Star Formation at Very Low Metallicity. IV. Fragmentation does not depend on metallicity fo
大小麦哲伦云的星际消光

大小麦哲伦云的星际消光薛梦瑶;姜碧沩;高健【摘要】We review the extinction laws of the Magellanic Clouds (MCs) from the ultravi-olet (UV), optical to infrared and the 2D extinction maps of the MCs. The dust properties of the MCs are also discussed. The color excess E (B-V ) of the Large Magellanic Cloud (LMC) is about 0.13 mag. If RV=2.6 is adopted, the AV value is about 0.34 mag. The color excess E(B-V ) of the Small Magellanic Cloud (SMC) is about 0.16 mag, which corresponds to an AV of about 0.45 mag (for RV=2.8). The star formation regions LMC 30 Dor and Super Shell LMC 2 could be considered as relatively dense regions. They both lack the 2175˚A extinction bump feature and their extinction rises steeply in the UV range, while in the diffuse regions of LMC, the 2175˚A bump is relatively strong. As for the SMC, the 2175˚A bump is absent in the SMC bar region, and the extinction rises even more steeply in the UV band, while in the SMC wing region, the 2175˚A bump feature is present. In the infrared band, the ratio of the color excess E(J-H)/E(H-K) is roughly between 1.03 and 1.36 for LMC, clearly lower than the ratio of the Milky Way that is about 1.73. In SMC, the infrared extinction is fairly small, and as a tiny error of photometry or intrinsic color will bring large uncertainty, there is no consistent conclusion yet. The extinction curves of the LMC and SMC can not be described by the simple CCM parameterization containing a single param-eter RV which is valid for the Milky Way. By fitting the extinction curve as well as the infrared radiationof the Magellanic Clouds, it is found that the carbonaceous dust grains, compared to the silicate dust, are less abundant than those in the Milky Way. In particular, the lack of small carbonaceous dust grains is suggested to explain the w eakness of the 2175˚A bump in the star-forming regions where the UV photons may destroy the very small carrier like graphite or PAH molecules. The dust-to-gas ratio in the MCs is significantly lower than that in the Milky Way very possibly due to the lower metallicity.%综述了大小麦哲伦云在紫外、可见光和红外波段的消光规律及其2D 消光图,并讨论了其尘埃的性质。
厦门“PEP”2024年11版小学5年级下册第十五次英语第2单元全练全测(含答案)

厦门“PEP”2024年11版小学5年级下册英语第2单元全练全测(含答案)考试时间:100分钟(总分:120)A卷考试人:_________题号一二三四五总分得分一、综合题(共计100题)1、填空题:I enjoy going for ______ (散步) in the evening. It helps clear my mind and relax.2、What is 6 + 2?A. 7B. 8C. 9D. 10答案:B3、What food do we get from cows?A. EggsB. MilkC. CheeseD. All of the above4、听力题:The flowers are ______ (beautiful) in the garden.5、填空题:The bee buzzes around the _______ (花).6、听力题:Star formation occurs in regions of space called molecular _______.7、How many days are there in a week?A. 5B. 6C. 7D. 8A base feels slippery and can turn __________ paper blue.9、听力题:She _____ (likes) to play soccer.10、填空题:My favorite way to celebrate is ______.11、听力题:A _______ is a combination of two or more substances that are not chemically combined.12、In what type of habitat do polar bears live?A. DesertB. ForestC. ArcticD. Savannah13、填空题:The __________ is a famous natural wonder in South America. (亚马逊雨林)14、填空题:The __________ is known for its unique rock formations. (约塞米蒂)15、填空题:A wildcat is a small ________________ (猫).16、How many days are in a week?A. FiveB. SixC. SevenD. Eight17、选择题:What is the name of the famous mountain in Oceania?A. Mount CookB. Mount KosciuszkoC. Mount TaranakiD. All of the above18、填空题:I want to create a ________ with my ideas.19、听力题:The chemical symbol for californium is _____.The owl has excellent _______ (夜间视力).21、听力题:The __________ is a region known for its rich biodiversity.22、听力题:The chemical formula for potassium chloride is ______.23、听力题:The chemical symbol for cobalt is _______.24、Which bird is known for its colorful feathers and ability to mimic sounds?A. CrowB. ParrotC. SparrowD. Eagle答案:B25、听力题:The _____ of a substance refers to its ability to conduct heat.26、听力题:Reptiles lay ______.27、听力题:I want to be a ______ (teacher) when I grow up.28、听力题:A chemical reaction can produce heat, light, or ______.29、填空题:I have a huge collection of _____ (玩具).30、听力题:The ______ is a common pet that barks.31、听力题:A covalent bond is formed when atoms __________ electrons.32、What do you call the place where you live?A. HouseB. HomeC. CityD. Street答案: BThe girl loves to ________.34、填空题:My favorite ________ is yellow.35、填空题:When light passes through a prism, it creates a _______. (光谱)36、What is the smallest continent?A. AfricaB. AsiaC. AustraliaD. Europe答案:C37、填空题:My brother is a _____ (学生) who loves history.38、填空题:A lizard can regenerate its ______ (尾巴).39、选择题:What color are school buses?A. BlueB. GreenC. YellowD. Red40、听力题:A solution is a homogeneous mixture of two or more ______.41、选择题:What is the color of the sun?A. BlueB. YellowC. RedD. Green42、听力题:The flowers are ___. (growing)43、What do you call a young duck?A. DucklingB. ChickC. GoslingD. Fawn44、听力题:The sun is ______ (shiny) during the day.45、听力题:Granite contains minerals like quartz, feldspar, and ______.46、oasis) is a fertile area in a desert. 填空题:The ____47、听力题:The __________ is a small area that is surrounded by a larger area.48、填空题:I like to __________ (动词) my __________ (玩具名) at night.49、听力题:The chemical formula for acetone is ______.50、填空题:The musician brings joy through _____ (音乐).51、听力题:I have a ___ (dream) of being an astronaut.52、填空题:I enjoy _______ (参加)社会活动。
Formation of Low Mass Stars in Elliptical Galaxy Cooling Flows

a r X i v :a s t r o -p h /9907364v 1 26 J u l 1999FORMATION OF LOW MASS STARS IN ELLIPTICAL GALAXYCOOLING FLOWSWilliam G.Mathews 2and Fabrizio Brighenti 2,32University of California Observatories/Lick Observatory,Board of Studies in Astronomy and Astrophysics,University of California,Santa Cruz,CA95064mathews@3Dipartimento di Astronomia,Universit`a di Bologna,viaZamboni 33,Bologna 40126,Italy brighenti@astbo3.bo.astro.itAbstractThermal X-ray emission from cooling flows in ellip-tical galaxies indicates that ∼1M ⊙of hot (T ∼107K)interstellar gas cools each year,accumulating ∼1010M ⊙over a Hubble time.Paradoxically,optical and radio frequency emission from the cooled gas is lacking,indicating that less than ∼10−3of the cooled gas remains.Many have speculated that the cooled gas has formed into relatively invisible low mass stars,particularly in the context of massive cooling flows in galaxy clusters.We focus here on cooling flows in el-liptical galaxies like NGC 4472where the cooled gas is made visible in emission lines from HII regions ionized and heated (T HII ∼104K)by stellar ultraviolet ra-diation.The low filling factor of HII gas requires that the hot gas cools at ∼106cooling sites within several kpc of the galactic center.HII mass slowly increases at each site at ∼10−6M ⊙yr −1until a neutral core develops.Neutral cores are heated (T HI ∼15K)and ionized (x ∼10−6)by thermal X-rays from the entire interstellar cooling flow.We show that the maximum mass of spherical HI cores that become gravitation-ally unstable is only ∼2M ⊙.No star can exceed this mass and fragmentation of collapsing cores produces stars of even lower mass.By this means we establish with some confidence that the hypothesis of low mass star formation is indeed correct –the IMF is bottom heavy,but may be optically luminous.Slightly more massive stars <∼4.5M ⊙can form near the effective ra-dius (r =8.57kpc in NGC 4472)if sufficient masses of interstellar gas cool there,producing a luminous pop-ulation of intermediate mass stars perhaps with radial orbits that may contribute to the stellar H βindex.The degree of ionization in gravitationally collapsing cores is sufficiently low to allow magnetic fields to dis-connect by ambipolar diffusion.Low mass star forma-tion is very efficient,involving ∼106M ⊙of galactic cold gas at any time,in agreement with observed up-per limits on cold gas mass.We discuss the cooling region surrounding a typical cooling site and show that the total X-ray absorption in cold and cooling gas is much less that that indicated by recent X-ray ing a mass dropout scheme consis-tent with X-ray observations and dynamical mass to light ratios,we plot the global H βsurface brightness profile in NGC 4472and compare it with the smaller contribution from HII gas recently ejected from red giant stars.The lifetime of cooled gas at each cooling site,∼105yrs,is too short to permit dust formation and perhaps also gas phase formation of molecules.Subject headings:galaxies:elliptical and lenticular –stars:formation –galaxies:cooling flows –galaxies:interstellar medium –X-rays:galaxies1.INTRODUCTION AND OVER VIEWStrong X-ray emission from luminous elliptical galaxies is clear evidence that the hot interstellar gas they contain is losing energy.Throughout most of the galactic volume,this loss of energy does not result in lower temperatures since the gas is continuously re-heated by compression in the galactic gravitational potential as it slowly moves inward.In this sense the galactic “cooling flow”is a misnomer.Ultimately,however,in the central regions of the flow the gas density becomes large enough for radiative losses to overwhelm dynamical compression and the gas cools catastrophically.For a typical galactic cooling rate,∼1M ⊙per year,the total amount of gas that cools in a massive elliptical over a Hubble time is large,several 1010M ⊙,a few percent of the total stellar mass.Remarkably,the amount of cold gas observed in el-lipticals,either in atomic or molecular form,is many orders of magnitude less than 1010M ⊙(Bregman,Hogg &Roberts 1992).The mass of central black holes in bright ellipticals is also relatively small,typi-cally less than a few 109M ⊙(Magorrian et al.1998),so the cooled gas cannot be in the holes.Soft X-ray absorption has been observed in some galactic cool-ing flows,indicating masses of cold gas comparable to the predicted value,but the quantitative significance or reality of this absorption is unclear at present.In addition to cold gas deposited by cooling flows,it is possible that additional cold,dusty gas is oc-casionally delivered to the centers of ellipticals as aresult of merging with gas-rich companion galaxies. While this is plausible for some gas-rich ellipticals having dusty clouds or lanes,if merging were an im-portant source of cold gas for all massive ellipticals, the merging rate would need to be carefully regulated in order to maintain the small amount of cold gas observed in normal ellipticals.For many years the standard theoretical explana-tion for this shortage of cooled gas is that it has been consumed in forming low mass stars(e.g.Fabian, Nulsen&Canizares1982;Thomas1986;Cowie& Binney1988;Ferland,Fabian,&Johnstone1994). Such young stars must have low masses since neither luminous OB stars nor Type II supernovae have been observed in normal ellipticals.Explaining the disap-pearance of cooled gas by invoking the poorly under-stood physics of star formation may seem contrived and the low stellar mass hypothesis has led to some ridicule.The fate of cooled gas in coolingflows associated with clusters of galaxies has received most of the ob-servational and theoretical attention because of the spectacularly large inferred mass deposition rates,˙M>∼100M⊙yr−1(Fabian1994).In addition to low mass stars,the apparent soft X-ray intrinsic ab-sorption of N∼1021cm−2indicates that∼1010M⊙of cold gas lies within∼100kpc of the cluster cores. Although enormous,this amount of gas would still be only a few percent of the total cooled gas based on the estimated˙M,so low mass stars are still the preferred endstate for most of the cooled cluster gas (Allen&Fabian1997).However,as with elliptical galaxies,this amount of cold gas should in general be detectable in HI or CO emission but has not(O’Dea et al.1994;Ferland,Fabian&Johnstone1994;Vogt &Donahue1995;Puy,Grenacher,&Jetzer1999), amounting to a colossal discrepancy between expec-tation and reality.In the discussion that follows we revisit the prob-lem of cooled gas from the perspective of the galactic coolingflow in NGC4472,a large,well-observed el-liptical galaxy.For relatively nearby ellipticals the threshold for radio detection is much lower and the ratio of observed to predicted cold gas masses is sim-ilar to that of more distant cluster coolingflows;e.g. H2and HI are undetected in NGC4472with an up-per limit107M⊙(Bregman,Roberts&Giovanelli 1988;Braine,Henkel&Wiklind1988),far below the ∼1010M⊙expected.Nevertheless,the intrinsic soft X-ray absorbing column in NGC4472,3×1021cm−2,and its relatively large covering factor indicates cool gas masses far in excess of the radio upper limit.The coolingflow in NGC4472clearly suffers from a minia-ture version of the same coolingflow problems of dis-tant clusterflows.But because of its proximity,there is a large available body of additional observational information for NGC4472that make it a more appro-priate venue to resolve or constrain theoretical possi-bilities for the fate of cooled gas.But the main advantage afforded by large,rela-tively nearby ellipticals emphasized here is that the cooled gas is heated,ionized and therefore illuminated by ultraviolet radiation from highly evolved galactic stars.We argue that the diffuse optical line emission from HII gas at T∼104K distributed across the cen-tral regions of most or all bright ellipticals is a direct tracer of cooled gas deposited by the hot interstellar gas.A simple analysis of this HII gas leads to the con-clusion that hot phase gas is not cooling into a single large cloud of neutral gas,but is cooling at a large number(∼106)of cooling sites located throughout the central regions of NGC4472.The HII gas pro-vides direct observational support for a distributed mass dropout that has been assumed by many authors in the past based on their interpretation of X-ray sur-face brightness profiles(e.g.Thomas1986;Sarazin &Ashe1989).As the mass of HII increases at each cooling site,a neutral core of very low temperature (T≈10−20K)eventually develops.We show that these neutral cores are very weakly ionized and can undergo gravitational collapse even in the presence of maximum strength magneticfields.Evidently this collapse results in local star formation.Another important conclusion from our study of the ionized gas in NGC4472is that only a very small amount,∼1M⊙,of neutral(or molecular)gas can accumulate at each cooling site before it undergoes gravitational collapse.The small mass of collaps-ing neutral cores is an essential requirement for low mass star formation.Previous studies(e.g.Ferland, Fabian&Johnstone1994)have shown that the gas temperature and Jeans mass are small deep within HI or H2gas irradiated by X-rays in cluster coolingflows, but this does not guarantee that massive stars can-not form.For example,the Jeans mass is often very low in Galactic molecular clouds but these clouds are also the birthsites for massive OB stars.The limited mass of cold gas at each cooling site in NGC4472and other similar ellipticals naturally prohibits stars more massive than about1M⊙from forming.The low mass star formation process we propose is also very efficient:the total mass of cold gas at all cooling sites in NGC4472at any time is very small, consistent with observed upper limit of cold gas(< 107M⊙).NGC4472is an excellent galaxy for study-ing this unique star formation process since very lit-tle alien gas or stars have been recently accreted into NGC4472by a merging process.Dusty,and there-fore accreted,gas is confined to within r<∼0.05r e (van Dokkum&Franx1995;Ferrari et al.1999) Afinal advantage of studying cooled gas in ellipti-cals like NGC4472is that the neutral gas formed in the cores of HII regions with temperatures T∼10K, only lasts for a time,<∼105years,that is too short for dust(and possibly many molecules)to form.Al-though dust and molecules are not required for low mass star formation to proceed,these components have complicated previous discussions of coolingflows in clusters of galaxies where,we assume,the cooling process resembles that in NGC4472.One shortcoming of our presentation–as with those of previous authors–is that we cannot reconcile the observed soft X-ray absorption in NGC4472with the small amount of cold gas indicated by null obser-vations of HI and CO emission.We assume,without much justification,that these contradictory observa-tions will be resolved in favor of the radio observations and that the soft X-ray absorption can be interpreted in another way.HII gas in elliptical galaxies can also arise from stellar mass loss which is ionized by hot central stars (planetary nebulae)and galactic UV radiation.We begin our discussion below with an argument that coolingflow dropout,not stellar mass loss,is likely to be the main contributor to internally-produced HII gas mass observed in nearby ellipticals.Then we dis-cuss an elementary model for the HII gas at a typi-cal interstellar cooling site and infer from this a low globalfilling factor for HII gas within the central re-gion of NGC4472.Next we model the cooling of hot gas from∼107to∼104K with a subsonicflow in pressure equilibrium–thisflow is useful in esti-mating the possible contribution of cooling gas to the X-ray absorption.This is followed by a discussion of the temperature,ionization level and gravitational instability of neutral cores at the centers of HII cool-ing site clouds.This part of our presentation fol-lows rather closely several previous discussions,but serves to illustrate that the more spatially concen-trated radiative transfer in spherical geometry still allows low temperatures and low mass star formation within these cores.We then show from observational considerations that magneticfields play little or no role in inhibiting the compression of interstellar gas as it cools from107to104K and from theoretical con-siderations that even the strongest observationally al-lowed magneticfields are unlikely to inhibit thefinal collapse of neutral cores to stellar densities.At the end of our presentation we discuss the effects of galac-tic gravitational forces and stellar collisions on cooling site clouds.Finally,to stimulate further observations of optical emission lines,we present a surface bright-ness map of NGC4472showing all the major com-ponents:stars,X-ray emitting gas,and HII gas from interstellar dropout and stellar ejecta.2.NGC4472:A PROTOTYPICAL ELLIP-TICAL GALAXYFor quantitative estimates in the following discus-sion,we use a specific galaxy,NGC4472,a massive E2 galaxy associated with the Virgo cluster.NGC4472 has been extensively observed at X-ray frequencies with Einstein HRI(Trinchieri,Fabbiano,&Canizares 1986)and with ROSAT HRI and PSPC(Irwin& Sarazin1996).The radial variations of hot gas density and temperature based on these X-ray data are illus-trated in Brighenti&Mathews(1997a).Although the outer region of the X-ray image of NGC4472is distorted,possibly by ram pressure interaction with ambient Virgo gas,the azimuthally averaged radial variation of electron density in NGC4472is typical of other bright ellipticals(Mathews&Brighenti1998).The most likely region for low mass star forma-tion in NGC4472is the volume within∼0.1r e where r e=8.57kpc is the effective or half-light radius at a distance of17Mpc.The gas that cools in NGC4472 cannot all collect at the origin,nor is it likely that most of the cooling occurs at very large galactic radii where the radiative cooling rate(∝n2)is inefficient. Brighenti&Mathews(1999a)have shown that if all the cooled gas accumulates at or near the very cen-ter of the galaxy,r<∼100pc,the remaining uncooled hot gas there is locally compressed and becomes very hot,but this is not observed.Alternatively,if most of the cooling and low mass star formation occurs in 0.1<∼r/r e<∼1,then the extremely close agreement between the total mass inferred from X-ray data and the known stellar mass in this region would be up-set(Brighenti&Mathews1997a).Finally,there isgood evidence from observed gas abundance and tem-perature gradients that hot interstellar gas isflowing inward within∼3r e through the optically bright re-gions of NGC4472(Brighenti&Mathews1999a),so it is unlikely that a significant number of low mass stars could form at r>∼3r e.It is most interesting therefore that HII optical line emission in Hα+[NII] lines is observed just in the region of NGC4472where low mass star formation is most expected,r<∼0.24r e (Macchetto et al.1996).3.SEVERAL SOURCES OF HII GASIn addition to interstellar gas cooling from the hot phase,cold gas is continuously expelled from stars throughout the galaxy as a result of normal stel-lar evolution.The total rate that mass is supplied by a population of old stars in NGC4472is˙M=α∗(t n)M∗t≈1M⊙yr−1where M∗t=7.26×1011 M⊙is the stellar mass in NGC4472andα∗(t n)≈1.7×10−12yr−1is the specific mass loss rate froma single burst of star formation after t n=13Gyrs (Mathews1989).The supply of gas from stars is com-parable to the rate that gas is observed to cool from the hot phase:˙M=(2µm p/5kT)L x,bol≈2.5M⊙yr−1,where m p is the proton mass and L x≈7.2×1041 ergs s−1is the bolometric X-ray luminosity of NGC 4472at a distance of17Mpc.Several lines of evidence suggest that most of the gas lost from stars in ellipticals is dissipatively and conductively heated and rapidly merges with the gen-eral hot interstellar coolingflow.Gas lost from orbit-ing stars inherits stellar velocities which,when dis-sipated in shocks or by thermal conductivity,equili-brates to the virial temperature of the stellar system, T∼1keV.However,the stellar virial temperature is about30percent lower than that of the more exten-sive dark halo.As coolingflow gas slowlyflows in-ward from the halo into the stellar region,it is cooled by∼0.3keV as it mixes with slightly cooler virial-ized gas ejected from local stars(Brighenti&Math-ews1999a).This produces the positive temperature gradients observed within a few r e.In addition,the iron,silicon and other elements supplied by the stars increases the metal abundance in the hot interstellar gas as it slowlyflows toward the galactic center within ∼r e,producing negative abundance gradients(Mat-sushita1997).These observations indicate that most or all of the gas ejected by stars merges with the hot gas phase.For simplicity,in the following discussion we ignore the HII contribution from stellar mass loss, but return to this question in§11.This is contrary to the hypothesis