 . The like that of a 1D simulation with the Mach number re-
average change of our Markov process should evolve as duced by a factor of 4. That is, the 3D simulations can
be modeled with our Markov model if the Mach num-
∆s(s2 , ∆t) = (1 − exp [−∆t/τA (s2 )]) (s2 − s∗ ) (16) ber M is replaced with an effective compressive Mach
number describing the system, which we point out is a
The relation approaches s2 − s∗ at long times; that is,
function of the sonic Mach number and the details of
after an eddy time the fluid of a given density s2 is
the turbulence driving. In our 1D modeling, there are
expected to lose memory of its past and on average re-
only compressive (shock) motions. In 3D, even if the
vert back to the mean log-density. We measured ∆s as
large-scale driving is compressive, shock collisions will
a function of time and density in our 1D simulations of
generate vorticity. If the driving is solenoidal, the vor-
varying Mach number, and fit it with Eqn. 16, see Fig. 1
ticity in the fluid is larger (by a factor of ∼ 2; Federrath
(bottom panel).
2013), because solenoidal motions are directly injected.
Then, we fit the term controlling the diffusion, S̃, to
But in the turbulent cascade, some of those motions are
match the resulting PDFs of the simulations. We find
converted into compressive motions and shocks form.
S̃ = 0.7/M for our 1D simulations.
The amount of compressive motion will be smaller in
In 1D, our model reduces to a single parameter model
solenoidally driven turbulence, and thus the strength of
(the sonic Mach number M). We shall see in § 5.1 that
shocks will be reduced.
in 3D our model still works remarkably well, if M is
replaced with the effective compressive Mach number.
5.2. Intermittent PDF statistics
5. RESULTS Higher-order moments of the turbulent density distri-
5.1. Applying the Markov model to turbulent bution can reveal its underlying physical conditions and
simulations the nature of turbulent driving (Burkhart et al. 2009).
We calculate the resulting mean, variance, and skewness
We investigate how well our intermittent Markov
of our model as a function of 1D (compressive) Mach
model can predict the density PDF of 1D and 3D tur-
number, and compare it to the lognormal model, shown
bulent simulations. In Fig. 3, we verify that our model,
in Fig. 4. Strong shocks cause intermittency, and evolve
calibrated to the 1D simulations, predicts the result-
on faster timescales than the rest of the fluid, leading to
ing density PDF of the 3D simulations accurately. We
a negative skewness that grows with Mach number.
plot the model and simulation data for Mach numbers
The variance of the PDF as a function of Mach num-
M = 2, 4, 8, 16. As the Mach number increases, the
ber is a highly discussed quantity in the study of tur-
density distribution becomes wider, has a more nega-
bulent statistics (see, e.g., the discussion around Fig. 6
tive mean, and a more negative skewness (due to inter-
of Hopkins 2013). Across various turbulent simulations
mittency). In addition, we have carried out 1D simu-
in the literature, it has been seen that the skewness
lations with Mach numbers M = 0.5, 1 to show that
as a function of Mach number does not vary simply as
our model is also applicable to the compressive subsonic
log(1 + M2compressive ), and it can in fact be larger, and
and transonic case, where intermittency is small and
there is large variance at high Mach numbers. The vari-
our model reduces to a near lognormal. We also show
ance is a function of the degree of intermittency in the
the PDFs of very high resolution (40963 ) 3D isother-
simulation, which can be affected by the details of the
mal turbulent simulations with M ∼ 17 from Federrath
turbulent driving. The measured variance in simulations
(2013). The tails of the PDF can be tricky to measure
may also be affected due to limitations of resolution and
due poor statistics at lower resolution, which is why we
finite box size (which bias the variance to slightly smaller
rely on this high resolution run here for our compari-
values). Interestingly, our Markov model predicts that
son2 . There are two different turbulent driving modes
when the turbulence is significantly intermittent, the
variance scales more strongly than log(1 + M2compressive )
2 see Federrath (2013) for a discussion of how finite resolution at high Mach numbers. In fact, it scales linearly with
biases the PDF the Mach number (Fig. 4). This linear scaling for vari-
6 Mocz & Burkhart

1
Federrath (2013) Mach 17 compressive
Federrath (2013) Mach 17 solenoidal
0.100 1D sim
model

0.010
Mach=16
8
0.001 4
2

10-4 1
0.5

10-5

10-6
-20 -15 -10 -5 0 5 10

Figure 3. Volume-average density PDFs of our Markov model and simulations. We compare our Markov model prediction
for the PDF (gray lines) to 1D ‘turbulence’ shock simulations (purple dots) with Mach numbers M = 0.5, 1, 2, 4, 8, 16. The
distributions are close to lognormal, but intermittency (strong shocks) leads to negative skewness that increases with Mach
number. Our models also do a good job at explaining the PDFs of 3D turbulent simulations. Shown here are two M ∼ 17
high-resolution (40963 ) simulations from Federrath (2013), one driven solenoidally, the other compressively. In the compressive
case, the PDF looks similar to a 1D simulation with a similar Mach number. In the solenoidal case, the PDF looks like that of
a 1D simulation with the Mach number reduced by a factor of 4.

