ebook img

Feedback Effects on Low-Mass Star Formation PDF

1.9 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Feedback Effects on Low-Mass Star Formation

The Astrophysical Journal, Accepted PreprinttypesetusingLATEXstyleemulateapjv.5/2/11 FEEDBACK EFFECTS ON LOW-MASS STAR FORMATION Charles E. Hansen 1, Richard I. Klein 1,2, Christopher F. McKee 1,3, and Robert T. Fisher 4 The Astrophysical Journal, Accepted ABSTRACT Protostellar feedback, both radiation and bipolar outflows, dramatically affects the fragmentation and mass accretion from star-forming cores. We use ORION, an adaptive mesh refinement (AMR) gravito-radiation-hydrodynamics code, to simulate low-mass star formation in a turbulent molecular cloud in the presence of protostellar feedback. We present results of the first simulations of a star- 2 formingclusterthatincludebothradiativetransferandprotostellaroutflows. Werunfoursimulations 1 to isolate the individual effects of radiation feedback and outflow feedback as well as the combination 0 of the two. We find that outflows reduce protostellar masses and accretion rates each by a factor of 2 threeandthereforereduceprotostellarluminositiesbyanorderofmagnitude. Thismeansthat,while radiation feedback suppresses fragmentation, outflows render protostellar radiation largely irrelevant n forlow-massstarformationaboveamassscaleof0.05M . Wefindinitialfragmentationofourcloud a (cid:12) J athalftheglobalJeanslength,around0.1pc. Withinsufficientprotostellarradiationtostopit,these 0.1 pc cores fragment repeatedly, forming typically 10 stars each. The accretion rate in these stars 3 scales with mass as predicted from core accretion models that include both thermal and turbulent 1 motions; the accretion rate does not appear to be consistent with either competitive accretion or ] accretion from an isothermal sphere. We find that protostellar outflows do not significantly affect R the overall cloud dynamics, in the absence of magnetic fields, due to their small opening angles and S poor coupling to the dense gas. The outflows reduce the mass from the cores by 2/3, giving a core h. to star efficiency, (cid:15)core (cid:39) 1/3. The simulations are also able to reproduce many observation of local star-formingregions. Oursimulationwithradiationandoutflowsreproducestheobservedprotostellar p luminosity function. All of the simulations can reproduce observed core mass functions, though we - o find they are sensitive to telescope resolution. We also reproduce the two-point correlation function r of these observed cores. Lastly, we reproduce IMF itself, including the low-mass end, when outflows st are included. a [ 1. INTRODUCTION number of stars with some efficiency, 0.2 < (cid:15) < 1.0, core 1 the IMF can be recreated. The actual conversion from The origin of the stellar initial mass function (IMF) v observed core masses to stellar masses may not be so is one of the most fundamental problems of star forma- 1 simple due to cores blending in projection (Kainulainen tion. The IMF can be described by single power law for 5 et al. 2009b; Michel et al. 2011), small cores that dis- stars above 0.5 M (Salpeter 1955), and a broken power 7 (cid:12) perse before making stars (Myers 2009; Padoan & Nord- 2 law (Kroupa 2002) for stars below this mass. Alterna- lund 2011) and cores accreting mass over time (Padoan . tively, it can be described as a log-normal distribution 1 &Nordlund2011). Nevertheless, theCMFlikelyplaysa with characteristic mass m = 0.2M that joins with 0 c (cid:12) strong role in creating the IMF theSalpeterpowerlawforstarsabove1.0M (Chabrier 2 (cid:12) The observed CMF provides support to core accre- 2005). Any theory of the IMF must explain both the 1 tion theories of star formation (Shu 1977; McKee & Tan functional form and the characteristic mass. A tantaliz- : 2003), which start with a bound core and produce a sin- v ing observational clue to the functional form lies in dust gle stellar system. Simulations of turbulence find the i observations in star-forming regions (Motte et al. 1998; X functional form of the core mass function (log-normal Testi&Sargent1998;Johnstoneetal.2000,2001;Motte r et al. 2001; Beuther & Schilke 2004; Stanke et al. 2006; plus power law) is the expected outcome of supersonic a turbulence (Padoan & Nordlund 2002; Padoan et al. Alvesetal.2007;Enochetal.2008;Sadavoyetal.2010). 2007). Analytic predictions of a turbulent density field These dust maps find many high density concentrations with self-gravity can also reproduce this functional form thatareconsistentwithprestellarandprotostellarcores. (Hennebelle & Chabrier 2008, 2009). The characteristic When the mass of these cores is calculated, the core coremassisthentheJeansmassatsomecriticaldensity mass function (CMF) has the same functional form as and temperature. However, choosing the correct den- the IMF, but with a higher characteristic mass, ranging sity and temperature is problematic. In purely isother- from0.2M to1M . Ifeachcoreisconvertedtoasmall (cid:12) (cid:12) mal turbulence, there is no characteristic Jeans mass. As objects collapse and the density increases, the Jeans 1Astronomy Department, University of California, Berkeley, mass decreases. There is no transition where this de- CA94720 2Lawrence Livermore National Laboratory, Livermore, CA creaseinJeansmasswillstopwithoutadditionalphysics. 