<<

Draft version August 20, 2021 Typeset using LATEX twocolumn style in AASTeX631

Radiative Turbulent Mixing Layers and the Survival of Magellanic Debris

Chad Bustard1 and Max Gronke2, ∗

1Kavli Institute for Theoretical Physics, University of California - Santa Barbara, Kohn Hall, Santa Barbara, CA 93107, USA 2Department of Physics & Astronomy, Johns Hopkins University, Bloomberg Center, 3400 N. Charles St., Baltimore, MD 21218, USA

Submitted to ApJ

ABSTRACT The Magellanic Stream is sculpted by its infall through the ’s circumgalactic medium, but the rates and directions of mass, momentum, and energy exchange through the Stream-halo interface are relative unknowns critical for determining the origin and fate of the Stream. Complementary to large-scale simulations of LMC-SMC interactions, we apply new insights derived from idealized, high- resolution “cloud-crushing” and radiative turbulent mixing layer simulations to the Leading Arm and Trailing Stream. Contrary to classical expectations of fast cloud breakup, we predict that the Leading Arm and much of the Trailing Stream should be surviving infall and even gaining mass due to strong radiative cooling. Provided a sufficiently supersonic tidal swing-out from the Clouds, the present-day Leading Arm could be a series of high-density clumps in the cooling tail behind the progenitor cloud. We back up our analytic framework with a suite of converged wind-tunnel simulations, finding that previous results on cloud survival and mass growth can be extended to high Mach number (M) flows with a modified drag time tdrag ∝ 1 + M and longer growth time. We also simulate the Trailing Stream; we find that the growth time is long (∼ Gyrs) compared to the infall time, and approximate Hα emission is low on average (∼few mR) but can be up to tens of mR in bright spots. Our findings also have broader extragalactic implications for e.g. galactic winds, which we discuss.

Keywords: hydrodynamics, Galaxy: halo, , turbulence

1. INTRODUCTION to drastically transform the structure of our galaxy and The Magellanic Stream is the dominant multiphase replenish our Galaxy’s fuel for (Cautun gas structure in the Milky Way halo. As the Large and et al. 2019). Small Magellanic Clouds (LMC and SMC, respectively) The Magellanic System, though, is not only a future fall through our Milky Way’s circumgalactic medium life-source for our quickly gas-consuming galaxy but also (CGM), their tidal forces strip loosely bound gas from a useful case-study on galaxy dynamics and evolution their disks, forming part of the massive Magellanic more broadly. The LMC and SMC, at only ∼ 50 and Stream (D’Onghia & Fox 2016). The Stream comprises ∼ 60 kpc away, respectively, provide a birds-eye view 9 of stellar feedback and gravitational and hydrodynamic the Trailing Stream, which contains at least 2 × 10 M arXiv:2108.08310v1 [astro-ph.GA] 18 Aug 2021 of gas (both neutral and ionized, Fox et al. 2014; Barger interactions that is unmatched by observations of other, et al. 2017) as well as the Leading Arm extending tens more distant galaxies (Mathewson et al. 1974, 1977; of kpc out in front of the Clouds and with HI mass Br¨unset al. 2005; Stanimirovi´cet al. 2008; McClure- 7 Griffiths et al. 2009; Nidever et al. 2010; Putman et al. ∼ 3 × 10 M (Br¨unset al. 2005). The Magellanic Sys- tem, if it collides with the Milky Way disk, promises 2012; Fox et al. 2014; Barger et al. 2017). Consequently, matching the Magellanic System’s present-day morphol- ogy, including the Leading Arm and Trailing Stream, Corresponding author: Chad Bustard has been a long-standing goal for simulations of galaxy [email protected] evolution (Gardiner et al. 1994; Connors et al. 2006; ∗ Hubble Fellow Besla et al. 2007, 2012; Hammer et al. 2015; Pardy et al. 2 Bustard & Gronke

2018; Tepper-Garc´ıaet al. 2019; Wang et al. 2019; Luc- and cooling the surrounding hot medium (Gronke & Oh chini et al. 2020). 2018; Ji et al. 2019; Gronke & Oh 2020a; Mandelker While tidal interactions between the LMC and SMC et al. 2020a; Fielding et al. 2020; Li et al. 2020; Sparre can account for much of the Stream’s shape and size, et al. 2020; Banda-Barrag´anet al. 2020; Kanjilal et al. enhanced further by an ionized Magellanic corona (Luc- 2021; Tan et al. 2021; Abruzzo et al. 2021; Farber & chini et al. 2020), it’s also clear from the fragmented Gronke 2021). While still a developing field, these in- structure of the Leading Arm and Trailing Stream sights have wide-ranging implications for the survival of (Br¨unset al. 2005), as well as the compressed lead- cold clouds embedded in hot galactic winds, as well as ing edge of the LMC (Salem et al. 2015), that the the filaments and high-velocity cloud (HVC) complexes Clouds and their debris are sculpted by the surrounding that comprise the Magellanic Stream. Milky Way corona. Some observations have attempted This paper is outlined as follows. In §2, we review to probe this in focused detail. Nigra et al.(2012) use some previous simulations of the interaction between pointed observations of an isolated cloud in the north- the Milky Way halo and the Magellanic System, and ern Magellanic Stream to probe the structure and kine- we outline the theoretical framework for turbulent mix- matics of the Stream-halo interface. They find that the ing. In §3, we apply this framework to the Leading cloud properties are best explained by a turbulent mix- Arm. As part of this effort, we present new simulations ing layer rather than a conduction dominated layer, and of a toy Leading Arm cylinder in a high Mach number they argue that cooling in the mixing layer should pro- flow (§4); this technical piece to extend the cloud crush- long the cloud lifetime by at least a factor of 2. Similarly, ing/survival framework beyond mildly transonic flows Putman et al.(2011) and For et al.(2013, 2014) analyze is necessary given the possibly high velocity achieved the discrete HI clouds and filaments of the Leading Arm by the Leading Arm upon swing-out from the Clouds. and Trailing Stream. In both environments, there are a In §5, we consider the Trailing Stream and estimate its large number of head-tail clouds with generally random prospects for survival and mass growth during infall. We pointing directions, suggesting a strong role played by discuss our results in §6, including a discussion of our turbulence. The number of head-tail clouds decreases work’s broader implications (§6.1) and the resolution cri- along the length of the Stream, consistent with a de- terion for mass flux convergence (§6.2). We conclude in crease in and hence a distance gradient §7. along the Stream. There is similarly strong evidence A set of simulation movies can be found at https:// for a distance variation in the Leading Arm (see e.g. bustardchad.wixsite.com/mysite/visualizations and will Figure 5 of Antwi-Danso et al. 2020 for a compilation be included in the online version of the article. of data points), with some clumps believed to be within ∼ 20 kpc of the Milky Way disk (McClure-Griffiths et al. 2. BACKGROUND 2008; Venzmer et al. 2012; For et al. 2016; Richter et al. 2.1. Previous Studies of the Magellanic - Milky Way 2018) in spite of ram pressure from the hot Milky Way Corona Interaction halo. To-date, only a few computational studies have While these observational studies clearly demonstrate probed the influence of the Milky Way halo on the Mag- the presence of turbulent mixing layers between the ellanic System, focusing on the dynamical role of ram Magellanic Stream and the surrounding hot halo, the pressure P = ρv2 due to the Clouds’ infall through rate and direction of the mass and momentum flux ram the background Milky Way halo. While strong ram pres- through these mixing layers (e.g. whether the Stream sure stripping of outer, loosely bound LMC and SMC is evaporating or growing in mass) is a major uncer- gas has been theorized by many authors as the forma- tainty and a key determinant of the origin and fate of the tion mechanism for the Trailing Stream (see D’Onghia Magellanic Stream. To that end, we aim to contribute & Fox 2016 and references therein), it alone doesn’t ex- an initial exploration here, motivated by significant the- plain the existence of the connecting oretical insights gained from high-resolution simulations the two Clouds (a sign of past tidal stripping), and the of cold gas structures embedded in hot, supersonic flows mass stripped from simulated Clouds falls far short of (e.g. Begelman & Fabian 1990; Klein et al. 1994; Scan- the observed Stream mass (Salem et al. 2015) unless the napieco & Br¨uggen 2015; Armillotta et al. 2017; Gronke Milky Way halo is significantly denser than expected out & Oh 2018; Mandelker et al. 2020a). A key finding is to the virial radius (Wang et al. 2019). that, in the presence of efficient radiative cooling, cold Ram pressure also affects the formation and sur- gas structures can survive for much longer than the clas- vival of the Leading Arm. Without including a Milky sical gas destruction time and even gain mass by mixing Way corona, simulations can more-or-less reproduce the Survival of Magellanic Debris 3 shapes and positions of the entire Magellanic System due to tidal interactions was fast enough. Thus far, stud- (Pardy et al. 2018); however, when otherwise identi- ies calibrating the swing-out systematically have not in- cal simulations include a gaseous Milky Way halo, the cluded a hot halo. Future, more comprehensive simula- Leading Arm disappears (Tepper-Garc´ıaet al. 2019; see tions are needed to explore whether such a large swing- also Wang et al. 2019; Lucchini et al. 2020; Williamson out is possible while simultaneously re-creating the mor- & Martel 2021). A plausible alternative is that the phology of the Trailing Stream, but until then, we can’t Leading Arm is the remnant of one or more fore-runner definitively rule out a Magellanic origin for the Lead- galaxies torn apart by ram pressure (Yang et al. 2014; ing Arm in favor of plausible but alternative fore-runner Hammer et al. 2015; Tepper-Garc´ıaet al. 2019). The theories. Now, we outline the framework that led us to recent discovery of dwarf galaxies belonging to a Mag- these conclusions. ellanic Group (D’Onghia & Lake 2008; Koposov et al. 2015; Drlica-Wagner et al. 2016; Pardy et al. 2020; Patel 2.2. Radiative Turbulent Mixing et al. 2020), of which the LMC and SMC are simply the The evolution of multiphase gas in galaxy halos is gov- largest, lend credence to this idea, but as pointed out erned by mass, momentum, and energy transfer at the in Tepper-Garc´ıa et al.(2019), there isn’t yet a clear interfaces between different phases. Interface properties candidate for the Leading Arm progenitor galaxy. are set by a combination of thermal conduction, mag- There are two possible, physical reasons why the netic fields, cosmic ray pressurization and heating, and present-day Leading Arm was not seeded by tidal strip- shear flows that drive the Kelvin-Helmholtz instability ping between the Magellanic Clouds: 1) The drag force and subsequent fluid mixing. While mixing layers are on the Leading Arm, during it’s swing out from the ubiquitous in astrophysics and are known to play impor- Clouds, was too strong, thereby blowing the gas back tant roles in the ISM, CGM, and IGM, most studies until into the Trailing Stream. Indeed, Tepper-Garc´ıaet al. recently have focused on the adiabatic Kelvin-Helmholtz (2019) note that, in their runs without Leading Arm for- instability, largely neglecting the important role played mation, the gas mass in the Trailing Stream increases by radiative cooling, which can be very efficient in in- relative to simulations with a clearly formed Leading termediate temperature gas, T ∼ 105 K, populating in- Arm. 2) The Leading Arm swung out in front of terface regions. Driven largely by attempts to explain the clouds but was disintegrated during infall. This the observed abundance of cold, T ∼ 104 K, gas em- scenario would seem plausible given the classical ex- bedded in hot galactic outflows, an explosion of papers pectation: a cloud subjected to strong ram pressure have been published in recent years revisiting the inter- would break apart within a few “cloud-crushing” times, play between turbulence and radiative cooling. Here, t ∼ χ1/2r /v (Klein et al. 1994) where v is the cc cl wind wind we will briefly review the most robust results from these velocity of the cloud relative to the background (in this studies before applying them to the Magellanic System. case, the infall velocity), r is the cloud radius, and χ cl Let’s envision a cold ∼ 104 K cloud with radius r is the overdensity of the cloud relative to the hot back- cl moving through a hot ∼ 106 K background at veloc- ground. For a typical size of 1 kpc and using 250 km/s, ity v . Let the overdensity of this cloud relative roughly the orbital velocity of the LMC, as the cloud wind to the background be χ, where χ ∼ 100 corresponds infall velocity, this timescale is < 50 Myrs. Unless the to pressure equilibrium between the cloud and back- Leading Arm formed very recently, as suggested by e.g. ground and will be the fiducial value throughout this estimates that match that of the SMC within paper1. A main quantity of interest is the ratio of shear- the last 200 Myrs (Richter et al. 2018), this destruction ing time (or cloud crushing time) to cooling time of the timescale is too short to allow for Leading Arm survival. intermediate temperature mixed gas at the interface be- In this paper, we will address both the drag time and tween hot and cold phases. This cooling time depends cloud destruction concerns; however, we do so using √ √ on T ≈ T T and n ≈ n n (Begelman & analytic arguments that scan a large parameter space mix cl h mix cl h Fabian 1990) where T and T are the cloud and hot and are motivated by sufficiently resolved simulations of cl h halo temperatures, and n and n are the cloud and cloud survival. In short, we find that reasonably mas- cl h hot halo number densities. If this cooling time, which sive Leading Arm complexes should quite easily survive is approximately an order of magnitude longer than the infall (and possibly grow in mass) due to radiative cool- ing at the turbulent mixing layer between the cold cloud and hot CGM. The only hurdle to the Leading Arm’s 1 This is not guaranteed to be true for multiphase structures with Magellanic origin, then, is the drag force, which can be long sound crossing times relative to other timescales, but it ap- pears to be true for at least HVCs in the direction of the Trailing overcome provided the Leading Arm’s swing-out speed Stream (Fox et al. 2005) 4 Bustard & Gronke cooling time of the cold gas, is shorter than the cloud tied to the increase in surface area resulting from cloud 1/2 crushing time (tcc = χ rcl/vwind), then the mixed gas stretching and disruption. can cool as quickly as it is produced by mixing, hence For cold streams, the mechanism for mass growth increasing the cold gas mass of the cloud. The ratio of is slightly different. For an infinitely long cylindrical cooling time to destruction time can be re-arranged to stream (as modeled in e.g. Mandelker et al. 2020a,b), give a critical cloud radius above which the destructive there is no elongation along the stream that can in- effects of turbulence are offset by cooling and subsequent crease the surface area (A = Aeff , fsa = 1). In- mass flux onto the cold cloud (Gronke & Oh 2018): stead, the growth in mass per unit length is tied to stream pulsation in the radial direction. Analogous to 5/2 Tcl,4M χ the survival criterion for clouds (Equation1) defined by rcl,crit ≈ 20pc 2 (1) nhTh/10 Λmix,−21.4 100 tcool,mix = tcc, survival of streams is determined by the The Mach number, M, is defined as the ratio of cold ratio of tcool,mix and an appropriate disruption time. cloud velocity to the hot gas sound speed, and the Detailed studies define the stream disruption time as cloud temperature and cooling function are normal- tshear = rcyl/αcylvcyl where rcyl and vcyl are the radius 4 and velocity of the infalling stream, and αcyl is a di- ized by their fiducial values: Tcl,4 = Tcl/(10 K) −21.4 −3 −1 mensionless quantity parameterizing the growth rate of and Λmix,−21.4 = Λmix/(10 erg cm s ) where Λ ∼ 10−21.4 is appropriate for solar metallicity gas at the mixing layer due to eddy mergers (see Padnos et al. 5 2018; Aung et al. 2019; Mandelker et al. 2019 for more Tmix ∼ 10 K. background). From Mandelker et al.(2020a), the stream Clouds with rcl > rcl,crit may undergo significant mor- phological changes, but the amount of cold gas should survival criterion is persist and grow. This process is, however, difficult to capture in the low-resolution CGM of galaxy evolution 5/2  χ 3/2 αcyl,0.1T M simulations. High-resolution cloud-crushing simulations r ≈ 20pc cyl,4 (3) cyl,crit 100 n /10−2Λ demonstrate that mass growth, even for clouds with cyl mix,−21.4 radii r  r , is only converged when the resolution cl cl,crit For a T ∼ 104 K, χ ∼ 100 stream in pressure equilib- is 8 cells per cloud radius. Unless refinement crite- ' rium with the hot halo, Equation3 gives an identical ria specifically target the CGM (Hummels et al. 2019; result to Equation1 when αcyl = 0.1, which is approxi- Peeples et al. 2019; van de Voort et al. 2019; Nelson mately satisfied for stream overdensities of χ ≈ 100 and et al. 2020), numerical diffusion dominates and multi- transonic or supersonic infall speeds (Mandelker et al. phase gas is over-mixed. To our knowledge, all cur- 2019, 2020a). rently published simulations of LMC-SMC interactions The growth time as defined in Equation2 is applicable lack sufficient resolution to capture cloud survival and to both clouds and streams and can be combined with converged mass growth. This is also true of isolated an equation for vmix (Equation 8 from Gronke & Oh LMC ram pressure stripping simulations (Salem et al. 2020a) to give 2015; Bustard et al. 2018, 2020). We will discuss the implications of this further in §6.2.  1/4 χtsc tcool,cold The instantaneous rate at which the cold gas mass tgrow ≈ (4) f t increases is given by the growth time (or entrainment sa sc,cold time) Note that the dependence on tcool,cold/tsc,cold is weak, m ρclrclA tgrow = ≈ (2) meaning the growth time is largely insensitive to the m˙ ρhvmixAeff value of the cooling curve at 104 K. Therefore, we expect where the numerator is the cloud mass (or for streams, (and simulations show) that these equations hold even the mass per unit length) and the denominator is the when the metallicity is very low or when photoionization change in mass per unit time, which depends on the from the UV background is included (Section 5.2 and mixing velocity, vmix, and Aeff , the total effective in- Ji et al. 2019; Mandelker et al. 2020a). In our mildly terface area where mixing occurs. From here on, we transonic simulations (§4) and consistent with previous define a fudge-factor fsa such that Aeff = fsaA to ac- literature, we find growth times as low as ∼ 20tsc in count for any changes in surface area. For a disrupt- the early phases of mass growth, while they asymptote ing cloud in the pre-entrainment phase, the surface area to values of order χtsc = 100tsc at late times when the steadily increases at a rate ∝ 1/tcc as the initial cloud cold gas is co-moving with the hot background. is redistributed into a tail of debris, increasing the over- Equations1 and4 describe the survival parameter all mixing. Therefore, the increase in cold gas mass is space and rate of growth, respectively. Of course, ram Survival of Magellanic Debris 5 pressure and interface mixing both impart a drag force, which would decrease Ncrit by a factor of ∼ 2 compared as well. The drag time for a cloud of radius rcl is to our estimate for Z = 0.1Z . tdrag ∼ χrcl/vwind. Under the assumption that the We can subsequently write the cloud crushing time, 1/2 2 cloud maintains its radius, tdrag ∼ χ tcc, which is drag time, and instantaneous growth time as approximately true for a transonic headwind; however,   supersonic flows change the cloud geometry, and in the 10rcl N19 tcc ∼ ∼ 14.4Myrs (7) strong cooling regime, momentum transfer also occurs vcl Mn−4 due to mixing. What we find in §4 is an increased 1/2 drag time for higher Mach number flows ∝ 1 + M tdrag ∼ α(1 + M)χ tcc (see also Scannapieco & Br¨uggen 2015; Gronke & Oh     N19 1 (8) 2020a) and an additional dependence on cooling effi- ∼ 28.8Myrs(α0.2) 1 + n−4 M ciency tcool,mix/tcc (see Figure9) that we subsume in a parameter α:  1/4 m χtsc tcool,s 1/2 t = ∼ tdrag ∼ α(1 + M)χ tcc (5) grow m˙ fsa tsc ! (9) N3/4 For our fiducial, strong cooling simulations, we find ∼ 739Myrs 19 1/4 α ∼ 0.2. This is consistent with Kanjilal et al. n−4Λ−23fsa (2021), which found that the simulated drag time was 1/2 where we use t ∼ Mt ∼ r /c and we define ≈ 0.4χ tcc for their M ∼ 1 simulations. This may also sc cc cl s,cl the normalized cloud column density N = N /(1019 depend on the form of the cooling curve near T ∼ Tcl 19 cl −2 −4 (Abruzzo et al. 2021), though we don’t explicitly test cm ), background halo density n−4 = nh/(10 −3 this. cm ), α0.2 = α/0.2 and for an estimate of the stream cooling time, the cooling function Λ−23 = −23 3 −1 2.3. Equations for the Survival and Growth of Cold Λ(1.5Ts)/10 erg cm s . We fiducially evaluate the Gas stream cooling time at T ∼ 1.5Ts, approximately where the cooling time is minimized. Note that the growth For applications to the Milky Way CGM, maybe the time has only a very weak dependence ∝ Λ−1/4 on this most powerful piece of this formulism is that, if we rea- choice, and therefore, Equations8 and9 are largely inde- sonably assume the fiducial parameters of Equation1 pendent of metallicity, though the onset of growth may (χ = 100, T = 104 K, T = 106 K), we can reduce cl h be affected by the low-temperature form of the cooling the equations for critical radius, growth time, and drag curve (Abruzzo et al. 2021). time to just three free parameters: the density of the hot Milky Way halo, nh, the cold gas column density, 3. THE LEADING ARM: ANALYTIC ESTIMATES N ∼ 2r n = 200r n and the Mach number, M. cl cl cl cl h Figure1 sketches the two main questions we have con- With these fiducial parameters, the critical radius cri- cerning the origin and evolution of the Leading Arm: teria from Equations1 and3 give identical results for streams and clouds and can be transformed into a criti- • Can the Leading Arm survive the (possibly quite cal column density criterion: strong) ram pressure upon swing-out from the Magellanic Clouds in the outer Milky Way halo? 18 −2 Or will the Leading Arm break apart and irre- Ncrit ∼ 1.2 × 10 M cm (Z = Z ) 18 −2 (6) versibly mix into the hot halo medium? Ncrit ∼ 4.8 × 10 M cm (Z = 0.1Z ) • Can the Leading Arm overcome drag from the Here, the difference is our estimate for Λ(Tmix), which background Milky Way halo and separate itself −21.4 3 −1 −22 we take to be 10 erg cm s for Z = Z and 10 from the LMC/SMC system to a reasonably large 3 −1 erg cm s for Z = 0.1Z . The latter value of 0.1 spatial extent to put it near it’s present-day posi- Z , which we take as our fiducial value for all analytic tion in front of the Clouds? estimates in this paper, is more appropriate for both the Leading Arm, with inferred generally in the 2 As written, there is no Mach number dependence here, but our 0.05-0.3 Z range (Fox et al. 2018; Richter et al. 2018), 3 simulations do show evidence for a ∼ M dependence of tgrow and the main SMC filament of the Trailing Stream with that warrants more detailed study (cf. Fig.6). We neglect it Z ∼ 0.1Z (Fox et al. 2013). The LMC filament is more here because, as we’ll see, tgrow is already long compared to the lookback time for plausible Leading Arm formation scenarios. metal-enriched with Z ∼ 0.5Z (Richter et al. 2013), 6 Bustard & Gronke