ance in 1D simulations is a theoretical upper limit for eled the mass-averaged PDF with timescales measured
3D simulations, where rotational motions in turbulence for Lagrangian volume elements, although the the PDFs
may reduce the variance. can easily be related to each other.
The mean of the PDF is simply set by mass conser- Federrath (2013); Squire & Hopkins (2017) provide an
vation. Our intermittent Markov model has decreased analytic model for non-lognormal PDFs, motivated by
mean log-density relative to the lognormal model. a phenomenological scaling of structure functions in in-
termittent turbulence, characterized by a parameter T ,
5.3. Comparison to other PDF models where T = 0 is lognormal, and T = ∞ is maximally in-
Scannapieco & Safarzadeh (2018) have also recently termittent. It was found that the parameter T increases
considered a Markov model for explaining the PDF in systematically with the compressive Mach number of
turbulence. There are a few important differences be- turbulent simulations. As such, the PDF is described
tween our model and theirs. We note that both models by a single parameter (the compressive Mach number),
choose time scales that are phenomenological, based on just as we have found. The model is complimentary to
simulation data. First, our model has an analytic solu- ours, where we have obtained the PDF by considering
tion, and describes the intermittency (strong deviations the timescales of transient structures, rather than their
from lognormal) of the turbulence, while Scannapieco spatial distributions.
& Safarzadeh (2018) choose τD (s) = τA (s) purposefully
to yield a near lognormal PDF, and their model has to 5.4. Applications of the PDF to star formation
be evaluated numerically through a Markov simulation. In the theoretical picture of a turbulent ISM, gas be-
Second, we modeled the volume-averaged PDF with the comes gravitationally unstable beyond a critical density.
timescales being measured from simulations in an Eule- Such a threshold has been investigated analytically and
rian frame, while Scannapieco & Safarzadeh (2018) mod- with simulations (Krumholz & McKee 2005; Hennebelle
Non-lognormal Markov model 7

In Eqn. 17 there is no fudge-factor to account for


the removal of gas by stellar winds or feedback. We
0
● will examine here how intermittency affects star forma-
-1 ● tion efficiency, which are observed to be a few percent

-2 ● -ln(1+M 2 )/2 (Krumholz et al. 2012).
mean(s)


● Our definition of the SFE in Eqn. 17 is similar to that
-3 ●
● of Burkhart & Mocz (2019), where the efficiency is esti-

-4 ● mated as the fraction of gas above the critical density,


● relative to the total gas. The primary difference between
-5 ●

● our approach and that of Burkhart & Mocz (2019) is

15 ● that the form of the density PDF is different. Burkhart


● & Mocz (2019) consider Eqn. 17 using a power law form

10 ● of the density PDF, which is the expected form consider-

var(s)

● ing gas that is influence primarily by its own self-gravity.




5 ● In our model, there is no self-gravity and we only take
● ln(1+M 2 )
● into account the intermittency of the PDF. We also do

● not include the ‘multi-free fall’ density inside the inte-
0
gral as is done in studies such as (Federrath & Klessen
1
2012), which give a very strong increase of the star for-
mation efficiency per free fall with sonic Mach number.
skew(s)

0●
0 The ‘multi-free fall’ scenario re-weights the integrand of
● ● ● ● ● ● ● ● ● ● ● ● ● ● ● Eqn. 17 by tff (ρ)/tff (ρ).
-1 Fig. 5 shows the star formation efficiency of our inter-
mittent model compared to the prediction for a lognor-
-2 mal, as a function of sonic Mach number. The star for-
2 4 6 8 10 12 14 16 mation efficiency is increased in our intermittent Markov
model for turbulence at high Mach numbers, due to the
Mach
negative skew of the PDF. If the degree of intermit-
Figure 4. Mean, variance, and skewness of the density PDF tency of star forming gas changes as a function of time,
as a function of 1D (compressive) Mach number (dotted),
it means that the average star formation efficiency can
and comparison to the lognormal model (thin). With in-
creased Mach number, the PDF becomes more negatively
vary by a factor of ∼ 2. The role of intermittency may
skewed as strong shocks make the turbulence more intermit- be incorporated into more detailed models of star for-
tent, and variance varies linearly with Mach number. mation efficiency.

& Chabrier 2011; Padoan & Nordlund 2011b; Federrath 1



& Klessen 2012; Burkhart 2018; Burkhart & Mocz 2019).
0.50
The star formation efficiency per freefall time is the frac- ●
● intermittent
tion of gas above this threshold. ● ●
● ● ● ● ● ● ● ● ● ● ●
For illustrative purposes, we take a transition density3
ϵff

lognormal
of scrit = log(M2 ). This is the post-shock density in the 0.10
isothermal limit. The post-shock density can be taken 0.05 Burkhar
t & Mocz
(2019) α=
as a critical density for collapse in virialized cores with- 2
out strong magnetic fields (Krumholz & McKee 2005;
Padoan & Nordlund 2011b; Burkhart & Mocz 2019). 0 5 10 15
The star formation efficiency (mass-weighted) per unit Mach
freefall time is estimated as:
Figure 5. Star formation efficiency (calculated as fraction
of gas above critical density scrit ) vs. sonic Mach number.
Z ∞
The efficiency is increased in our intermittent model for tur-
= PM ds (17)
scrit
bulence at high Mach number, due to the density PDF be-
ing negatively skewed. For comparison, we show efficiencies
3 Various treatments of the critical density in the literature may calculated from a lognormal model and the ‘lognormal plus
vary with order unity factors, which we ignore. (α = 2) powerlaw’ model of Burkhart & Mocz (2019).
8 Mocz & Burkhart