94550 This means the core masses are either infinitely small or 3PhysicsDepartment, UniversityofCalifornia, Berkeley, CA functions of the global Jeans mass of the host molecu- 94720 4Physics Department, University of Massachusetts, Dart- lar cloud. Observations are consistent with a universal mouth,MA IMF, however, even over a range of cloud Jeans masses 2 (Kroupa 2002; Chabrier 2003; Bastian et al. 2010). This TABLE 1 means the characteristic core mass must be set by lo- Table of simulations calphysics,whichisothermalturbulencecannotprovide. Name ThermalPhysics Winds? Star-forming regions are approximately isothermal be- B Barotropic No cause the thermal time scales are much shorter than the BW Barotropic Yes dynamical time scales, but there are ways to break this R Radiation No isothermality. RW Radiation Yes One approach is to focus on the coupling between gasanddustinstar-formingenvironments(Larson2005; alytical estimates of mass loss from winds can fully ex- Elmegreen et al. 2008). At low densities, gas-dust cou- plain the range of mass loss expected from observations pling is poor and the gas is theoretically slightly sub- 0.2 < (cid:15) < 1.0 depending on the details of the cores core isothermal (temperature decreases with increasing den- andtheoutflows(Matzner&McKee2000). Theoutflows sity). At higher densities, gas and dust are well coupled travelbeyondtheirstarsoforiginanddepositenergyinto and the gas is theoretically slightly super-isothermal. parsec-scale turbulent motions. Evidence suggests that This yields a critical density and temperature at the molecular cloud turbulence appears on the scale of the transitionthatcanbeconvertedintoaJeansmass. This entire cloud (Ossenkopf & Mac Low 2002; Brunt et al. criticaldensity,ρ∼10−19 gcm−3,islowerthantheden- 2009),soismostlikelydrivenbysourcesotherthanpro- sities of large star-forming regions like Orion, however, tostellar outflows. Nonetheless, the dynamics on parsec and unlikely to explain the characteristic core mass in scales can be significantly altered by outflows (Norman these regions. &Silk1980;McKee1989;Li&Nakamura2006;Banerjee One critical mass is the point when dust becomes et al. 2007; Nakamura & Li 2007; Swift & Welch 2008; opaque to its own thermal radiation (Low & Lynden- Carroll et al. 2010; Arce et al. 2010; Wang et al. 2010). Bell 1976). At that density, the gas will heat up and In this paper, we investigate the fragmentation of a raise the Jeans mass, creating a minimum Jeans mass of parsec-scale molecular cloud into cores and then into fragmentation. A barotropic simplification of this effect stars. This requires refinement to capture the fragmen- sets the mass in many simulations (e.g. Bate & Bon- tation and radiative transfer to fragment at the correct nell 2005; Bonnell et al. 2006; Offner et al. 2008; Bate mass scale, similar to Offner et al. (2009b). This also 2009a; Hennebelle et al. 2011). The density of this tran- requires simulation of protostellar outflows, to capture sition is extremely high (∼ 10−13 g/cm3) (Masunaga the CMF to IMF efficiency, similar to the work of Cun- et al. 1998) and the resulting Jeans mass (∼ 0.004M ) ningham et al. (2011) for high-mass star formation. The (cid:12) is much lower than the characteristic core mass (Low & goal of this paper is to explain both the observed CMF Lynden-Bell 1976; Whitworth et al. 2007). In addition, and the IMF while self-consistently finding (cid:15) . In or- core the barotropic approximation is inaccurate when com- der to do this, we have performed the first simulation of pared to simulations that include dust radiation (Boss a star-forming cluster to include both radiative transfer et al. 2000; Krumholz et al. 2007a; Offner et al. 2009b; and protostellar outflows. Bate2009b;Price&Bate2009;Tomidaetal.2010). The We describe our numerical method and simulation importanceofdustradiationcanbeseeninBate(2009b) setup in §2. In §3, we report the results of our sim- andPrice&Bate(2009),whofoundthattheinclusionof ulations. We discuss the implications of our results on radiationsignificantlysuppressestheformationofbrown starformationtheoryandcomparetoobservationsin§4. dwarfsdespitethenearabsenceofprotostellarluminosity We summarize our conclusions in §5. in the simulations. The most powerful break from isothermality comes 2. SIMULATIONS from protostellar radiation. Massive protostars are ca- Weperformfourprimarysimulationswithnearlyiden- pable of heating an entire cloud (Krumholz et al. 2007a; tical initial conditions but different physics as controlled Cunningham et al. 2011; Myers et al. 2011; Krumholz numerical experiments in order to isolate the impor- et al. 2011). Low-mass stars do not have the same long tance of feedback effects. These simulations all include range influence, but simulations show they can still dra- hydrodynamics, gravity and basic sink particle physics matically reduce fragmentation in the disk and recover (Krumholz et al. 2004), but may also include radiation a 1 M characteristic core mass (Offner et al. 2009b; (cid:12) (Krumholz et al. 2007a) and/or sink particle outflows Krumholz et al. 2011). Protostellar radiation does not (Cunningham et al. 2011). The simulations are shown create a unique critical density, but it does weaken the in Table 1. Barotropic simulations are labeled with a density dependence of the effective Jeans mass (Bate B, simulations with radiation are labeled with R, and 2009b). simulation with protostellar winds are labeled with W. Given a core mass function, there is still the question Simulation labels can contain multiple letters. of CMF to IMF efficiency, (cid:15) . The primary mecha- core nism for reducing the core mass is protostellar outflows. 