Can the Leading Arm overcome the drag Or is tdrag short? from ram pressure? i.e. is t long? drag Leading Arm

Leading Arm Ram Pressure

Does the Leading Arm survive or dissolve into the Milky Way halo? SMC Trailing Stream LMC i.e. is N > Ncrit

Figure 1. Sketch of our simplified theoretical experiment, in which we envision the Leading Arm swinging out in front of the LMC and SMC at a velocity exceeding the infall velocity of the Clouds. To recreate the present-day position of the Leading Arm, the Leading Arm must survive infall through the Milky Way halo (with survival criterion N > Ncrit) and maintain its position ahead of the Clouds while being pushed back by ram pressure (i.e. the drag time tdrag must be sufficiently long).

The scenario we envision is a race between the where v0 = Mcs,h is the initial swing-out velocity of the LMC/SMC complex moving at velocity vLMC ∼ 300 Leading Arm, M is the Mach number of that swing-out, 6 km/s and a Leading Arm cloud moving at velocity vLA and cs,h ∼ 111 km/s is the sound speed in the ∼ 10 K 3 relative to the Milky Way halo . In this section, we will Milky Way halo. We will assume that tdrag is constant assume that the Leading Arm is initially infalling faster during infall and equal to its value at the swing-out posi- than the Clouds at lookback time tlookback, but it’s ve- tion. As we will show in Figure2, for plausible Leading locity can decrease dramatically due to drag, thereby Arm masses, tdrag is longer than the lookback time, im- allowing the Clouds to pass the Leading Arm. Realistic plying that the Leading Arm is not entrained during scenarios, then, are those in which the Leading Arm is its infall. While not modeled here, two competing ef- out in front of the Clouds at present-day, i.e. after time fects will happen in a realistic scenario: will pull tlookback. Note that this exercise serves to illuminate the the Leading Arm towards the Milky Way, and a grad- parameter space of allowable scenarios to zeroth order ual increase in density will decrease the drag time. Since −1/3 but neglects other dynamical effects such as gravity from tdrag ∝ nh if we consider a constant mass cloud main- the Milky Way. Eventually, full hydrodynamic simula- taining pressure equilibrium with the halo, the decrease tions of the Magellanic System at sufficient resolution in drag time is expected to be small, and changes to our (§6) will be needed. conclusions are minor. The distance between the Leading Arm and LMC is To obtain the lookback time, we created a polynomial fit to the LMC distance as a function of time (as was Z tlookback done in Bustard et al. 2020) taking the LMC orbit from d = (vLA(t) − vLMC)dt (10) 0 Kallivayalil et al.(2013); Salem et al.(2015). While other orbits are possible, we choose this one because it is where we assume vLMC = 300 km/s is a constant, and motivated by recent estimates and preva- vLA(t) follows from the drag equation lently used in first-passage simulations where many of dv v2 the Magellanic System features are well-recreated (e.g. LA = − (11) dt tdragv0 Besla et al. 2012; Pardy et al. 2018; Lucchini et al. 2020). One can easily go through this same exercise for other which is derived by assuming that the ram pressure force orbits. ρ v2 decelerates the cloud. h The Leading Arm distance is then mapped to a given The solution to Equation 10 is halo density, assuming either the best-fit β-profile from   Salem et al.(2015) tlookback d = v0tdragln 1 + − vLMCtlookback (12) tdrag  2!−3β/2 3 We assume for simplicity that the Milky Way halo is non-rotating r nh(r) = n0 1 + (13) rc Survival of Magellanic Debris 7

Looback Time (Myrs) −3 800 700 600 500 400 300 200 100 with n0 = 0.46 cm , rc = 0.35 kpc, and β = 0.559, or

Salem+ 2015 the profile from Faerman et al.(2020) 4 10 Faerman+ 2020  −a ) r 3 nh(r) = n0 (14)

m rCGM c (

h −5 −3 n with n0 = 1.3 × 10 cm , rCGM = 283 kpc, and a = 0.93. 10 5 These density profiles are plotted together in Figure 220 200 180 160 140 120 100 80 Distance from Milky Way (kpc) 2, where we’ve reversed the x-axes to show progression over time going from left to right. Both profiles are sim- Lookback Time (Myrs) 800 700 600 500 400 300 200 100 ilar to other recent models (Bregman et al. 2018) but differ most notably at large radii where observational constraints are somewhat less certain.4 For the remain- ) s

r 4

y 10 der of this section, we will assume the profile from Salem M ( et al.(2015), and in the Appendix, we show our results w o r g t are not sensitive to this choice. M = 3e6 M M = 3e7 M We consider Leading Arms with a range of masses, M = 3e8 M 103 related to the column density and cloud radius by 220 200 180 160 140 120 100 80 Distance from Milky Way (kpc) 3 3 3 N19 Looback Time (Myrs) Mcl ∼ 4/3πmn¯ clrcl ∼ 4 × 10 2 M (15) 800 700 600 500 400 300 200 100 n−4 104 M = 3e6 M M = 3e7 M M = 3e8 M = 3.0 = 4.5 here assuming a spherical mass distribution for simplic- = 6.0 ity. For reference, the present-day HI gas mass of the ) s r 7 y Leading Arm is estimated to be 3×10 M (Br¨unset al. M (

g 3

a 10 2005). While there must be some additional mass in r d t ionized hydrogen, the Leading Arm has been notoriously difficult to observe in Hα (Antwi-Danso et al. 2020), and

220 200 180 160 140 120 100 80 there is no clear estimate of the ionized hydrogen mass. Distance from Milky Way (kpc) We consider plausible Leading Arm masses at swing-out Looback Time (Myrs) to be those within a factor of 10 of the present-day HI 800 700 600 500 400 300 200 100 mass from Br¨unset al.(2005). This allows for the pos- M = 3e6 M M = 3e7 M M = 3e8 M 102 sibility that the Leading Arm grew by a factor up to 10 k

c compared to its swing-out mass, or that the observed a b k o

o Leading Arm is only one piece of the otherwise hidden, l 1 t 10

/

g full structure, e.g. a set of clumps in the Leading Arm’s a r d t stretched, cooling tail.