of Thomas(1986)that gas ejected from stars remains largely neutral and collapses into (very)low mass stars without joining the hot phase.The assimilation of stellar ejecta into the hot in-terstellar gas is greatly accelerated by dynamical and thermal processes resulting from the orbital motion of mass-losing stars through the coolingflow(Mathews 1990).Rayleigh-Taylor and other instabilities shred the ejected gas into many tiny cloudlets,greatly in-creasing the surface area presented to the hot cool-ingflow gas and their rapid dissipation by conductive heating.In addition,neutral clumps of gas expelled from stars always have ionized outer layers which are easily ablated and reformed;this results in a rapid and complete disruptive heating of the clump.In con-trast,cold gas produced as gas cools directly from the hot interstellar phase is necessarily formed in the local rest frame of the coolingflow gas so the violent dy-namical instabilities that accompany stellar mass loss are not expected.After∼106years,however,the denser cooling region may begin to fall in the galac-tic gravitationalfield(see§10),possibly leading to some disruption at the cloud boundary(Malagoli et al.1990;Hattori&Habe1990).Assuming that ra-diative cooling from the hot interstellar phase occurs, as gas cools through HII temperatures it is thermally protected by surrounding gas at intermediate tem-peratures(104<T<107K)where the thermal con-ductivity is very low.The global kinematics of the two HII gas components of internal origin are quite different.HII regions produced by stellar ejecta will initially tend to mimic local random and systematic stellar motions while HII gas arising from cooling gas will initially share the velocity of local hot gas.A third source of HII gas in ellipticals are the ion-ized parts of gas acquired in recent merging events such as the small dusty clouds within∼0.05r e of the center of NGC4472(van Dokkum&Franx1995). This gas is spatially disorganized and is dynamically unrelaxed.Dust is another clue of its external ori-gin since gas formed by cooling from the hot phase should be nearly dust-free due to sputtering(Draine &Salpeter1979;Tsai&Mathews1995;1996)and may not have time to grow dust in the gas phase(§7).Our interest here is with the HII component pro-duced as gas cools from the hot phase and we assume that this component dominates the optical line emis-sion in NGC4472.4.THE INVERSE HII REGIONA small cloud of HII gas that has cooled from the hot interstellar medium is photoionized by stellar UV radiation arriving at its outer boundary;this is the inverse geometry of normal HII regions ionized by a central star.We suppose that the HII cloud is spheri-cal and that the electron density n e and temperature T=104K are uniform throughout the HII gas.The spatial uniformity of the HII density is essentially un-affected by small local gravitationalfields due to inter-nal stars,the neutral core in the cloud if one exists, or the HII gas itself.The mass of these HII clouds located at the centers of local cooling sites slowly in-creases with time.Thefirst step in understanding the evolution of HII clouds is to determine the maximum size and mass that can be ionized by stellar UV in the central regions of NGC4472.This size depends on the HII electron density and the mean intensity of galactic UV starlight J uv(r).The intensity of ionizing radiation can be deter-mined by an appropriate integral over the galactic stellar distribution.For this we assume a de Vau-couleurs stellar distribution similar to that in NGC 4472,with an effective radius of r e=8.57kpc and an outer maximum stellar radius of25r e.The stellar density and mass are given to a good approximation byρ∗=ρo(b4r/r e)−0.855exp[−(b4r/r e)1/4]andM(r)=M oγ(8.58,[b4r/r e]1/4)where b=7.66925andγ(a,z)is the incomplete gamma function(Mellier&Mathez1987).If the de Vaucouleurs distribution extends to in-finity,the total mass would be M t=M oΓ(8.58)= 1.6582×104M o where M o=16πρo(r e/b4)3.It is nat-ural to normalize the density coefficientρo tofit the de Vaucouleurs distribution for NGC4472in the re-gion0.1<∼r/r e<∼1where the X-ray and stellar mass determinations agree,i.e.ρo=3.80×10−18gm/cm3. When the stellar distribution is truncated at25r e the total mass6.97×107M⊙is only about1percent less than an infinite stellar distribution having the same ρo.At every radiusx=r/r ein the de Vaucouleurs distribution the mean stellar column density˜J can be found by integrating over solid angle,˜J=1n2αB.The density of HII gas isρ=nMfρwhere fρ= 5µ/(2+µ)=1.20,assumingµ=0.61for the molec-ular weight.The total mass of the Stromgren sphereism s=4n5α3Bwhere m p is the proton mass.Some imprecision is ex-pected since we have ignored those ionizing photons that pass through the HII cloud unabsorbed.How-ever,calculations of the transfer of ionizing radiation in the inverse HII region indicate that Equation(1) is accurate to<∼5percent.Figure2illustrates the radial variation of electron density n e=(ρ/m p)(2+µ)/5µin HII gas(solid line) with galactic radius in NGC4472and the correspond-ing local inverse Stromgren radius r s(long dashed line).Within the radius where Hαis observed in NGC4472,x=r/r e<∼0.24wefind r s≈0.3−0.8 pc,n e≈20−90cm−3,and the mass of a typical Stromgren cloud is m s≈2M⊙.The radial col-umn density in an HII Stromgren cloud is typically N s=n e r s≈1.2×1020cm−2.Hot gas is assumed to be cooling at numerous sites throughout this central region of NGC4472and the cooling is made visible by optical line emission from the HII regions.The mass of any particular HII cloud increases slowly with time,supplied by local cool-ing from the hot gas phase or by dissipative merging of clouds.Presumably,new HII clouds are contin-uously forming from the cooling interstellar gas at newly-formed cooling sites and old sites and associ-ated clouds are disappearing.But we suppose that the mean age of cooling sites is long compared to the time required for typical HII clouds to reach the Stromgren mass;in this case the average cloud can be approximated with Stromgren parameters.Notice also that r s≪r so that even the largest HII clouds are very small compared to their distance to the cen-ter of the galaxy.5.GEOMETRY OF HII AND COLD GASWe propose that most of the extended Balmer line emission in ellipticals arises from a multitude of HII clouds at or near their Stromgren radii.If so,the to-tal volume within clouds occupies only a tiny fraction f F of the galactic volume within the Hα-emitting re-gion of NGC4472,r<∼2kpc.Thefilling factor f F can be estimated by comparing the total volume of HII required to produce the observed Balmer line lu-minosity to the apparent volume from which optical line emission is observed.In most optical observations,such as those of Mac-chetto et al(1996),Hα(6562˚A)is blended with two nearby[NII]lines at6584and6548˚A.The totalflux observed by Macchetto et al.(1996)in all three lines is F lines=17.30×10−14ergs cm−2 s−1.Observations at higher resolution reveal that the F([NII]6584)/F(Hα)≈1.38and F(6584)/F(6548)=bining all these ratios,and adopting CaseB conditions F Hα/F Hβ=2.86,the Hβflux from NGC4472is F Hβ=2.13×10−14ergs cm−2s−1 and the total luminosity from all HII emission is L Hβ=4πD2F Hβ=7.34×1038erg s−1,assuming a distance of D=17Mpc to Virgo.How many dust-free Stromgren clouds are required to produce this total luminosity?The Hβluminosity of a single Stromgren cloud isℓβ,s=n2eǫβ(4π/3)r3s=1.5×1031n2e r3spc ergs s−1 whereǫβ=1.0×10−25erg cm3s−1is the Hβemis-sivity at T=104K.For typical values of n e and r spc(in parsecs)in the central galaxy x<∼0.25,ℓβ≈3−7.5×1033ergs s−1.Therefore,about N cl= 105−106Stromgren clouds are required to account for the Balmer line luminosity observed.The HIIfill-ing factor is found by comparing the volume of all HII gas V cl=L Hβ/ n e 2ǫHβ=2.5×1060cm3(assuming n e =50cm−3)with the total volume of the Hβ-emitting region,V tot=(4/3)π(0.24r e)3=1.1×1066 cm−3.Thefilling factor of HII gas f F=2×10−6is very small,consistent with our proposition that HII emission arises from many small clouds and with ear-lier estimates of f F(Baum1992).If∼1M⊙of hot gas cools each year in NGC4472,then the mass of each cloud will grow quite slowly,∼10−6−10−5M⊙yr−1,requiring t s∼2×105−2×106years to form a typical Stromgren cloud.The total mass of all the HII emitting gas in NGC4472is M II= n Mf F V cl≈1.2×105M⊙,similar to values in the literature but here evaluated using a consistent physical model.A smallfilling factor also implies that each HII cloud is exposed to the unabsorbed stellar UV emis-sion from the entire galaxy,provided the cloud sys-tem is approximately spherical.The“optical depth”for intersecting a Stromgren cloud across the opti-cal line-emitting region within r t=0.24r e isτ=πr2s r t N cl/V tot≈0.006−0.06.Sinceτ≪1the clouds do not shadow each other.In realityτcould be larger (i)if the typical cloud crossection is much less than πr2s(τ∝r−1s)or(ii)if the cloud system were not spherical;a disk-like configuration could result from galactic rotation.In any case,the assumption thateach HII cloud is exposed to the full,unabsorbed stel-lar UV emission is likely to be a reasonably good ap-proximation.Combining previous results,the average column depth that HII gas presents to X-radiation through-out the galactic core,N∼N sτ∼1018−1019cm−2,is much less than the value N∼3×1021cm−2that best fits the observed X-ray continuum(Buote1999).The size that an HII cloud presents to absorbing X-rays is larger than r s since we have ignored the extended cooling region around each cloud with temperatures between106and104K in which X-rays can still be absorbed.This assumption will be justified below.The total mass of HI or H2gas observed in the cen-tral regions of NGC4472,M cold<107M⊙,is another potential source of X-ray absorption.If this mass of cold gas were arranged in a disk of thickness of the X-ray absorbing column N=3×1021cm−2,located in the galactic core and oriented with its symmetry axis along the line of sight,it would have a radius<370 pc,somewhat larger than the faint patch of dust ob-served by van Dokkum&Franx(1996).However a cloud of size370pc obscures only∼0.007of the total X-ray luminosity of NGC4472and would therefore produce negligible X-ray absorption.The true X-ray absorption is very probably much less than3×1021 cm−2.The observation of NGC4472by Buote us-ing the∼4’beam of ASCA also included the nearby gas-rich dwarf irregular galaxy UGC7636(Irwin& Sarazin1996;Irwin,Frayer&Sarazin1997)which may be the source of the X-ray absorbing column at-tributed to NGC4472if its covering factor is suffi-ciently large.Although it seems likely that interstellar magnetic fields are important in the centers of ellipticals(Math-ews&Brighenti1997;Godon,Soker&White1998), it is remarkable that the observed optical line emis-sion does not indicate strongfields in the HII gas. Typical HII densities in bright ellipticals determined from[SII]6716/6731line ratios are∼100−200cm−3 (Heckman et al.1989;Donahue&Voit1997),similar to(or even a bit larger than)the values found here for NGC4472(Figure2).(Unfortunately,we have been unable tofind a determination of the HII density spe-cific to NGC4472.)This suggests that the HII gas density is not being diluted by magnetic support,i.e. B2/8π<2nkT or B<70µG in the HII gas.HII densities of∼100are also supported by comparing the ionization parameter U=n iph/n e for pressure equilibrium HII gas in NGC4472(short dashed line in Figure2)with values that characterize the entire observed line spectrum.Within∼r e in NGC4472log U≈−3.3,very similar to values of U required to reproduce LINER type spectrum typically observed in ellipticals(e.g.Johnstone&Fabian1988);thisprovides an independent check on our HII gas density and J near the center of NGC4472.The apparent absence of magnetic support in the HII gas is interesting since the hot phase gas is re-quired to havefields of at least severalµG at largegalactic radii to explain Faraday depolarization of ra-dio sources and distant quasars(Garrington&Con-way1991).Interstellarfields>∼1µG can be generated in a natural way by stellar seedfields and turbulent dynamo action in the hot gas(Mathews&Brighenti1997).As the gas density increases by∼1000when it cools from the hot phase to HII temperatures,a field of1µG would grow to100µG ifflux is con-served,B∝ρ2/3.The initialfield in the hot gas in r<∼0.24r e would need to be surprisingly small, <∼0.7µG,to evolve into the rather smallfields allowed in HII clouds,B<∼70µG,implied by typical electron densities.Small HIIfields can be understood if localcooling sites form in interstellar regions having lower than averagefields;in pressure balance,lowerfields require higher hot gas densities which cool preferen-tially.Alternatively,it is possible thatfield reconnec-tion has been very efficient during cooling,implying a disorganizedfield and considerable stirring motion during the cooling process.6.COOLING SITE GAS DYNAMICSCooling sites in the hot interstellar gas are initiated in regions of low entropy(i.e.low temperature,high density)which cool preferentially by radiative losses. Entropyfluctuations can be generated by a variety of complex events:stellar mass loss,occasional Type Ia supernovae,sporadic mergers with other nearby (dwarf)galaxies,and differential SNII heating events and outflows that occurred in pregalactic condensa-tions.Due to the complicated nature of these inter-actions,it is difficult to predict the amplitudes and mass scales of the entropy inhomogeneities so the de-tailed nature of the initial cooling process remains unclear.However,once cooling commences,the gas flow toward the cooling site may evolve toward a sim-ple profile provided entropyfluctuations in the hot gas are not too severe over theflow region.We now describe such a model for cooling site。
The Formation of Massive Stars from Turbulent Cores

a r X i v :a s t r o -p h /0206037v 2 13 N o v 2002astro-ph/0206037Preprint typeset using L A T E X style emulateapj v.14/09/00THE FORMATION OF MASSIVE STARS FROM TURBULENT CORESChristopher F.McKee 1and Jonathan C.Tan 21.Departments of Physics &Astronomy,University of California,Berkeley,CA 94720,USA.2.Princeton University Observatory,Princeton,NJ 08544,USA.Submitted 31st May 2002,Accepted 12th November 2002ABSTRACTObservations indicate that massive stars in the Galaxy form in regions of very high surface density,Σ∼1g cm −2.Clusters containing massive stars and globular clusters have a column density comparable to this.The total pressure in clouds of such a column density is P/k ∼108−109K cm −3,far greater than that in the diffuse interstellar medium or the average in giant molecular clouds.Observations show that massive star-forming regions are supersonically turbulent,and we show that the molecular cores out of which individual massive stars form are as well.The protostellar accretion rate in such a core is approximately equal to the instantaneous mass of the star divided by the free-fall time of the gas that is accreting onto the star (Stahler,Shu,&Taam 1980).The star-formation time in this Turbulent Core model for massive star formation is several times the mean free-fall time of the core out of which the star forms,but is about equal to that of the region in which the core is embedded.The high densities in regions of massive star formation lead to typical time scales for the formation of a massive star of about 105yr.The corresponding accretion rate is high enough to overcome the radiation pressure due to the luminosity of the star.For the typical case we consider,in which the cores out of which the stars form have a density structure ρ∝r −1.5,the protostellar accretion rate grows with time as ˙m ∗∝t .We present a new calculation of the evolution of the radius of a protostar and determine the protostellar accretion luminosity.At the high accretion rates that are typical in regions of massive star formation,protostars join the main sequence at about 20M ⊙.We apply these results to predict the properties of protostars thought to be powering several observed hot molecular cores,including the Orion hot core and W3(H 2O).In the Appendixes,we discuss the pressure in molecular clouds and we argue that “logatropic”models for molecular clouds are incompatible with observation.Subject headings:hydrodynamics —ISM:clouds —stars:formation1.introductionMassive stars are fundamental in the evolution of galaxies since they produce the heavy elements,energize the interstellar medium,and possibly regulate the rate of star formation.Remarkably little is known about how massive stars form,however:the problem is difficult observationally because massive star formation occurs in distant,highly obscured regions,and it is difficult theoretically because of the many processes that must be included.Even such a basic parameter as the time it takes to form a massive star has been uncertain.This time scale,or equivalently,the protostellar accretion rate,affects the luminosity of the protostar (particularly for masses m ∗ 10M ⊙(Palla &Stahler 1992)and the strength of protostellar outflows (Richer et al.2000).Arguments based on extrapolating from low-mass star formation lead to formation times t ∗f >106yr,a significant fraction of the main-sequence lifetime of the star (Bernasconi &Maeder 1996;McLaughlin &Pudritz 1997,hereafter MP97;Stahler et al.2000).Comparison with observations of hot molecular cores (Osorio,Lizano &D’Alessio 1999;Nakano et al.2000)suggest substantially smaller time scales,t ∗f 105yr.An analysis based on observations of protostellar outflows suggests t ∗f ∼3×105yr (Behrend &Maeder 2001).The small (∼1×106yr)spread in ages of stars in the Orion Nebula Cluster (Palla &Stahler 1999),where there is no evidence that the higher mass stars have formed systematically later compared to the lower-mass population,sets an upper limit of t ∗f 1Myr in this case.What has been lacking is an adequate understanding of how the formation time is governed by the conditions in the gas out of which the star forms.Our understanding of low-mass star formation is on a far better footing,since it has received much more observational and theoretical attention (Shu,Adams &Lizano 1987).Low-mass stars form by accreting gas from a molecular “core”in which gravity overcomes thermal and nonthermal (magnetic and turbulent)pressure gradients.Shu (1977)considered the collapse of a singular isothermal sphere,finding˙m ∗=0.975c 3th 20K3/2M ⊙yr −1,(1)where c th is the isothermal sound speed and the numerical evaluation assumes n He =0.2n H 2.Observed temperatures of 10to 20K in regions of low-mass star formation imply accretion rates of about 10−6to 10−5M ⊙yr −1,consistent with the inferred values of t ∗f for low-mass stars in these regions (Lada 1999).There are two difficulties in extending this theory to high-mass stars.The first,discussed in some detail by Stahler et al.(2000),is that the predicted accretion rate depends only on the temperature of the gas.Once massive stars form,12the gas may be heated to temperatures∼50to100K,but thefirst massive stars that form in a region will emerge from gas at10−20K and will have low accretion rates and correspondingly long formation times.The second difficulty is feedback from the massive stars.Since the Kelvin-Helmholtz contraction time is less than the accretion time for massive stars,they evolve along the main sequence while accreting.Massive protostars are thus very luminous,and it has been suggested that the radiation pressure and ionization they produce can halt the accretion and determine the upper limit of the stellar mass function(Larson&Starrfield1971;Kahn1974;Wolfire&Cassinelli1987;Jijina&Adams1996).This feedback is so strong that it is impossible to form stars as massive as those observed if the accretion is assumed to be spherical,and the discrepancy grows as the accretion rate is reduced.These considerations have motivated the radical suggestion that massive stars form via the coalescence of low-mass stars in order to achieve a more rapid build up of the final stellar mass(Bonnell,Bate&Zinnecker1998).Recently,McKee&Tan(2002;hereafter MT)addressed the accretion-rate problem.Thefirst step in resolving the problem of the apparently low accretion rates is to realize that,although equation(1)was derived for an isothermal gas, it should hold approximately when nonthermal support due to magneticfields and turbulence is included as well(Stahler, Shu&Taam1980;Shu et al.1987).Observed cores have turbulent motions that increase systematically with radius (Larson1981;Caselli&Myers1995),and this leads to an increase in the accretion rate with time(Myers&Fuller1992; Caselli&Myers1995;MP97).Larger signal speeds allow for the hydrostatic support of denser gas cores,which then have shorter free-fall times and thus greater accretion rates once they become unstable.We term this the turbulent core model for massive star formation.The second step in resolving the accretion rate problem is the recognition that massive stars form in regions of very high pressure and density.MT showed that for typical pressures in regions of massive star formation(Plume et al.1997), stars form in a time of order105yr.This result for the time scale is somewhat longer than that of Osorio et al.(1999), who inferred stellar masses and accretion rates by comparing calculated spectra with observations.The purpose of this paper is several fold.First,we determine the relation between the pressure in a molecular cloud and its surface densityΣ.We show that observed regions of massive star formation,both Galactic and extragalactic, haveΣ∼1g cm−2,corresponding to mean pressures¯P/k∼4×108K cm−3.Second,we extend the self-similar theory presented by MT to allow for magneticfields and for a thermally supported core.Finally,we determine the radius and luminosity of accreting protostars,and use observed hot core luminosities to predict accretion rates and masses of several nearby massive protostars.2.the pressure and surface density of regions of high-mass star formationA cloud in hydrostatic equilibrium has a total pressure that is proportional to the square of the surface density,P∼GΣ2, whereΣ≡M(R)/πR2(e.g.,Elmegreen1989).The Appendix contains a detailed discussion of this relation,and for our fiducial case—including the effects of magneticfields and allowing for embedded stars—wefind that the mean pressure in a cloud is typically¯P≃0.88GΣ2.What is the value of the surface density in regions of current and past massive star formation?Regions of high-mass star formation studied by Plume et al.