2.1. Initial Conditions Stars of all masses show bipolar outflows during their All simulations have the same initial conditions, also formation (Richer et al. 2000; Shepherd 2003) and are used in Offner et al. (2009b). The initial gas tempera- recreated in MHD simulations with sufficient resolution ture is T = 10 K, the box length is L = 0.65 pc and (Ciardi & Hennebelle 2010; Tomida et al. 2010). These g outflows remove mass that would otherwise accrete onto the average density is ρ¯ = 4.46×10−20 g cm−3, corre- stars, thereby reducing the final mass (Matzner & Mc- sponding to nH =1.91×104 cm−3. The total box mass Kee 2000; Arce & Sargent 2006; Wang et al. 2010). An- is 185 M(cid:12). For radiative simulations, the radiation tem- perature, T is initialized to 10 K. The radiation energy r 3 density is thus E =aT4 =7.56×10−11 erg cm−3. dynamics including self-gravity, radiative transfer, pro- r Toobtaintheturbulentinitialconditions,webeginour tostellar outflows, and radiating star particles, all on an simulations without self-gravity and apply velocity per- adaptive grid. Every cell in the grid has four conserved turbationstoaninitiallyconstantdensityfieldusingthe quantities: mass density, ρ, vector momentum density, method described in Mac Low (1999). These perturba- ρv, gas energy density, ρe, and radiation energy density, tions correspond to a Gaussian random field with a flat E. These conserved quantities can be used to calculate powerspectrumintherange1≤k ≤2. Theapplication derivedquantitiessuchasvelocity,v,andpressure,P. In of these perturbations continues for three cloud cross- addition to the gas quantities, we evolve point-mass star ing times and then stops. At this point the turbulence particles, each with a position x , mass M , momentum i i follows a Burgers power spectrum, P(k) ∝ k−2, charac- p , angular momentum, j and luminosity L . The sub- i i i teristic of supersonic hydrodynamic turbulence. The 3D script i refers to the star particle number. The particle turbulent Mach number is M = 6.6, which gives a 3D methodisexplainedinKrumholzetal.(2004)(hereafter rms velocity dispersion, σ =1.2 km/s. With this Mach KKM04),withtheadditionofradiation(Krumholzetal. v number the cloud is approximately virialized: 2007a) and outflows (Cunningham et al. 2011). The full set of evolution equations for gas and particles is 5σ2 α = (cid:39)1. (1) vir GM/R ∂ρ+∇·(ρv)+(cid:88)[M˙ W(r )−M˙ W (r )ξ(θ )]=0, This is slightly above the linewidth-size relation ∂t KKM04 i w,i w i i σ (cid:39) 0.7(R/1 pc)1/2 km s−1 (Solomon et al. 1987; i (5) Heyer & Brunt 2004), and is equivalent to σ = 1.2(R/1 pc)1/2 km s−1,whichiswellwithintheobserved range (e.g. Falgarone et al. 2009) ∂ρv +∇·(ρvv)=−∇P −ρ∇φ−(cid:88)(p˙W(r )− (6) After driving for three cloud crossing times, we then ∂t i i turn off driving, turn on gravity and follow the subse- M˙ v W (r )ξ(θ )·ˆr ), quentgravitationalcollapseforapproximatelyoneglobal w,i w,i w i i i free fall time: (cid:114) 3π ∂(ρe) tff = 32Gρ¯ =0.315 Myr, (2) ∂t +∇·[(ρe+P)v]=ρv∇φ−κRρ(4πB−cE)− where ρ¯is the mean density of the box. The simulations (cid:18) ρ (cid:19)2 Λ(T )− (7) withradiationbecomeprohibitivelycomputationallyex- µm g H pensive at late times and are stopped at t = 0.83 t ff (cid:88) withatotalstellarmass of30M forsimulationR.The [ε˙ W(r )− (cid:12) KKM04 i barotropicsimulationsarecontinuedtot=1.05tff before i they are stopped. At this time the total stellar mass in k T K simulation B is 50 M(cid:12) compared to the total simulation M˙w,iWw(ri)ξ(θi)µ(Bγ−w1)], mass of 185 M . There is still gas bound to protostars (cid:12) totaling 11 M when the simulations end. Our stellar (cid:12) (cid:18) (cid:19) mass estimates may therefore be too low by 20%. ∂ cλ E−∇· ∇E =κ ρ(4πB−cE)+ (8) Given our temperature of 10 K, the Jeans length at ρ¯ ∂t κ ρ P R is λJ =(cid:18)πGcρ¯2s(cid:19)1/2 =0.20 pc, (3) (cid:18)µmρH(cid:19)2Λ(Tg)+(cid:88)LiW(ri), i and the Jeans mass is ∇2φ=−4πG[ρ+(cid:88)M δ(r )], (9) i i 4π (cid:18)λ (cid:19)3 i M = J ρ¯=2.7 M . (4) J 3 2 (cid:12) 1 M˙ = M˙ , (10) i 1+f KKM04 The turbulent Jeans mass, at density ρ = M2ρ¯, is 0.4 w MT(cid:12)h.e calculations have a 2563 base grid with 4 levels of M˙w,i =fwM˙i = 1+fwf M˙KKM04, (11) refinement by factors of 2, giving an effective resolution w of 40963. This resolution corresponds to ∆x4 = 32 AU p˙i =p˙KKM04, (12) at the finest refinement level. r =x−x , (13) i i 2.2. Evolution Equations θ =acos(ˆr ·ˆj). (14) We use the parallel adaptive mesh refinement code i i i ORION for our simulations. The numerical method is The quantities entering these equations are defined nearly identical to what we have used in previous pa- below. Equations (5) and (6) are the fluid equations pers (Krumholz et al. 2007a; Offner et al. 2009b; Cun- for mass and momentum, modified to include particles. ningham et al. 2011; Krumholz et al. 2011; Myers et al. Equations (7) and (8) are the energy equations for gas 2011). ORION solves the equations of compressible gas and radiation respectively. The Poisson equation for the 4 gravitational potential, φ is given by equation (9). The overthe4-cellaccretionzoneoftheparticleandW rep- w particle evolution is given by equations (10), (11) and resentstheoutflowwindowfunction. Outflowsareimple- (12). We use periodic boundary conditions for all gas mentedasinCunninghametal.