100 The second panel of Figure2 shows the growth time 6 220 200 180 160 140 120 100 80 for Leading Arms with masses between 3 × 10 and Distance from Milky Way (kpc) 8 3 × 10 M . The growth times (Gyrs or longer) are sig- nificantly greater than the considered lookback times, Figure 2. From top to bottom row, First: Milky Way halo number density vs distance from Salem et al.(2015) and so it’s unlikely that the Leading Arm mass has grown Faerman et al.(2020). To relate the lookback time to ap- since its swing-out from the Clouds. The third and proximate distance from the Milky Way, we use the LMC fourth panels of Figure2 show the drag time and ra- orbit provided by Kallivayalil et al.(2013), more specifically tio of drag to lookback times as a function of distance a polynomial fit as a function of time as was used in Bus- in the halo. Note the very weak dependence on Mach tard et al.(2020). Second: Growth time (Equation9) for number, following from Equation8. Like the growth 6 7 Leading Arms of constant masses 3 × 10 M , 3 × 10 M times, drag times are consistently longer than lookback (present-day estimate of HI mass from Br¨unset al. 2005), 8 and 3×10 M . We’ve here used the Salem et al.(2015) halo profile, and we’ve assumed spherical symmetry to relate the 4 The Faerman et al.(2020) profile is isentropic and includes non- Leading Arm parameters to a total mass. Different linestyles thermal pressure contributions. How cosmic rays and magnetic denote varying M. Third: Drag time (Equation8 assuming fields affect the mass entrainment and deceleration of the Leading α = 0.2) Fourth: Ratio of drag time to lookback time, which Arm are fascinating topics beyond the scope of this paper. for all reasonable Leading Arm masses, is > 1, meaning the present-day Leading Arm is in the non-entrained phase. 8 Bustard & Gronke

= 3.0, vLMC = 300 km/s = 4.5, vLMC = 300 km/s 1021 100 1021 100

50 50

25 25 8 M 8 M 20 3e 05) 20 3e 05) 10 s+ 20 10 s+ 20 Brun Brun M ( 10 M ( 10 ) 7 ) 7 3e e6 M 3e e6 M 2 3 2 3

m 0 m 0 c c

( ( Ncrit(Z 0.1Z )

N Ncrit(Z 0.1Z ) 10 N 10 1019 1019

25 Ncrit(Z 1.0Z ) 25 Separation (in kpc) Separation (in kpc) after lookback time after lookback time Ncrit(Z 1.0Z ) 50 50

1018 100 1018 100 220 200 180 160 140 120 100 80 60 220 200 180 160 140 120 100 80 60 Distance from Milky Way (kpc) Distance from Milky Way (kpc)

Figure 3. Final separation between the Leading Arm and LMC/SMC (Equation 10) if the Leading Arm with column density N (y-axis) is swung-out from the Clouds at a certain distance from the Milky Way (x-axis) for two different swing-out velocities v = Mcs,hot (with M denoted in the title of the left and right panel). Mappings between distance, lookback time, and Milky Way halo density are shown in Figure2. Here, we assume the Salem et al.(2015) halo density profile, and in the Appendix, we show nearly identical plots using the profile from Faerman et al.(2020). We also overplot lines of constant Leading Arm mass, including the present-day HI estimate (Br¨unset al. 2005), assuming spherical symmetry to relate the mass to column density. White lines denote the critical column densities for the Leading Arm to survive swing-out (Equation6). Note that for 7 all reasonable Leading Arm masses ∼ 3 × 10 M , the column density is above the survival criterion. For a M = 4.5 swing-out, corresponding to an initial Leading Arm velocity of ∼ 500 km/s relative to a static Milky Way halo, a large parameter space exists where the Leading Arm could out-run the Clouds instead of being blown back into the Trailing Stream. For a gentler swing-out (e.g. M = 3), the separation is everywhere reduced. Leading Arm formation in the inner Milky Way halo especially requires larger Mach number swing-outs. times, meaning the present-day Leading Arm is in a is at least tens of kpc. More separation is better, pre-entrainment, stretching phase during which it either since Equation 10 actually gives us the distance trav- gains mass very slowly or is in a transient period of cold eled by the head of the Leading Arm, while the long mass reduction. This leads us to conclude that the Lead- cooling tail could trail behind for tens or hundreds of ing Arm must have been at least as massive at swing-out kpc. In our strong cooling simulations, we especially as it is today. find this to be true, as the distance traveled (Equation Nevertheless, while significant mass growth during in- 10) is actually quite comparable to the total tail length fall appears unlikely, plausible Leading Arms fit com- since the tail grows as fast as the cloud is destroyed: fortably within the strong cooling (survival) regime. ltail(t)/rcl,0 ∼ α(1 + M)χln(1 + t/tdrag) (cf. §4.2.1). Figure3 shows the separation d (Equation 10) as a func- Some regions of the parameter space in Figure3 tion of initial Leading Arm column density N and dis- clearly forbid a sufficient separation, resulting in just tance from the Milky Way, using Equation8 to calculate enough Leading Arm slow-down to allow the Clouds to tdrag. Overplotted are lines of constant Leading Arm “catch up”. This is especially true for a M = 3 swing- cloud mass for a given column density and halo density. out, corresponding to v0 ≈ 333 km/s, where earlier for- We also overplot lines corresponding to the critical col- mation scenarios further out in the Milky Way halo are umn density Ncrit (Equation6) for both solar metallicity preferred, as the drag force is lower in the outer halo. (Z = Z ) and Z = 0.1Z . These lines fall well below On the other hand, even drag in the inner 75-100 kpc the regime of plausible Leading Arm masses. Instead halo can be overcome if the swing-out is fast enough. A of dissolving into the Milky Way halo, the Leading Arm M = 4.5 swing-out, corresponding to v0 ≈ 500 km/s, should survive infall, with its mass redistributed into a is sufficient even if the Leading Arm was formed only clumpy, cooling tail (as in Figure4) . ≈ 200 Myrs ago. Consistent with Tepper-Garc´ıaet al.(2019), however, So can the Leading Arm originate in the Clouds? Yes, the Leading Arm may be instead blown into the Trail- as long as the swing-out is fast enough, and the initial ing Stream. We consider the Leading Arm formation Leading Arm mass is at least comparable to the present- scenario to roughly “work” when the swing-out mass is day mass (Br¨unset al. 2005). If in the strong cooling close to the Br¨unset al.(2005) estimate and when the regime, which seems applicable, the present-day Lead- final separation between Leading Arm and LMC/SMC ing Arm is likely a series of clumps in the long cooling Survival of Magellanic Debris 9 tail behind the progenitor cloud. This is consistent with from Equation4. While not of significant consequence observations, which find a clear distance variation along for the Leading Arm scenario we study (tgrow is already the Leading Arm (Venzmer et al. 2012; For et al. 2016; large compared to tlookback), this additional dependence Antwi-Danso et al. 2020). The clumpy, fragmented na- may have broader implications, especially for galactic ture of the Leading Arm, rather than a coherent tail, winds (§6.1). Those interested primarily in implications could be due to cut-offs in observational sensitivity, i.e. for the Magellanic system can skip directly to §5. that we observe only the highest column density clumps formed through thermal instability (see the distinctly 4.1. Numerical Methods clumpy morphology in Figure4) or disruptive turbu- Our simulations largely follow the setup from Gronke lence in the Milky Way halo. The latter may affect fu- & Oh(2018, 2020a) albeit with critical differences. We ture mass growth (Gronke et al. 2021), especially for used Athena 4.0 (Stone et al. 2008) to solve the cou- high-M flows which redistribute cloud mass into longer, pled equations of hydrodynamics on a 3D cartesian grid more diffuse tails that could be more easily disrupted by using the HLLC Riemann solver, second-order recon- external turbulence. struction, and the van Leer unsplit integrator (Gardiner Why does the Leading Arm disappear in LMC-SMC & Stone 2008). For radiative cooling, we employ the simulations? It’s likely that the swing-out speed is sim- Townsend(2009) “exact” cooling routine, which is effi- ply too low. This is to be expected since the same simu- cient and accurate. While the cooling curve we use as- lations run without a Milky Way corona, hence no drag sumes solar abundances, the survival and mass growth force, roughly reproduce the present-day Leading Arm picture we probe is robust to different cooling curves configuration (Pardy et al. 2018; Tepper-Garc´ıaet al. (see Gronke & Oh 2018; Ji et al. 2019), and our results 2019). Another possibility is that refinement schemes can be scaled to a given t /t – we fiducially as- in LMC-SMC simulations lead to over-mixing and over- cool,mix cc sume Λ(T ) ∼ 10−21.4 erg cm3 s−1 for Z = Z and destruction of the Leading Arm. Cooling, cloud coagula- mix Λ(T ) ∼ 10−22 erg cm3 s−1 for Z = 0.1Z . Following tion, and subsequent mass growth are most important in mix the setup in Gronke & Oh(2018, 2020a), we assume the the wake behind the Leading Arm, which may be under- background medium and cloud are at T = 4×106 K and resolved if traditional density-based refinement criteria T = 4 × 104 K, respectively. To ensure that the back- are used (see §6.2). ground medium doesn’t cool over the long timescales of A task for future simulations, we believe, is to test our simulations and to emulate the effect of heating, we whether tidal interactions can simultaneously produce a turn off cooling for temperatures above 2.4×106 K; pre- large enough swing-out speed while still reproducing the vious studies have verified that this has no appreciable morphology of the Trailing Stream. The simulations of effect on the cloud evolution (Gronke & Oh 2020a; Kan- Williamson & Martel(2021), for instance, suggest that jilal et al. 2021; Abruzzo et al. 2021). We also enforce Leading Arm features can generally be formed by tidal a temperature floor at the cloud temperature of 4 × 104 interactions even when including a gaseous Milky Way K to mimic the effects of photoionization from the ex- halo. While those simulations are not direct LMC-SMC tragalactic UV background; again, the results are not analogs, we echo their sentiment that more simulations overly sensitive to this choice. are needed before invoking alternate Leading Arm for- To mimic infall through the Milky Way halo, we place mation scenarios. ourselves in the frame of the Leading Arm and initial- ize a hot background with velocity v = vxxˆ. To reduce 4. THE LEADING ARM: SIMULATIONS computational cost, we employ a cloud-tracking routine In §3, we estimated the survival criterion, growth time, that shifts the reference frame to track the cold gas cen- and drag time for the Leading Arm assuming that the ter of mass. The left boundary (−xˆ) is an inflow or theoretical framework of cloud crushing and radiative “wind-tunnel” boundary condition that enforces a pre- turbulent mixing is applicable for higher Mach number scribed density and pressure but varies the velocity as flows, which has not been rigorously tested. Here, we the Leading Arm is accelerated by the headwind, i.e. as present a suite of idealized, hydrodynamic simulations the reference frame shifts. Along the cylinder axis (ˆy), of a 2.5-dimensional cylinder akin to an “arm” of gas in we use periodic boundary conditions, while at all other a hot, laminar flow. We demonstrate that the equations boundaries (the ±zˆ and +ˆx directions) we use outflow of §3 do apply for high Mach number flows (we test up to conditions where interior quantities are copied over to M = 6) and demonstrate reasonable convergence, which the ghost cells. is not a priori clear. We do, however, find evidence for Two very technical points are in order here. First, a Mach number dependence on tgrow, which is omitted simulations with a smooth object such as this (or a 10 Bustard & Gronke

4 6 Table 1. Simulation parameters. For all simulations, we choose Tcl = 4 × 10 K, Thot = 4 × 10 K, and χ = 100 (following −21.4 3 −1 Gronke & Oh(2020a)). All conversions to physical units assume Λ( Tmix) ∼ 10 erg cm s appropriate for gas that is not too far from solar metallicity.

−2 −2 tcool,mix rcl rcl(pc) mres(M ) Ncl(cm ) Ncrit(cm ) t tcc (Myrs) M ∆x n 3 2 1019 1019 cc −4 10 n−4 0.1 15 1.5 8, 16, 32 291 0.73, 0.36, 0.18 1.8 0.18 0.1 15 3 8, 16, 32 582 0.73, 0.36, 0.18 3.6 0.36 0.1 15 4.5 8, 16, 32 873 0.73, 0.36, 0.18 5.4 0.54 0.1 15 6.0 8, 16 1164 0.73, 0.36 7.2 0.72 1.0 1.5 1.5 16 29 0.36 0.18 0.18 2.0 0.75 1.5 16 15 0.36 0.09 0.18 0.2 7.5 4.5 16 437 0.36 2.7 0.54 0.4 3.75 4.5 16 218 0.36 1.4 0.54 1.0 1.5 4.5 16 87 0.36 0.54 0.54 2.0 0.75 4.5 16 44 0.36 0.27 0.54 10.0 0.15 4.5 16 8.7 0.36 0.054 0.54 more typically simulated sphere) aligned with the grid cloud radii, where we found through low-resolution runs show clear signs of the carbuncle instability, whereby that the M = 6 run produced a tail longer than 240rcl, strong planar shocks propagating along a grid become necessitating a somewhat longer box. unstable and exhibit numerical artefacts on the grid The suite of simulations we analyzed is shown in Table scale. To counter this, we tried the “H-correction” in 1, along with conversions to physical units for the cloud Athena, which adds numerical dissipation in the direc- radius, cloud crushing time, mass resolution, cloud col- tion perpendicular to the shock; however, we found that umn density, and critical column density for cloud sur- this correction changed the mass growth in our test vival. Unless otherwise noted, all results shown are for cases. Instead, we found that a more robust fix was to simulations with resolution of 16 cells per cloud radius. break the cylinder’s symmetry by applying lognormally- 4.2. Results distributed perturbations to the cylinder’s shape. These small perturbations, in some sense, make the simulations Figure4 shows the average projected overdensity less idealized and more realistic, while they also appear ρ/ρhot for the M = 1.5, 3, 6 simulations, each with to mitigate the carbuncle issue. tcool,mix/tcc = 0.1. Snapshots are viewed down the axis Second, we tested a number of grid dimensions and of the cylinder (y-axis) and taken at the end of each sim- found that the grid length along the cylinder axis gen- ulation, when the cold gas is approximately entrained in erally didn’t matter; the simulations were converged in the hot flow. Each simulation results in clouds that lose mass growth, drag time, etc. all the way down to a their morphology to strong Kelvin-Helmholtz instabili- grid length of just two cells. Even a purely 2D sim- ties after a few cloud crushing times. The overall cold ulation showed similar trends, albeit with slower mass gas mass grows significantly, though, as the cloud debris growth than in 3D, but was morphologically very dif- increases the surface area at which the hot gas shearing past the cold gas produces a radiative turbulent mix- ferent. We settled on a domain width of 3 × rcl in the yˆ direction. In thez ˆ direction, we found good conver- ing layer, facilitating mass transfer from the hot to cold phases. After entrainment, as shown in each panel of gence using a width of 24 × rcl (twice as wide as in Gronke & Oh 2020a), which ensured that all cloud frag- Figure4, there is a clear tail of cold gas in the leading ments stayed within the domain for the M = 1.5 and cloud’s wake. 3 simulations. For the M = 4.5 and 6 simulations, we In Figure5, we quantify the mass growth and rela- noticed a wider dispersal of cloud fragments in thez ˆ di- tive velocity ∆v13 of the cold gas (where the subscript denotes ρ > ρ /3) for each of the t /t = 0.1 rection and therefore extended the box width to 36rcl cl cool,mix cc simulations with varying Mach number. Time is nor- and 48rcl, respectively, in those runs. It was also im- 1/2 portant to use a sufficiently long box in the stream-wise malized by tdrag/α = (1 + M)χ tcc to show that the drag time clearly follows a 1 + M scaling. The veloc- xˆ direction. Since tdrag/tcc ∝ M, the box length also needs to scale ∝ M in order to enclose the full gaseous ity in each case falls to ∼ 1/2 it’s initial value, v0 after 1/2 tail that forms behind the Leading Arm. The box length approximately 0.2(1 + M)χ tcc, which motivates our for the M = 1.5, 3, 4.5, 6 runs are 60, 120, 180, and 360 choice of α = 0.2 in Equation8. We will explore the reason for this tdrag scaling in more detail in §4.2.3. To Survival of Magellanic Debris 11