(1997)are characterized by virial masses M vir∼3800M⊙and radii R∼0.5pc.As discussed in the Appendix,the virial mass M vir is related to the actual mass M by the virial parameter,αvir=M vir/M.The virial parameter is about1.3−1.4for GMCs,whereas it is quite close to unity for cores that are actively forming stars.Although there is no direct determination ofαvir for massive-star forming clumps,Plume et al.(1997)regard the virial mass as the most accurate estimate for the mass(αvir≃1).Since this is consistent with the results for low-mass star formation discussed in the Appendix,we shall adoptαvir=1for our numerical estimates.The mean column density in the clumps studied by Plume et al.is thenΣcl≃1g cm−2=4800M⊙pc−2;the corresponding visual extinction is A V=(N H/2×1021cm−2)mag=214Σcl mag.From maps of a larger number of sources,Shirley et al(2002)find a median surface density(which they regard as a better characterization of their sample)from the virial theorem ofΣcl=0.60g cm−2.Mueller et al.(2002a,b)have determined the gas masses of a number of these sources from observations of the350µm dust ing the Ossenkopf and Henning(1994)opacities,theyfind gas masses that are on average3.4times lower than the virial mass.The mean surface densityΣcl=0.19±0.12g cm−2for the sources they were able to model in detail,while their total sample gives a higher value,Σcl=0.73±1.7g cm−2.Had they used the Pollack et al.(1994)opacities instead,the inferred masses and surface densities would have been larger by a factor of about2.5.These data show that there is a considerable dispersion in the mean surface density of the sources.These column densities are far greater than those associated with GMCs(0.035g cm−2—Solomon et al.1987)or regions of low mass star formation(the average column density in the C18O cores in Taurus is0.032g cm−2—Onishi et al.1996).We now compare the column density observed in regions of massive star formation with the surface density of stars in both Galactic and extragalactic star clusters.Before doing so,we note that there are several effects that can alter the relation between the observed stellar surface density and that of the molecular gas out of which the cluster formed.First, since not all the gas in the cloud goes into stars,some of the gas is ejected,causing the cluster to expand.For example, if the star formation were50%efficient and the gas were lost slowly,then the cluster would expand by approximately a factor of two from its initial size(Hills1980).Including the mass lost from the cluster implies that the initial surface density of gas would have been8times greater than thefinal surface density of stars.On the other hand,if magnetic fields were important in the support of the cloud,then the stellar velocity dispersion would have been less than the virial velocity(Patel&Pudritz1994),which would lead to an increase in the stellar surface density.If the size of the cluster is estimated from observations of the massive stars,then the resulting surface density will be larger than that of the total3star cluster insofar as the massive stars form preferentially in the inner regions of clusters (Bonnell &Davies 1998).This effect would be amplified if the cluster is old enough to have experienced significant mass segregation.Overall,we can expect the stellar surface density to be at best within a factor of a few of the initial gas surface density.First,consider Galactic star clusters.The virial mass of the Orion Nebula Cluster (ONC),including both stars and gas,is about 4650M ⊙(averaging the two models of Hillenbrand &Hartmann 1998).About half the mass of the cluster is within 0.8pc,giving Σ≃0.24g cm −2.The Arches is a large cluster near the Galactic Center that contains many massive stars (Figer et al.1999).Kim et al.(2000)have analyzed this cluster,taking into account the strong tidal effects experienced by a cluster so near the Galactic Center.They conclude that the initial mass of the cluster was about 2×104M ⊙.From their results,we estimate that the half-mass radius is about 0.4pc,giving Σ≃4g cm −2.Figer et al.infer a very flat IMF for the Arches (d N ∗/d ln m ∗∝m −0.6∗vs.m −1.35∗for the Salpeter IMF).If the actual IMF of the Arches is closer to the Salpeter value and the heavy extinction has led to an underestimate of the number of low-mass stars,then the surface density would be larger than this estimate.Globular clusters in the Galaxy no longer contain any massive stars,but they presumably did in their youth.There is a considerable dispersion in the properties of globular clusters;here we estimate the surface density of a typical one.The typical mass of a Galactic globular is about 2×105M ⊙(inferred from Binney &Merrifield 1998).The data summarized by van den Bergh,Morbey,&Pazder (1991)give a median half-light radius for Galactic globulars of R proj ,1/2=2.6pc,where the subscript “proj”emphasizes that this is a projected radius.If the light traces the mass,then the spherical half-mass radius is R 1/2≃1.31R proj ,1/2(Spitzer 1987),which is 3.4pc.The surface density inside the spherical half-massradius is then 0.5M/(πR 21/2)≃0.6g cm−3.The corresponding value of the pressure,∼3×108K cm −3,is similar to the estimates of Elmegreen &Efremov (1997)based on the same line of reasoning.It has been suggested that super-star clusters (SSCs)in external galaxies are globular clusters in the process of formation (e.g.,Ho &Filippenko 1996).The SSCs NGC 1569–A1and A2each have masses of about 4×105M ⊙(Gilbert &Graham 2001)and projected half-light radii of about 1.7pc (DeMarchi et al 1997),corresponding to Σ≃2.7g cm −2.A particularly dramatic SSC is the one observed by Turner,Beck,&Ho (2000)in NGC 5253:they estimate that a star cluster with total ionizing luminosity of ∼4×1052s −1is confined within a region about 1×2pc in diameter.For zero age main sequence models (Schaerer &de Koter 1997),this corresponds to a total stellar mass of 0.6−1.5×106M ⊙,for Salpeter extrapolation to 1and 0.1M ⊙,respectively.If we take the projected half-mass radius to be 0.75pc,then Σ≃20−50g cm −2.These results are collected in Table 1.It is striking that both the molecular regions where massive stars are forming and the star clusters that formed from them all have surface densities Σ∼1g cm −2(with the exception of the cluster in NGC 5253,which is somewhat larger).We first characterize the properties of molecular gas in such an environment and then determine the time scale for star formation.3.self-similar cores and clumpsMolecular clouds are inhomogeneous on a wide range of scales (Williams,Blitz,&McKee 2000),and numerical sim-ulations show that this is a natural outcome of supersonic turbulence,both with and without magnetic fields (Vazquez-Semadeni et al.2000).Following Williams et al.,we define a star-forming clump as a massive region of molecular gas out of which a star cluster is forming;a core is a region of molecular gas that will form a single star (or multiple star systemTable 1Characteristic Surface Densities of Regions of High-Mass Star FormationObject [Ref.]aM (M ⊙)R 1/2(pc)Σ(g cm −2)¯Pcl /k (K cm −3)aReferences:(1)Plume et al.(1997);(2)Hillenbrand &Hartmann (1998);(3)Figer et al.(1999);(4)Kim et al.(2000);(5)Binney &Merrifield (1998);(6)van den Bergh et al.(1991);(7)Gilbert &Graham (2001);(8)DeMarchi et al.(1997);(9)Turner,Beck,&Ho (2000)b Virial mass estimatescThe half-mass radius is not well-defined for the Plume et al.(1997)clouds,since the mass distribution on larger scales is not known.We therefore evaluate Σ=M/πR 2using the typical radius and virial mass that they observe.dExtrapolation from inferred LyC luminosity of H II region based on Salpeter IMF with a lower mass limit m ℓ=1,0.1M ⊙.4such as a binary).Star-forming clumps are observed to be approximately gravitationally bound (Bertoldi &McKee 1992),whereas cores are necessarily gravitationally bound.Our basic assumption is that a star-forming clump and the cores embedded within it are each part of a self-similar,self-gravitating turbulent structure on all scales above that of the thermal Jeans mass (MT),an assumption that appears to be in accord with observation (Williams et al.2000).Except on small scales,the turbulence is therefore necessarily supersonic,and we shall show that this is self-consistent below.We further assume that the structure is approximately stationary in time (i.e.,it is not in a state of dynamical collapse or expansion);it follows that the star-forming clump and most of the cores within it are in approximate hydrostatic equilibrium.However,some of the cores are gravitationally unstable and are collapsing on a dynamical time scale—these are the cores that are in the process of forming stars.Finally,we assume that the clump and the cores within it are approximately spherical.As shown by Bertoldi &McKee (1992),deviations from sphericity of a factor of a few are readily accounted for;in any case,observations of high-mass star forming clumps show that they are approximately spherical (Shirley et al.2002).This model for regions of massive star formation is necessarily highly approximate.Regions of massive star formation are observed to have highly supersonic velocity dispersions (Plume et al.1997)and are presumably rge amplitude density fluctuations in such regions continually form,grow,and dissipate.Since we have assumed that a clump has a lifetime that is at least several times greater than its free-fall time,it follows that only a small fraction of the mass of the clump can be contained in density fluctuations that are gravitationally bound and are undergoing gravitational collapse—i.e.,in cores that are in the process of star formation.The same conclusion applies to massive cores,which are themselves turbulent;if on the contrary,most of the mass of a core were in density fluctuations that were undergoing gravitational collapse,then the core would form a cluster of low-mass stars instead of a massive star.Calculations of turbulent,self-gravitating clouds indicate that indeed only a small fraction of the mass is gravitationally unstable at any given time (e.g.,Vazquez-Semadeni,Ballesteros-Paredes,&Klessen 2002).In this section we shall work out the properties of this self-similar structure,leaving consideration of the effects of thermal pressure to §5.The results in this section apply equally well to an individual core or to a clump.For clumps,we allow for the possibility that only a fraction f g of the mass of the clump is gaseous,with the rest in stars.If stars are present,we assume that their spatial distribution is identical to that of the gas.We do not attempt to evaluate the mass distribution of the cores,which is related to the initial mass function (IMF)of the stars that form;instead,this mass distribution is assumed to be such that it leads to the observed IMF.Presumably this mass distribution is determined by the characteristics of the turbulence in the star-forming clump (e.g.,Henriksen &Turner 1984;Elmegreen &Falgarone 1996).In a self-similar,spherical medium,the density and pressure each have a power-law dependence on radius,ρ∝r −k ρand P ∝r −k P .It follows that the sphere is a polytrope with P ∝ργp .In hydrostatic equilibrium we have (MP96;McKee &Holliman 1999)k ρ=22−γp =2(k ρ−1).(2)Let c ≡(P/ρ)1/2be the effective sound speed.The equation of hydrostatic equilibrium then givesM =k P c 2r2πGr 2,(4)whereA =(3−k ρ)(k ρ−1)f g =γp (4−3γp )f gAk 2P1/8,(6)r =AGM 22πG 3M 21/4.(8)In this subsection,A can refer either to clumps,which generally have f g <1,or to cores,which are assumed to havef g =1;in the remainder of the paper,however,we shall use A to refer only to cores.5 Observe that the only dependence of these properties on the polytropic indexγp,or equivalently,on the power-law indices k P and kρ,is in the factors A and k P,which are raised to relatively low powers.Studies of low-mass star formation often adopt the singular isothermal sphere as a model(Shu1977),which appears to be approximately consistent with observation(Andre,Ward-Thompson,&Barsony2000).Such a model hasγp=1,so that kρ=k P=2and A=1. No data are available on the structure of cores that are forming very massive stars.MP96and MP97have discussed an alternative model,the logatrope,in which the pressure varies as the logarithm of the density,but as discussed in Appendix B,this model is not physically realistic.Observations of molecular clouds show that the velocity dispersionσincreases outwards(Larson1981).Since we expect thatσ∝c,and since c2=P/ρ∝r2−kρ,this implies that kρ<2andγp<1 (Maloney1988).(Ifγp is written asγp=1+1/n,this corresponds to a negative index n;such a polytrope is therefore referred to as a“negative index polytrope.”)For“high-mass”cores in Orion,Caselli&Myers(1995)find kρ≃1.45with a dispersion of±0.2.According to van der Tak et al.(2000),the clumps in which high-mass cores are embedded have values of kρranging from1to2,centered around1.5.1Beuther et al.(2002)find kρ≃1.6±0.5for their sample of massive star-forming regions,while Mueller et al.(2002b)find kρ≃1.8±0.4for a sample of31sources(see Evans et al.2002for a review).For our numerical estimates of the properties of both high-mass cores and clumps,we shall adopt kρ=1.5,which corresponds toγp=2/3,k P=1, and A=3/4.For comparison,the value kρ=1.8found by Mueller et al.(2002b)corresponds toγp=8/9,k P=1.6,and A=0.96.The value ofγp determines the structure of a polytrope.As discussed by McKee&Holliman(1999),the stability of the polytrope depends on the adiabatic indexγas well as onγp.If the polytrope is isentropic(γ=γp),then the maximum density contrast between the center and edge of a stable molecular cloud is quite limited(≤14forγp≤1).On the other hand,for non-isentropic clouds,an infinite density contrast is possible in a stable cloud providedγexceeds a critical value;for ourfiducial case ofγp=2/3,this critical value ofγis unity.The value ofγfor turbulent magneticfields is 4/3,significantly above this critical value.The effective value ofγfor the turbulent motions in molecular clouds is not known;this is a thorny problem,since the motions are inferred to damp very rapidly(Vazquez-Semadeni et al.2000). The large density contrasts observed in molecular clouds suggest that,insofar as polytropic models are applicable,the effective value ofγis large enough to render the clouds stable against dynamical collapse.3.1.The Mean Pressure in ClumpsThe mean pressure in a clump is¯P cl≡(1/V cl) P dV.As discussed in Appendix A,this pressure is proportional to GΣ2cl,so we write¯Pcl≡φ¯P GΣ2cl,(9) whereφ¯P is a numerical factor of order unity.From equation(A5),wefindφ¯P= 3π63.2.Properties of CoresWe are now in a position to express the properties of a core out of which a massive star forms in terms of the pressure of the clump in which it is embedded and the mass of the star that will form.The key point is that,on average,the pressure at the surface of a core will be the same as the ambient pressure in the clump there,P s,core=P cl(r).We define the parameterφP,core as the ratio of this pressure to the mean pressure in the clump.Since the clump is self-similar by assumption,we haveφP,core≡P cl(r)3 r3 M(r)3[M(r)/M cl]−2;for ourfiducial case of kρ=1.5,we haveφP,core=2φ1/2B=1.61 φP,coreφ¯P30M⊙ 1/4Σ1/4cl km s−1,(17)→1.27 m∗fk2 P ǫ2coreφP,coreφ¯P 1/4m∗f30M⊙ 1/2Σ−1/2cl pc.(20)7Finally,the density of H atoms at the surface of a core is (eq.8)n H ,s =1.10×106Ak 2P ǫ2core φ3P,core φ3¯P 1/4 m ∗f 30M ⊙−1/2Σ3/2clcm −3.(22)Note that equations for the clump velocity dispersion,radius,and density can be obtained from equations (17),(19),and (21)by making the replacements m ∗f /ǫcore →M cl and A →A core f g ;by setting φP,core =(3−k P )/3so as to set the pressure equal to its value at the surface of the clump (eq.12);and by setting αvir φB =αSPS so as to ensure that the clump is in hydrostatic equilibrium (eq.A8).How are the core properties related to those of the clump in which it is embedded?Consider first the surface densityof a core,Σcore ≡(m ∗f /ǫcore )/πR 2core .Equations (7)and (13)giveΣcoreAπφP,core φ¯P 1/2→1.22.(23)The cores thus have a column density comparable to that of the clump as a whole.Note that Σcl includes the mass of the stars in the clump;the ratio of Σcore to the column density of gas in the clump is larger by a factor 1/f g .One of the important questions that must be addressed in any model for the formation of dense star clusters is how a large number of protostellar cores can fit into the small volume of the clump in which the stars are forming.This issue is automatically solved in our model since the cores are much denser than the clumps.We compare the mean density in a core,¯ρcore ,to the mean gas density in the central part of a clump,ρcl =φρ,core ¯ρcl =φρ,core 3f g M/(4πR 3),where ¯ρcl isthe mean overall clump density and the factor φρ,core =(3−k ρ)/3·(r/R cl )−k ρallows for central concentration.As before we assume the typical high-mass star-forming core is located at r =R cl /3(Bonnell &Davies 1998),so that for k ρ=1.5,φρ,core ≃2.6.With the aid of equations (7)and (13)we find¯ρcore φρ,core 3φP,core f 1/4gǫcore M4000M ⊙1/230M ⊙r 3R t =GM coreR core =f 1/4gM core1/6,(27)is then somewhat greater than unity,as required.Note that this ratio increases weakly with the clump mass,so that the omission of effects associated with tidal gravitational fields becomes a better approximation for more massive star-forming clumps.4.self-similar high-mass star formationHaving determined the properties of massive star-forming clumps and the cores embedded within them,we now consider the rate at which a protostar grows once the core from which it is forming becomes gravitationally unstable.Dimensional arguments indicate that the accretion rate should be ˙m ∗∼M core /t ff,where t ff=(3π/32Gρ)1/2is the free-fall time (Stahler,Shu,&Taam 1980;see also Shu et al 1987).Following MT,we write this as˙m ∗=φ∗m ∗π√G,(29)8which is the generalization of equation(1)to the non-isothermal case.Equation(28)could be violated in the sense thatφ∗is much greater than unity only in the unlikely case that the star forms from a coherent spherical implosion;if the star formation is triggered by an external increase in pressure,φ∗could be increased somewhat,but deviations from spherical symmetry in the triggering impulse and in the protostellar core will generally preventφ∗from becoming too large.The equation could be violated in the opposite sense thatφ∗is much less than unity if the core is magnetically dominated,so that collapse could not begin until the magneticfield diffused out of the core.However,the magnetic energy in star-forming cores is generally less than or about equal to the gravitational energy(Crutcher1999),so this effect is unlikely to be significant.Note that since the core out of which a massive star forms is highly turbulent,it will be clumpy and the accretion rate can be expected to show largefluctuations—i.e.,in realityφ∗will be a random function of time.MT showed that if the collapse is spherical and self-similar,thenφ∗is quite close to unity provided that the value ofρentering tffis evaluated at the radius in the initial core that just encloses the gas that goes into the star when its mass is m∗.The resulting value ofφ∗averages over thefluctuations that would occur in an actual turbulent core.For a polytropic sphere,ρ∝r−kρand M∝r3−kρ,which impliesρ∝M−kρ/(3−kρ)=M2/(4−3γp).Since we have assumed that the stellar mass is a constant fractionǫcore of the gas mass(see eq.15),the accretion rate varies as˙m∗∝m∗ρ1/2∝m1−1/(4−3γp)∗.(30) Integration yieldsm∗=m∗f tt =(4−3γp)m∗ft∗f 3−3γp.(32)Note that forγp<1the accretion rate accelerates(MP97);in particular,for thefiducial caseγp=2/3,we have˙m∗∝t. From equations(28)and(32),the star-formation time,which is the time from thefirst formation of a protostellar core to the time the star reaches itsfinal mass,is thent∗f=(4−3γp)t∗f,(34a)=π√3 (2−γp)2(4−3γp)(7−6γp)/2m03/8=0.663.For other values ofγp in the range0≤γp≤1,which is equivalent to1≤kρ≤2,the approximationφ∗non≃1.62−0.48kρ(35) is accurate to within about1%(MT).For thefiducial case(kρ=1.5),we haveφ∗non=0.90.We conclude that˙m∗≃m∗/tffto within a factor1.5for non-magnetic,spherical cores in which the effective sound speed increases outward.We can estimate the effect of magneticfields on the accretion rate from the work of Li&Shu(1997),who considered collapse of self-similar,isothermal,magnetized,toroidal clouds.The equilibrium surface density isΣ=(1+H0)c2th/(πG̟), where̟is the cylindrical radius and H0is a parameter that increases from zero as the magneticfield is increased.They show that the accretion rate is˙m∗=1.05(1+H0)c3th/G,which is larger than that in equation(1)by about a factor (1+H0).However,equation(28)predicts˙m∗∝Mρ1/2∝M3/2/̟3/2∝Σ3/2̟3/2∝(1+H0)3/2.To reconcile this result with the correct value,we requireφ∗≃φ∗non。
Heavy Elements in QSOs Star Formation and Galaxy Evolution at High Redshifts

f (ai)
a⊙
where (b/a)⊙ is the solar abundance ratio, and N and f are respectively the column densities and ionization fractions of elements a and b in ion states i and j. Ideally, one has abundance-
into relative abundances. For example, the abundance ratio for any two elements a and b can be
written as,
a = log N (ai) + log f (bj) + log b
(1)
b
N (bj)
arXiv:astro-ph/9809386v1 29 Sep 1998
HEAVY ELEMENTS IN QSOS: STAR FORMATION AND GALAXY EVOLUTION
AT HIGH REDSHIFTS
F. Hamann University of California, San Diego, USA
independent constraints on the ionization fractions from the column densities of different ions of
the same element. Otherwise, we can also constrain the ionization by comparing column densities
Three general, independent probes of QSO abundances are readily observable at all redshifts: the broad emission lines (BELs), the broad absorption lines (BALs) and the intrinsic narrow absorption lines (NALs). Each of these probes has its own theoretical and observational uncertainties, so it is essential to consider as many of them as possible. I am now involved in several projects to examine a wide range of abundance diagnostics in QSOs at different redshifts and luminosities. My principle collaborators are Drs. T. Barlow, F. Chaffee, G. Ferland, C. Foltz, V. Junkkarinen, K. Korista and J. Shields.