(2011),andsummarized and particle quantities. here. For the radiative runs, we adopt Marshak boundary Our outflow model is specified by the dimensionless conditions for the radiation field. This allows radiation parameter f , which sets the mass flux of outflow as w to escape from the box as it would from a molecular a fraction of the accretion onto a star, and v , the w,i cloud. The equation of state for the gas is given by wind launch speed. The mass fraction in our simula- tions is f = 0.3. The wind speed is set by the Keple- ρk T (cid:18) v2(cid:19) w (cid:112) P = B g =(γ−1)ρ e− , (15) rian speed at the surface of the star, vk,i = GMi/r∗,i µmH 2 where r∗,i is the protostellar radius, but is is capped at 60 km s−1 for computational speed. Specifically, where µ=2.33 is the mean molecular weight for molec- v =min(v ,60 km s−1). The velocity cap has a sim- ular gas of Solar composition and γ is the ratio of spe- w,i k,i ilar effect to the choice of Cunningham et al. (2011) to cific heats. Since most of the simulation domain is too use v = v /3 for the most massive stars in the cal- coldtorotationallyexcitemolecularhydrogen, weadopt w,i k,i γ = 5/3, representing a monatomic ideal gas. The term culation. Wind gas is injected at temperature Tw =104 κ ρ(4πB−cE) in equations (7) and (8) represents en- K. P ergy exchanged between the radiation field and the dust The wind is injected over a window function Ww, in our gas, with B = caTg4/4π representing the Planck 1 (cid:26)r−2 if 4∆x<r ≤8∆x emission function integrated over all frequencies. The W = . (17) w C 0 otherwise opacities κ and κ are Planck and Rosseland means 1 P R given by the dust opacities from the iron normal, com- which represents a shell just outside of the accretion re- posite aggregates dust model of Semenov et al. (2003). gion. The normalization constant C is computed nu- 1 We assume that the gas and the dust are thermally cou- merically to avoid numerical aliasing effects that occur pled. When the gas temperature exceeds the dust de- from injecting a spherical wind into a Cartesian grid. struction temperature, the energy exchange term goes Theexactangulardistributionofthewindisdescribed to zero and the gas and radiation unrealistically decou- in the function ξ. The functional form is taken from ple. Toaddresscoolingfromgasabovethedustdestruc- Matzner & McKee (1999), tion temperature, we use the line cooling function Λ(T ) g from Cunningham et al. (2006). This removes energy (cid:20) (cid:18) 2 (cid:19) (cid:21)−1 from the gas and adds that energy to the radiation field ξ(θ)= ln (sin2θ+θ2) , (18) θ 0 (see Cunningham et al. (2011) for further details). The 0 radiation flux limiter is given by λ = 1 (cid:0)cothR− 1(cid:1), where θ is a flattening parameter that sets the opening R R 0 whereR=|∇E|/κ ρE (Levermore&Pomraning1981). angleofthewind. Weusethefiducialvalueofθ =0.01. R 0 It should be noted that we have excluded the radiation Equation(18)isaveragedoverthesolidanglesubtended pressure and radiation enthalpy advection terms from by a grid cell at the outer radius of W . This averaging w equations (6), (7) and (8) that appear in the analogous is particularly important near θ = 0. In addition, ξ is equationinKrumholzetal.(2007a). Thisapproximation set to zero for θ ∼ π/2, so that winds are not injected is justified in the formation of low-mass stars, as shown directly into the plane of any equatorial disks. in Offner et al. (2009b). We update the luminosity of each star, L , using the i Whenradiationisneglected,theenergyexchangeterm protostellar evolution model described in Offner et al. fromequation(7)disappears,andweclosethesystemof (2009b). Inthismodel,75%oftheaccretionenergyisra- equations with a barotropic equation of state for the gas diatedawaywhile25%isnominallyusedtopowerwinds. pressure: The energy of winds in our simulations is already deter- (cid:34) (cid:18) ρ (cid:19)γ−1(cid:35) minedbyfw andvw,i,whichistypically15%oftheaccre- P =ρc2 1+ , (16) tion energy. The remaining 10% of the accretion energy s0 ρ c is effectively lost. When winds are not present, we still only radiate 75% of the accretion energy for consistency (cid:112) wherecs0 = kBT0/µmH istheisothermalsoundspeed across simulations. at temperature T0 = 10 K, γ = 5/3 and ρc is the crit- The evolution equations can be described as the fluid ical density. The critical density determines the transi- and radiation equations from Offner et al. (2009b) com- tion from isothermal to adiabatic regimes and we adopt binedwiththeparticleequationsandlinecoolingofCun- ρc = 2×10−13 g cm−3 to agree with the collapse so- ningham et al. (2011), but there is one important mod- lution from Masunaga et al. (1998) prior to H2 dissoci- ification. In the KKM04 sink particle