Figure 4. Snapshots of average overdensity projected along the axis of the cylinder (y-axis) at the end of each simulation when the cold gas is effectively entrained in the hot flow. Snapshots are shown for M = 1.5, 3, 6 simulations with tcool,mix/tcc = 0.1. Length scales are shown to highlight that higher Mach number flows lead to longer cool tails, hence the need for longer simulation boxes. Note that our simulation box for the M = 6 run is actually longer than 240rcl; we cut it here for presentation. quantify the total cold gas mass, we use two definitions: ∝ M3. We’ll revisit the origins of this scaling in more ρ > ρcl/3 (solid lines in Figure5) T < 2Tcl (dashed detail in §4.3. lines). Both show clear increases in mass over time, with The right panel of Figure6 shows the change in sur- M/M0 ≈ 10-20 by the time the cloud is entrained in the face area vs ∆v13/cs,h, i.e. the average cold gas velocity hot flow. The cold mass measured from the temperature divided by the hot gas sound speed. At early times, criterion (T < 2Tcl) is consistently higher than that us- d(A/A0)/dt is approximately the same for each Mach ing the density criterion (ρ > ρcl/3), which reflects that number, but the curves begin to diverge around 10-20tcc the cooling tail lacks thermal pressure support. This as the simulations reach entrainment at different rates. appears to be more prominent at higher Mach number, The higher Mach number simulations reach entrainment likely due to increased support from turbulent pressure the slowest, hence the surface area continues to in- (Ji et al. 2019). crease for a longer duration. Qualitatively, the late-time To unravel these behaviors further, we quantify the cold gas morphologies are very similar, but stretched growth time tgrow = m(t)/m˙ (t) and the surface area stream-wise to lengths ∝ M when measured in units increase of the cold gas, A(t)/A0 for each of these four of the cloud radius. By entrainment (∆v13/cs,h ≈ 0), runs. We calculate the surface area using a marching A/A0 ∝ M. The picture, then, is that mixing and mass cubes algorithm and a threshold for ‘cold gas’ defined growth occur regardless of Mach number, but it takes as T < 2Tcl. The results are shown in Figure6. From longer for the destructive effects of high-M flows to be the left panel, we see that tgrow is greatest at late times offset by the increased surface area (due to increased when the cloud is more-or-less entrained, but during the tdrag ∝ 1 + M). initial cloud stretching phase, the surface area increases rapidly, leading to shorter growth times. For M = 1.5, 4.2.1. Tail Length t is as short as 20t , but there is clearly a Mach grow sc A consequence of the increased drag time is that the number dependence that we find to approximately scale duration of ‘stretching’ is longer. This implies an in- 12 Bustard & Gronke

Cold Gas Mass Evolution quickly entrained in the hot halo, the tail length is ap- proximately the full distance traveled by the cloud head, > cl/3

T < 2Tcl which could be hundreds of kpc for the Leading Arm in- 101 fall scenario. One relevant consequence of this is that, in the pre- 0 entrainment phase (t < tdrag), when the total cold gas M / mass has not yet significantly increased, the mass per M = 1.5 unit length ∼ Mcl,0/(vwindt) is decreasing. Since this = 3.0 transient phase is prolonged for higher Mach number = 4.5 100 flows, the low-density tail may be disrupted by external = 6.0 turbulence before mass growth can begin. Simulations 0 20 40 60 with a turbulent rather than a laminar flow are needed Time / tcc to assess the implications of this.

Cold Gas Velocity 4.2.2. Varying Cooling Efficiency 100 = 1.5 Finally, having analyzed a set of simulations where = 3.0 the parameters are comfortably in the survival / mass = 4.5 growth regime of tcool,mix/tcc < 1, we explore a set

d = 6.0 n i of M = 4.5 simulations with varying tcool,mix/tcc to w v

/ test whether the cloud survival criterion changes at high 3 /

1 Mach number. To run these simulations, we uniformly v scale the cloud size and box dimensions to change tcc, 1 10 while tcool,mix is the same for each simulation. Figure 8 compares snapshots at the end of the simulation for the tcool,mix/tcc = 0.1, 1.0 simulations. For the smaller 0.0 0.2 0.4 0.6 0.8 1.0 cloud (t /t ∼ 1.0) not in the strong cooling 1/2 cool,mix cc Time / (1 + ) tcc (strong survival) regime, the downstream tail is less co- herent, as the mass fluxes into and out of the cold phase Figure 5. Varying M and keeping tcool,mix/tcc = 0.1. Top: are roughly balanced. The result is that the cold gas Cold gas mass as a function of time normalized to the initial mass stays roughly constant, with only a small uptick cold mass. Both M(ρ > ρcl/3) (sold lines) and M(T < 2Tcl) (dashed lines) are shown. Bottom: Average velocity of cold at late times (see Figure9), and the cold gas is more dis- gas (with ρ > ρcl/3) normalized to the wind velocity. Time persed rather than concentrated in a monolithic stream. 1/2 is expressed in units of tdrag/α = (1 + M)χ tcc (Equation In Figure9, we show the mass (left) and velocity 8) to show the 1+M dependence. The velocity drops to half (right) evolution for tcool,mix/tcc varying from 0.1 to 1/2 it’s initial value after roughly 0.2(1 + M)χ tcc, motivating 10.0. The behavior follows the survival criterion well, our choice of α = 0.2. with the tcool,mix/tcc = 10.0 simulation showing a clear decrease in cold gas mass, while the tcool,mix/tcc < 1.0 creased tail length, and thus, increased surface area simulations all grow at a rate that monotonically de- for higher Mach number runs. One can immediately creases with increasing tcool,mix/tcc. Interestingly, even see this in Figure4, where we mark the scale on each the tcool,mix/tcc = 1.0 simulation shows a slight up-tick panel. For these simulations in the strong cooling regime in cold gas mass at late times when the shear veloc- (tcool,mix/tcc = 0.1), cooling happens so rapidly that the ity has fallen below ∼ 1/2 its initial value. In this case, tail forms immediately upon cloud breakup. We show the dispersed chunks of cold gas coagulate back together a rough quantification of this in Figure7, which plots and, by the time they are entrained, have even increased the amount the simulation frame has shifted over time in total mass compared to the initial cloud mass. This, (solid lines), as well as the largest distance between cold as well as the finite growth times seen even at late times (ρ > ρ/3) gas cells (dot-dashed lines). The former tells in Figure6, would appear to be consistent with mass us how far the front of the cloud has traveled, which entrainment by mixing and cooling due to continuous we find to fit the drag law distance very well. The lat- ‘pulsations’ (Gronke & Oh 2020a), which occurs at a ter gives a rough estimate of the tail length, which is mixing rate independent of Mach number. only marginally less than the distance traveled by the cloud head (Equation 10). Since from the initial shock 4.2.3. Drag Time Dependence on Cooling and Mach onwards, cloudlets peel off the cold gas cloud and are Number Survival of Magellanic Debris 13

101 Surface Area = 1.5 = 4.5 = 3.0 = 6.0 = 1.5 = 3 30 = 4.5

c = 6.0 s t 0 / 100 20 A w / o r A g t 10

> cl/3

T < 2Tcl 10 1 0 0.0 0.2 0.4 0.6 0.8 1.0 0 1 2 3 4 5 6 1/2 Time / (1 + ) tcc v13/cs, h

Figure 6. Varying M and keeping tcool,mix/tcc = 0.1. Left: Cold gas growth time normalized to the sound crossing time, tsc, defining cold gas as ρ > ρcl/3 (solid lines) or T < 2Tcl (dashed lines). The expectation from Equation4 is that tgrow ≈ 100tsc, which holds approximately true in the late-time entrained phase, but the growth time is much shorter in the initial expansion phase when the cloud surface area grows rapidly. Right: The change in cold gas surface area as a function of cold gas velocity normalized by the initial wind speed. This shows that higher Mach number flows lead to proportionally greater increases in surface area since the cloud is stretched length-wise for a longer period of time (tdrag ∝ M). By entrainment (when ∆v13 ∼ 0), A/A0 ∝ M, as the M = 3 and 4.5 runs with 16 cells per cloud radius have surface areas ≈ 2 and 3 times greater than the M = 1.5 run, although we note the absolute values of the surface area are not converged even at 32 cells per cloud radius.

Frame shift (solid); xmax - xmin (dot-dashed) advected tracer fluid that is initially set to 1 inside the 350 = 1.5 cloud and 0 elsewhere. The average cloud velocity is 300 = 3.0 then vcl = hcvxi/hci, where c is the tracer concentration = 4.5 and cvx is the concentration-weighted velocity along the l

c 250 r = 6.0 headwind direction. With the slight exception of the

/ 200 M = 1.5 run, the drag times for adiabatic simulations e

c are independent of M, consistent with other adiabatic n 150 a cloud-crushing simulations (e.g. Goldsmith & Pittard t s i 100 2017). D The cooling simulations, however, have drag times 50 ∝ 1 + M, and interestingly, the drag times are longer 0 than those for the adiabatic runs. This initially seems 0 20 40 60 80 to be in conflict with Figure9, which shows stronger Time / tcc cooling generally leads to faster entrainment. However, cooling plays a dual role: compared to an adiabatic sim- Figure 7. Solid lines: The simulation frame shift, repre- ulation, it increases the drag time by inhibiting cloud senting the distance traveled by the front of the cold cloud. dispersal (Scannapieco & Br¨uggen 2015), but it also de- Dot-dashed lines: The distance between the front and back creases the drag time by facilitating momentum trans- of the cold tail, which is only marginally smaller than the full fer. Which effect dominates depends on the strength of distance traveled by the cloud head in these strongly cooling cooling. Figure9 suggests an inflection point between simulations (tcool,mix/tcc = 0.1). tcool,mix/tcc = 2 and 10 where the increased drag time due to compression becomes offset by the decreased drag The right panel of Figure9 shows that the drag time time due to momentum transfer in the growing cloud depends on the cooling efficiency tcool,mix/tcc. To un- 1/4 tails. Naively, we expect tdrag ∝ tgrow ∝ (tcool,cl/tsc) derstand this better, we also ran a set of simulations (see Equation9). Comparing our simulations in Figure with the same parameters as our tcool,mix/tcc = 0.1 sim- 9 with tcool,mix/tcc = 0.1 and 1.0, where the only dif- ulations but without cooling. The velocity profiles are ference is an increased cloud radius (and sound crossing shown in Figure 10. Since the adiabatic (tcool,mix/tcc = time) by a factor of 10, tdrag should vary by a factor of ∞) runs destroy the cloud and mix all cold gas into the 101/4 = 1.77, which appears consistent with Figure9 if hot background within a few tcc, we no longer quan- tify the cold gas velocity but instead use a passively 14 Bustard & Gronke

Figure 8. Snapshots of average overdensity projected along the axis of the cylinder (y-axis) at the end of each simulation when the cold gas is effectively entrained in the hot flow (∆v13/vwind ≈ 0). Snapshots are shown for M = 4.5 simulations with tcool,mix/tcc = 0.1 and tcool,mix/tcc = 1. Simulations in the weak or marginal growth regime (tcool,mix/tcc ∼ 1) show more cloud dispersal, a delay in cloud coagulation and mass growth, and therefore a less coherent tail by the time the cold gas is entrained.

Cold Gas Mass Evolution: = 4.5 Cold Gas Velocity: = 4.5 100 101

0 10 d n i w 0 v M / /

1 3 10 tcool, mix/tcc = 0.1 / tcool, mix/tcc = 0.1 M 1 1 tcool, mix/tcc = 0.2 v 10 tcool, mix/tcc = 0.2

tcool, mix/tcc = 0.4 tcool, mix/tcc = 0.4 2 10 tcool, mix/tcc = 1.0 tcool, mix/tcc = 1.0 > cl/3 tcool, mix/tcc = 2.0 tcool, mix/tcc = 2.0

tcool, mix/tcc = 10.0 T < 2Tcl tcool, mix/tcc = 10.0 10 3 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 1/2 1/2 Time / (1 + ) tcc Time / (1 + ) tcc

Figure 9. Left: Cold gas mass as a function of time normalized to the initial cold mass. Right: Average velocity of cold gas (with ρ > ρcl/3) normalized to the wind velocity. In the left panel, both M(ρ > ρcl/3) (sold lines) and M(T < 2Tcl) (dashed lines) are shown. In all cases, M = 4.5, but the cloud size, hence tcc, is varied. tcool,mix/tcc / 1.0 appears to be a suitable cut-off for mass growth, as the tcool,mix/tcc = 10 simulation shows a precipitous drop in cold gas mass, while the tcool,mix/tcc = 1 1/2 simulation is marginal. The right panel shows that the velocity drops to ∼ 1/2 its initial value after a time ∼ 0.2(1 + M)χ tcc but monotonically trends towards higher drag times as tcool,mix/tcc increases from 0.1 to 2.0. Clearly the drag time is correlated with cooling strength, as cooling leads to momentum transfer beyond that due to ram pressure. we eye-ball tdrag as the time for the velocity to drop to the cloud primarily orthogonal to the headwind, mak- 1/2 its initial value. ing the cloud’s total column in the headwind direction Now, what is the reason for the ∝ 1 + M scaling? a factor of 1 + M higher. Scannapieco & Br¨uggen(2015) attribute this to the A related possibility is that the shock is isothermal, oblique bow shock that compresses the cloud and raises thereby inducing a density increase ∝ M2 instead of its density by a factor of 1+M. Since the cloud main- ∝ M. This would happen if tcool,cl  tshock−cross ∼ tcc 4 tains a constant temperature of order 10 K, this den- (which is the case for all surviving clouds as tcool,mix > sity increase induces a pressure gradient between the tcool,cl). If accompanied by a uniform (mass conserving) head and tail of the cloud that stretches the cloud in compression rcl → rcl/M of our cylinder, rather than the headwind direction. The net effect is to compress Survival of Magellanic Debris 15