Oxygen in the Galactic thin and thick disks

a rXiv:as tr o-ph/21229v12Dec22CNO in the Universe ASP Conference Series,Vol.**VOLUME***,**YEAR OF PUBLICATION**C.Charbonnel,D.Schaerer &G.Meynet,eds.Oxygen in the Galactic thin and thick disks T.Bensby,S.Feltzing,and I.Lundstr¨o m Lund Observatory,Box 43,SE-22100Lund,Sweden Abstract.First results from a study into the abundance trends of oxy-gen in the Galactic thin and thick disks are presented.Oxygen abun-dances for 21thick disk and 42thin disk F and G dwarf stars based on very high resolution spectra (R ∼215000)and high signal-to-noise (S/N >400)of the faint forbidden oxygen line at 6300˚A have been de-termined.We find that [O /Fe]for the thick disk stars show a turn-down,i.e.the “knee”,at [Fe/H]between −0.4and −0.3dex indicating the onset of SNe type Ia.The thin disk stars on the other hand show a shallow decrease going from [Fe /H]∼−0.7to the highest metallicities with no apparent “knee”present indicating a slower star formation history.1.Introduction The Galactic thin and thick disks are two distinct stellar populations in terms of age distributions and kinematics.The chemical trends in the two systems are also most likely different although recent works give conflicting results,see e.g.Chen et al.(2000)and Fuhrmann (1998).We show that the abundance trends for oxygen are different for the thin and thick disks.2.Observations The selection of thin and thick disk stars was based on kinematics and is fullydescribed in Bensby et al.(2003a,in prep).We calculated Gaussian probabil-ities for each star that it belongs to the thin and thick disk respectively,using the galactic velocity components U ,V ,and W of the stars.Stars with high probabilities of belonging to either the thin or the thick disk were then selected.The sample consists of 21thick disk stars and 42thin disk stars.Spectra were obtained with the CES spectrograph on the ESO 3.6m tele-scope with a a resolution of R ∼215000and a signal-to-noise S/N >400.Telluric lines were divided out using spectra from fast rotating B stars.Further details are given in Bensby et al.(2003b,in prep).3.Abundances and resultsOxygen abundances were determined through fitting of synthetic spectra to the observed spectra.The forbidden oxygen line at 6300˚A that has a blend of nickel in its right wing.At low metallicities this blend is often negligible,but12Bensby et al.becomes severe at higher metallicities.This is illustrated in Fig.1where we plot synthetic and observed spectra for three stars at different metallicities.Fe and Ni abundances have been determined from our FEROS spectra(R∼48000)by measuring equivalent widths of approximately140Fe i,30Fe ii,and50Ni i lines for each star(Bensby et al.2003a in prep.)Figure1.The forbidden oxygen line at6300˚A for three stars withdifferent metallicities;HIP103458(thick disk),HIP96124(thick disk),and HIP78955(thin disk).The observed spectra are plotted with solidcircles.Three different synthetic spectra are shown for each star:onlythe forbidden oxygen line(dashed line),only the blending nickel line(dotted line),combination of the two(solid line).The two plots in Fig.2presents our results.These are ourfindings:1.The thin and thick disk stars clearly show different abundance trends.This is a strong indication of their disparate origin and different epochs of formation.2.A turn-down for[O/Fe]at[Fe/H]∼−0.35for the thick disk stars,from being roughlyflat,continuing down to solar values.This feature is most likelya signature of the onset of SNIa.3.The thin disk stars show a shallow decrease when going from the lowest metallicities to solar values,not showing a knee.This implies that the star for-mation rate in the thin disk was quite low compared to that in the thick disk.4.At super-solar metallicities the trend found at sub-solar metallicities contin-ues linearly for the thin disk stars.In contrast Nissen and Edvardsson(1992) found[O/Fe]to level out at these metallicities.However,they did not take the Ni i blend in the[O i]line into account,which becomes important at these metal-licities,see Fig.1.This result has implications for different models of supernova yields,and will be investigated further.Oxygen in the Galactic thin and thick disks3Figure2.Abundance trends for oxygen.Thick disk stars are markedbyfilled circles and thin disk stars by empty circles.All stars have also been observed with the FEROS spectrograph and abun-dances for other elements have been determined(Na,Mg,Al,Si,Ca,Sc,Ti, V,Cr,Mn,Fe,Co,Ni,Zn,Y,Ba,Eu).For theα-elements wefind the same signature from the onset of SNIa in the thick disk which appears to be absent in the thin disk,see Feltzing et al.(2002)and Bensby et al.(2003a,in prep),in good agreement with the trends wefind for oxygen.A few stars merits,due to their positions in Fig.2,further comments:two thick disk stars at[Fe/H]∼−0.3and one thin disk star at[Fe/H]∼−0.6.The latter may be due to the fact that the thick disk also contain stars with“cold”kinematics.Thefirst two are a bit harder to understand but their kinematics might have been heated through close encounters or they might have been kicked-out from a double or multiple stellar system.ReferencesBensby,T.,Feltzing,S.,&Lundst¨o m,I.2003a and2003b,both in preparation Chen,Y.Q.,Nissen,P.E.,Zhao,G.et al.2000,A&AS,141,491Feltzing,S.,Bensby,T.,&Lundst¨o m,I.2002,A&A,in press(astro-ph/0211589) Fuhrmann,K.1998,A&A,338,161Nissen,P.E.,&Edvardsson,B.1992,A&A,261,255。
The Star Formation History of Field Galaxies
a rXiv:as tr o-ph/978220v125Aug1997THE STAR FORMATION HISTORY OF FIELD GALAXIES Piero Madau Space Telescope Science Institute,3700San Martin Drive,Baltimore MD 21218;madau@ Lucia Pozzetti Dipartimento di Astronomia,Universit`a di Bologna,via Zamboni 33,I-40126Bologna;lucia@astbo3.bo.astro.it and Mark Dickinson 1,2Department of Physics and Astronomy,The Johns Hopkins University,Homewood Campus,Baltimore MD 21218;med@ ReceivedABSTRACTWe develop a method for interpreting faint galaxy data which focuses on the integrated light radiated from the galaxy population as a whole.The emission history of the universe at ultraviolet,optical,and near-infrared wavelengths is modeled from the present epochto z≈4by tracing the evolution with cosmic time of the galaxy luminosity density,as determined from several deep spectroscopic samples and the Hubble Deep Field(HDF) imaging survey.In a q0=0.5,h50=1cosmology,the global spectrophotometric properties of field galaxies can be wellfit by a simple stellar evolution model,defined by a time-dependent star formation rate(SFR)per unit comoving volume and a universal initial mass function (IMF)extending from0.1to125M⊙.While a Salpeter IMF with a modest amount of dust reddening or a somewhat steeper mass function,φ(m)∝m−2.7,can both reproduce the data reasonably well,a Scalo IMF produces too much long-wavelength light and is unable to match the observed mean galaxy colors.In the best-fit models,the global SFR rises sharply,by about an order of magnitude,from a redshift of zero to a peak value at z≈1.5in the range0.12–0.17M⊙yr−1Mpc−3,to fall again at higher redshifts.After integrating the inferred star formation rate over cosmic time,wefind a stellar mass density at the present epoch ofΩs h250∼>0.005,hence a mean stellar mass-to-light ratio∼>4in the B-band and ∼>1in K,consistent with the values observed in nearby galaxies of various morphological types.The models are able to account for the entire background light recorded in the galaxy counts down to the very faint magnitude levels probed by the HDF.Since only∼20%of the current stellar content of galaxies is produced at z>2,a rather low cosmic metallicity is expected at these early times,in good agreement with the observed enrichment history of the damped Lyman-αsystems.The biggest uncertainty is represented by the poorly constrained amount of starlight that was absorbed by dust and reradiated in the IR at early epochs.A “monolithic collapse”model,where half of the present-day stars formed at z>2.5and were shrouded by dust,can be made consistent with the global history of light,but overpredicts the metal mass density at high redshifts as sampled by QSO absorbers.Subject headings:galaxies:evolution–galaxies:formation1.IntroductionIn the past few years two different approaches have been widely used to interpret faint galaxy data (see Ellis1997for a recent review).In the simplest version of the“traditional”scheme,a one-to-one mapping between galaxies at the present epoch and their distant counterparts is assumed:one starts from the local measurements of the distribution of galaxies as a function of luminosity and Hubble type and models their photometric evolution assuming some redshift of formation and a set of parameterized star formation histories(Tinsley1980;Bruzual&Kron1980;Koo1985;Guiderdoni&Rocca-Volmerange 1990;Metcalfe et al.1991;Gronwall&Koo1995;Pozzetti,Bruzual,&Zamorani1996).These,together with an initial mass function(IMF)and a cosmological model,are then adjusted to match the observed number counts,colors,and redshift distributions.Beyond the intrinsic simplicity of assuming a well defined collapse epoch and pure-luminosity evolution thereafter,the main advantage of this kind of approach is that it can easily be made consistent with the classical view that ellipticals and spiral galaxy bulges formed early in a single burst of duration1Gyr or less(see,e.g.Ortolani et al.1995 and references therein).Because much of the action happens at high-z,however,these models predict far more Lyman-break“blue dropouts”than are seen in the Hubble Deep Field(HDF)(Ferguson& Babul1997;Pozzetti et al.1997),and cannot reproduce the rapid evolution–largely driven by late-type galaxies–of the optical luminosity density with lookback time observed by Lilly et al.(1996)and Ellis et al.(1996).Less straighforward models which include,e.g.,a large population of dwarf galaxies that begin forming stars at z≈1(Babul&Ferguson1996),or do not conserve the number of galaxies due to merger events(Broadhurst,Ellis,&Glazebrook1992;Carlberg&Charlot1993)also appear unable to match the global properties of present-day galaxies(Ferguson1997;Ferguson&Babul1997).A more physically motivated way to interpret the observations is to construct semianalytic hierarchical models of galaxy formation and evolution(White&Frenk1991;Lacey&Silk1991; Kauffmann&White1993;Kauffmann,White,&Guiderdoni1993;Cole et al.1994;Baugh et al. 1997).Here,one starts ab initio from a power spectrum of primordial densityfluctuations,follows the formation and merging of dark matter halos,and adopts various prescriptions for gas cooling,star formation,feedback,and dynamical friction.These are tuned to match the statistical properties of both nearby and distant galaxies.In this scenario,there is no period when bulges and ellipticals form rapidly as single units and are very bright:rather,small objects formfirst and merge continually tomake larger ones.While reasonably successful in recovering the counts,colors,and redshift distributions of galaxies,a generic difficulty of such models is the inability to simultaneously reproduce the observed local luminosity density and the zero-point of the Tully-Fisher relation(White&Frenk1991).In this paper we shall develop an alternative method,which focuses on the emission properties of the galaxy population as a whole.It traces the cosmic evolution with redshift of the galaxy luminosity density–as determined from several deep spectroscopic samples and the HDF imaging survey–and offers the prospect of an empirical determination of the global star formation history of the universe and initial mass function of stars independently,e.g.,of the merging histories and complex evolutionary phases of individual galaxies.The technique relies on two basic properties of stellar populations:a)the UV-continuum emission in all but the oldest galaxies is dominated by short-lived massive stars,and is therefore a direct measure,for a given IMF and dust content,of the instantaneous star formation rate (SFR);and b)the rest-frame near-IR light is dominated by near-solar mass evolved stars that make up the bulk of a galaxy’s stellar mass,and can then be used as a tracer of the total stellar mass density.By modeling the“emission history”of the universe at ultraviolet,optical,and near-infrared wavelengths from the present epoch to z≈4,we will shed light on some key questions in galaxy formation and evolution studies:Is there a characteristic epoch of star and metal formation in galaxies?What fraction of the luminous baryons observed today were already locked into galaxies at early epochs?Are high-z galaxies obscured by dust?Do spheroids form early and rapidly?Is there a universal IMF?Let us point out some of the limitations of our approach at the outset.(1)We shall study the emission properties of“normal”,optically-selectedfield galaxies which are only moderately affected by dust–a typical spiral emits30%of its energy in the far-infrared region(Saunders et al.1990).Starlight which is completely blocked from view even in the near-IR by a large optical depth in dust will not be recorded by our technique,and the associated baryonic mass and metals missed from our census.The contribution of infrared-selected dusty starbursts to the present-day total stellar mass density cannot be very large,however,for otherwise the current limits to the energy density of the mid-and far-infrared background would be violated(Puget et al.1996;Kashlinsky,Mather,&Odenwald1996;Fall,Charlot, &Pei1996;Guiderdoni et al.1997).Locally,infrared luminous galaxies are known to produce onlya small fraction of the IR luminosity of the universe(Soifer&Neugebauer1991).(2)Our method bypasses the ambiguities associated with the study of morphologically-distinct samples whose physicalsignificance remains unclear,but,by the same token,it does not provide any direct information on the processes which shaped the Hubble sequence.Similarly,this approach does not specifically address the evolution of particular subclasses of objects,like the oldest ellipticals or low-surface brightness galaxies, whose star formation history may have differed significantly from the global average(e.g.Renzini1995; McGaugh&Bothun1994).(3)Although in our calculations the IMF extends from0.1to125M⊙,by modeling the rest-frame galaxy luminosity density from0.15to2.2µm we will only be sensitive to stars within the mass range from∼0.8to about20M⊙.This introduces non-negligible uncertainties in our estimate of the total amount of stars and metals produced.(4)No attempt has been made to include the effects of cosmic chemical evolution on the predicted galaxy colors.All our population synthesis models assume solar metallicity,and thus will generate colors that are slightly too red for objects with low metallicity,e.g.truly primeval galaxies.(5)The uncertanties present in our estimates of the UV luminosity density from the identification of Lyman-break galaxies in the HDF are quite large,and the data points at z>2should still be regarded as tentative.This is especially true for the faint blue dropout sample at z≈4,where only one spectroscopic confirmation has been obtained so far(Dickinson 1997).On the other hand,there is no evidence for a gross mismatch at the z≈2transition between the photometric redshift sample of Connolly et al.(1997)and the Madau et al.(1996)UV dropout sample.The initial application of this method was presented by Lilly et al.(1996)and Madau et al.(1996, hereafter M96).A complementary effort–which starts instead from the analysis of the evolving gas content and metallicity of the universe–can be found in Fall et al.(1996).Unless otherwise stated,we shall adopt in the following aflat cosmology with q0=0.5and H0=50h50km s−1Mpc−1.2.The Evolution of the Galaxy Luminosity DensityThe integrated light radiated per unit volume from the entire galaxy population is an average over cosmic time of the stochastic,possibly short-lived star formation episodes of individual galaxies,and will follow a relatively simple dependence on redshift.In the UV–where it is proportional to the global SFR–its evolution should provide information,e.g.,on the mechanisms which may prevent the gas within virialized dark matter halos from radiatively cooling and turning into stars at early times,or on the epoch when galaxies exhausted their reservoirs of cold gas.From a comparison between different wavebands it should be possible to set constraints on the average IMF and dust content of galaxies.The comoving luminosity density,ρν(z),from the present epoch to z≈4is given in Table1in five broad passbands centered around0.15,0.28,0.44,1.0,and2.2µm.The data are taken from the K-selected wide-field redshift survey of Gardner et al.(1997),the I-selected CFRS(Lilly et al.1996)and B-selected Autofib(Ellis et al.1996)surveys,the photometric redshift catalog for the HDF of Connolly et al.(1997)–which take advantage of deep infrared observations by Dickinson et al.(1997)–and the color-selected UV and blue“dropouts”of M96.3They have all been corrected for incompleteness by integrating over the best-fit Schechter function in each redshift bin,ρν(z)= ∞0Lνφ(Lν,z)dLν=Γ(2+α)φ∗L∗.(1) As it is not possible from the Connolly et al.(1997)and M96data sets to reliably determine the faint end slope of the luminosity function,a value ofα=−1.3has been assumed at each redshift interval for comparison with the CFRS sample(Lilly et al.1995).The error bars are typically less than0.2in the log,and reflect the systematic uncertainties introduced by the assumption of a particular value ofαand,in the HDF z>2sample,in the volume normalization and color-selection region.In the K-band, the determination by Gardner et al.(1997)agrees to within30%with Cowie et al.(1996),and we have assigned an error of0.1in the log to the estimate of the local luminosity density at2.2µm.Despite the obvious caveats due to the likely incompleteness in the data sets,different selection criteria,and existence of systematic uncertainties in the photometric redshift technique,the spectroscopic, photometric,and Lyman-break galaxy samples appear to provide a remarkably consistent picture of the emission history offield galaxies.The UV luminosity density rises sharply,by about an order of magnitude,from a redshift of zero to a peak at z≈1.5,to fall again at higher redshifts(M96;Lilly et al.1996;Connolly et al.1997).This points to a rapid drop in the volume-averaged SFR in the last8–10 Gyr,and to a redshift range1∼<z∼<2in which the bulk of the stellar population was assembled.The decline in brightness at late epochs is shallower at longer wavelengths,as galaxies becomes redder with cosmic time,on the average.3.Indicators of Past and Present Star Formation ActivityStellar population synthesis has become a standard technique to study the spectrophotometric properties of galaxies.Here,we shall make extensive use of the latest version of Bruzual&Charlot(1993) isochrone synthesis code,optimized with an updated library of stellar spectra(Bruzual&Charlot1997), to predict the time change of the spectral energy distribution of a stellar population.The uncertanties linked to the underlying stellar evolution prescriptions and the lack of accurateflux libraries do not typically exceed35%(Charlot,Worthey,&Bressan1966).Shortward of the Lyman edge,however,the differences in the predicted ionizing radiation from model atmospheres of hot stars can be quite large (Charlot1996a).We shall consider three possibilities for the IMF,φ(m)∝m−1−x:a Salpeter(1955) function(x=1.35),a Scalo(1986)function,which isflatter for low-mass stars and significantly less rich in massive stars than Salpeter,and an intermediate case with x=1.7.In all models the metallicity is fixed to solar values and the IMF is truncated at0.1and125M⊙.3.1.Birthrate–Ultraviolet RelationThe UV continuum emission from a galaxy with significant ongoing star formation is entirely dominated by late-O/early-B stars on the main sequence.As these have masses∼>10M⊙and lifetimes t MS∼<2×107yr,the measured luminosity becomes proportional to the stellar birthrate and independent of the galaxy history for t≫t MS.This is depicted in Figure1,where the power radiated at1500˚A and2800˚A is plotted against the instantaneous SFR for a model stellar population with different star formation laws,SFR∝exp(−t/τ),whereτis the duration of the burst.After an initial transient phase where the UVflux rises rapidly and the turnoffmass drops below10M⊙,a steady state is reached where one can writeSFRL UV=const×4The luminosities at1500˚A and2800˚A have been averaged over a rectangular bandpass of width ∆λ/λ=20%in order to approximate the standard broadbandfilters used in the observations.Fig.1.—SFR-UV relation for models with various exponentially declining star formation rates at ages between0.01and15Gyr.Solid lines:τ=20Gyr.Dot-dashed lines:τ=10Gyr.Dashed lines:τ=5 Gyr.Dotted lines:τ=1Gyr.The set of curves on the left-hand side of the plot assume a Scalo IMF,the ones on the right-hand side a Salpeter function.durations∼<1Gyr and a Scalo IMF,the luminosity at2800˚A becomes a poor SFR indicator after a few e-folding times,when the contribution of intermediate-mass stars becomes significant.After averaging over the whole galaxy population,however,we willfind that the UV continuum is always a good tracer of the instantaneous rate of conversion of cold gas into stars.3.2.Mass–Infrared RelationIf we assume a time-independent IMF,we can then use the results of stellar population synthesis modeling,together with the observed UV emissivity,to infer the evolution of the star formation activity in the universe(M96).As already mentioned,the biggest uncertainty in this procedure is due to dust reddening,as newly formed stars which are completely hidden by dust would not contribute to the UV luminosity.The effect is potentially more serious at high-z,as for afixed observer-frame bandpass, one is looking further into the ultraviolet with increasing redshifts.On the other hand,it should be possible to test the hypothesis that star formation regions remain,on average,largely unobscured by dust throughout much of galaxy evolution by looking at the near-infrared light density.This will be affected by dust only in the most extreme,rare cases,as it takes an E(B−V)∼>4mag to produce an optical depth of order unity at2.2µm.Although different types of stars–such as supergiants,AGB,and red giants–dominate theK-band emission at different ages in an evolving stellar population,the mass-to-infrared light ratio is relatively insensitive to the star formation history(Charlot1996b).Figure2shows M/L K(in solar units)as a function of age for models with various exponentially declining SFR compared to the values observed in nearby galaxies of early to late morphological types.5As the stellar population ages,the mass-to-infrared light ratio remains very close to unity,independent of the galaxy color and Hubble type. We can use this interesting property to estimate the baryonic mass in galaxies from the local K−band luminosity density,logρK(0)=27.05±0.1h50ergs s−1Hz−1Mpc−3(Gardner et al.1997).The observed range0.6h50∼<M/L K∼<1.9h50translates into a visible(stars+gas)mass density at the present day of2×108∼<ρs+g(0)∼<6×108h250M⊙Mpc−3(3)(0.003∼<Ωs+g∼<0.009).We shall see in the next section how the observed integrated UV emission,with or without the addition of some modest amount of reddening,may account for the bulk of the baryons traced by the K-band light,and how initial mass functions with relatively few high-mass stars(such as the Scalo IMF),or models with a large amount of dust extinction at all epochs will tend to overproduce the near-infrared emissivity.4.Stellar Population Synthesis ModelingAn interesting question now arises as to whether a simple stellar evolution model,defined by a time-dependent SFR per unit volume and a constant IMF,may reproduce the global UV,optical,and near-IR photometric properties of the universe as given in Table1.In a stellar system with arbitrary star formation rate,the luminosity density at time t is given by the convolution integralρν(t)= t0Lν(τ)×SFR(t−τ)dτ,(4) where Lν(τ)is the specific luminosity radiated per unit initial mass by a generation of stars with ageτ. In the instantaneous recycling approximation(Tinsley1980),the total stellar mass density produced at time t isρs(t)=(1−R) t0SFR(t)dt,(5) where R is the mass fraction of a generation of stars that is returned to the interstellar medium,R≈0.3,0.15,and0.2for a Salpeter,x=1.7,and Scalo IMF,respectively.6In computing the time evolution of the spectrophotometric properties of a stellar population in comoving volumes large enough to be representative of the universe as a whole,ourfirst task is to relate the observed UV emission to a SFR density.We assume a universal IMF andfit a smooth function to the UV continuum emissivity at various redshifts.By construction,all models will then produce,to within the errors,the right amount of ultraviolet light.We then use Bruzual-Charlot’s synthesis code to predict the cosmic emission history at long wavelengths.Fig. 2.