6. CONCLUSIONS • Star formation efficiencies calculated from density


We have developed a simple Markov model with an an- PDFs are expected to increase at high Mach num-
alytic solution to describe the non-lognormal density dis- bers when the PDF exhibits strong intermittency.
tribution (intermittent behavior) in supersonic, isother-
mal turbulence. The model was calibrated against 1D
‘turbulent’ shock simulations, and shown to fit PDFs of • Our model also applies to subsonic compressive
3D supersonic turbulence as well. Our main conclusions turbulence, where the model reduces to a near log-
are as follows: normal density distribution.
• High density structures in supersonic turbulence
are produced by strong shocks that have shorter Our model provides a simple way to quantify the role
transient timescales than low density structures. of intermittency on density structures in turbulent envi-
ronments, such as star forming molecular clouds. It can
• These strong shocks lead to deviations from Gaus-
be applied to factor in intermittency in studies where
sianity in the natural log of the density PDF, par-
the density PDF is used to predict star formation effi-
ticularly an increasing negative skewness as a func-
ciencies.
tion of turbulent Mach number.

• We have characterized these features of supersonic


turbulence with a simple Markov model.
The authors would like to thank Eve Ostriker and Jim
• Our results can be applied to both 1D (shocks Stone for valuable discussions. Support for this work
only) and 3D simulations (shocks plus turbu- (P.M.) was provided by NASA through Einstein Post-
lence) with various driving methods (compressive doctoral Fellowship grant number PF7-180164 awarded
vs solenoidal). This strongly suggests that the by the Chandra X-ray Center, which is operated by
simplified 1D picture of shocks holds in 3D (which the Smithsonian Astrophysical Observatory for NASA
is a system of quasi-1D, interacting shocked sheet under contract NAS8-03060. B.B. thanks the Simons
fragments with turbulence). Foundation for generous support.

REFERENCES
Burkhart, B. 2018, ApJ, 863, 118 Larson, R. B. 1981, MNRAS, 194, 809
Burkhart, B., Falceta-Gonçalves, D., Kowal, G., & Mocz, P., & Burkhart, B. 2018, MNRAS, 480, 3916
Lazarian, A. 2009, ApJ, 693, 250 Molina, F. Z., Glover, S. C. O., Federrath, C., & Klessen,
Burkhart, B., & Lazarian, A. 2012, ApJL, 755, L19 R. S. 2012, MNRAS, 423, 2680
Burkhart, B., & Mocz, P. 2019, ApJ, 879, 129 Nordlund, Å. K., & Padoan, P. 1999, in Interstellar
Turbulence, ed. J. Franco & A. Carraminana, 218
Falceta-Gonçalves, D., Lazarian, A., & Kowal, G. 2008,
Padoan, P., & Nordlund, Å. 2011a, ApJ, 730, 40
ApJ, 679, 537
—. 2011b, ApJL, 741, L22
Federrath, C. 2013, MNRAS, 436, 1245
Pan, L., Padoan, P., & Nordlund, Å. 2019, arXiv e-prints,
Federrath, C., & Klessen, R. S. 2012, ApJ, 761, 156 arXiv:1905.00923
Federrath, C., Klessen, R. S., & Schmidt, W. 2008, ApJ, Passot, T., & Vázquez-Semadeni, E. 1998, PhRvE, 58, 4501
688, L79 Price, D. J., Federrath, C., & Brunt, C. M. 2011, ApJL,
Hennebelle, P., & Chabrier, G. 2011, ApJL, 743, L29 727, L21
Hopkins, P. F. 2013, MNRAS, 430, 1880 Robertson, B., & Goldreich, P. 2018, ApJ, 854, 88
Konstandin, L., Schmidt, W., Girichidis, P., et al. 2016, Scannapieco, E., & Safarzadeh, M. 2018, ApJL, 865, L14
MNRAS, 460, 4483 Shu, F. H., Adams, F. C., & Lizano, S. 1987, ARA&A, 25,
Kowal, G., Lazarian, A., & Beresnyak, A. 2007, ApJ, 658, 23
Springel, V. 2010, MNRAS, 401, 791
423
Squire, J., & Hopkins, P. F. 2017, MNRAS, 471, 3753
Krumholz, M. R., Dekel, A., & McKee, C. F. 2012, ApJ,
Uhlenbeck, G. E., & Ornstein, L. S. 1930, Physical Review,
745, 69
36, 823
Krumholz, M. R., & McKee, C. F. 2005, ApJ, 630, 250 Vazquez-Semadeni, E. 1994, ApJ, 423, 681

Potrebbero piacerti anche