100 +M scaling from this mixing picture? This could follow t cool, mix = if t ∝ t ∝ M. At least in the early evolution of tcc drag mix tcool, mix our clouds, it does appear that t ∼ 1/v ∝ M, 1 = 0.1 mix mix 6 × 10 tcc as seen in the bottom panel of Figure 10, which shows

d vmix =m ˙ /ρhA for each of our tcool,mix/tcc = 0.1 runs

n 1 i 4 × 10 0 w with 16 cells per cloud radius. Note that this scaling v

/ 1 l 3 × 10 also appears in adiabatic wind simulations, but tmix has c

v the opposite sign, i.e. the cloud loses mass. For exam- 2 × 10 1 = 1.5 ple, in Goldsmith & Pittard(2017), tmix ∝ M for their = 3.0 wind-tunnel simulations with χ = 10 (see their Figure = 4.5 5), although it does level off at very high Mach num- = 6.0 bers and appears constant at all Mach numbers if the 10 1 0 5 10 15 20 25 30 overdensity is χ = 1000 instead (Goldsmith & Pittard Time / tcc 2018). Note that albeit detailed simulations of turbulent mixing layers in a plane-parallel setup find a somewhat −3/4×4/5 1.0 different Mach scaling of tmix ∝ M , they com- 4 /

1 ment that this is dependent on the geometry employed )

l (Ji et al. 2019; Fielding et al. 2020; Tan et al. 2021).

c 0.8 ,

c Our analysis here, then, appears to explain the seem- s t /

l ingly disparate dependencies of tdrag on M found by c 0.6 , l Goldsmith & Pittard 2017 (no dependence) and Scan- o o

c napieco & Br¨uggen 2015( tdrag ∝ 1 + M). The key dif- t (

0.4

l ference is cooling, not necessarily because of its ability c

, = 1.5 s to increase cloud survival and reshape the cloud but also c

( = 3

/ 0.2 because of momentum transfer via mixing, which is more x i = 4.5 efficient for stronger cooling and seemingly dominant m = 6.0 v when t /t ∼ 0.1. While the effects of cloud com- 0.0 cool,mix cc 0 20 40 60 pression increase the drag time, leading to e.g. slower Time / tcc acceleration of cold clouds in galactic winds (Zhang et al. 2017) or longer deceleration of infalling clouds, addi- Figure 10. Top: Average gas velocity, weighted by a cloud tional momentum transfer in the strong cooling regime tracer, for our strongly cooling simulations tcool,mix/tcc = 0.1 somewhat offsets this, shortening the drag time by a and identical simulations run without cooling tcool,mix/tcc = 1/4 factor we expect is ∝ (tcool,cl/tcc) . ∞. The adiabatic simulations show similar drag times independent of Mach number. Bottom: Mixing velocity vmix =m/Aρ ˙ h normalized to its expected scaling (Equa- 4.3. Growth Time Dependence on Cooling and Mach tion 8 from Gronke & Oh 2020a). vmix shows a clear inverse Number Mach number dependence, only asymptoting to similar val- ues after entrainment, which takes longer for higher Mach Following similar arguments as above, we can now rea- number flows. son why we find a Mach number dependence for the growth time. From Figure 10, we see that vmix asymp- totes to similar values regardless of M, but the paths to a predominantly transverse compression, then tdrag ∼ 2 get there are very different. vmix is initially quite low χM r/Mvwind ∝ M. While we do observe a transient, elevated cloud den- for high M, likely due to two effects: 1) high-M flows sity, the dominant and most persistent geometric change induce a strong shock that crushes the cloud more dra- in our simulations is the eventual formation of a cool matically than low-M flows. Indeed, we observe that our Leading Arm is fragmented into more chunks that tail. Combined with the apparent scaling of tdrag with cooling efficiency, it appears that momentum transfer disperse to larger distances away from the initial cloud; via cooling and mixing is at least as important as mo- this led us to increase our simulation box width for the mentum transfer via ram pressure.5 Can we recover a 1 M = 4.5 and 6 runs. That is, the destructive effects of the wind that cause energy loss are strong. This is seemingly consistent with the findings of Gronke & 5 −1/2 This makes sense as tdrag/tgrow ∼ vmix/vwind ∼ χ /M Oh(2020b), where clouds thrown violently out of pres- where we used conservatively vmix ∼ cs,cold. sure equilibrium shatter into many clumps. 2) Mixing 16 Bustard & Gronke is hindered at high-M since the Kelvin-Helmholtz in- identical for our assumed parameters to Equation1 de- stability is suppressed for high-M (vmix ∼ 1/M; Gold- rived for clouds. smith & Pittard 2017). That is, the effect that causes The Trailing Stream’s infall velocity can’t exceed the mass gain is weaker. Only when the clouds have be- infall velocity of the LMC/SMC, which is M ∼ 2 − 3 come partially entrained does the mixing velocity rise depending on lookback time (Besla et al. 2012). We will to considerable values. The net effect is to create longer assume the Stream infall is M ∼ 2 through a halo with 6 growth times for higher Mach number flows, with a de- constant temperature Thot ∼ 10 K. We also assume the 3 pendence that we find to scale as tgrow ∝ M (Figure Stream is in pressure equilibrium with its surroundings 6), but we defer a more detailed analysis of this exact and therefore has an overdensity χ ∼ 100 and temper- 4 scaling to future study. While this scaling is mainly in- ature Tcyl ∼ 10 K. For such a setup, with metallicity consequential for the Leading Arm, which we estimate is ∼ (0.1−0.3)Z , all Stream sections with column density 19 −2 well before the partial entrainment stage and, therefore, N ' 10 cm should maintain or grow in mass, rather not considerably growing in mass in any case, this scal- than dissolve (Equation6). This criterion is satisfied for ing could have interesting consequences for entrainment a large portion of the Stream (Br¨unset al. 2005; Nidever and growth of cold clouds in highly supersonic galactic et al. 2010). winds. We speculate on these consequences in §6.1. Following Mandelker et al.(2020b), one can also show that if the Stream satisfies this criterion at some point, 5. THE TRAILING MAGELLANIC STREAM it will stay above this survival threshold for its entire In this section, we apply the theoretical framework of infall through denser parts of the halo. Take a Stream 2 radiative turbulent mixing layers to the Trailing Stream. with radius rcyl(r) and line-mass m(r) = πrcyl(r)ρ(r). We begin with quick analytic estimates (§5.1) and then Assuming both the halo and Stream are isothermal and present a set of simulations where we corroborate our that the Stream maintains pressure equilibrium with its predictions for the Stream survival criterion and mass surroundings, the overdensity χ is constant so ρ(r) = influx rate (§5.2). In §5.2.1, we briefly show predic- χρh(r). One can quickly show that the column density 1/2 1/2 tions for Hα emission, comparing to recent observations N(r) = 2rcyl(r)ncyl(r) ∝ m (r)ρh (r). If the initial (Barger et al. 2017) and previous simulations. Stream is within the growth regime, and the halo density increases during infall, then both terms increase with 5.1. Analytic Estimates decreasing r; therefore, N(r) increases while the Stream becomes denser and narrower. The main difference between this analysis and that of This picture of Stream survival is quite different from §3 is the morphology of the Trailing Stream vs the Lead- previously published results. Primarily to study the ori- ing Arm. Based on simulations of LMC-SMC tidal in- gin of Hα emission along the Stream, Bland-Hawthorn teractions (Besla et al. 2012; Pardy et al. 2018; Lucchini et al.(2007) and Tepper-Garc´ıaet al.(2015) simulate a et al. 2020), we assume the Stream is formed pre-infall supersonic headwind slamming into a turbulent Stream. into the Milky Way halo, and we model the geometry This interaction sends a cascade of shocks from the front as a cylinder with major axis parallel to the infall di- to the back of the Stream that generates Hα bright rection. This geometry is unstable to a number of both spots. This simulated Stream is initialized as under- surface and body modes of the Kelvin-Helmholtz insta- pressurized compared to the background by a factor of bility; combined with thermal instability, these modes 10, however, which results in a quick destruction of the destroy the axisymmetry of the Stream and naturally Stream within 100-200 Myrs. Instead, if the Stream is form the turbulent, clumpy Stream we observe (Put- closer to pressure equilibrium, we show that cooling is man et al. 1998; Br¨unset al. 2005; Stanimirovi´cet al. strong enough – or, in other words, the column density 2008; Nidever et al. 2010). of the Magellanic Stream is sufficiently large – to allow This initial geometry is fundamentally different from the survival of the stream. that of the Leading Arm, where the cylinder axis was The Stream mass, however, is not likely to grow a perpendicular to the headwind. In that case, mass considerable amount during infall. From Equation4, growth ensues after the initial cloud is stretched par- the growth time is of order a Gyr or greater, even at allel to the headwind, thereby increasing the surface the present-day position of the Clouds at r ∼ 50 − 60 area available for mixing. In the Stream case, no such kpc with n ∼ 10−4. Unless the LMC and SMC orbits stretching occurs, and any increase in surface area, then, h are such that they have spent more time in the inner is tied to radial growth of the cylinder. Equation4 still halo (and so has the Stream), t  t , and applies, but with A ≈ A , and the survival criterion grow lookback eff the factor of a few gap in simulated (e.g. Pardy et al. is described by Equation3, which as discussed in §2 is Survival of Magellanic Debris 17

2018) vs observationally inferred (D’Onghia & Fox 2016) Table 2. Stream simulation parameters. For all simulations, 4 6 Stream mass cannot be closed by mass influx through a we choose M ∼ 2, Tcyl = 10 K, Thot = 10 K, and χ = radiative turbulent mixing layer. Note that the stream 100 such that the Stream and halo are initially in pressure equilibrium. might fragment and the individual pieces separate far enough to increase Aeff (cf. §2), and thus, increase the rcyl nh tsc N cold gas growth rate. However, we expect this to be an (kpc) (cm−3) (Myrs) (cm−2) order unity effect for the Stream. 0.5 10−5 44.6 3.1 × 1018 2.0 10−5 172 1.2 × 1019 5.2. Simulations 0.5 10−4 44.6 3.1 × 1019 −5 19 In this section, we present a set of simulations whereby 5.0 10 446 3.1 × 10 −5 19 a hot (T = 106 K) background shears past a cold (T 10.0 10 892 6.2 × 10 −4 20 = 104 K), cylindrical Stream. We use the FLASH 2.0 10 172 1.2 × 10 −4 20 MHD code (Fryxell et al. 2000), which uses a direction- 5.0 10 446 3.1 × 10 ally unsplit staggered mesh solver (Lee & Deane 2009; Lee 2013) based on a finite-volume, high-order Godunov scaled down by this metallicity. Note once again that scheme. the growth time and drag time (Equations8 and9) are Our setup is, in many ways, identical to that of (Man- insensitive to metallicity, while small changes between delker et al. 2020a) but with a few differences. We setup 0.1 and 0.5 Z affect the critical column density for sur- a 3D, axisymmetric cylinder with radius rcyl and length vival (Equation6) by factors less than a few. Subcycling 24rcyl in a domain with periodic boundary conditions is utilized to better resolve the cooling time. A tempera- along the axis of the cylinder and inflow/outflow bound- ture floor of 300 K is included, although most of the cold aries elsewhere. The cylinder is initially static, and the Stream gas is maintained at close to 104 K by photoion- background has velocity v = 228 km/s, i.e. a Mach num- ization (note that we don’t include self-shielding). As ber M ∼ 2. This value is motivated by the LMC infall for the Leading Arm simulations, we also cut off cooling velocity in the outer halo from (Besla et al. 2012; Salem 5 for gas temperatures above 0.6Thot = 6 × 10 K. et al. 2015; Bustard et al. 2020). The box dimensions Table2 shows the cylinder radius, background halo are fiducially 24rcyl × 24rcyl × 24rcyl, and we employ density, sound crossing time tsc, and column density for a simple density-based static mesh refinement to focus each simulation we consider. Figure 11 shows a column resolution on the area surrounding the Stream. In units density snapshot after t ∼ 11.5tsc perpendicular to the of rcyl/∆x, our base resolution is 1.25, and our highest cylinder axis for our simulation with initial radius rcyl = resolution is 20, which is achieved in the dense Stream −4 −3 2 kpc in a background halo density of nh ∼ 10 cm . and most of the surrounding mixing layer. As in Man- This gives an initial Stream column density of 1.2 × delker et al.(2020a), we find converged Stream growth 1020 cm−2, well within the survival regime. Overplotted rates at this resolution; in higher resolution runs, the on- contours show estimated Hα emission (see §5.2.1) set of the Kelvin-Helmholtz instability simply happens Figure 12 shows the mass evolution for each of our sim- earlier (see bottom panel of Figure 12). ulations (top panel). The cut-off in growth vs destruc- For radiative cooling, we assume the gas is in pho- tion, calculated from Equation6 to be ∼ 1019cm−2 for toionization equilibrium with the metagalactic UV back- low-metallicity gas appears robust: after a brief delay, in ground (Haardt & Madau 2012) with no additional ion- which the Kelvin-Helmholtz instability is growing from izing source from e.g. the Clouds or Milky Way. The numerical noise, Streams with N < 1019cm−2 gradually cooling rate is a tabulated function of density and tem- lose mass, while those with N > few ×1019cm−2 gain perature, while the gas ionization state (which feeds into −5 −3 mass. Even the rcyl = 5 kpc, nh = 10 cm simu- the equation of state) is a tabulated function of density lation shows an uptick in cold gas mass at late times, and internal energy (Wiersma et al. 2009). Motivated but we stop the simulation too early to follow the full by HST-COS observations of the Stream’s SMC filament evolution. Those with N ∼ 1019cm−2 appear marginal. (with Z ∼ 0.1Z ) and LMC filament (with Z ∼ 0.5Z ), The bottom panel of 12 shows the growth time for the as well as X-ray observations probing the metallicity of two most quickly growing simulations with rcyl = 2, 5 the Milky Way halo (Miller & Bregman 2015; Marty- −4 −3 kpc and nh = 10 cm . Simulations at our fiducial nenko 2021), we assume that both the Stream and Milky resolution (rcyl/∆x ∼ 20) develop the Kelvin-Helmholtz Way halo have metallicity Z = 0.3Z . The cooling instability the fastest, but those with rcyl/∆x ∼ 10 con- (or heating rate) is simply the contribution from hy- verge eventually to the same tgrow ∼ (0.1 − 0.2)χtsc. drogen and helium plus the contribution from metals, 18 Bustard & Gronke