—Total(processed gas+stars)mass-to-K band light ratio versus age for models with various exponentially declining star formation rates.Solid lines:Salpeter IMF.Dotted lines:Scalo IMF.From top to bottom,each set of curves depict the values forτ=1,5,10,and20Gyr,respectively.The data points show the visible mass-to-infrared light ratio versus B−V color(top axis)observed in nearby galaxies of various morphological types(Charlot1996b).In one of the scenarios discussed in the next section,the effect of dust attenuation will be taken into account by multiplying equation(4)by p esc,a time-independent term equal to the fraction of emitted photons which are not absorbed by dust.For purposes of illustration,we assume a foreground screen model,p esc=exp(−τν),and SMC-type dust.7This should only be regarded as an approximation,since hot stars can be heavily embedded in dust within star-forming regions,there will be variety of extinction laws,and the dust content of galaxies will evolve with redshift.While the existing data are too sparse to warrant a more elaborate analysis,this simple approximation well highlights the main features and assumptions of the model.It is possible to gauge the luminosity-weighted amount of dust extinction at the present epoch by looking at the observed local far-infrared luminosity density.For normal galaxies, the emissivity from8to115µm is estimated from the IRAS survey to be around30%of their integrated emission in the B-band(Saunders et al.1990).Assuming that a negligible fraction of the ionizingflux emerges in nebular lines,and a Salpeter IMF,this value implies a small luminosity-weighted color excess, E(B−V)≈0.012(τB≈0.025).The Calzetti,Kinney,&Storchi-Bergmann(1994)empirical extinction law for starbursts,normalized to A(V)/E(B−V)=4.88,yields a mean dust opacity that is similarly low.4.1.Salpeter IMFFigure3shows the model predictions for the evolution ofρνat rest-frame ultraviolet to near-infrared frequencies.The data points with error bars are taken from Table1,and a Salpeter IMF has been assumed.In the absence of dust reddening,this relativelyflat IMF generates spectra that are too blue to reproduce the observed mean(luminosity-weighted over the entire population)galaxy colors.The shape of the predicted and observedρν(z)relations agrees better to within the uncertainties if some amount of dust extinction,E(B−V)=0.1,is included.In this case,the observed UV luminosities must be corrected upwards by a factor of1.4at2800˚A and2.1at1500˚A.As expected,while the ultraviolet emissivity traces remarkably well the rise,peak,and sharp drop in the instantaneous star formationFig.3.—Evolution of the luminosity density at rest-frame wavelengths of0.15(dotted line),0.28(solid line),0.44(short-dashed line),1.0(long-dashed line),and2.2(dot-dashed line)µm.The data points with error bars are taken from Lilly et al.(1996)(filled dots at0.28,0.44,and1.0µm),Connolly et al.(1997) (empty squares at0.28and0.44µm),Madau et al.(1996)and Madau(1997a)(filled squares at0.15µm), Ellis et al.(1996)(empty triangles at0.44µm),and Gardner et al.(1997)(empty dot at2.2µm).The inset in the upper-right corner of the plot shows the SFR density(M⊙yr−1Mpc−3)versus redshift which was used as input to the population synthesis code.The model assumes a Salpeter IMF,SMC-type dustin a foreground screen,and a universal E(B−V)=0.1.rate(the smooth function shown in the inset on the upper-right corner of thefigure),an increasingly large component of the longer wavelengths light reflects the past star formation history.The peak in the luminosity density at1.0and2.2µm occurs then at later epochs,while the decline from z≈1to z=0is more gentle than observed at shorter wavelengths.The total stellar mass density at z=0is ρs(0)=3.7×108M⊙Mpc−3,with a fraction close to65%being produced at z>1,and only20%at z>2.In the assumed cosmology,about half of the stars observed today are more than9Gyr old,and only20%are younger than5Gyr.84.2.x=1.7IMFFigure4shows the model predictions for a x=1.7IMF and negligible dust extinction.While able to reproduce quite well the B-band emission history and consistent within the error with the local K-band light,this model slightly underestimates the1µm luminosity density at z≈1.The total stellar mass density today is larger than in the previous case,ρs(0)=6.2×108M⊙Mpc−3.4.3.Scalo IMFFigure5shows the model predictions for a Scalo IMF.Thefit to the data is now much poorer,since this IMF generates spectra that are too red to reproduce the observed mean galaxy colors,asfirst noted by Lilly et al.(1996).Because of the relatively large number of solar mass stars formed,it produces too much long-wavelength light by the present epoch.The addition of dust reddening would obviously make thefit even worse.The total stellar mass density produced is similar to the Salpeter IMF case.5.Clues to Galaxy Formation and EvolutionThe results shown in the previous section have significant implications for our understanding of the global history of star and structure formation.Here we discuss a few key issues which will assist inFig.4.—Same as in Figure 3,but assuming an IMF with φ(m )∝m −1−1.7and no dust extinction.Fig.5.—Same as in Figure 3,but assuming a Scalo IMF and no dust extinction.This model overproduces the local K -band emissivity by a factor of 2.interpreting the evolution of luminous matter in the universe.5.1.Extragalactic Background LightOur modeling of the data points to the redshift range where the bulk of the stellar mass was actually produced:1∼<z∼<2.The uncertainties in the determination of the luminosity density at that epoch are,however,quite large.At z≈1,the increase in the“estimated”emissivity(i.e.,corrected for incompleteness by integrating over the best-fit Schechter function)over that“directly”observed in the CFRS galaxy sample is about a factor of2(Lilly et al.1996).Between z=1and z=2,the peak in the average SFR is only constrained by the photometric redshifts of Connolly et al.(1997)and by the HDF UV dropout sample,both of which may be subject to systematic biases.An important check on the inferred emission history offield galaxies comes from a study of the extragalactic background light(EBL),an indicator of the total optical luminosity of the universe.The contribution of known galaxies to the EBL can be calculated directly by integrating the emittedflux times the differential galaxy number counts down to the detection threshold.We have used a compilation of ground-based and HDF data down to very faint magnitudes(Pozzetti et al.1997;Williams et al. 1996)to compute the mean surface brightness of the night sky between0.35and2.2µm.The results are plotted in Figure6,along with the EBL spectrum predicted by our modeling of the galaxy luminosity density,Jν=1dzρν′(z)(6)whereν′=ν(1+z)and dl/dz is the cosmological line element.The overall agreement is remarkably good,with the model spectra being only slightly bluer,by about20–30%,than the observed EBL.The straightforward conclusion of this exercise is that the star formation histories depicted in Figures3and 4appear able to account for the entire background light recorded in the galaxy counts down to the very faint magnitudes probed by the HDF.Fig.6.—Spectrum of the extragalactic background light as derived from a compilation of ground-based and HDF galaxy counts(see Pozzetti et al.1997).The2σerror bars arise mostly fromfield-to-field variations.Solid line:Model predictions for a Salpeter IMF and E(B−V)=0.1(star formation history of Figure3).Dotted line:Model predictions for a x=1.7IMF and negligible dust extinction(starformation history of Figure4).5.2.The Stellar Mass Density TodayThe best-fit models discussed in§4generate a present-day stellar mass density in the range between 4and6×108M⊙Mpc−3(0.005∼<Ωs h250∼<0.009).Although one could in principle reduce the inferred star formation density by adopting a top-heavy IMF,richer in massive UV-producing stars,in practice a significant amount of dust reddening–hence of“hidden”star formation–would then be required to match the observed galaxy colors.The net effect of this operation would be a large infrared background (see below).The stellar mass-to-light ratios range from4.5in the B-band and0.9in K for a Salpeter function,to8.1in B and1.5in K for a x=1.7IMF.Note that these values are quite sensitive to the lower-mass cutoffof the IMF,as very-low mass stars can contribute significantly to the mass but not to the integrated light of the whole stellar population.A lower cutoffof0.2M⊙instead of the0.1M⊙adopted would decrease the mass-to-light ratio by a factor of1.3for a Salpeter function,1.6for x=1.7, and1.1for a Scalo IMF.5.3.The Star Formation Density TodayThe predicted local rate of star formation ranges between0.95×10−2(Salpeter)and1.6×10−2M⊙yr−1Mpc−3(x=1.7).According to Gallego et al.(1995),the Hαluminosity density of the local universe is logρHα=39.1±0.04ergs s−1Mpc−3.Let us assume case-B recombination theory,and adopt the mean unreddened spectrum at z=0corresponding to the Salpeter IMF model, i.e.,assume the emitted ionizing photons are converted to Hαradiation before being absorbed by dust. The Hαluminosity density can then be related to the local SFR per unit volume according tologρHα=39.2+log SFR。
Radio Triggered Star Formation in Cooling Flows
a rXiv:as tr o-ph/9911129v18Nov1999Presented at “Life Cycles of Radio Galaxies”,July 15-17,1999,STScI,Balti-more Radio Triggered Star Formation in Cooling Flows B.R.McNamara a ,a Harvard-Smithsonian Center for Astrophysics,60Garden St.Cambridge,MA 021381IntroductionMore than half of clusters within redshift z ∼0.1contain bright,central X-ray emission from ∼keV gas that appears to be cooling at rates of ∼10−1000M ⊙yr −1(Fabian 1991).Commonly referred to as cooling flows,persistent accretion of this cooling material onto the bright,central galaxies in clusters (CDGs)at even a fraction of these rates would be capable of fueling vigorous star formation and the central engines generating their radio sources.Enhanced levels of cold gas and star formation are indeed seen in cooling flows (see McNamara 1997for a review).However,the inferred star formation rates are only <∼1−10%of the cooling rates derived from X-ray observations,and the amounts of cold gas detected outside of the X-ray band would account for <∼108yr of accumulated material.Between 60–70%of CDGs in coolingFigure 1(left):Correlation between central U −B continuumformation to the presence of a coolingflow.The star formation rates associated with the objects in Figure1range from<∼1−100M⊙yr−1(McNamara& O’Connell1989,McNamara1997;Cardiel et al.1998).Beyond the central regions,the spatially averaged surface brightness profiles usually follow the de Vaucouleurs r1/4law well into their halos.If the CDG has the characteristic envelope of a cD galaxy(Schombert1987),the profile rises above the r1/4 law extrapolated outward from the halo.In Figure2I show U and R surface brightness profiles for the CDG in the distant,z=0.1386coolingflow cluster Abell1068,whose cooling rate is estimated to be˙m x∼400M⊙yr−1(Allen et al.1995).The U-band profile rises above the r1/4profile in the inner several kpc of the galaxy.Beyond the inner few arcsec,both the U and R profiles follow the r1/4profile until reaching the cD envelope atµ(R)≃25mag.arcsec−2, where the surface brightness rises above the r1/4profile with an amplitude of ∼0.5mag.Apart from the blue core,this surface brightness profile is typical for cD galaxies in clusters with and without coolingflows(Porter et al.1991). There is little evidence to suggest that the average halo structure and colors of coolingflow galaxies have recent star formation in excess of what is seen in non coolingflow galaxies.The blue inner regions appear to be the result of accretion concentrated onto the core of a preexisting galaxy,but evidently not throughout its volume.3Radio Triggered Star FormationMost coolingflows harbor luminous∼1040−42ergs s−1emission line nebulae extending several to tens of kpc around the CDG nuclei(Heckman et al. 1989;Baum1991).The line emission and blue optical continuum are usually extended on similar spatial scales(Cardiel et al.1998),and the radio and emission line morphologies and powers are correlated,although with a large degree of scatter(Baum1991,but also see Allen1995).The tendency for strong line emission from warm,104K gas to lie along the edges of radio sources is particularly germane to understanding star formation in these objects.An early example was seen in the Abell1795CDG(van Breugel et al.1984),and a more striking example is seen in Hαimagery of the Abell2597CDG with the Hubble Space Telescope(Koekemoer et al.1999).Furthermore,the radio jets in Abell1795and Abell2597bend at roughly90degree angles and inflate into radio lobes at the locations of dust clouds embedded in the emission-line nebulae(Sarazin et al.1994;McNamara et al.1996).Their disrupted(i.e. bending)radio morphologies are almost surely the result of collisions between the radio jets and cold,dense clouds associated with the line-emitting gas. At the same time,Abell2597and Abell1795have bright blue optical con-tinuum(blue lobes)along their radio lobes(McNamara&O’Connell1993; McNamara1997),much like the so-called alignment effect seen in distant ra-dio galaxies(McCarthy1993).That this phenomenon is seen in a relatively small sample of CDGs is particularly interesting.Unlike distant radio galaxies, the coolingflow CDGs were selected on the basis of their X-ray properties, rather than their radio properties.Upon their discovery,two models emerged to explain the blue lobes:jet-induced star formation(De Young1995)and scat-tered light from an obliquely directed active nucleus(Sarazin&Wise1993; Murphy&Chernoff1993;Crawford&Fabian1993).The scattered light hy-pothesis predicts the blue lobe light should be polarized,as is found in many distant radio galaxies exhibiting the alignment effect(Jannuzi&Elston1991; di Serego Alighieri1989).U-band continuum polarization measurements for the Abell1795and Abell2597CDGs obtained with the KPNO4m Mayall telescope gave upper limits of<∼6%to the degree of polarization in both ob-jects,which effectively excluded the scattering hypothesis(McNamara et al. 1996;1999).Subsequent HST images of both objects resolved the blue lobes into knots of young star formation(McNamara et al.1996;Pinkney et al.1996;Koekemoer et al.1999).The HST R-band image of Abell1795’s blue knots are shown against a contour map of the radio source in Figure3.The stellar knots are found along the edges of the radio lobes and near the collision sites of the radio plasma and cold gas.They are not found primarily along the radio jets,as would be expected if the triggering mechanism were shocks traveling transverse to the jet trajectory,as predicted in jet-induced star formation models(De Young1995;Daly1990,Begelman&Cioffi1989).The observations suggest that momentum transferred through direct collisions between the radio plasma and cold gas clouds may be a more suitable triggering mechanism.(D.De Young pointed out that the strongest shocks would occur at the point of impact,and these shocks provide a possible triggering mechanism.) Although star formation at rates of∼10−40M⊙yr−1appears to be occurring in these objects,the radio sources may not have triggered all star formation.In addition to the blue light along the radio lobes,a more diffuse blue component that accounts for more than half the blue light is seen.Therefore,the radio source may be augmenting star formation in preexisting star bursts.4A Burst Mode of Star Formation in Cooling FlowsTracing the history of a stellar population,even in isolation,is difficult.The problem is further complicated when the population is embedded in a bright background galaxy.The blue lobes in the Abell1795and Abell2597CDGs are thefirst clear-cut evidence for a burst mode of star formation in coolingflows. The blue lobes cannot be old because the the alignment between the radio and optical structures can last only a fraction of the radio source lifetime and theFigure3(left):HST image in V+R of Abell1795’s blue lobes(greyscale) resolved into knots of star formation along the3.6cm radio lobes(contours). Figure4(right):Radio power plotted against the polarized luminosity for alignment effect radio galaxies.The3C radio galaxies are grouped to the upper right,and the upper limits to the polarized luminosity for Abell1795 and Abell2597are to the lower left of the plot.The solid line,normalized to the median of the3C points,represents L(U)pol∝P rad.The vertical dashed line indicates approximately the transition between FR I and FR II radio luminosities.stellar diffusion time scale,both∼107yr.Additional evidence supporting a burst mode of star formation in coolingflows has accumulated in recent years. Cardiel et al.(1998)have argued using the Mg II absorption line index,the 4000˚A break,and far UV colors that short duration bursts(<∼107yr)or constant star formation with ages≪1Gyr bestfit Bruzual model isochrones. While acknowledging the large uncertainties in the population isochrones,a burst mode of star formation is unexpected in simple,continuous coolingflow models(e.g.Fabian1991).If star formation is indeed being fueled by cooling flows,it would seem that gas is not accreting continuously.Transient sources of fuel,such as mergers or stripping,may also be contributing.5Are CDGs in Cooling Flows Low Radio Power Siblings of High Redshft Radio Galaxies?The premise that blue lobes are sites of star formation is supported by several facts.The absence of a polarized signal from the blue lobes effectively excludes the scattered light hypothesis.Synchrotron radiation can be excluded by theabsence of a detailed correlation between the radio source and blue lobes,and the nebular continuum is insufficiently strong to account for the blue color excesses.However,Balmer absorption is seen in the spectra of some objects (Allen1995),and the emission line luminosities and H II region characteristics are often consistent with powering by young stars(Shields&Filippenko1990; Voit&Donahue1997),so star formation is almost certainly the primary source of the color excesses in CDGs.The situation is more complex in the high redshift powerful radio galaxies(HzRGs)exhibiting the alignment effect.The aligned optical continuum in HzRGs is often strongly polarized,which has been interpreted as the signature of scattered light from an obliquely-directed active nucleus(di Serego Alighieri et al.1989;Jannuzi&Elston1991).In Figure4, I plot our polarizedflux upper limits for the blue lobes in Abell1795,Abell 2597,and the alignment regions of several HzRGs against rest frame20cm radio power(see McNamara et al.1999).The polarizedfluxes are measured in the rest frame U-band,and can be compared directly.Although the HzRGs are2–3orders of magnitude more powerful in their radio and polarizedfluxes, a linear extrapolation downward between radio power and polarizedflux from the mean HzRG value to the coolingflows would predict a lower polarized flux than is observed.Assuming similar host galaxy properties and scattering environments in both types of object,and further assuming the polarized flux scales approximately in proportion to radio power(see McNamara et al. 1999),at the precision of our measurements,we should not have detected a polarizedflux in Abell2597and Abell1795.In addition,it would seem that the polarizedfluxes of HzRGs generally account for a large but incomplete fraction of the blue light,and occasionally unpolarized star light dominates(e.g.van Breugel et al.1998).It is possible then that the blue lobes in coolingflows and the alignment effect in powerful radio galaxies are similar phenomena. But while starlight dominates the aligned continuum in lower radio power CDGs,scattered light dominates in HzRGs owing to their more powerful nuclei (McNamara et al.1999).6An Analysis of New Imagery for the Abell1068CDGIn this section I discuss new optical imagery of the Abell1068central cluster galaxy.The data provide new clues to the relationship between star formation and the radio source,and raise new questions regarding the mechanism fueling star formation.U-band CCD imaging is the most sensitive means of isolat-ing and studying the bluest galaxy populations from the ground.The blue populations in CDGs often contribute more than half of the central U-band light,while the fraction decreases to∼10%or less in the R and I bands.The blue populations can therefore be isolated by modeling and subtracting the background galaxy leaving the blue regions in residual.By doing so in two orFigure5:Imagery of Abell1068:U-band image(upper left);U−R color map(grayscale)superposed on R-band contours(upper right);U-band con-tours,after subtracting a smooth U-band model CDG galaxy,on the20cm FIRST radio grayscale image(lower left);Hαmap(lower right).The panels are registered to the same scale;north is at top and east is to the left.more pass bands,intrinsic colors of the blue population can be estimated.I applied this procedure to the z=0.1386Abell1068CDG,one of the most distant and largest coolingflows(˙m x∼400M⊙yr−1)discovered in the ROSAT All Sky Survey(Allen et al.1995).It is also one of the bluest CDGs in my sample.Figure5presents4-panels showing the U-band image to the upper left,a U−R color map(grayscale)superposed on R-band contours to the upper right,U-band contours,after subtracting a smooth U-band model CDG galaxy,on the20cm FIRST radio grayscale image(F W HM=5.4arc-sec),lower left,and an Hαmap,lower right.The panels are registered to the same scale;north is at top and east is to the left.Gray regions in the color map are abnormally blue.Several features are noteworthy.First,the central region within a13kpc diameter is∼0.5−0.9mag bluer than normal.The nuclear colors,after Kcorrection,range between(U−R)K,0≃1.5−2.3(the foreground reddening is negligible).An arc of blue light lies8arcsec(25kpc)in projection to the north-west of the nucleus,and a large wisp or arc of blue light extends to the south-west,until meeting a bright blue patch of light13arcsec to the east of the nucleus,and about8arcsec to the north of the bright neighboring galaxy to the south-west of the nucleus.This feature is nearly as blue as the nucleus with(U−R)K,0≃1.6.Finally,several blue knots,15–30arcsec north-west of the nucleus,appear along a line between the nucleus and a disturbed galaxy 35arcsec to the north-west of the nucleus.The remaining colors of the off-nuclear features range from(U−R)K,0≃2.0to the normal background color (U−R)K,0≃2.4.After subtracting a model galaxy from the U and R CDG images,Ifind an intrinsic nuclear blue population color(U−R)K,0∼−0.2.This color is consistent with Bruzual-Charlot population model colors for a∼107yr old burst population or continuous star formation for∼0.1Gyr.The colors are bluer than expected for star formation in a coolingflow that has been accreting continuously for>∼1Gyr.The accretion population’s luminosity mass is∼2×108M⊙,which would correspond to a star formation rate of∼80M⊙yr−1. The off-nuclear colors,being a few tenths of a magnitude redder than the nuclear colors,are consistent with a several107yr old burst or continuous star formation for<∼1Gyr.The offnuclear blue regions are apparently not in dynamical equilibrium.They appear to be stripped debris,possibly from the bright neighboring galaxies to the north-west and south-west of the nucleus.The disturbed appearance of the north-west galaxy’s R-band isophotes support the stripping hypothesis. The blue regions are considerably bluer than their putative parent galaxies, which would be consistent with blue material being composed primarily of young stars that formed out of cold material stripped from the galaxies.6.1Radio Triggered Star Formation in Abell1068?Both the Abell1068CDG and the bright galaxy to the south west of the CDG are radio sources.Each have radio powers of∼8.5×1024W/Hz,which are typical for FR I radio sources.In addition,the nucleus is embedded in a luminous emission line nebula with an Hαluminosity>∼2×1042ergs s−1(Allen et al.1992).Although only a low resolution radio map is available,the radio source appears extended to the north-west in the same direction as a tongue of Hαemission extending from the nucleus.Both the radio source and the tongue of Hαemission terminate8arcsec(25kpc)to the north-west of the nucleus at the location of the bright blue arc.Such a close spatial relationship between the radio source,nebular emission,and knots of star formation are commonin powerful radio galaxies in general,and in coolingflows in particular.It is tempting to speculate that,with high resolution radio maps in hand,the radio and optical morphologies will again be consistent with radio triggered star formation in the blue arc to the north-west,much like Minkowski’s Object (van Breugel et al.1985).7The Fueling MechanismThe origin of the material fueling star formation is of fundamental interest.A coolingflow origin is supported by the correlation between central blue color excess in CDGs and the cooling rate of the intracluster gas,derived independently from X-ray observations,shown in Figure1(e.g.McNamara 1997;Cardiel et al.1998).Were major galaxy mergers supplying the fuel, this correlation would be difficult to to explain.I would then expect CDGs experiencing significant bursts of star formation to be observed with equal frequency in coolingflow and non-coolingflow clusters alike,but they are not.Nonetheless,the evidence supporting periodic bursts of star formation implies an intermittent source of fuel.Ram pressure stripping of cold gas from neighboring cluster galaxies may be such a source of fuel,and might account for the˙m x–blue color correlation.The cooling rate˙m x∝ρ2gas,and the ram pressure force on a parcel of gas isρgas v2.Therefore,the dense coolingflow regions provide a large stripping cross section capable of sweeping cold,dense molecular gas from cluster dwarf galaxies and spirals,which would rain onto the parent CDG.Abell1068may be a case in point,as might the Abell1795 CDG(McNamara et al.1996).8Cooling Flows and the Chandra X-ray ObservatoryAs I wrote this article,Chandra was launched and began sending astonishingly crisp images of cosmic X-ray sources.During the next few years,many of Chandra’s targets will be clusters of galaxies,and the coolingflows promise some of the most interesting and productive cluster science.Their bright cores–the characteristic signature of a coolingflow–afford Chandra the opportunity to take full advantage of its nearly perfect,half arcsecond mirrors.For thefirst time,we will be capable of mapping structure in the X-ray-emitting gas on angular scales smaller than the radio sources and star formation regions.The temperature and density maps on these small scales will provide local cooling rates that can be compared directly to optically-derived star formation rates. Perhaps more than any other X-ray telescope planned or in queue,Chandra will advance our understanding of the dynamical and thermal state of clustercores,which hopefully will bring the long-standing coolingflow problem to resolution.9Summary•Unusually blue colors associated with young,massive stars frequent the central regions of coolingflow CDGs.The probability of detecting a blue population increases sharply with˙m x derived from X-ray observations.