methodology, all ation. Simulations that use the barotropic equation of particleswithoverlappingaccretionzonesaremergedto- state achieve maximum densities of 5×10−15 g cm−3, gether. Thisgivesaneffectivemergerradiusof8cells,or with an effective temperature at ρmax of 10.8 K. Most of 256AUatagridresolutionof32AU.Tolimitthiseffect, thegasinanygivensimulationsiseffectivelyisothermal. wechangedthemergerradiusto4cells,representingthe The particle quantities M˙ , p˙ and point when a particle is in the accretion zone of another KKM04 KKM04 ε˙ in Equations (5 - 7) represent the sink parti- particle. This gives an effective merger radius of 128 AU KKM04 cle accretion rates of mass, momentum and energy from at our resolution. Even with this improvement, our par- the surrounding gas in the absence of winds as given by ticle algorithm will unrealistically merge stars that pass KKM04. The function W represents a window function within 128 AU. To address this, we have implemented a 5 mass limit of m = 0.05 M , above which stars do merge (cid:12) notmerge. Thislimitischosentocorrespondtothemass of second collapse in the formation of a star with final mass1M (Masunagaetal.1998;Masunaga&Inutsuka (cid:12) 2000). Second collapse occurs when the protostar’s core temperature becomes high enough to dissociate molec- ular hydrogen. Before second collapse, protostars are extended balls of gas with radii of a few AU and have a muchhighercollisionalcross-sectionthanmainsequence stars. Aftersecondcollapse,stellarmergersshouldbeex- tremelyrareandwedonotallowthem. Thisapproachis alsousedinMyersetal.(2011). Themassofsecondcol- lapse depends on the accretion history of the protostar and is necessarily less than 0.05 M for brown dwarfs (cid:12) (Stamatellos&Whitworth2009;Bate2011). Theeffects of numerical merger suppression is explored in §4.9. ORION utilizes a second order Godunov scheme to solve the equations of compressible gas dynamics (Tru- elove et al. 1998; Klein 1999). These are equations (5)- (7), excluding terms from stars and radiation. The Pois- son equation (9) is solved using a multi-grid iteration scheme (Truelove et al. 1998; Klein 1999; Fisher 2002). Theflux-limiteddiffusionradiationequation(8)andthe radiation terms in equation (7) are solved using the con- servative update scheme from Krumholz et al. (2007b) modified to include the pseudo-transient continuation of Shestakov & Offner (2008). We use the Truelove criterion (Truelove et al. 1997) to determinetheadditionofnewAMRgridssothatthegas density in the calculation always satisfies J2πc2 ρ< s , (19) G(∆x )2 l where ∆x is the cell size on level l. We adopt a Jeans l numberof0.125. Inthesimulationswithradiativetrans- fer, it is necessary to resolve the spatial gradients in the radiationfield. Areasofhighradiationgradientsarenear accreting stars, which tend to already be refined under the Truelove criterion. This is not always true for more evolved stars, which have higher luminosities and have accretedthedensegasthatwouldtriggerrefinement. We find that the radiation gradients are adequately resolved Fig. 1.—ColumndensityoftheentiresimulationdomainforBW (left)andB(right)attimes0,0.2,0.4,0.6,0.8and1.0t fromtop by refining whenever |∇E|∇x /E >0.25. ff l to bottom. Star particles are marked with white circles. There is very little difference on the domain scale with and without winds 3. RESULTS for the barotropic simulations. The color bar is gcm−2 and the entiredomainis0.67pcacross. 3.1. Large Scale Evolution The evolution of the barotropic simulations with and without winds is depicted in Figure 1. Figures 2 and 3 in the core and an additional group of low-mass stars depict the evolution of the radiative simulations without forms, totaling ∼ 10 stars per core. These cores with and with winds, respectively. In all simulations, for multiple stars generally resemble observed high-stellar- t(cid:46)0.4t ,therearecloud-scalefilamentsthatslowlycon- density cores (Teixeira et al. 2007; Chen & Arce 2010). ff tract, allowing 3 turbulent cores of width ∼ 20,000 AU There are an additional 20 stars in cores that are still and density ∼10−19 g cm−3 to form. This length is half forming at the end of the simulation, giving a global to- the Jeans length at the average cloud density. The first tal of 80 stars. Three of the cores coalesce by the end of coretoformisatthecenterofeachpanelinFigures1-3, the simulation to form a single group of 30 stars. the second core is left of the central core, near the left The evolution of the 3D rms velocity dispersion, σ , is v edge,andthethirdcoreisneartherightedge,attheend shown in Figure 4. The global turbulence decays un- ofasimulation-widefilament. Atthispoint,thecoresbe- til star formation ramps up at t ∼ 0.5t . There are ff gintofragment,whilenewcoresform,eventuallyforming two main mechanisms for star formation to increase σ . v 6 fully developed cores. These cores each have a cen- First, as stars accrete mass and deepen their gravita- tral stellar system. In simulation R, this central system tionalpotential,thesurroundinggascanconvertgravita- is all that forms in each core. In all other simulations, tionalenergyintokineticenergyasitfallsintothestars. the central system represents ∼ 75% of the stellar mass This is shown in the gradual increase in Mach number 6 Fig. 2.