21 r 10 cyl nh N kpc cm 3 cm 2 2.5 0.5 10 5 3.1 × 1018 2.0 10 5 1.2 × 1019 20 0.5 10 4 3.1 × 1019 ) l 2.0 5.0 10 5 3.1 × 1019 y c 10.0 10 5 6.2 × 1019 T 4 20

2 2.0 10 1.2 × 10 1.5 5.0 10 4 3.1 × 1020 < T (

0 1.0

10 M / M 0.5 ) 2 − m c (

0.0 ) y t c i s p 0 5 10 15 20 25 n

k 20 ( 0 10 e t/tsc D

z n m u

l 4 3 rcyl = 2.0 kpc, nh = 10 cm rcyl/ x = 20 o

C 1 4 3 r / x = 10 10 rcyl = 5.0 kpc, nh = 10 cm cyl rcyl/ x = 5 10 c s t / w 0 o 10 r g t 20 t = 2.0 Gyr 19 10 5 0 5 10 10 x (kpc) 10 1 0 5 10 15 20 25 Time / tsc Figure 11. Total Stream column density for the rcyl = −4 −3 2.0 kpc, nh = 10 cm simulation. Contours denote Hα Figure 12. Top: M/M0 for the cold gas (defined as surface brightnesses (estimated as SBcool/100) of 1 (blue), 5 T < 2Tcyl) for each simulation we ran. Lines are color- (cyan), 10 (yellow), and 15 (red) mR. This snapshot is taken coded by Stream column density projected perpendicular to after 2 Gyrs ∼ 11.5tsc when the cold gas mass has grown by the Stream axis. Streams with column densities above a few a few tens of percent. ×1019cm−2 eventually start to grow in mass, roughly con- sistent with the expectation from Equation6 with M ∼ 2. Bottom: Growth times for the simulations with the most This corresponds to tgrow ∼ 1.7 and 4.2 Gyrs for the r = 2 and 5 kpc simulations, respectively. These rates rapid growth onset, and with varying resolution (different cyl linestyles). Lower resolution leads to a delay in growth, but are consistent with our analytic estimates (Equation4) the growth times for rcyl/∆x ∼ 10, 20 eventually converge but are, again, too long to give any considerable mass to ∼ (0.1 − 0.2)χtsc, consistent with Equation4. growth within a lookback time.

5.2.1. Hα Emission – was proposed in the literature (Bland-Hawthorn et al. 2007; Bland-Hawthorn 2009; Tepper-Garc´ıaet al. 2015). Previous work focused on the Stream - halo interaction Figure 17 of Barger et al.(2017) shows the possible con- has been largely motivated by the peculiarly large Hα tribution of this shock-cascade to the measured Hα in- emission observed along the Stream, 30-50 milliRayleigh tensity at various Magellanic Stream longitudes. Since (mR) along many sightlines, which is too high to origi- the elevation in Hα emission depends on the ambient nate from photoionization from the Clouds, Milky Way, density, this model is distance-dependent: Barger et al. and extragalactic background (Weiner & Williams 1996; (2017) estimate that if the Trailing Stream is positioned Bland-Hawthorn et al. 2013; Barger et al. 2017). To 75 kpc above the South Galactic Pole, the shock-cascade partially explain this, a slow “shock-cascade” – where model could elevate the Hα emission by 50 mR near the stripped gas collides with Stream clouds further down- LMC and by 20 mR at the tail of the Stream. stream, forming a cascade of shocks, ionizes the Stream Survival of Magellanic Debris 19

As discussed previously, though, an implication of this shock cascade model, in which the Stream is initialized Z (shock) (Hα) 2 far from pressure equilibrium, is a quick dissolution of µHα = (1 + (Y/4X))KRαβ (nHII) ds Stream gas into the Milky Way halo; therefore, it’s of 4 b interest to check if our simulations, which instead retain (Hα) (1 + c)(T/10 K) αβ (T ) ≈ α0 4 d and even grow the Stream’s cold gas mass, can produce 1 + c(T/10 K) similar Hα bright spots at shocks in the radiative tur- K = 1.67 × 10−4cm2smR, the hydrogen and helium bulent mixing layer. R mass fractions are X = 0.7154, Y = 0.2703 appro- The standard relationship between isobaric cooling lu- priate for solar abundance, fully ionized gas (Asplund minosity and mass growth is (Fabian 1994) (Hα) et al. 2009), and αβ (T ) is the effective Hα recom- 5 m˙ 2 bination coefficient following the fitting formula of Pe- L = kBThot(1 + M ) (16) 2 µmp quignot et al.(1991). The constants are α0 = 1.169 × 10−13cm3s−1, c = 1.315, b = -0.6166, d = 0.523. To get 2 where the factor of M arises due to dissipation of the the H II density, we employ Trident (Hummels et al. infalling cold stream’s kinetic energy, in addition to the 2017) assuming photoionization equilibrium with the dissipation of thermal energy of the background cooling same metagalactic UV background we assume for our gas. cooling function. 2 Usingm ˙ ∼ m0/tgrow, where m0 = πrcyl`ρ for a cylin- Contours in Figure 11 show that Hα emission is, in der of length `, and assuming tgrow ∼ 0.1χtsc = 10tsc, large regions, only of order 1 mR. During the period of the total cooling surface brightness over a cylinder of t ' 7.5tsc, we measure the volume-averaged Hα emis- area A = 2πrcyl` is sion to be 0.5-1 mR with a standard deviation ∼ 1 − 2  n  mR. Maxima fluctuate between 15 and 20 mR, some- SB = L/A ∼ 100.5 h (1 + M2) mR cool 10−4cm−3 what lower than the findings of Tepper-Garc´ıa et al. (17) (2015), where detectable (> 30 mR) surface bright- Our simulations appear to reproduce this quite well. nesses appear in small bright spots. As in Tepper-Garc´ıa As in Gronke & Oh(2020a); Mandelker et al.(2020a), et al.(2015), our measured average and maximum sur- we output the total cooling luminosity and find that it face brightnesses are dependent on background halo den- matches the expectation givenm ˙ to within a factor of a sity, and with a higher halo density, we may achieve few during the mass growth stage. Note that this lumi- detectable maxima. nosity decreases with increasing Stream distance from We caution, though, that we don’t include self- the Milky Way. For the same Stream radius, if we de- shielding, so much of the Stream interior has unreal- crease the background halo density by a factor of 10, istically high H II densities. Our grid resolution is also we find a uniform decrease in emission by a factor ∼ 10 higher than the 1◦ resolution of WHAM (Barger et al. consistent with Equation 17. 2017), and we don’t account for beam smearing that The turbulent mixing layer is populated with gas at a would decrease the surface brightness. Nevertheless, variety of temperatures between 104 K and 106 K. We this may be worth future study given that our evolved confirm that a large contribution to the cooling emis- Stream is clearly turbulent, with bright Hα knots pro- sivity comes from T ∼ 104 K, as seen in other mix- duced in a similar manner (via shock ionization) to ing layer simulations (Tan & Oh 2021), including those Bland-Hawthorn et al.(2007) and Tepper-Garc´ıaet al. that include photoionization (Ji et al. 2019; Mandelker (2015) but without a fast disruption of the Stream. Fu- et al. 2020a). This is because, even though photoioniza- ture work could also include the contribution from colli- tion significantly decreases the hydrogen bump in the sional excitation that peaks near T ∼ 2×104 K, whereas cooling curve Λ (Wiersma et al. 2009), the emissivity here we’ve only considered Hα recombination. The dif- is ∼ n2Λ, and the n2 term offsets the low Λ. It’s been ference is likely a small factor, with still only of order suggested, then, that radiation in mixing layers of cold 1% of the hydrogen cooling luminosity being due to Hα; infalling streams can power Lyα blobs in massive ha- the greatest contribution is Lyα (Goerdt et al. 2010). los at high redshift (Goerdt et al. 2010; Mandelker et al. While shock ionization may be sufficient to explain the 2020b); naturally, one should wonder whether detectable underlying Hα emission along the Stream, these values Hα emission can arise too. of tens of mR fall far short of the observed Stream emis- We post-process our simulations to produce maps of sion above the South Galactic Pole, where there is an Hα surface brightness, calculated using Equation 13 especially large bump in Hα emission (> 300 mR). The from Tepper-Garc´ıaet al.(2015): most plausible explanation for this, so far, is recent ac- 20 Bustard & Gronke tivity from the . A bipolar, radiative ure6). These aspects deserve future study since they ionization cone from a Seyfert nucleus associated with call into question the feasibility of cold cloud accelera- SGR A* can not only explain the elevated Hα emission tion and growth in highly supersonic outflows. We spec- but also higher hydrogen ionization fraction, higher C ulate that this is why the high resolution, global galaxy IV/ CII and Si IV/ Si II ratios, and high C IV and Si simulations of Schneider et al.(2020), which model a IV column densities along that segment of the Stream highly supersonic M82-like wind, don’t find clear evi- (Bland-Hawthorn et al. 2013, 2019; Fox et al. 2020). dence of mass influx from the hot to cold phase. Instead, especially at large radii, mass flows from the cold to hot 6. DISCUSSION phase. This may be explained by the higher drag times While the main purpose of our work was to apply and higher growth times we find. Interestingly, the cold results from radiative turbulent mixing to the Magel- mass outflow rate in Schneider et al.(2020) rises un- lanic Stream in particular, our results also inform our til a distance of r ∼ 3 kpc, which could reflect either broader understanding of extragalactic multiphase gas an increase in total mass outflow rate or efficient cloud more broadly. We begin our discussion with implica- growth in this region where the hot phase is instead only tions for cold filaments, galactic winds, and ram pres- mildly supersonic (see their Figure 6). sure stripped tails behind infalling dwarf galaxies, specif- These additional dependencies on the Mach number ically the LMC and SMC (§6.1. We then discuss refine- are particularly important for subgrid models of cold gas ment guidelines to accurately capture mass entrainment in galactic winds such as Huang et al.(2020) as they in large-scale LMC-SMC simulations (§6.2) and discuss affect the cold gas mass and velocity, and hence most the caveats of our hydrodynamic treatment (§6.3). potential observables. Furthermore, if geometry depen- 6.1. Broader Extragalactic Implications dent effects such as the ‘shielding’ can indeed change the behavior dramatically (as suggested in this work), they While we focused in this work on the Milky Way halo, would need to be taken into account by such models. our findings also have implications for extragalactic sys- Ram pressure stripped tails: In addition to LMC-SMC tems. Most notably, they can be applied to multiphase tidal stripping, a popular explanation for the twisting, galactic winds, ram pressure stripping of galaxies, and bifurcated filaments of the Stream is ram pressure strip- cosmological filamentary inflows. As the latter has been ping of the Clouds by the Milky Way corona (e.g. Moore discussed extensively in the literature (Mandelker et al. & Davis 1994; Hammer et al. 2015; Wang et al. 2019). 2020a,b), we focus on the former two. While this is achieved in some simulations with dense Galactic winds: Observations of fast-moving, cold gas Milky Way coronae, it’s problematic that detailed simu- in galactic winds (Veilleux et al. 2020) have driven a lations of LMC ram pressure stripping, which are tuned surge of interest in the cloud-crushing problem. A main to fit the truncation radius of the LMC’s leading edge, development in recent years has been the aforemen- find that the mass stripped from the galaxy is small tioned survival and growth of cold gas in certain regimes compared to the Stream mass (Salem et al. 2015). It in- (t < t ), but most studies to-date have focused cool,mix cc creases somewhat when LMC outflows are flung into the on transonic or mildly supersonic flows. For a spheri- headwind (Bustard et al. 2018, 2020), but seemingly not cally symmetric cloud, previous simulations in the mass enough to create the observed LMC filament (Nidever growth regime showed poor convergence in high-M flows et al. 2008). However, these simulations, which focus (Gronke & Oh 2020a), possibly due to shattering of cold resolution on the dense LMC disk, fall short of the re- gas into smaller and smaller clumps down to the cool- finement criterion (§6.2) to capture cold gas survival be- ing length c t (McCourt et al. 2018) expected for s cool hind the LMC. The result is that recent ejecta is quickly M 1.6 flows at χ 100 (Gronke & Oh 2020b). This & & mixed into the Milky Way halo, but this situation should scale is not resolved in our work or in any published actually be quite ripe for mass growth: the LMC/SMC cloud-crushing simulations for overdensities of χ ∼ 100. infall is only M ∼ 2 − 3, and stripped gas forming a Nevertheless, with a slight shift from a spherically sym- long, cooling tail undergoes areal expansion that is key metric cloud to an axisymmetric 2.5D cloud, we find to rapid mass growth. Such an effect is, indeed, realized reasonable convergence, likely due to a “shielding” ef- in recent simulations of “jellyfish” galaxies ram pressure fect, i.e., a more quiescent region in the tail leading to stripped by the intracluster medium (Tonnesen & Bryan the re-coagulation of the clumps instead of further frag- 2021). mentation (as seen in McCourt et al. 2018; Melso et al. Similar coagulation and mass growth in long, cool- 2019, see also the setup of Banda-Barrag´anet al. 2020). ing tails behind the Clouds may be an enticing way to High-M flows do, however, have longer entrainment produce the bifurcated structure of the Stream without times ∝ 1+M and longer growth times (§4.2.3 and Fig- Survival of Magellanic Debris 21 requiring more direct ejection of gas from the Clouds Similarly, Tepper-Garc´ıaet al.(2019) use a grid-based via an increase in halo density (i.e. an increase in code and quote a maximum spatial resolution of ∼ 60 pc, 2 Pram ∼ ρv ). which is sufficient to resolve (by l = 8 cells) a column 20 −2 of N ∼ 10 cm at n−4 ∼ 1 or a column of N ∼ 19 −2 6.2. Refinement Guidelines 10 cm at n−4 ∼ 0.1. While they do not mention As discussed in § 3, one posibility for the inability of their refinement criterion, we can presume for the sake larger scale simulations to reproduce the Leading Arm is of argument that they have a density threshold in order numerical overmixing due to insufficient resolution. In to keep the mass per cell ∼constant. this section, we estimate what the resolution or refine- For the cloud-crushing problem, in particular, these ment requirements are in order to see the survival and refinement schemes are likely problematic. Much of growth of the Leading Arm discussed in this work. the mass growth occurs in the tail behind the disrupt- ing cloud, and with coarse-resolution in low-density gas, A cloud of radius rcl and density ρcl has a mass 3 the cloud fragments that would coagulate and grow at Mcl = 4/3πρclrcl. From here on, let’s assume the ap- proximately 104 K cloud is in pressure equilibrium with higher resolution would instead irreversibly mix into its surroundings, meaning its overdensity χ ∼ 100 rel- the background. This issue of numerical convergence ative to the background. For a mean mass ofm ¯ = plagues all hydrodynamic, galaxy-scale simulations to −24 −4 some extent, prompting the adoption of new refinement 1.67 × 10 g, ρcl = 100m ¯ × 10 n−4 where n−4 is the background halo density normalized to 10−4 cm−3. schemes (van de Voort et al. 2019; Hummels et al. 2019; We then re-write the cloud radius in terms of the total Peeples et al. 2019; Nelson et al. 2020; Mandelker et al. gas column density and the background halo density: 2021) that force uniform refinement in the CGM or are based on pressure or cooling length instead of density. rcl ≈ N/2ncl = 50N19/n−4 where N19 is the column density normalized to 1019 cm−2. The cloud mass is Such schemes are key to more accurate modeling of cold then gas in galaxy halos, and they may have great utility for simulations of the Magellanic System as well. 3 3 N19 Mcl ≈ 4 × 10 M 2 (18) n−4 6.3. Missing Physics To resolve such a cloud by l cells per dimension at a In this study, we used a simplified setup and omit- Mcl 4 3 mass resolution of mres, one needs = πl . We ted several physical effects – most notably the impact mres 3 re-write this in terms of the column density, i.e. what of magnetic fields and thermal conduction. While it has column density can be resolved by l cells at a mass res- been shown in similar ‘wind tunnel’ studies that mag- olution of mres: netic draping can facilitate entrainment through mag- netic drag (McCourt et al. 2015), for overdensitites of 2/3 1/3 2/3 1/3 N19 ≈ 0.1n−4 lmres = n−4 lmres,3 (19) χ ∼ 100 fairly strong magnetic fields (β < 1 in the hot medium, cf. equation (23) in Gronke & Oh 2020a) are 3 where mres,3 = mres/10 M . required to allow entrainment. Such values of β, while Simulations suggest convergence of mass growth when likely not realized on average in the CGM (Prochaska the cloud is resolved by a minimum of l = 8 cells (Gronke & Zheng 2019), may be created by magnetic field am- & Oh 2020a; Kanjilal et al. 2021); therefore, at the plification at cloud interfaces. For instance, McClure- 3 mass resolution of ∼ 4 × 10 M typical of state-of-the- Griffiths et al.(2010) measure a coherent field strength art Magellanic system simulations (Pardy et al. 2018; > 6µG in a Leading Arm HVC, which is strong enough Wang et al. 2019; Lucchini et al. 2020), only gas struc- to have an influence, but the corrections to drag time 20 −2 tures with column densities ' 10 cm at n−4 ∼ 1 and cold mass growth appear to be order unity even in 19 −2 −2 or column densities ' 2.5 × 10 cm at n−4 ∼ 0.1 are simulations with β approaching 10 due to compres- expected to have converged mass flux through the cloud- sional amplification (Gronke & Oh 2020a; Butsky et al. halo interface. In the outer halo, where the latter value 2020; Ji et al. 2016). of n−4 ∼ 0.1 is probably appropriate, this means the Similarly, thermal conduction does impact the dynam- 6 evolution of clouds with masses ' 6 × 10 M , less than ics and can prolong the lifetime of cold gas by compress- 7 e.g. the current Leading Arm mass of 3 × 10 M , are ing the clouds (Br¨uggen& Scannapieco 2016). Here, we well-reproduced; however, for higher background den- focused on the rcl > rcl,crit regime in which cold gas sities deeper into the Milky Way halo, convergence is can survive altogether. The Leading Arm and Trailing unlikely even for structures as massive as the Leading Stream comfortably fit into this regime. Furthermore, Arm. we note that it has recently been shown that the in- 22 Bustard & Gronke clusion of thermal conduction does not affect the cold momentum transfer via mixing – both critically gas mass growth rate in turbulent mixing layers (Tan depend on cooling (see Appendix and Figure9), et al. 2021) which also explains the convergence we find unlike adiabatic simulations where tdrag is inde- in tgrow. This is because the mass transfer is not domi- pendent of M (§4.2.3). nated by numerical diffusion but rather by mixing, and • Longer drag times for high-M flows are accom- we do resolve the outer mixing scale ∼ rcl reasonably well. panied by longer cloud tails in the strong cool- However, we note that both the inclusion of mag- ing regime: ltail(t)/rcl,0 ∼ α(1 + M)χln(1 + netic fields as well as thermal conduction can affect the t/tdrag). This increases the surface area and morphology of the cold gas (e.g., Gronke & Oh 2020a; eventually offsets the destructive effects of high Br¨uggen& Scannapieco 2016) as well as observables Mach number cloud crushing, leading to consid- (Tan & Oh 2021) and would need to be considered in a erable mass growth. In the weak cooling regime study focusing on the long-term evolution of the Lead- (tcool,mix/tcc ∼ 1), mass growth is slower, and the ing Arm. Naturally, in such future work one would also tail length is shorter (Figure8). need to take into account the hot gas halo profile as well • In addition to longer drag times for high-M flows, as the effect of gravity. 3 we also find longer growth times tgrow ∝ M (Fig- ure6). This suggests that cloud entrainment and 7. CONCLUSIONS growth in highly supersonic outflows, such as star- In this paper, we studied the effects of radiative tur- burst winds like M82, is more difficult than in bulent mixing on the ∼ 104 K gas originating from the transonic flows typically studied in many cloud- Magellanic clouds. We were motivated by two ques- crushing simulations. Cloud growth along a wind tions concerning the formation, survival, and possible flux tube is, in fact, likely distance-dependent. growth of Magellanic components: 1) Can the Leading This would need to be taken into account when Arm survive its (possibly high Mach number) “swing- comparing to larger scale galactic wind simula- out” from the Clouds and sufficiently separate itself from tions. the Clouds in opposition to ram pressure? 2) What are the rate and sign of the mass flux through the Trail- 7.2. Implications for the Leading Arm ing Stream’s radiative turbulent mixing layer, and is These results have implications, most clearly, for the this sufficient to grow previously simulated Streams (e.g. Leading Arm, whose Magellanic origin has recently been Pardy et al. 2018) to the observationally estimated mass questioned based on arguments of short drag times 9 > 2 × 10 M (D’Onghia & Fox 2016)? and fast Leading Arm breakup in the Milky Way halo (Tepper-Garc´ıaet al. 2019); the proposed alternative is 7.1. Survival and Growth of Cold Gas in High Mach that the Leading Arm is debris from a fore-runner dwarf Number Flows galaxy. Our complementary findings, based on idealized Our findings from the simulations (§4) are as follows: but high-resolution simulations, are as follows:

7 • For Mach numbers tested (up to M = 6), we find • Reasonably massive (∼ 3 × 10 M , Br¨unset al. cloud survival and subsequent growth via radia- 2005) Leading Arms have sufficient column den- 18 −2 tive, turbulent mixing that has been previously sities (N > Ncrit ≈ 4.8 × 10 M cm ) to sur- shown for lower Mach number flows. The survival vive infall through the Milky Way halo. In this criterion (rcl > rcrit, or tcool,mix/tcc ∼ 1) (Equa- scenario, the filaments and clumps of the Leading tion1) appears robust, and we find good conver- Arm could be part of a long debris field in the gence likely due to the “shielding” effect induced wake of the progenitor cloud (see Figure4), which by our assumed 2.5D cloud morphology (see Ap- survives infall and even grows in mass due to ra- pendix). This is contrary to 3D, spherical cloud diative cooling at the cloud-halo interface. The simulations in the mass growth regime, which show distance variation along the Leading Arm inferred poor convergence at high Mach number (Gronke from observations supports this picture. & Oh 2020a). • Whether the Leading Arm can overcome drag and 1/2 • We find tdrag ∼ α(1 + M)χ tcc, with α ∼ 0.2 − stay ahead of the Clouds is most sensitive to the 0.3 for our strongly cooling (tcool,mix/tcc ∼ 0.1) Leading Arm’s velocity at swing-out. A more vi- simulations. The Mach number dependence fol- olent tidal stripping event (M > 4), especially in lows from a combination of geometric effects and the lower-density outer halo, is preferable (Figure Survival of Magellanic Debris 23

3). Some more recent formation scenarios are per- • The associated cooling luminosity and estimated missible but require higher M swing-outs. Hα emission in the turbulent mixing layer are like- wise fairly small on average, but bright spots (tens • The drag time and mass doubling time for these of mR) may partially explain the observed Hα allowed scenarios are at least as long as the infall emission along pointings separated from the main time, however, meaning we could be seeing the HI stream (Barger et al. 2017). This shock ioniza- Leading Arm in its more destructive, stretching tion is qualitatively similar to the shock cascade phase rather than its rapid growth phase (Figure model put forth by Bland-Hawthorn et al.(2007) 2). The Leading Arm, then, likely had to be at but here operates in a Stream that grows in mass least as massive at swing-out as it is today. instead of quickly dissolves. While our study focuses on the origin of the Leading Arm due to an earlier LMC-SMC interaction, the results presented here would also affect the ‘forerunner’ theory (in which the Leading Arm is due to gas stripped from ACKNOWLEDGMENTS other dwarf galaxies). In each case, the Leading Arm The authors gratefully acknowledge Nir Mandelker comprises the disrupted clouds and filaments of a larger for detailed feedback and discussions, as well as Elena progenitor object. However, our results suggest that D’Onghia, Andy Fox, Scott Lucchini, Peng Oh, Ellen the Leading Arm may not need to originate in a dwarf Zweibel and the organizers and participants of the KITP galaxy protected by a dark matter halo, as suggested “Fundamentals of Gaseous Halos” workshop. CB was by Tepper-Garc´ıa et al.(2019). Instead, cooling and supported in part by the National Science Founda- coagulation allows the clumps to survive and even grow tion under Grant No. NSF PHY-1748958 and by the in mass despite fast breakup of the progenitor cloud. Gordon and Betty Moore Foundation through Grant Sufficiently resolved simulations of Leading Arm infall No. GBMF7392. MG was supported by NASA subject to gravity, a changing density profile, etc. are a through the NASA Hubble Fellowship grant HST-HF2- logical next step to confirm this. 51409 and acknowledges support from HST grants HST- GO-15643.017-A, and HST-AR15039.003-A. This work 7.3. Implications for the Trailing Magellanic Stream made use of the Athena MHD code (Stone et al. 2008), We also briefly estimate the effects of radiative turbu- the yt analysis and visualization package (Turk et al. lent mixing layers on the Trailing Stream (§5). 2011), and the Trident package (Hummels et al. 2017). Computations were performed on the Stampede2 su- • We estimate that the majority of the Stream (with percomputer under allocations TG-AST190019 and TG- 19 −2 N > Ncrit ≈ 10 cm ) is surviving infall and AST180036 provided by the Extreme Science and Engi- gaining mass, but the mass growth rate is too long neering Discovery Environment (XSEDE), which is sup- (a few Gyrs) to significantly increase the Stream ported by National Science Foundation grant number mass. ACI-1548562 (Towns et al. 2014).

APPENDIX

.1. Varying Halo Density Profile In §3, we probed the survival of the Leading Arm and its separation from the Clouds using a toy analytic problem. Our results are presented in Figure3 assuming the Salem et al. 2015 Milky Way halo density profile (Equation 13. Here, Figure 13 shows the same analysis using instead the Faerman et al. 2020 density profile (Equation 14), which gives a denser Milky Way corona beyond ∼ 100 kpc. The resulting figure is almost identical to Figure3, showing our results are insensitive to small changes in halo density; instead, the separation is most sensitive to the velocity of the Leading Arm swing-out.

.2. Resolution Study

We ran each of our M = 1.5, 3, and 4.5 simulations with tcool,mix/tcc = 0.1 at resolutions of 8, 16, and 32 cells per cloud radius. Figure 14 shows the resulting cold gas masses and average cold gas velocities. The M = 6 simulations were only run at 8 and 16 cells per cloud radius and aren’t shown here. Somewhat surprisingly, our results show good convergence, especially in the cold gas mass measured as ρ > ρcl/3. The T > 2Tcl curves trend higher with 24 Bustard & Gronke

= 3.0, vLMC = 300 km/s = 4.5, vLMC = 300 km/s 1021 100 1021 100

50 50

25 25 3e8 M 3e8 M 20 05) 20 05) 10 s+ 20 10 s+ 20 Brun Brun M ( 10 M ( 10 ) 3e7 ) 3e7 2 2 3e6 M 3e6 M

m 0 m 0 c c

( ( Ncrit(Z 0.1Z )

N Ncrit(Z 0.1Z ) 10 N 10 1019 1019

25 Ncrit(Z 1.0Z ) 25 Separation (in kpc) Separation (in kpc) after lookback time after lookback time Ncrit(Z 1.0Z ) 50 50

1018 100 1018 100 220 200 180 160 140 120 100 80 60 220 200 180 160 140 120 100 80 60 Distance from Milky Way (kpc) Distance from Milky Way (kpc)

Figure 13. Final separation between Leading Arm and LMC/SMC, now assuming the Milky Way halo density profile from Faerman et al.(2020). Mappings between LMC/SMC radius and lookback time are unchanged. The results are nearly identical to that shown in Figure3. increased resolution, likely reflecting increased (better-resolved) turbulent pressure support. The drag time becomes shorter at high resolution, which is possibly another indication that momentum transfer in the tail crucially affects the drag time. When compared to previous simulations of high-M cloud-crushing (Gronke & Oh 2020a), though, these differences seem miniscule. We believe this is related to the choice of “cloud” geometry. Gronke & Oh(2020a) modeled spherical 3D clouds and found poor convergence for simulations with M > 1.5 (see their Figure 13). Our “2.5D” cloud geometry, motivated by our envisioned geometry of the Leading Arm at swing-out from the Clouds, appears to facilitate a “shielding” effect that prevents shear instabilities from further stripping cloud fragments. This effect is seen as well in Melso et al.(2019) in simulations of infalling gas clouds and in Banda-Barrag´anet al.(2020), which models shock-multicloud interactions and reports that most bulk properties are converged even for M = 10 simulations. While the mass growth rate and entrainment time are reasonably well converged in these simulations, we caution that the mixing layer itself and any associated observables are not converged. These depend also on the inclusion of thermal conduction, which does not affect the mass influx rate in turbulent mixing layer simulations (Tan et al. 2021) but does affect ion column densities (Tan & Oh 2021).