•Star formation in coolingflows apparently occurrs in repeated,short du-ration(<∼1Gyr)bursts,not continuously as would be expected in standard coolingflow models.•Bursts of star formation are often triggered by the radio sources.•Cold material stripped from neighboring galaxies may feed the the radio source and fuel some star formation in CDGs.ReferencesAllen,S.W.1995,MNRAS,276,947Allen,S.,Edge,A.,Fabian,A.,B¨o hringer,H.,Crawford,C.,Ebeling,H.,Johnstone, R.,Naylor,T.,Schwarz,R.1992,MNRAS,259,67Allen,S.,Fabian,A.,Edge,A.,B¨o hringer,H.,White,D.1995,MNRAS,275,741 Baum,S.A.1992,in Clusters and Superclusters of Galaxies,ed.A.C.Fabian (Dordrecht:Kluwer),171Begelman,M.C.,&Cioffi,D.F.1989,ApJ,345,L21Burns,J.O.,Loken,C.,Gomez,P.,Rizza,E.,Bliton,M.,&Ledlow,M.in Galactic and Cluster Cooling Flows,ed.N.Soker(San Francisco:PASP),21Cardiel,N.,Gorgas,J.,&Aragon-Salamanca,A.1998,MNRAS,298,977 Crawford,C.S.,&Fabian,A.C.1993,MNRAS,265,431Daly,R.A.1990,ApJ,355,416De Young,D.S.1995,ApJ,446,521di Serego Alighieri,S.,Fosbury,R.A.E.,Quinn,P.J.,&Tadhunter,C.N.1989, Nature,341,307Fabian,A.C.1991,in Clusters and Superclusters of Galaxies,ed.A.Fabian(Kluwer: Dordrecht),151Heckman,T.M.,Baum,S.A.,van Breugel,W.J.M.,&McCarthy,P.J.1989, ApJ,338,48Jannuzi,B.T.,&Elston,R.1991,ApJ,366,L69Johnstone,R.M.,Naylor,T.,Fabian,A.C.,1991,MNRAS,248,18 Koekemoer,A.M.,O’Dea,C.P.,Sarazin,C.L.,McNamara,B.R.,Donahue,M., Voit,G.M.,Baum,S.A.,&Gallimore,J.F.1999.ApJ,in pressMcCarthy,P.J.1993,ARAA,31,639McNamara,B.R.1997,in Galactic and Cluster Cooling Flows,ed.N.Soker(San Francisco:PASP),109McNamara,B.R.&O’Connell,R.W.1989,AJ,98,2018McNamara,B.R.,&O’Connell,R.W.1993,AJ,105,417McNamara,B.R.,Wise,M.,Sarazin,C.L.,Jannuzi,B.T.,&Elston,R.1996,ApJ, 466,L9McNamara,B.R.,Jannuzi,B.T.,Sarazin,C.L.,Elston,R.,&Wise,M.1996,ApJ, 469,66McNamara,B.R.,Jannuzi,B.T.,Sarazin,C.L.,Elston,R.,&Wise,M.1999,ApJ, 518,167Murphy,B.W.,&Chernoff,D.F.1993,ApJ,418,60Pinkney,J.,et al.1996,ApJ,468,L13Porter,A.C.,Schneider,D.P.,&Hoessel,J.G.1991,AJ,101,1561Sarazin,C.L.,Burns,J.O.,Roettiger,K.,&McNamara,B.R.1994,Ap J,447,559 Sarazin,C.L.,&Wise,M.W.1993,ApJ,411,55Schombert,J.M.1987,ApJS,64,643Shields,J.C.,Filippenko,A.V.1990,ApJ,353,L7Uson,J.M.,Boughn,S.P.,Kuhn,J.R.1991,ApJ,369,46van Breugel,W.,Stanford,A.,Dey,A.,Miley,G.,Stern,D.,Spinrad,H.,Graham, J.,McCarthy,P.1998,astro-ph/9809186van Breugel,W.,Heckman,T.,&Miley,G.1984,ApJ,276,79van Breugel,W.,Filippenko,A.,Heckman,T.,&Miley,G.1985,ApJ,293,83 Voit,G.M.,&Donahue,M.1997,ApJ,486,24211。
Color Dependence in the Spatial Distribution of Satellite Galaxies
a r X i v :0711.0989v 2 [a s t r o -p h ] 8 A p r 2008Astronomy &Astrophysics manuscript no.chen cESO 2008April 8,2008Color Dependence in the Spatial Distribution of Satellite GalaxiesJacqueline Chen 1Argelander-Institut f¨u r Astronomie,Universit¨a t Bonn,Auf dem H¨u gel 71,D-53121Bonn,Germany;jchen@astro.uni-bonn.de Preprint online version:April 8,2008ABSTRACTAims.We explore the color dependence of the radial profile of satellite galaxies around isolated parent galaxies.Methods.Samples of potential satellites selected from large galaxy redshift surveys are significantly contaminated by interlopers –objects not bound to the parent galaxy.We use the Sloan Digital Sky Survey to estimate the interloper fraction in samples of candidate satellite galaxies.Results.We show that samples of red and blue satellites have di fferent interloper populations:a larger fraction of blue galaxies are likely to be interlopers compared to red galaxies.Both with and without interloper subtraction,the radial profile of blue satellites is significantly shallower than that of red satellites.In addition,while red and blue primaries have di fferent interloper fractions,the slope of the corrected radial profiles are consistent after interloper correction.We discuss the implications of these results for galaxy formation models.Key words.cosmology:theory –dark matter –galaxies:formation –galaxies:fundamental parameters –galaxies:structure1.IntroductionIn cold dark matter (CDM)models of the universe,large num-bers of dark matter (DM)subhalos lie within the virial radius of larger dark matter halos.In some halos and subhalos,bary-onic material has cooled and formed stars,resulting in a central galaxy and satellite galaxies.The spatial distribution of satel-lite galaxies in galaxy-sized halos,then,reflects the evolution of satellite galaxies and the mass accretion history of their par-ent halos.For example,dark matter simulations suggest that the angular distribution of subhalos follows the shape of the DM halo,which is indicative of infall of subhalos along filaments (e.g.,Zentner et al.2005;Libeskind et al.2005),while observa-tions suggest that satellites lie along the major axis of the light distribution for early-type galaxies (e.g.,Sales &Lambas 2004;Brainerd 2005;Yang et al.2006;Azzaro et al.2007).In dark matter simulations,the radial distribution of subhalos is biased with respect to the density profile of DM halos;at small separations from the center of the halo (within ∼20−50%of the virial radius),the distribution of DM subhalos has a lower con-centration,but it follows the dark matter density profile at larger radii (Ghigna et al.1998;Col´ın et al.1999;Ghigna et al.2000;Springel et al.2001;De Lucia et al.2004;Diemand et al.2004;Gao et al.2004;Nagai &Kravtsov 2005;Macci`o et al.2006).Studies that include baryons,star formation and cooling,how-ever,show that the distribution of galaxies associated with sub-halos has a steeper inner profile than the subhalo distribution,both at cluster and at galaxy scales (Nagai &Kravtsov 2005;Macci`o et al.2006).For samples of subhalos selected by tidally-truncated mass,objects near the halo center lose a greater per-centage of their dark matter mass than objects near the virial radius.Stellar mass selected samples of satellite galaxies are resistant to this e ffect since baryonic components are located in the centers of dark matter subhalos and are tightly bound.The observed radial profile of satellites in galaxy-sized halos is generally more centrally concentrated than the subhalo dis-tribution (Chen et al.2006)but consistent with the dark matter profile (van den Bosch et al.2005;Chen et al.2006),although Sales &Lambas (2005)have contrary results.The color dependence in the spatial distribution of satellite galaxies has been studied extensively in angular distributions.The angular distribution of satellites is anisotropic and aligned with the major axis for red host galaxies and is consistent with isotropic for blue hosts (Yang et al.2006;Agustsson &Brainerd 2008;Azzaro et al.2007).These results are consistent with the picture where the orientation of the major axis of elliptical galax-ies is determined by the direction along which subhalos are falling into the galaxy –along filaments.The orientation of spiral galaxies is determined by the angular momentum vector,which,in simulations,has shown only poor alignment with the mi-nor axis of the dark matter halo.A secondary result found in these studies is that red satellites of red hosts show a stronger anisotropy than blue satellites.This result might be explained by a scenario where satellite color is determined by accretion time;red satellites were accreted earlier,while blue satellites represent recent infall.If the major axis of the host galaxy is set early and the orientation of infalling subhalos with respect to the host galaxy changes over time,then objects which accreted early would show a stronger alignment with respect to the host galaxy.These observational results have been analyzed by com-parison to mock galaxy catalogs derived from semi-analytic models –simulating galaxies using the mass accretion histo-ries of dark matter halos in a simulation and a variety of as-sumption about the physics of galaxy formation and evolution.Agustsson &Brainerd (2008)use mock galaxy catalogs to sug-gest that the di fference in degree of anisotropy between blue and red host galaxies is real and not due to di fferences in the inter-loper populations,while Kang et al.(2007)come to an oppo-site conclusion using a very di fferent technique –the color de-pendence is due to interlopers in the group catalog.Kang et al.(2007)also conclude that the di fference in alignment between red and blue satellites is due to the masses of the satellites;red satellites are larger and associated with subhalos which were2Jacqueline Chen:Color Dependence in the Spatial Distribution of Satellite Galaxiesmore massive at the epoch of accretion and which in simula-tions are accreted more preferentially along the major axis of the halo(Libeskind et al.2005;Wang et al.2005).The semi-analytic model of Agustsson&Brainerd(2008)suggests that the amount of interloper contamination is much greater for blue satellites than red satellites(∼50%to∼15%,defining inter-lopers as objects with a physical separation greater than500kpc from the parent galaxy),although the color dependence in the anisotropy of satellites is still observed after subtracting inter-lopers.The color dependence in the radial distribution of satel-lites has been studied by Sales et al.(2007)using semi-analytic galaxy catalogs constructed from the Millennium Simulation. They note that the radial distribution of red satellites is signif-icantly more centrally concentrated than the distribution of blue satellites.This trend is attributable to accretion time in a scenario similar to one explanation for greater anisotropy in the angular distribution of red satellites compared to blue satellites:early ac-creting satellites are stripped of hot gas,stop forming stars,and become redder.Early accreters are also likely to orbit closer to the center.Observationally,the effects of color selection on the radial distribution of satellite galaxies has been previously examined by Sales&Lambas(2004)using data from the Two Degree Field Galaxy Redshift Survey(2dFGRS).Theyfind a steeper outer slope for satellites of blue parent galaxies than for red pri-maries.In addition,red satellites have a distribution that requires a larger core radius than blue satellites and they attribute this result to correlations between the primary and satellite prop-erties.However,in observational searches for satellites,candi-date satellites are chosen based upon their projected separation and velocity difference from the parent galaxy.Samples of ob-jects chosen in this manner are heavily contaminated by ob-jects which are not satellites–interlopers.In Chen et al.(2006), we discussed the importance of interlopers in making relatively unbiased estimates of the projected radial distribution of satel-lite galaxies and developed a reliable method of interloper sub-traction.Other methods to account for interlopers have been investigated.For example,van den Bosch et al.(2004)exclude interlopers using an iterative,adaptive selection criterion for satellites,and Chen(2008)model satellites and interlopers to-gether using a halo occupation distribution(HOD)based ana-lytic model for the galaxy correlation function.The Chen et al. (2006)approach has the advantage that it may easily be com-pared to previous results.In this paper,we add color selection to the interloper estima-tion method described in Chen et al.(2006)and apply it to data from the Sloan Digital Sky Survey(SDSS)spectroscopic sam-ple.We show that samples of red and blue candidate satellites have different levels of contamination by interlopers,while the model interloper samples for red and blue primaries are similar in color distribution.We discuss these results in terms of the en-vironmental dependence of satellite galaxies and wider applica-tions to the angular distribution of satellite galaxies and to galaxy formation models.2.Observational Data2.1.The Sloan Digital Sky SurveyThe Sloan Digital Sky Survey(SDSS)includes imaging of the the northern Galactic cap infive bands,u,g,r,i,z,down to r∼22.5using a dedicated2.5m telescope at Apache Point Observatory in New Mexico in addition to spectro-scopic observations for a subsample of objects from the imag-ing catalog(York et al.2000;Fukugita et al.1996;Gunn et al. 1998;Hogg et al.2001;Smith et al.2002;Strauss et al.2002; Blanton et al.2003;Gunn et al.2006;Tucker et al.2006).The SDSS spectroscopy is carried out using opticalfibers posi-tioned in pre-drilled holes on a circular plate,with mini-mum separation betweenfibers of55′′,thefiber collision dis-tance.Reobservations of afield can result in observed spec-tra with separations less than thefiber collision distance,down to thefiber diameter of3′′.The spectroscopic targets are se-lected with r-band Petrosian magnitudes r≤17.77and r-band Petrosian half-light surface brightnessesµ50≤24.5mag arcsec−2.An automated pipeline measures the redshifts and classifies the reduced spectra(Stoughton et al.2002;Pier et al. 2003;Ivezi´c et al.2004,D.J.Schlegel et al.2008,in prepara-tion).1We use the spectroscopic Main galaxy catalog available as Data Release Six(DR6;Adelman-McCarthy et al.2008),cov-ering an area of7425deg2.Because the SDSS spectroscopy is taken through circular plates with afinite number offibers offinite angular size,the spectroscopic completeness varies across the survey area.The resulting spectroscopic mask is represented by a combination of disks and spherical polygons (Tegmark et al.2004).Each polygon also contains the com-pleteness,a number between0and1based on the fraction of targeted galaxies in that region which were observed.We ap-ply this mask to the spectroscopy and include only galaxies from regions where completeness is at least90%.We use r-band magnitudes in DR6,built from the NYU Value-Added Galaxy Catalog(Blanton et al.2005),normalized to h=1,such that M r=M0.1r−5log10h,where M0.1ris the absolute magni-tude K-corrected to z=0.1(kcorrect v4.1.4)as described in Blanton&Roweis(2007).2.2.Volume-Limited Galaxy SamplesFollowing the procedure in Chen et al.(2006),we use a volume-limited galaxy sample with a depth of13,500km s−1,corre-sponding to the limiting redshift of z=0.045.This limit is cho-sen as a trade-offbetween the volume of the sample and the ab-solute magnitude limit for our satellites,which would need to be decreased to brighter magnitudes for more distant primaries.In addition,the limiting magnitude sets a minimum separation at whichfiber collisions become important,which increases with distance.To include more distant primaries we would have to sacrifice the ability to probe density distributions at small sepa-rations.From our volume-limited sample,we construct a primary sample of isolated host galaxies and a sample of potential satel-lites that are projected close to primaries and refer to these two samples as the primary sample and the satellite sample.Isolated host galaxies are chosen in order to reduce the number of galaxy groups selected and eliminate the contamination from satellites of galaxy group members.Parameters for the criteria follow Prada et al.(2003)and Chen et al.(2006)and are listed in Table 1.The isolation criterion requires that a primary have only neigh-bors at least two magnitudes fainter within∆R=0.5h−1MpcJacqueline Chen:Color Dependence in the Spatial Distribution of Satellite Galaxies3 Table1.Selection&Isolation CriteriaParameters Valueand∆V=1000km s−1.2Potential satellites of any isolated pri-mary must be at least2magnitudes fainter than the primaryand withinδr=0.5h−1Mpc andδv=500km s−1.The max-imum absolute magnitude for satellites is set by M r,lim−5log h=17.77−DM−K0.1in the r-band,where the17.77is thefluxlimit in this band,DM is the distance modulus,and K0.1is theK-correction at z=0.1.This gives a limiting absolute magnitudeof M r=−17.77.The satellites are thus limited to the bright-est satellite galaxies,∼0.1L∗.In order to avoid biasing froma deficit of close pairs of objects,the minimum separation be-tweenfibers is32.9h−1kpc–thefiber collision separation at theredshift of the furthest point in our sample.Finally,we choosegalaxies that are in areas that are at least90%complete.For therange−23<M r<−20,there are1602primary galaxies and690objects in the satellite sample with projected radii greaterthan the minimum separation and less than0.5h−1Mpc.2.3.Interloper SubtractionThere is a fraction of objects in our satellite samples that arenot gravitationally bound to the primaries but are included inthe sample because of projection effects.Throughout this pa-per,we call such objects interlopers(in turn,satellites sampleswithout interloper contamination are referred to as true satellitesamples).Interloper subtraction is discussed in greater detail inChen et al.(2006).In semi-analytic galaxy catalogs and in dark matter simula-tions,interlopers are significant in samples of satellites or DMsubhalos.For example Agustsson&Brainerd(2008)find in asemi-analytic galaxy catalog constructed from the MillenniumSimulation–using a stricter isolation criteria than we use–thatthe interloper contamination fraction is∼30%.In DM-only sim-ulations,the fraction of interlopers as a function of projected ra-dius rises from a few percent at R∼50h−1kpc to nearly100%atR=0.5h−1Mpc(see Fig.1,Chen et al.2006).Chen et al.(2006)developed and tested several methods ofsubtracting interlopers from the satellite sample statistically.Theprojected surface density of candidate satellites, Σ(R) sat is es-timated in bins and normalized by the total number of primarygalaxies in the sample.Simple methods of interloper subtractionestimate the corresponding projected surface density in interlop-3In our data there is a total of3325model interlopers.4Jacqueline Chen:Color Dependence in the Spatial Distribution of SatelliteGalaxiesFig.1.The color distribution for the primary sample (solid),satellite sample (dotted),and model interloper sample (dashed).Red satellites are defined to have M g −M r >0.65,while red primaries have colors M g −M r >0.9.Poisson errors are shown for the primary and satellite sample.The corresponding Poisson errors for the model interloper sample are significantly smaller than for the satellite sample.our parent galaxy halos,making fitting for the scale radius im-practical.3.ResultsThe M g −M r color distributions for theprimary,satellite,and model interloper samples are shown in Fig.1.The primary sam-ple is redder than the satellite sample with a large red peak and a bluer tail,while the satellite sample has a vaguely bimodal distri-bution,split at M g −M r ∼0.65.The model interloper distribution appears tilted to bluer objects compared with the distribution of the satellite sample.3.1.Red and Blue SatellitesThe di fferences in color distribution between the satellite sample and the model interloper sample suggests that a greater fraction of blue objects in the satellite sample will be interlopers than of red objects.Defining red satellites to M g −M r >0.65,of the 690objects in the satellite sample,285are red and 405are blue.By comparison to the model interloper sample,the fraction of the objects that are interlopers in the whole sample is 24%,while the percentages for the red and blue satellite sample are 16%and 30%,respectively.This di fference a ffects the estimates of the slope of the power-law fit to the radial distribution.In Figure 2,the biasing of the best-fit power-law is shown for all satellites and blue and red satellites.The distribution of blue satellites is significantly flatter than the distribution of red satellites,and the distribution of blue satellites is more e ffected by interlopers than the distribution of red satellites.In the interloper-contaminated satellite samples,the best-fit slopes for blue and red satellites (shown in Table 2and Figure 3)are inconsistent,with a much shallower slope for blue satellites.After interloper subtraction,although the di fference in slopes isFig.2.The surface density of the satellite sample (squares)and the interloper-subtracted satellite sample (circles)for all satel-lites (top),blue satellites (center),and red satellites (bottom).The best fit power-laws are also plotted with solid lines for the satellite sample and dotted lines for the interloper-subtracted sample.smaller,the blue slope is still significantly shallower than the red slope,and the slopes are inconsistent given both the marginal-ized errors and their 1σconfidence intervals (see Fig.3).The projected radial distribution of red satellites is as steep as might be expected for the dark matter density distribution in halos which host galaxies such as are found in our primary sample.The distribution of blue satellites,on the other hand,is as shal-low as might be expected for the subhalo distribution in the same host halos (see,for comparison,Chen et al.2006).Jacqueline Chen:Color Dependence in the Spatial Distribution of Satellite Galaxies5 Table2.Best-Fit Power-Law Slopes to the Surface DensityProfile of SatellitesInput Data Satellite Sample InterloperSubtracted Fig.3.The68%confidence intervals for the normalization(A)and slope(α)of the power-law,Σ(R)=ARα.Shaded contours are for blue satellites;solid line contours are for red satellites. The contour for interloper-subtracted samples are in each case atsteeper slopes than their satellite sample counterparts.In general,we expect blue satellites to be fainter than red satellites,so it is useful to attempt todisentangle the effects of luminosity from those of color.While Chen et al.