—Columndensity(left)anddensityweightedtemperature Fig. 3.— Same as Figure 2, but for simulation RW. The high (right) for simulation R at times 0, 0.2, 0.4, 0.6, and 0.8 t from temperature regions are the paths of outflows. It only takes a ff toptobottom. Starparticlesaremarkedwithwhitecircles. small amount of gas at 104 K to move the average temperature above30Kthroughthatlineofsight. for t > 0.6t in the B simulation. This effect is strong ff injected by the winds is greater than the characteris- enough to return σ to near its original virialized value v tic cloud momentum. At this point, the total amount by itself. In rare cases, a many-body close encounter be- of turbulent momentum that has been dissipated (in- tween stars will eject some gas at high velocities. There cluding dissipation of wind momentum) is roughly the is not much momentum injected this way and the en- total amount of momentum that has been injected by ergy quickly dissipates, but it causes spikes in σ for the v the winds. By the end of the BW and RW simulations, barotropicsimulations,whichhavemoresmall-scalefrag- the total wind momentum injected into the cloud is over mentation and therefore more many-body close encoun- twice the characteristic cloud momentum. The kinetic ters. The second mechanism occurs when protostellar energyinjectedfromthewindsdissipatesovertime,sug- winds are included. Somemass accreted ontostars isdi- gesting a steady-state solution where the velocity dis- rectly injected around the stars at high velocities. This persion of the cloud is constant with time as the winds causesthesmoothincreaseinσ forsimulationsBWand v replenish energy as quickly as it can dissipate. RW as well as spikes from events with particularly high accretion rates that lead to bursts in wind momentum. 3.2. Evolution of the Protostellar Population ThetotalmomentuminjectedbywindsformodelBW is shown in Figure 5. For comparison, a characteris- The total mass in stars and the total number of stars tic value of the magnitude of the momentum associated as functions of time is shown are Figures 6 and 7. The withinternalmotionsinthecloud,M σ isalsoplot- realizationoftheinitialturbulenceisslightlydifferentbe- cloud v ted. For t > 0.8t , the total momentum that has been tween the radiative and barotropic simulations, so star ff 7 however, because protostellar luminosity inhibits frag- mentation and the winds reduce that luminosity. The threesimulationsotherthanRshowadramaticincrease in the number of stars at 0.6 t <t<0.8 t . ff ff Fig. 4.—TimeevolutionofglobalrmsMachnumberforsimula- tionswithandwithoutwindsandwithandwithoutradiation. The turbulentenergyinallsimulationsdecaysforhalfaglobalfreefall time, at which point gravitational potential energy from stars is convertedintokineticenergy,whichraisesthermsvelocity. When winds are included, they contribute over twice as much energy as gravityitself. Fig. 6.— Total mass in stars as a function of time for the four simulations. Themassoftheentiresimulationdomainis180M(cid:12). Fig. 5.—Totalmomentumthathasbeeninjectedbywindsover timeforthebarotropicsimulations. Forcomparison,thetotalmass ofthesimulationmultipliedbythevelocitydispersionisalsoplot- ted. Thetotalwindmomentumintegratesallinjectedmomentum over time, even from winds that have decayed. As a result, the amountofinjectedmomentumiseventuallyhigherthantheactual momentumofthecloud. formation begins at different times. This is the result of a slightly higher maximum density due to turbulence anditisunrelatedtothechoiceofradiativeorbarotropic thermodynamics. Thefirststarsformatt=0.2t forthe ff radiative simulations and t = 0.35t for the barotropic ff simulations. The turbulence overlaps enough between theradiativeandbarotropiccases,however,thatatlater times the total mass in stars at any given time is similar for the two cases when winds are not included. Winds reduce the mass in stars by about a factor of 3 in both the radiative and barotropic cases. The number of stars Fig. 7.—Numberofstarsasafunctionoftimeforthefoursim- does not change between BW and B, implying winds do ulations. Inthebarotropiccase,thenumberofstarsisunaffected not cause or suppress fragmentation by themselves. The by winds. In the radiative case, radiation suppresses the number number of stars in RW is significantly greater than in R, ofstarsunlesswindsarepresent. 8 The evolution of median stellar mass is shown in Fig- ure8. Thisisaroughproxyforthecharacteristicmassof the protostellar mass function. Note that it will always be lower than the median mass of the IMF, because not all of the stars have finished accreting. Both BW and RW maintain a median around 0.05 M (similar to that (cid:12) of the protostellar mass functions in McKee & Offner 2010) throughout the simulation. The median does in- creaseforsimulationBWaroundt>t astheformation ff rate of new stars decreases. The median of B fluctuates more, but is around 0.2 M . Lastly, R maintains a me- (cid:12) dian around 0.5M . This general behavior should be (cid:12) expected. Themedianmassislowestwhenwindsarein- cludedandhighestwhenradiationisallowedtosuppress fragmentation. The case with both winds and radiation endsupsimilartoBWbecausewindsreduceprotostellar luminosities. Fig. 9.