REFERENCES

Abruzzo, M. W., Bryan, G. L., & Fielding, D. B. 2021, Besla, G., Kallivayalil, N., Hernquist, L., et al. 2007, ApJ, arXiv e-prints, arXiv:2101.10344 668, 949 Antwi-Danso, J., Barger, K. A., & Haffner, L. M. 2020, —. 2012, MNRAS, 421, 2109 ApJ, 891, 176 Bland-Hawthorn, J. 2009, in The Galaxy Disk in Armillotta, L., Fraternali, F., Werk, J. K., Prochaska, J. X., Cosmological Context, ed. J. Andersen, Nordstr¨oara, & Marinacci, F. 2017, MNRAS, 470, 114 B. m, & J. Bland-Hawthorn, Vol. 254, 241–254 Bland-Hawthorn, J., Maloney, P. R., Sutherland, R. S., & Asplund, M., Grevesse, N., Sauval, A. J., & Scott, P. 2009, Madsen, G. J. 2013, ApJ, 778, 58 ARA&A, 47, 481 Bland-Hawthorn, J., Sutherland, R., Agertz, O., & Moore, Aung, H., Mandelker, N., Nagai, D., Dekel, A., & B. 2007, ApJL, 670, L109 Birnboim, Y. 2019, MNRAS, 490, 181 Bland-Hawthorn, J., Maloney, P. R., Sutherland, R., et al. Banda-Barrag´an,W., Br¨uggen,M., Heesen, V., et al. 2020, 2019, ApJ, 886, 45 arXiv e-prints, arXiv:2011.05240 Bregman, J. N., Anderson, M. E., Miller, M. J., et al. 2018, Barger, K. A., Madsen, G. J., Fox, A. J., et al. 2017, ApJ, ApJ, 862, 3 851, 110 Br¨uggen,M., & Scannapieco, E. 2016, Begelman, M. C., & Fabian, A. C. 1990, MNRAS, 244, 26P doi:10.3847/0004-637X/822/1/31 Survival of Magellanic Debris 25

Cold Gas Mass Evolution: = 1.5 Cold Gas Velocity: = 1.5 101 1.0 > cl/3 rcl/ x = 8

T < 2Tcl rcl/ x = 16 0.8 rcl/ x = 32 d n i w 0 0.6 v M / / 3 / M 1 0.4 v

rcl/ x = 8 0.2 100 rcl/ x = 16 rcl/ x = 32 0.0 0 5 10 15 20 0 5 10 15 20 Time / tcc Time / tcc

Cold Gas Mass Evolution: = 3.0 Cold Gas Velocity: = 3.0 1.0 > cl/3 rcl/ x = 8 101 T < 2Tcl rcl/ x = 16 0.8 rcl/ x = 32 d n i w 0 0.6 v M / / 3 / M 1 0.4 v

rcl/ x = 8 0.2 rcl/ x = 16 100 rcl/ x = 32 0.0 0 10 20 30 0 10 20 30 Time / tcc Time / tcc

Cold Gas Mass Evolution: = 4.5 Cold Gas Velocity: = 4.5 1.0 > cl/3 rcl/ x = 8

T < 2Tcl rcl/ x = 16 0.8 101 rcl/ x = 32 d n i w 0 0.6 v M / / 3 / M 1 0.4 v

rcl/ x = 8 0.2 rcl/ x = 16 0 10 rcl/ x = 32 0.0 0 10 20 30 40 50 0 10 20 30 40 50 Time / tcc Time / tcc

Figure 14. Resolution study for the cold gas mass growth (left) and cold gas velocity (right). All simulations have tcool,mix/tcc = 0.1 but have different Mach numbers M = 1.5, 3, 4.5. Contrary to previous high Mach number simulations for a 3D spherical cloud, which were clearly not converged (Gronke & Oh 2020a), we find that M (ρ > ρcl/3) is well-converged for these simulations with an effectively 2.5-D morphology. On the other hand, M (T < 2Tcl) continues to rise with increasing resolution, likely reflecting the greater number of cells that lie at the temperature floor. The cold gas velocities appear reasonably converged at 32 cells per cloud radius, though there is a clear trend for slightly shorter drag times at higher resolution. A separate study is needed to assess convergence at even higher resolution, for higher M flows, and for spherical cloud geometries.

Br¨uns,C., Kerp, J., Staveley-Smith, L., et al. 2005, A&A, Cautun, M., Deason, A. J., Frenk, C. S., & McAlpine, S. 432, 45 2019, MNRAS, 483, 2185 Bustard, C., Pardy, S. A., D’Onghia, E., Zweibel, E. G., & Connors, T. W., Kawata, D., & Gibson, B. K. 2006, Gallagher, J. S., I. 2018, ApJ, 863, 49 MNRAS, 371, 108 Bustard, C., Zweibel, E. G., D’Onghia, E., Gallagher, J. S., D’Onghia, E., & Fox, A. J. 2016, ARA&A, 54, 363 I., & Farber, R. 2020, ApJ, 893, 29 D’Onghia, E., & Lake, G. 2008, ApJL, 686, L61 Butsky, I. S., Fielding, D. B., Hayward, C. C., et al. 2020, Drlica-Wagner, A., Bechtol, K., Allam, S., et al. 2016, ApJ, 903, 77 ApJL, 833, L5 26 Bustard & Gronke

Fabian, A. C. 1994, ARA&A, 32, 277 Ji, S., Oh, S. P., Ruszkowski, M., & Markevitch, M. 2016, Faerman, Y., Sternberg, A., & McKee, C. F. 2020, ApJ, MNRAS, 463, 3989 893, 82 Kallivayalil, N., van der Marel, R. P., Besla, G., Anderson, Farber, R. J., & Gronke, M. 2021, arXiv e-prints, J., & Alcock, C. 2013, ApJ, 764, 161 arXiv:2107.07991 Kanjilal, V., Dutta, A., & Sharma, P. 2021, MNRAS, 501, Fielding, D. B., Ostriker, E. C., Bryan, G. L., & Jermyn, 1143 A. S. 2020, ApJL, 894, L24 Klein, R. I., McKee, C. F., & Colella, P. 1994, ApJ, 420, 213 For, B. Q., Staveley-Smith, L., Matthews, D., & Koposov, S. E., Belokurov, V., Torrealba, G., & Evans, McClure-Griffiths, N. M. 2014, ApJ, 792, 43 N. W. 2015, ApJ, 805, 130 For, B.-Q., Staveley-Smith, L., & McClure-Griffiths, N. M. Lee, D. 2013, Journal of Computational Physics, 243, 269 2013, ApJ, 764, 74 Lee, D., & Deane, A. E. 2009, Journal of Computational For, B. Q., Staveley-Smith, L., McClure-Griffiths, N. M., Physics, 228, 952 Westmeier, T., & Bekki, K. 2016, MNRAS, 461, 892 Li, Z., Hopkins, P. F., Squire, J., & Hummels, C. 2020, Fox, A. J., Frazer, E. M., Bland-Hawthorn, J., et al. 2020, MNRAS, 492, 1841 ApJ, 897, 23 Lucchini, S., D’Onghia, E., Fox, A. J., et al. 2020, Nature, Fox, A. J., Richter, P., Wakker, B. P., et al. 2013, ApJ, 772, 585, 203 110 Mandelker, N., Nagai, D., Aung, H., et al. 2020a, MNRAS, Fox, A. J., Wakker, B. P., Savage, B. D., et al. 2005, ApJ, 494, 2641 —. 2019, MNRAS, 484, 1100 630, 332 Mandelker, N., van den Bosch, F. C., Nagai, D., et al. Fox, A. J., Wakker, B. P., Barger, K. A., et al. 2014, ApJ, 2020b, MNRAS, 498, 2415 787, 147 Mandelker, N., van den Bosch, F. C., Springel, V., et al. Fox, A. J., Barger, K. A., Wakker, B. P., et al. 2018, ApJ, 2021, arXiv e-prints, arXiv:2107.03395 854, 142 Martynenko, N. 2021, arXiv e-prints, arXiv:2105.02557 Fryxell, B., Olson, K., Ricker, P., et al. 2000, ApJS, 131, Mathewson, D. S., Cleary, M. N., & Murray, J. D. 1974, 273 ApJ, 190, 291 Gardiner, L. T., Sawa, T., & Fujimoto, M. 1994, MNRAS, Mathewson, D. S., Schwarz, M. P., & Murray, J. D. 1977, 266, 567 ApJL, 217, L5 Gardiner, T. A., & Stone, J. M. 2008, Journal of McClure-Griffiths, N. M., Madsen, G. J., Gaensler, B. M., Computational Physics, 227, 4123 McConnell, D., & Schnitzeler, D. H. F. M. 2010, ApJ, Goerdt, T., Dekel, A., Sternberg, A., et al. 2010, MNRAS, 725, 275 407, 613 McClure-Griffiths, N. M., Staveley-Smith, L., Lockman, Goldsmith, K. J. A., & Pittard, J. M. 2017, MNRAS, 470, F. J., et al. 2008, ApJL, 673, L143 2427 McClure-Griffiths, N. M., Pisano, D. J., Calabretta, M. R., —. 2018, MNRAS, 476, 2209 et al. 2009, ApJS, 181, 398 Gronke, M., & Oh, S. P. 2018, MNRAS, 480, L111 McCourt, M., Oh, S. P., O’Leary, R., & Madigan, A.-M. —. 2020a, MNRAS, 492, 1970 2018, MNRAS, 473, 5407 —. 2020b, MNRAS, 494, L27 McCourt, M., O’Leary, R. M., Madigan, A.-M., & Gronke, M., Oh, S. P., Ji, S., & Norman, C. 2021, arXiv Quataert, E. 2015, MNRAS, 449, 2 e-prints, arXiv:2107.13012 Melso, N., Bryan, G. L., & Li, M. 2019, ApJ, 872, 47 Haardt, F., & Madau, P. 2012, ApJ, 746, 125 Miller, M. J., & Bregman, J. N. 2015, ApJ, 800, 14 Hammer, F., Yang, Y. B., Flores, H., Puech, M., & Moore, B., & Davis, M. 1994, MNRAS, 270, 209 Fouquet, S. 2015, doi:10.1088/0004-637X/813/2/110 Nelson, D., Sharma, P., Pillepich, A., et al. 2020, MNRAS, Huang, S., Katz, N., Scannapieco, E., et al. 2020, MNRAS, 498, 2391 497, 2586 Nidever, D. L., Majewski, S. R., & Burton, W. B. 2008, Hummels, C. B., Smith, B. D., & Silvia, D. W. 2017, ApJ, The Astrophysical Journal, 679, 432. 847, 59 http://arxiv.org/abs/0706.1578%5Cnhttp: Hummels, C. B., Smith, B. D., Hopkins, P. F., et al. 2019, //stacks.iop.org/0004-637X/679/i=1/a=432 ApJ, 882, 156 Nidever, D. L., Majewski, S. R., Butler Burton, W., & Ji, S., Oh, S. P., & Masterson, P. 2019, MNRAS, 487, 737 Nigra, L. 2010, ApJ, 723, 1618 Survival of Magellanic Debris 27

Nigra, L., Stanimirovi´c,S., Gallagher, John S., I., et al. Stone, J. M., Gardiner, T. A., Teuben, P., Hawley, J. F., & 2012, ApJ, 760, 48 Simon, J. B. 2008, ApJS, 178, 137 Padnos, D., Mandelker, N., Birnboim, Y., et al. 2018, Tan, B., & Oh, S. P. 2021, arXiv e-prints, arXiv:2105.11496 MNRAS, 477, 3293 Tan, B., Oh, S. P., & Gronke, M. 2021, MNRAS, 502, 3179 Pardy, S. A., D’Onghia, E., & Fox, A. J. 2018, ApJ, 857, Tepper-Garc´ıa,T., Bland-Hawthorn, J., Pawlowski, M. S., 101 & Fritz, T. K. 2019, MNRAS, 488, 918 Pardy, S. A., D’Onghia, E., Navarro, J. F., et al. 2020, Tepper-Garc´ıa,T., Bland-Hawthorn, J., & Sutherland , MNRAS, 492, 1543 R. S. 2015, ApJ, 813, 94 Patel, E., Kallivayalil, N., Garavito-Camargo, N., et al. Tonnesen, S., & Bryan, G. L. 2021, ApJ, 911, 68 2020, ApJ, 893, 121 Towns, J., Cockerill, T., Dahan, M., et al. 2014, Computing Peeples, M. S., Corlies, L., Tumlinson, J., et al. 2019, ApJ, in Science & Engineering, 16, 62. 873, 129 doi.ieeecomputersociety.org/10.1109/MCSE.2014.80 Pequignot, D., Petitjean, P., & Boisson, C. 1991, A&A, Townsend, R. H. D. 2009, ApJS, 181, 391 251, 680 Turk, M. J., Smith, B. D., Oishi, J. S., et al. 2011, ApJS, Prochaska, J. X., & Zheng, Y. 2019, MNRAS, 485, 648 192, 9 Putman, M. E., Peek, J. E. G., & Joung, M. R. 2012, van de Voort, F., Springel, V., Mandelker, N., van den ARA&A, 50, 491 Bosch, F. C., & Pakmor, R. 2019, MNRAS, 482, L85 Putman, M. E., Saul, D. R., & Mets, E. 2011, MNRAS, Veilleux, S., Maiolino, R., Bolatto, A. D., & Aalto, S. 2020, 418, 1575 A&A Rv, 28, 2 Putman, M. E., Gibson, B. K., Staveley-Smith, L., et al. Venzmer, M. S., Kerp, J., & Kalberla, P. M. W. 2012, 1998, Nature, 394, 752 A&A, 547, A12 Richter, P., Fox, A. J., Wakker, B. P., et al. 2018, ApJ, 865, Wang, J., Hammer, F., Yang, Y., et al. 2019, MNRAS, 486, 145 5907 —. 2013, ApJ, 772, 111 Weiner, B. J., & Williams, T. B. 1996, AJ, 111, 1156 Salem, M., Besla, G., Bryan, G., et al. 2015, The Wiersma, R. P. C., Schaye, J., & Smith, B. D. 2009, Astrophysical Journal, doi:10.1088/0004-637X/815/1/77 Monthly Notices of the Royal Astronomical Society, 393, Scannapieco, E., & Br¨uggen,M. 2015, ApJ, 805, 158 99 Schneider, E. E., Ostriker, E. C., Robertson, B. E., & Williamson, D., & Martel, H. 2021, ApJ, 907, 9 Thompson, T. A. 2020, ApJ, 895, 43 Yang, Y., Hammer, F., Fouquet, S., et al. 2014, MNRAS, Sparre, M., Pfrommer, C., & Ehlert, K. 2020, MNRAS, 442, 2419 499, 4261 Zhang, D., Thompson, T. A., Quataert, E., & Murray, N. Stanimirovi´c,S., Hoffman, S., Heiles, C., et al. 2008, ApJ, 2017, MNRAS, 468, 4801 680, 276