(2006)tests luminosity dependence(which is further discussed in Section 4),they do not have sufficient numbers of objects to test color and luminosity together.Figure4shows the luminosity distri-butions for the blue and red satellites and primaries.We split each color sample of satellites into a faint and a bright sample at M r=−18.3.Table2shows that the interloper-subtracted sur-face density profiles of faint and bright samples are consistent for color-selected samples,i.e.,color is the dominant attribute in the radial distribution of satellites.3.2.Red and Blue PrimariesWe split the1602primary galaxies by color at M g−M r=0.9. The591red primary galaxies have358objects in their satellite sample,while the1011blue primary galaxies have332candi-Fig.4.The luminosity distribution with corresponding Poisson errors of primaries(top)and objects in the satellite sample(bot-tom),where red objects and blue objects are shown separately, in dotted and solid lines respectively.Bright and faint satellites are split at M r=−18.3,while bright and faint primaries are split at M r=−20.8.date satellites.The interloper percentages for the red and blue primary sample are21%and28%,respectively.Red primaries of M g−M r>0.9have a larger fraction of red objects in their satellite sample than than blue primaries(52%to30%)and,sub-sequently,more true satellites(see Fig.5).This can also be seen in Fig.6,where the amplitude of the satellite profile of red pri-maries is greater than that of the blue primary profile.After inter-loper subtraction,the slopes of the density profiles of satellites for blue and red primaries are consistent.Figure4additionally shows that red primaries are in gen-eral brighter than blue primaries.Splitting the color-selected pri-maries into faint and bright samples at M r=−20.8,Table2 shows significant differences for the distribution of satellites of faint and bright primaries,depending on the color of the primary. Red,brighter primaries have an interloper-subtracted satellite profile that is shallower than the distribution of satellites in red, fainter primaries,α=−1.34±0.11toα=−1.82+0.15−0.14.This is not unexpected;from numerical simulations we expect the satel-lite distribution to scale with the mass distribution of the primary,6Jacqueline Chen:Color Dependence in the Spatial Distribution of SatelliteGalaxiesFig.5.The color distribution for the satellite sample and model interloper sample for red primaries (thick and thin solid lines,respectively)and the color distribution for the satellite sample and model interloper sample for blue primaries (thick and thin dashed lines,respectively).and larger primaries have smaller concentrations leading to shal-lower slopes at the radii at which we measure.Interestingly,blue primaries show the opposite dependence:the distribution of satellites around brighter primaries seems to have a steeper slope than around fainter primaries.This trend is accompanied by a change in the fraction of red satellites:38%of the satel-lite sample for bright,blue primaries are red,while 23%of the corresponding sample for faint,blue primaries are red.The er-rors,however,are nearly as large as the discrepancy between the samples,and larger samples will be required to confirm this re-sult.In addition,the slope for satellites of bright,blue primaries is steeper than that for bright,red primaries,which may reflect color dependence in using luminosity as a proxy for mass.Similar results can be seen if wesplit both the primary and the satellite samples by color.In Table 3,we show that red satel-lites around blue primaries have a larger interloper fraction and a steeper power-law slope than found for red satellites of red primaries.The di fference in power-law slope can be attributed to the same luminosity dependence discussed previously:red primaries are more massive and have mass and satellite dis-tributions described by smaller concentrations.Interloper con-tamination increases from small separations to large,and for larger primaries,the range of radii we probe preferentially sam-ple areas with smaller level of interloper contamination.On the other hand,for blue satellites of red and blue primaries,we find very similar interloper fractions and shallow power-law slopes.Unfortunately,in all cases,statistics are poor and future larger samples will be required for statistical significance to be as-cribed.3.3.Environmental DependenceDespite the di fference in satellite samples,the color distribution of the model interloper samples of red and blue primaries are similar to each other and to the color distribution of the satel-lite sample of blue primaries (see Fig.5).Restated,although red primaries are found in more clustered environments,the envi-Fig.6.The surface density of the satellite sample (squares)and the interloper-subtracted satellite sample (circles)for satellites of blue primaries (top)and of red primaries (bottom).The best fit power-laws are also plotted with solid lines for the satellite sample and dotted lines for the interloper-subtracted sample.Table 3.Primaries and Satellites Selected by ColorInput DataNumber in Interloper Power-Law (Primary-Satellite)Satellite SampleFractionSloperonments of red and blue primaries are not noticeably di fferent as measured by the color distribution of model interlopers.Red primaries live in more clustered environments than blue primaries as measured in the average surface number density of model interlopers:0.159h 2Mpc −2for red primaries and 0.117h 2Mpc −2for blue primaries.In the previous section,we noted that these slightly di fferent environments have similar color dis-tributions of faint objects,even though we do not use fixed lumi-nosity criteria for these objects.We test one less clustered envi-ronment,sampling isolated points that lie outside of a 2h −1Mpc radius from our isolated galaxies (outside the criteria for our in-terloper subtraction method).Here,the average surface density is 0.042h 2Mpc −2.All three model interloper distributions are plotted in Fig.7,in which all environments have similar colorJacqueline Chen:Color Dependence in the Spatial Distribution of Satellite Galaxies7Fig.7.The color distribution for the model interloper samples for red primaries (solid),blue primaries (dashed),and for sam-pling isolated points greater than 2h −1Mpc from primaries (dot-ted).The average surface number density of objects in these samples is 0.159,0.117,and 0.0415h 2Mpc −2.dependences with a possible trend to bluer objects in less dense fields.4.ConclusionsIn Chen et al.(2006),we constrained the projected radial dis-tribution for isolated galaxies and found that their power-law slopes are steeper than the expected slopes for the distribution of dark matter subhalos and may be as steep as the density pro-file of the host dark matter halos.We reproduce this result with a survey area that is ∼50%larger.However,the distribution of satellite galaxies shows some significant dependence on color.When samples of candidate satellites are split by color,we see that blue objects are more likely to be interlopers than red objects.The observed estimated interloper contamination of red objects in the satellite sample is 16%,while that of blue ob-jects is 30%.Agustsson &Brainerd (2008)produce a di ffer-ence of 15%and 50%,respectively,in the interloper contami-nation of red and blue satellites samples,using a semi-analytic galaxy formation model.It is,then,not appropriate to assume that red and blue satellites have the same level of interloper con-tamination and this must be taken into account in testing the color dependence in the angular distribution of satellite galaxies.This result also suggests that the prevalence of faint red galax-ies could be developed into a method to find small groups of galaxies in a manner similar to the method by which the red se-quence of early-type galaxies is used to find galaxy clusters (e.g.,Gladders &Yee 2000).Both with and without interloper subtraction,the radial pro-file of blue satellites is significantly shallower than that of red satellites.The best-fit power-law slope of interloper-subtracted blue satellites is α=1.34±0.10and the that of red satellites is α=−1.80±0.10.This result is consistent with the trend found by Sales et al.(2007)who test color dependence in semi-analytic galaxy catalogs constructed from the Millennium Simulation,finding a more centrally concentrated radial distribution of red satellites than of blue satellites.This trend is repeated when theyselect by a proxy for accretion time –whether the satellite retains a DM halo or whether it has been tidally destroyed.Blue satellites are generally fainter than red satellites.Correspondingly,Chen et al.(2006)found that the best-fit power-law slope for bright satellites is steeper than that of faint satellites (cut at M r =−18.28)–although without statistical sig-nificance –in a volume-limited sample.In a flux-limited sample,the reverse relation was found;however,this sample used a lu-minosity cut as faint as the faintest satellites in our sample,at M r =−17.76,and so is not directly comparable.When we cut the satellites samples by color and luminosity,we find that the dominant e ffect is from color;the best-fit power-law slopes of bright and faint samples of red satellites are consistent as are those of bright and faint blue satellites.The power-law slope for red satellites is as steep as might be expected for the dark matter density distribution of halos which host galaxies like those found in our primary sample.On the other hand,the power-law slope for blue satellites is as shal-low as the expected subhalo distribution.The shallowness of the subhalo profile is attributed to tidal stripping.This is unlikely to e ffect satellite galaxies,since they are located at the centers of the dark matter subhalos.The shallower profile for blue satel-lites as compared to red satellites,then,might be interpreted as consistent with the scenario where satellite color is determined by accretion time:red satellites were accreted earlier,while blue satellites represent more recent infall.Satellites are expected to be redder in the inner regions of parent halos due to environ-mental processes that shut o ffstar formation (ram pressure strip-ping,strangulation,etc.).This morphological segregation has also been observed in more massive clusters and in the fainter satellites found in the Local Group.When splitting the satellite sample by primary color,red primaries have a significantly larger fraction of red satellites and somewhat smaller interloper fractions than blue primaries.After interloper subtraction,the best-fit power-law slopes of satellites of red and blue primaries are consistent within errors.Red primaries are,on average,more luminous than blue galax-ies.Correspondingly,Chen et al.(2006)found that the best-fit power-law slope for satellites bright primaries is consistent with that of satellites of faint primaries (cut at M r =−21).When we cut the primary sample by luminosity and color,the trend with luminosity is di fferent for red and blue primaries.The slope of satellites of bright,red primaries is shallower than that of faint,red primaries,a relation probably dominated by the mass of the parent halo as brighter primary galaxies reside in bigger par-ent halos which have mass distributions characterized by smaller concentrations.The slope of satellites of bright,blue primaries is steeper than that of faint,blue primaries,as the fraction of red satellites drops with primary luminosity.The color distribution of objects in model interloper sam-ples are similar,regardless of their environment (as measured by average surface density).While robust conclusions cannot be drawn as to how similar these distributions are,it suggests that there are fundamental di fferences between satellite galax-ies and faint galaxies in the field.Intriguingly,this blue-tilted color distribution also resembles that of the satellites sample of blue primaries.Better understanding of the processes that e ffect the color of faint objects and the radial distribution of satellite galaxies will require further studies.Acknowledgements.We would like to thank Andrey Kravtsov,Francisco Prada,Michael Blanton,and Erin Sheldon for their suggestions and invaluable contri-butions to understanding interloper contamination and systematics in the data.In addition,we would like to acknowledge the anonymous referee for many helpful suggestions.。
Star Formation in Clusters
a rXiv:as tr o-ph/4821v111Aug24Star Formation in Clusters By S ØR E N S.L A R S E N ESO /ST-ECF,Karl-Schwarzschild Strasse 2,D-85748Garching bei M¨u nchen,Germany The Hubble Space Telescope is very well tailored for observations of extragalactic star clusters.One obvious reason is HST’s ability to recognize clusters as extended objects and measure sizes out to distances of several Mpc.Equally important is the wavelength range offered by the instruments on board HST,in particular the blue and near-UV coverage which is essential for age-dating young clusters.HST observations have helped establish the ubiquity of young massive clusters (YMCs)in a wide variety of star-forming environments,ranging from dwarf galaxies and spiral disks to nuclear starbursts and mergers.These YMCs have masses and structural properties similar to those of old globular clusters in the Milky Way and elsewhere,and the two may be closely related.Several lines of evidence suggest that a large fraction of all stars are born in clusters,but most clusters disrupt rapidly and the stars disperse to become part of the field population.In most cases studied to date the luminosity functions of young cluster systems are well fit by power-laws dN (L )/dL ∝L −αwith α≈2,and the luminosity of the brightest cluster can (with few exceptions)be predicted from simple sampling statistics.Mass functions have only been constrained in a few cases,but appear to be well approximated by similar power-laws.The absence of any characteristic mass scale for cluster formation suggests that star clusters of all masses form by the same basic process,without any need to invoke special mechanisms for the formation of “massive”clusters.It is possible,however,that special conditions can lead to the formation of a few YMCs in some dwarfs where the mass function is discontinuous.Further studies of mass functions for star clusters of different ages may help test the theoretical prediction that the power-law mass distribution observed in young cluster systems can evolve towards the approximately log-normal distribution seen in old globular cluster systems.rsen:Star Formation in Clustersexcellent blue and UV quantum efficiency have become available,a better understanding of“dome”seeing has led to improved image quality,and8–10m ground-based telescopes have made it possible to obtain high-quality spectra of faint objects detected in HST images.2.Why star clusters?While star clusters have been the subject of substantial interest for many years,it may be worth recalling some of the main motivations for studying them.First,there are a number of problems which make clusters interesting in their own right. These involve both their formation,subsequent dynamical evolution and ultimate fate. Atfirst glance,clusters appear deceptively simple:they are aggregations of a few hundred to about a million individual stars,generally constituting a gravitationally bound system (although the latter may not be true for some of the youngest systems).Yet,constructing realistic models of their structure and dynamical evolution has proven to be a major challenge,and it is only now becoming possible to carry out reasonably realistic N-body simulations including the effects of stellar evolution,external gravitationalfields,and the rapidly varying gravitational potential in the early phases of cluster evolution during which gas is expelled from the system(Joshi et al.2000;Giersz2001;Kroupa&Boily 2002).The models must be tested observationally,and HST data currently represent the only way to reliably measure structural parameters for extragalactic star clusters. Second,there is growing evidence that a significant fraction of all stars form within clusters,although only a small fraction of these stars eventually end up in bound clusters (Lada&Lada2003;Fall2004).Therefore,the problem of understanding star formation is intimately linked to that of understanding cluster formation,and a theory of one cannot be complete without the other.It is of interest to investigate how the properties of star clusters might depend on environment,as this might provide important clues to any differences in the star formation process itself.In particular,HST has made important contributions towards establishing the presence of“young massive clusters”(YMCs†)in a variety of environments,which appear very similar to young versions of the old globular clusters which are ubiquitous around all major galaxies.Globular cluster formation was once thought to be uniquely related to the physics of the early Universe(e.g.Peebles& Dicke1968;Fall&Rees1985)but it now seems to be an ongoing process which can be observed even at the present epoch.Third,star clusters are potentially very useful as tracers of the stellar populations in their host galaxies.Clusters can be identified and studied at much greater distances than individual stars.In most cases,they are composed of stars which,to a very good approximation,formed at the same time and have the same metallicity.This is in contrast to the integrated light from the galaxies,which may originate from an unknown mix of stellar populations with different ages and metallicities.Although the effects of stellar evolution alone cause a cluster to fade by5-6magnitudes(in V-band)over10Gyrs (Bruzual&Charlot2003),it is in principle possible to detect clusters which have formed during the entire lifetime of galaxies,out to distances of several Mpc.In particular, globular clusters have been used extensively in attempts to constrain the star formation histories of early-type galaxies.†Also known as“super star clusters”,“populous clusters”or“young globular clusters”.rsen:Star Formation in Clusters3 3.HST and Extragalactic Star ClustersHST is almost ideally tailored for studies of extragalactic star clusters.Three main reasons for this are:•Angular resolution:clusters typically have half-light radii of2–4pc(see Section5), and can thus be recognized as extended objects out to distances of10–20Mpc with the ∼0′′.05resolution offered by WFPC2or ACS.With careful modeling of the point spread function(PSF)this limit may be pushed even further.•Field size:At10Mpc,the ACS200′′×200′′field-of-view corresponds to about10 kpc×10kpc,making it possible to cover a significant fraction of a typical galaxy in a single pointing.•Spectral range:For studies of young stellar populations,optical and near-UV spectral coverage is essential,as discussed below.There is currently no alternative to HST on the horizon which offers a similar combi-nation of capabilities.The James Webb Space Telescope(JWST),while offering vastly improved efficiency in the IR,will offer no significant gain in resolution over HST,and will be limited to longer wavelengths.Ground-based adaptive optics(AO)can provide similar,or even better resolution than HST,but only within a small(∼20′′)isoplanatic field of view.Furthermore,AO lacks the stable PSF of HST which is critical for many purposes(e.g.when measuring structural parameters for star clusters at the limit of the resolution),and is in any case limited to the IR(at least for now).The GALEX mission offers wide-field UV imaging,but with a spatial resolution that is inferior by far to that of HST(about5′′).The need for optical and near-UV imaging in particular deserves some additional com-ments.Figure1shows simple stellar population(SSP)model calculations(Bruzual& Charlot2003)for the evolution of the U−B,B−V and V−K broad-band colors of a single-burst stellar population.The models are shown for metallicities Z=0.02(Solar) and Z=0.004between ages of106years and1010years.As seen from thefigure,the U−B color is an excellent age indicator in the range from107to a few times108years, increasing by more than0.5mag and with little metallicity dependence over this age range.The B−V color,in contrast,remains nearly constant over the same age range, and offers little leverage for age determinations.In practice,there are complicating prob-lems such as dust extinction,which in general will make it difficult to obtain accurate age estimates from a single ing a combination of two colors,such as U−B and B−V,will make it possible to constrain both age and reddening,while at the same time being relatively insensitive to metallicity effects.The relation between age and location of a cluster in the(U−B,B−V)two-color plane has been calibrated with clusters in the Large Magellanic Cloud through the so-called‘S’-sequence(Elson&Fall1985;Girardi et al.1995).For ages younger than about107years,line emission becomes important(An-ders&Fritze-v.Alvensleben2003),while the age-metallicity degeneracy(Worthey1994) becomes a difficulty at older ages.A more recent discussion of photometric age indica-tors,with emphasis on the importance of blue and UV data,is in Anders et al.(2004b). Use of e.g.the V−K color can help put further constraints on the metallicity and may also help constrain the ages of stellar populations in the range∼200Myr to∼500Myr (Maraston et al.2002),although the models are more uncertain and depend strongly on the stellar evolutionary tracks used in the construction of the SSP models(Girardi2000). High-resolution,wide-field imaging in the blue and/or UV will be especially important for attempts to constrain not only the luminosity function,but also the mass function of clusters.For a long time,WFPC2was the“workhorse”on HST,and it remains the only wide-field imager on board HST with U-band imaging capability.However,the sensitivityrsen:Star Formation in ClustersFigure1.Evolution of broad-band colors as a function of age and metallicity according to Bruzual&Charlot(2003)simple stellar population models.of WFPC2in the U-band is rather low and the detectors are steadily degrading.The Wide Field Camera3,with its panchromatic coverage,would be an ideally suited instrument for such studies.4.Setting the stage:early developmentsEven within the Local Group,it has long been known that the traditional distinction between open and globular clusters that can be applied fairly easily in the Milky Way breaks down in some other galaxies.The classical example is the“blue globular”clusters in the Large Magellanic Cloud,which are not easily classified as either open or globular clusters.The most massive of these objects have masses up to∼105M⊙(Elson&Fall 1985;Fischer et al.1992;Richtler1993;Hunter et al.2003)similar to the median mass of old globular clusters and about an order of magnitude more massive than any young open cluster known in the Milky Way.Yet,these objects have young ages,and are still being produced today by the LMC.Similar clusters have been found in M33(Christian &Schommer1982,1988).A good indication of the status of research in extragalactic young star clusters shortly prior to HST is provided by Kennicutt&Chu(1988;hereafter KC88).These authors compiled observations of what they refer to as“young populous clusters”(PCs)in14 galaxies for which data was available at that time.Half of the galaxies studied by KC88 were Local Group members(Milky Way,LMC,SMC,M33,M31,NGC6822and IC1613). As noted by KC88,a severe difficulty in comparing observations of PCs in different galaxies,made by different authors,is the widely variable completeness of the surveys, and the different definitions of such clusters.KC88adopted a(somewhat arbitrary) definition of a young PC as an object with an estimated mass>104M⊙and a colorrsen:Star Formation in Clusters5 B−V<0.5.They noted a conspicuous deficiency of populous clusters in the Milky Way and M31,the two only large Sb/Sbc-type spirals in the sample,and suggested that this might be linked to the deficiency of giant H ii regions in the same two galaxies.By comparing the relative numbers of PCs and giant H ii regions in their sample of galaxies, KC88concluded that PCs may indeed form inside such regions,but not all giant H ii regions produce bound clusters.This is very much in line with recent indications that only a small fraction of star clusters of any mass remain bound(Fall2004).The galaxies which did contain PCs were all late-type,though not all late-type galaxies were found to contain PCs.A significant exception is the Local Group dwarf irregular IC1613,which contains few if any star clusters at all(van den Bergh1979;Hodge1980)in spite of some on-going star formation.The near-absence of star clusters in IC1613may be as important a clue to the nature of the cluster formation process as the abundant cluster systems in starbursts and merger galaxies(Section5).To a large extent,research in old globular clusters(GCs)remained detached from that of YMCs until fairly recently.It was well-known that early-type galaxies typically have many more GCs per unit host galaxy luminosity(Harris&van den Bergh1981)than spirals and irregulars,a fact that was recognized as a problem for the idea that early-type galaxies form by mergers of gas-rich spirals(van den Bergh1982).Schweizer(1987) proposed that this problem might be solved if new GCs form during the merger.This idea was further explored by Ashman&Zepf(1992),who predicted that the resulting merger product should contain two distinct GC populations:one metal-poor population inherited from the progenitor galaxies,and a new metal-rich population formed in the merger.The two GC populations should be identifiable in the color distributions of the resulting GC systems.Two highly influential discoveries soon followed:Bimodal color distributions were discovered in several GC systems around early-type galaxies(Zepf &Ashman1993;Secker et al.1995;Whitmore et al.1995),and highly luminous,com-pact young star clusters were found in ongoing or recent mergers like the Antennae and NGC7252(Whitmore et al.1993;Whitmore&Schweizer1995).In retrospect,it had al-ready been known for a long time that even the metallicity distribution of the Milky Way GC system is strongly bimodal(Zinn1985).The mean metallicities of the two modes in the Milky Way are in fact quite similar to those seen in early-type galaxies.The Milky Way is unlikely to be the result of a major merger,and there are also other indications that not all properties of GC systems in early-type galaxies can be explained by a naive application of the merger model.Alternative scenarios have later been put forward to ex-plain the presence of multiple GC populations(e.g.Forbes et al.1997;Cˆo t´e et al.1998), but it is beyond the scope of this paper to discuss any of these in prehensive discussions can be found e.g.in Harris(2001)and Kissler-Patig(2000).Nevertheless,the discovery of young globular cluster-like objects in ongoing mergers was a tantalizing hint that it might be possible to study the process of globular cluster formation close-up at the present epoch,and not just from the fossil record.5.Extragalactic Star Clusters in Different EnvironmentsTables1–5are an attempt to collect a reasonably complete list of galaxies where YMCs have been identified(until∼May2004),along with some pertinent references.For each galaxy,the main facilities used for the observations are listed,although in many cases it is impossible to give a comprehensive listing.Standard abbreviations(ACS,FOC,GHRS, STIS,NICMOS,WFPC,WFPC2)are used for HST instruments.Other abbreviations are WIYN(Wisconsin Indiana Yale NOAO3.5m),UKIRT(United Kingdom Infra-Red Telescope),NTT(ESO3.5m New Technology Telescope),CFHT(3.6m Canada-Francersen:Star Formation in ClustersHawaii Telescope),DK154(Danish1.54m at ESO,La Silla)and NOT(2.56m Nordic Optical Telescope).The level of detail provided in different studies varies enormously–in some cases,identifications of YMCs are only a byproduct of more general investigations of galaxy properties(e.g.Meurer et al.1995;Scoville et al.2000)while other studies are dedicated analyses of cluster systems in individual galaxies.Galaxies marked with an asterisk(⋆)are used later(Section6.2)when discussing luminosity functions.In the following I briefly discuss a few illustrative cases from each table and then move on to discuss more general properties of young cluster systems.5.1.Starburst galaxiesThe richest populations of YMCs are often found in major mergers(Section5.2).How-ever,there are also examples of YMCs in starbursts which are not directly associated with major mergers,although they may in some cases be stimulated by more benign interactions or accretion of companion satellites(Table1).In the case of NGC7673, for example,Homeier&Gallagher(1999)argue that the morphological features of the galaxy point toward a minor merger,while the starburst in M82may have been triggered by tidal interactions with M81.M82is also noteworthy for being thefirst galaxy in which the term‘super star cluster’was used.It was introduced by van den Bergh(1971), who was careful to point out that the nomenclature was not intended to imply that these objects are necessarily bound.The presence of SSCs in M82was confirmed by O’Connell et al.