—Totalstellarluminosityversustimeforsimulationswith radiation,bothwithandwithoutwinds. Windsdramaticallylower theluminosity. As was seen in the plot of total luminosity, the mean and median values of protostellar luminosity are much lower when winds are included. The disparity in aver- age luminosity is even greater than the disparity in to- tal luminosity because there are fewer stars when winds are excluded. The average luminosity in simulation R is heavily influenced by a single 6.6 M star that accounts (cid:12) for over half the total luminosity in the simulation. Un- like the low-mass stars, most of this luminosity is pow- eredbynuclearfusionratherthanaccretion. Protostellar luminosities will be discussed further in section 4.3 Fig. 8.— Median mass of stars as a function of time for the foursimulations. Thetwocaseswithwindsmaintainlowmedians throughoutthesimulations. Thecasewithradiationwithoutwinds (caseR)isabletosuppressfragmentationandnewstarformation largelystopsastheoriginalstarsaccretemass. Thegloballuminosityevolutionfortheradiativesimu- lationsisplottedinFigure9. Thewindsreducethetotal luminositybyafactorofupto10atanygiventime. This is expected since the global accretion luminosity is L= 3 (cid:88)GM(cid:63)iM˙(cid:63)i ∼ GM(cid:63),totM˙(cid:63),tot; (20) 4 R (cid:104)R (cid:105) (cid:63)i ∗ i given that the total mass in stars M is reduced by (cid:63),tot a factor of 3 when winds are included, and the total accretion rate of stars M˙ is also reduced by 3, the Fig. 10.— Mean and median stellar luminosity versus time for (cid:63),tot total luminosity is therefore reduced by a factor of 9 as- simulationswithradiation. Thetoppanelisthesimulationwithout windsandthebottompanelisthesimulationwithwinds. suming the characteristic stellar radius, (cid:104)R (cid:105), does not (cid:63) change. Main sequence stars typically have a positive correlation between mass and radius, suggesting the fac- 3.3. Thermal Evolution tor of 9 is an upper limit. However, at any given point in time, many stars in our simulations are in the de- All simulations start at a background temperature of generate regime where mass and radius are negatively 10 K and are bathed in 10 K radiation. Stellar radiation correlated (Chabrier et al. 2009), which counteracts the and mechanical energy from protostellar outflows, can positivecorrelationfromthehighermassstarsandkeeps raise this temperature. We have identified gas heated the total luminosity ratio near 9. above 12 K as thermally affected by stellar feedback. The average stellar luminosity is shown in Figure 10. Turbulent dissipation alone heats almost no gas above 9 12K,leavingstellarfeedbackastheonlyexplanationfor this heating. The total mass of this gas is shown in Fig- ure 11. The simulation with winds has significantly less Fig. 11.—Totalgasmassheatedabove12Kversustime, com- paredtothebackgroundvalueof10K.Thetotalmassinstarsis alsoplottedforreference. Fig. 12.— Phase plot showing total gas mass as a function of temperature(y-axis)anddensity(x-axis)forradiativesimulations heated gas than the simulation without winds. This is with (left) and without (right) winds. Phase plots are taken at duetothereducedluminositycausedbythewindsshown times of 0.25, 0.5 and 0.75 t from top to bottom. The high- ff in Figure 9. In each simulation, the mass in heated gas temperature, low-density gas on the left part of the wind phase plotsisoutflowgas. Warm,high-densitygasisgasnearaluminous roughly follows the mass in stars. When winds are in- star. cluded, the mass in stars drops by a factor of 3, and the mass of heated gas also falls due to the reduced stellar luminosity. Someofthelostradiativeheatingisreplaced mentally different from the isotropic, homogeneous hy- by mechanical heating from outflows. This is evident in drodynamic turbulence it replaces. This result is also the ratio of heated gas mass to stellar mass. The mean seen in Nakamura & Li (2007), who find that the late value of this ratio is 0.9 without winds, but rises to 1.5 timeturbulentstatisticsdonotmatchexpectedisotropic with them. Wind gas is injected at a temperature of 104 hydrodynamic results. One key difference is the energy K,butcoolsquickly. Themassingaswithtemperatures from outflows is highly anisotropic. Outflow cavities are above 1000 K is less than 2% of the mass in stars at any marked by long channels with high velocity shear be- given time. tweenthefastoutflowgasandtheslowambientgas. This Tofurtherexploretheheatedgas,temperature-density shear is detectable as solenoidal energy. There is some phase plots with and without winds are shown in Figure compressiveenergyattheheadoftheoutflowcavity,but 12. The phase plots with and without winds are notably most of the surface area is the side walls of the cavity different in two areas. First, the wind gas fills the high- and not the head. To measure the relative importance temperature,low-densitydomain,whilethesamedomain of solenoidal and compressive motions, we use the ratio is empty without winds. Second, high-density gas with (cid:104)|∇×v|2(cid:105)/(cid:104)(∇·v)2(cid:105), shown in Figure 13. In the case of ρ>10−16g/cm3hasahighertemperaturerangewithout isotropicturbulence,thisratiois2,whichisalsotheratio ofsolenoidaltocompressiveenergies(Elmegreen&Scalo windsthanwithwindsbecausetheextrastellarluminos- 2004). Wind injection greatly increases the solenoidal ity heats that gas. When winds are included, that dense velocities, steadily increasing the solenoidal to compres- gas is less common in addition to being colder; there is sive ratio over the course of the simulation. Bursts in morefragmentation,whichturnsdensegasintostars. In outflows injection around t = 0.5t and t = 0.65t are addition,someofthegasisalsoblownawaybythewinds ff ff also visible in the solenoidal velocity. Anisotropic tur- themselves. bulence behaves differently than isotropic turbulence; in 4. DISCUSSION particular, it takes longer to decay since it decays on the crossing time calculated from the smallest velocity 4.1. Supporting a Cloud with Outflows dispersion (Hansen et al. 2011). It is difficult to mea- The turbulent evolution shown in Figures 4 and 5 sure this increase in the decay time in our simulations, roughly agrees with previous simulations of molecular however, becausethewindsdriveturbulenceoverawide clouds with outflows, such as those in Nakamura & Li range of scales. (2007) and Wang et al. (2010). Turbulent energy decays The other major difference between the initial turbu- initially,onlytobereplacedbykineticenergyfromwinds lence and wind-driven turbulence is seen in the rms ve- and from gravity. While these sources can increase the locity, σ , of gas with ρ > ρ¯. Even in isotropic, dense total kinetic energy of gas, the new turbulence is funda- homogenous, hydrodynamic turbulence, there is a neg- 10 a centrally concentrated cloud can achieve a quasi-static balance between outflows and gravity. 4.2. Comparison to the IMF Our simulations are marginally able to capture bina- ries, so most star particles should represent a single stel- larobjectinsteadofastellarsystem. Thetypicalbinary separation for main sequence stars is 50 AU (Mathieu 1994), just slightly larger than our resolution of 32 AU. We cannot form stars within 128 AU of another star due to merging star particles, but stars that form be- yond 128 AU can spiral in to 32 AU through interaction with gas. At any given time, about 2/3 of our star par- ticles are in stellar multiples. The multiple properties are dynamic due to unstable high-order multiples and we cannot compare to the observed system properties. We can, however, compare to observed stellar proper- ties. The mass functions of the four main simulations Fig. 13.— Ratio of solenoidal to compressive velocities squared are shown in Figure 15 and compared to the stellar IMF ((cid:104)|∇×v|2(cid:105)/(cid:104)(∇·v)2(cid:105)) versus time with and without winds for in Chabrier (2005) as well as protostellar mass functions the radiative simulations. For pure hydrodynamic, isotropic tur- from McKee & Offner (2010). The mass functions in bulence,thisratioisaround2. Thisratiostaysnear2whenwinds are excluded. When winds are included, the turbulence is much moreanisotropic,leadingtohighersolenoidalfractions. ative correlation between density and velocity, causing σ <σ (Offner et al. 2009a). The winds themselves dense v are collimated, very low density gas and have difficulty transmitting energy into high density gas. This means that while σ is much greater with winds than without, v σ does not change much when winds are included. dense The evolution of σ compared to σ is shown in Fig- dense v ure 14. Fig. 15.—Themassfunctionofallstarsineachsimulationare shown in blue histograms. The stellar IMF from Chabrier (2005) is plotted as the solid green line. The protostellar mass functions fortheturbulentcoremodelandtheisothermalspheremodelfrom McKee&Offner(2010)arethedashedlinesanddashdottedlines, respectively. Top left: RW at t = 0.83t . Top right: R at t = ff 0.83t . Bottom left: BW at t = 1.09t . Bottom right: B at ff ff t=1.03t ff Figure 15 are shown at the latest time available for each simulation. The barotropic simulations could be evolved to later times due to the computational expense of flux- limited diffusion with many stellar sources. The mass Fig. 14.—TimeevolutionofrmsMachnumberofdensegasand functions in Figure 15 are not exactly comparable to an allgaswithandwithoutwinds. WindssignificantlyraisetheMach number of the light gas, but do not strongly influence the dense IMF because some of the stars are still accreting. The gasturbulence. barotropic mass functions have evolved to a later time and should be similar to the IMF. The radiative simu- Because the dense gas is relatively unaffected by the lations are still actively accreting and are closer to the outflows, if our cloud had been centrally concentrated protostellar mass functions in McKee & Offner (2010). likethatofNakamura&Li(2007)orWangetal.(2010), Among the protostellar mass functions, the turbulent the dense part of the cloud would have most likely col- core and competitive accretion (not shown in Figure 15) lapsedonitselfevenwiththesupportofprotostellarout- models both roughly match RW. The isothermal sphere flows. Magnetic fields may have an effect, as shown by protostellar mass function is the best fit to mass func- Wang et al. (2010). The magnetic fields help transmit tionsofthesimulationswithoutwinds,butthisismerely outflowenergytoamuchlargersolidangle, sothateven a function of both mass functions being too top-heavy.

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.