(1995)who identified about100clusters in WFPC images.An example of a starburst which is unlikely to be triggered by an interaction is NGC5253,which is located about600kpc from its nearest neighbor,M83(Harris et al.2004).One of thefirst surveys to provide a systematic census of star clusters in a sample of starburst galaxies was the work by Meurer et al.(1995),who observed9galaxies with HST’s Faint Object Camera(FOC).Meurer et al.noted that a high fraction,on average about20%,of the UV luminosity in these starbursts originated from clusters or compact objects,and a hint of a trend for this fraction to increase with the underlyingrsen:Star Formation in Clusters7UV surface brightness.They also measured cluster sizes similar to those of Galactic globular clusters,and found the luminosity functions to be well represented by a power-law dN(L)/dL∝L−αwithα≈2.YMCs have been identified in several nuclear and circumnuclear starburst regions,often associated with barred spiral galaxies(Table2).Maoz et al.(1996)studied5circumnu-clear starbursts and found that as much as30%–50%of the UV light came from compact, young star clusters with half-light radii<5pc and estimated masses up to about105M⊙. Again they found the luminosity functions to be well approximated by power-laws with slopeα≈2.Buta et al.(2000)found a much steeper slope(α=3.7±0.1)in their study of the circumnuclear starburst in NGC1326,but noted that their sample might be contaminated by individual supergiant stars.In some cases the ring-like structure of the nuclear starburst is not quite so evident.Watson et al.(1996)discovered4lu-minous clusters in the central starburst region of NGC253,the brightest of which has M V=−15,an inferred mass in excess of1.5×106M⊙and a half-light radius of2.5pc. However,these clusters may be part of a compact ring-like structure with a radius of about50pc(Forbes et al.2000).Most of the clusters in the nuclear starburst of M83 are also located within a semicircular annulus(Harris et al.2001),but again the ring is more poorly defined.5.2.MergersMany of the most spectacular YMC populations have been found in merger galaxies. NGC1275was one of thefirst galaxies in which HST data confirmed the existence of YMCs,although at least one object in this galaxy was already suspected to be a massive cluster based on ground-based data(Shields&Filippenko1990).With the Planetary Camera on HST,Holtzman et al.(1992)identified about60cluster candidates with ab-solute magnitudes up to M V=−ing WFPC2data,Carlson et al.(1998)identified about3000clusters,of which about1200have blue integrated colors and estimated ages between0.1and1Gyr.The young clusters had estimated masses and sizes similar to those of old globular clusters,although Brodie et al.(1998)found that the Balmer line equivalent widths measured on spectra of5clusters were too strong to be consistent with standard SSP models,unless a stellar mass function truncated at2M⊙−3M⊙wasrsen:Star Formation in Clustersadopted.With accurate modeling of the HST point spread function and high dispersion spectroscopy with8–10m class telescopes,it might be possible to constrain the virial masses of some of the brightest clusters,and thereby provide independent constraints on their stellar IMF.While NGC1275may have experienced a recent merger/accretion event(Holtzman et al.1992),it is hardly one of the classical“Toomre”mergers(Toomre&Toomre 1972).One of the nearest ongoing,major mergers is the“Antennae”NGC4038/39, where HST observations have revealed a rich population of luminous,compact young star clusters with typical half-light radii∼4pc(Whitmore&Schweizer1995;Whitmore et al.1999).The brightest of them reach M V≈−14and have estimated masses close to 106M⊙(Zhang&Fall1999).Similar rich populations of YMCs have been found in many other mergers,like NGC3256where Zepf et al.(1999)identified about1000compact bright,blue objects on WFPC2images within the central7kpc×7kpc region.Again, the young clusters contribute a very significant fraction(15%–20%)of the blue light within the starburst region.Zepf et al.(1999)estimated half-light radii of5–10pc for the clusters in NGC3256,somewhat larger than for the Antennae,but note that1PC pixel corresponds to a linear scale of8pc at the distance of NGC3256,so that thersen:Star Formation in Clusters9clusters are only marginally resolved.Interestingly,only a shallow trend of cluster sizeversus luminosity was found,with radius r scaling with luminosity L roughly as r∝L0.07.NGC7252is a somewhat more advanced system than NGC3256or the Antennae.Miller et al.(1997)date the cluster system at between650Myr and750Myr.Remark-ably,both photometry and dynamical measurements yield a mass of about8×107M⊙for the most massive object(W3)(Maraston et al.2004),making it about an order of magnitude more massive than any old globular cluster in the Milky Way.With a half-light radius of17.5±1.8pc,this object is much larger than a normal star cluster,andmay be more closely associated with the“Ultra Compact Dwarf Galaxies”in Fornax (Hilker et al.1999;Drinkwater et al.2003).5.3.Dwarf/Irregular galaxiesThe bright“central condensations”in NGC1569were noted already by Mayall(1935)onplates taken with the36inch Crossley reflector at Lick Observatory,though Arp&Sandage(1985) were probably thefirst to recognize them as likely star clusters.At a distance of only∼2Mpc(Makarova&Karachentsev2003),these clusters appear well resolved on HST images with half-light radii of about2pc(O’Connell et al.1994;de Marchi et al.1997).One of the clusters,NGC1569-A,is actually a double cluster,and STIS spectroscopyhas shown that one component exhibits Wolf-Rayet features while the other componentis devoid of such features,suggesting an age difference of a few Myrs between the two components(Maoz et al.2001b).Using high-dispersion spectroscopy from the NIRSPEC spectrograph on the Keck II telescope,Gilbert&Graham(2003)derived dynamical mass estimates of about0.3×106M⊙for each of the two components of NGC1569-A,and0.18×106M⊙for NGC1569-B,again very similar to the typical masses of old globular clusters,and consistent with the clusters having“normal”stellar mass functions(see also Section6.4).A peculiar feature of the NGC1569cluster population is that the next brightest clustersafter NGC1569-A and NGC1569-B are more than2magnitudes fainter(O’Connell etrsen:Star Formation in Clustersal.1994).An even more dramatic discontinuity in the luminosity function is seen in NGC1705which has only a single bright cluster,and in NGC4214there is a gap of about1.5mag from the brightest2clusters down to number3(Billett et al.2002). Interestingly,while the clusters in NGC1569and NGC1705are young(∼107years),the two clusters in NGC4214are both about250Myrs old(Billett et al.2002),demonstrating that massive clusters are capable of surviving for substantial amounts of time at least in some dwarf galaxies.5.4.Spiral galaxy disksMost of the YMCs discussed in the preceding sections are located in environments that are peculiar in some way,or at least different from what we see in the solar neighborhood. Thus,it is tempting to speculate that the absence of YMCs in the Milky Way indicates that their formation somehow requires special conditions.There is,however,increasing evidence that YMCs can form even in the disks of spiral galaxies.Table5lists a number of nearby spirals in which YMCs have been identified.A few(e.g.M51)are clearly involved in interactions,but none of them are disturbed to a degree where they are not clearly recognizable as spirals.The nuclear starburst in M83was already mentioned in Section5.1,but there is also a rich population of young star clusters throughout the disk (Bohlin et al.1990;Larsen&Richtler1999),the most massive of which have masses of several times105M⊙.An even more extreme cluster is in NGC6946,with a dynamical mass estimate of about1.7×106M⊙(Larsen et al.2001).The disks of spiral galaxies can evidently form star clusters with masses as high as those observed in any other environment,including merger galaxies like the Antennae and starbursts like M82. Most of the spirals in Table5are type Sb or later,but one exception is NGC3081.In this barred S0/Sa-type spiral,Buta et al.(2004)detected a number of luminous young clusters in the inner Lindblad resonance ring at5kpc.Buta et al.(2004)found rather large sizes for these clusters,with estimated half-light radii of about11pc.This is much larger than the typical sizes of Milky Way open and globular clusters and indeed of YMCs found in most other places,and raises the question whether these objects might be related to the“faint fuzzy”star clusters which are located in an annulus of similar radius in the lenticular galaxy NGC1023(Larsen&Brodie2000;Brodie&Larsen2002), but have globular cluster-like ages.rsen:Star Formation in Clusters11 6.General properties of cluster systemsJust how similar are the properties of star clusters in different environments,and what might they tell us about the star formation process?Objects like NGC1569-A appear extreme compared to Milky Way open clusters or even to young LMC clusters: O’Connell et al.(1994)estimate that NGC1569-A has a half-light surface brightness over 65times higher than the R136cluster in the LMC,and1200times higher than the mean rich LMC cluster after allowing for evolutionary fading.Do such extreme objects constitute an altogether separate mode of star/cluster formation,or do they simply represent a tail of a distribution,extending down to the open clusters that we encounter locally?And are YMCs really young analogs of the old GCs observed in the Milky Way and virtually all other major galaxies?6.1.Luminosity-and mass functionsOne of the best tools to address these questions is the cluster mass function(MF).In the Milky Way and the Magellanic Clouds,the MF of young star clusters is well approximated by a power-law dN(M)/dM∝M−αwhereα≈2(Elmegreen&Efremov1997;Hunter et al.2003).This is deceptively similar to the luminosity functions derived in many young cluster systems,but it is important to recognize that luminosity functions are not necessarily identical,or even similar to the underlying MFs(unless the age distribution is a delta function).Unfortunately,MFs are difficult to measure directly.The only practical way to obtain mass estimates for large samples of clusters is from photometry,but because the mass-to-light ratios are strongly age-dependent masses cannot be estimated without reliable age information for each individual cluster.As discussed in Section3,this is best done by including U-band imaging,which is costly to obtain in terms of observing time. So far,MFs have only been constrained for a few,well-studied systems.In the Antennae, Zhang&Fall(1999)found a power-law shape with exponentα≈2over the mass range 104M⊙to106M⊙,similar to the MF of young LMC clusters.Bik et al.(2003)find α=2.1±0.3over the range103M⊙to105M⊙for M51,and de Grijs et al.(2003a)find α=2.04±0.23andα=1.96±0.15in NGC3310and NGC6745.The many studies which have found similar power-law luminosity functions are of course consistent with these results,but should not be taken as proof that the MF is as universal as the LF.Conversely,any differences in the LFs observed in differ-ent systems would not necessarily imply that the MFs are different.There are some hints that slight LF variations may be present:Elmegreen et al.(2001)find LF slopes ofα=1.58±0.12andα=1.85±0.05in NGC2207and IC2163,while Larsen(2002) and Dolphin&Kennicutt(2002)find somewhat steeper slopes(α=2.0−2.5)in several nearby spiral galaxies.While Whitmore et al.(1999)findα=2.12±0.04for the full sample of Antennae clusters,there is some evidence for a steepening at brighter mag-nitudes withα=2.6±0.2brighter than M V=−10.4.However,measurements of LF slopes are subject to many uncertainties,as completeness and contamination effects can be difficult to fully control,and it is not presently clear how significant these differences are.More data is needed.Another important question is how the MF evolves over time.While the evidence avail-able so far indicates that the MF in most young cluster systems is well approximated by a uniform power-law with slopeα≈2down to the detection limit,old GC systems show a quite different behavior.Here,the luminosity function is wellfit by a roughly log-normal distribution with a peak at M V∼−7.3(about105M⊙for an age of10–15 Gyr)and dispersion∼1.2mag(e.g.Harris&van den Bergh1981).Thus,old globu-lar clusters appear to have a characteristic mass of about∼105M⊙,while there is no characteristic mass for young clusters.This difference might seem to imply fundamen-。
- 1、下载文档前请自行甄别文档内容的完整性,平台不提供额外的编辑、内容补充、找答案等附加服务。
- 2、"仅部分预览"的文档,不可在线预览部分如存在完整性等问题,可反馈申请退款(可完整预览的文档不适用该条件!)。
- 3、如文档侵犯您的权益,请联系客服反馈,我们会尽快为您处理(人工客服工作时间:9:00-18:30)。
a rXiv:079.353v1[astro-ph]21Se p27Star Formation at Very Low Metallicity.IV.Fragmentation Does Not Depend on Metallicity for Cold Initial Conditions.Anne-Katharina Jappsen Canadian Institute for Theoretical Astrophysics,Toronto,ON,Canada jappsen@cita.utoronto.ca Ralf S.Klessen Institut f¨u r Theoretische Astrophysik,Zentrum f¨u r Astronomie der Universit¨a t Heidelberg,Heidelberg,Germany rklessen@ita.uni-heidelberg.de Simon C.O.Glover Astrophysikalisches Institut Potsdam,Potsdam,Germany sglover@aip.de and Mordecai-Mark Mac Low 1American Museum of Natural History,New York,NY,USA mordecai@ ABSTRACTPrimordial star formation appears to result in stars at least an order of mag-nitude more massive than modern star formation.Bromm et al.proposed that the transition from primordial to modern initial mass functions occurs due to the onset of effective metal line cooling at a metallicity Z/Z ⊙=10−3.5.However,their simulations neglected molecular cooling.We here perform simulations using the same initial conditions,but including molecular cooling,using both a simpli-fied chemical network that primarily follows molecular hydrogen formation,and amore complex network that also directly follows carbon monoxide and water.Wefind in both cases that molecular cooling allows roughly equivalent fragmentationto proceed even at zero metallicity for the Bromm et al.initial conditions.Theapparent transition just represents the point where metal line cooling becomesmore important than molecular cooling.In all cases,the fragments are mas-sive enough to be consistent with models of primordial stellar masses,suggestingthat the transition to the modern initial mass function may be determined byother physics such as dust formation.We conclude that such additional coolingmechanisms,combined with the exact initial conditions produced by cosmologi-cal collapse are likely more important than metal line cooling in determining theinitial mass function,and thus that there is unlikely to be a sharp transition inthe initial mass function at Z/Z⊙=10−3.5.Subject headings:stars:formation–stars:mass function–early universe–hydrodynamics–equation of state–methods:numerical1.IntroductionThe formation of stars is a key process in the early universe with far-reaching conse-quences for cosmic reionization and galaxy formation(Loeb&Barkana2001;Bromm&Larson2004;Glover2005).The physical processes that govern stellar birth in a metal-free or metal-poor environment,however,are still very poorly understood.Numerical simulations of thethermal and dynamical evolution of gas in primordial protogalactic halos indicate that themetal-freefirst stars,the so called Population III,are expected to be very massive,withmasses anywhere in the range20–2000M⊙(Abel,Bryan,&Norman2002;Bromm et al.2002;Yoshida et al.2006;O’Shea&Norman2007a).In contrast,the stellar mass spec-trum in the present-day universe is dominated by low-mass stars,peaking at around0.2M⊙(Scalo1998;Kroupa2002;Chabrier2003).This means that at some stage of cosmic evolutionthere must have been a transition from primordial,high-mass star formation to the“nor-mal”mode of star formation that dominates today.The discovery of extremely metal-poorsubgiant stars in the Galactic halo with masses below one solar mass(Christlieb et al.2002;Beers&Christlieb2005)indicates that this transition occurs at abundances considerablysmaller than the solar value.At the extreme end,these stars have iron abundances less than10−5times the solar value,and carbon or oxygen abundances that are still<∼10−3times the solar value.These stars are thus strongly iron deficient,which could be due to unusual abun-dance patterns produced by enrichment from pair-instability supernovae(Heger&Woosley2002)from Population III stars or due to mass transfer from a close binary companion(Ryan et al.2005;Komiya et al.2007).There are hints for an increasing binary fraction with decreasing metallicity for these stars(Lucatello et al.2005).In a seminal paper,Bromm et al.(2001)proposed that the transition from high to low mass star formation should occur at a critical metallicity Z crit≈10−3.5Z⊙as a result of atomicfine-structure line cooling from alpha elements,such as carbon and oxygen,becoming important.Their study is based on smoothed particle hydrodynamics simulations of the collapse of cold gas in a top-hat potential at densities below n=106cm−3.It assumes that cooling from molecular hydrogen is negligible in the density and metallicity range considered. The value of Z crit can also be estimated analytically by calculating the metallicity required to produce a cooling rate equal to the rate of adiabatic compression heating for given halo properties(Bromm&Loeb2003;Santoro&Shull2006;Frebel,Johnson&Bromm2007). However,Bromm et al.(2001)noted that neglecting H2cooling could significantly influence the resulting fragmentation pattern.Along the same lines,Omukai et al.(2005)argued, based on the results of detailed one-zone calculations,that molecular cooling would indeed dominate the cooling over many orders of magnitude in density.They argued that dust-induced cooling at densities above n≈1013cm−3may be a more viable mechanism to explain fragmentation in very low metallicity gas,a conjecture that has recently been confirmed by Clark et al.(2007).The effects of molecular cooling at densities up to n≈500cm−3have been discussed by Jappsen et al.(2007a,hereafter Paper II)in three-dimensional collapse simulations of warm ionized gas in protogalactic halos for a wide range of environmental conditions.This study used a time-dependent chemical network running alongside the hydrodynamic evolu-tion as described in Glover&Jappsen(2007,hereafter Paper I).The physical motivation was to investigate whether small protogalaxies that formed within the relic H ii regions left by neighboring Population III stars could form subsequent generations of stars themselves, or whether the elevated temperatures and fractional ionizations found in these regions sup-pressed star formation until larger protogalaxies formed.We found in Paper II that molecular hydrogen dominates the cooling of the gas for abundances up to at least10−2Z⊙.In addition, there was no evidence for fragmentation at densities below500cm−3.To understand whether this apparent discrepancy with the Bromm et al.(2001)re-sults comes from including molecular hydrogen cooling or from adopting vastly different initial conditions—here ionized halo gas in a dark matter halo with a NFW density profile (Navarro et al.1997),there cold atomic gas in a top-hat halo—we apply our time-dependent chemical network with molecular hydrogen cooling to Bromm et al.(2001)-type initial con-ditions.Our study is structured as follows.In the next section,§2,we give the details on the numerical set-up and the adopted initial conditions.In§3we report our results,and wesummarize in§4.2.Numerical Set-UpTo adequately describe star formation in the early universe,it is necessary to follow the cooling and possible fragmentation of the gas in the central regions of dark-matter halos over many orders of magnitude in density.Due to the nonlinear nature of theflow,it is not known in advance where and when gravitational collapse will occur,and consequently stars will form.To compute the time evolution of the system we therefore resort to smoothed particle hydrodynamics(SPH),which is a Lagrangian scheme to solve the equations of hydrodynamics (Benz1990;Monaghan1992,2005)coupled with the simple time-dependent chemical network discussed in Paper I,or with the more extended network discussed in Glover(2007,hereafter, Paper III).Thefluid is represented by an ensemble of particles,andflow quantities are obtained by averaging over an appropriate subset of SPH particles.The method is able to resolve high density contrasts as particles are free to move,and so the particle concentration increases naturally in high-density regions.We use a modified version of the parallel SPH code GADGET(Springel2005).Once the central parts of a collapsing fragment exceed a density n=105cm−3we introduce a“sink particle”(Bate et al.1995;Jappsen et al.2005), which is able to accrete gas from its surroundings while keeping track of the mass and linear and angular momentum of the infalling material.Replacing collapsing high-density cores by accreting sink particles allows us to follow the dynamic evolution of the system over many local free-fall timescales.2.1.Chemistry and CoolingWe present results from simulations performed using two different chemistry and cooling models.Thefirst model used in our simulations is the detailed model of low-metallicity gas chemistry outlined in Paper III.This model includes cooling both from atomicfine structures lines and also from the molecules H2,HD,CO,OH and H2O.It was designed to accurately follow the formation and destruction of these molecular species over a wide range of temperatures and densities.The other model used is the simpler one described in Paper I.This model follows the chemistry of just seven key coolants:H,H2,C+,C,O,Si,and Si+. It was designed to be used in conditions in which the majority of the gas phase metals are in atomic or ionized form,and so does not include cooling from species such as CO,OH or water,or any chemistry involving these species.In Paper I,we showed that the neglect of molecular species including metals can be justified in simulations of gravitationally collapsing protogalactic gas provided that n< 500/t char cm−3,where where t char is the characteristic physical timescale of interest,in Myr. At higher gas densities,molecule formation from metals may become more significant,al-though its ultimate importance depends on a number of factors,such as the temperature and metallicity of the gas,and the present(or absence)of an ultraviolet radiationfield.The second model is similar in design to that used in Bromm et al.(2001);the major differences between their model and ours are that we do not make any a priori assumptions regarding chemical equilibrium,and,more significantly,the fact that we include the formation of H2 and cooling from H2line emission.The majority of our modelling was done using the more complex network.However, for the purposes of comparison,we also ran a few simulations using our simpler model. Comparison between the results of these two sets of simulations helps to highlight how accurate our chemical modelling needs to be in order to properly model the thermodynamics and dynamical evolution of the gas at the densities treated in this paper.Although our chemical models are capable of accounting for the effects of dust,cosmic rays,and photodissociation by an external radiationfield,we omit these complications in the present study.A full list of the chemical reactions included in the versions of the simple network (reactions1–47)and complex network used in this paper is given in Table2,along with the corresponding rate coefficients and references to the source of the data.Given these rate coefficients,calculation of the appropriate chemical rates is for the most part straightforward. One minor exception is the case of H2collisional dissociation,which has a rate coefficient that depends on both temperature and density.For the collisional dissociation of H2by H or H2(reactions9&10),we calculate the rate coefficients using a function of the formlog k i= n/n cr1+n/n cr log k i,l,(1)where k i,l and k i,h are the low density and high density limits for the collisional dissociation rate due to collisions with species i(which can be found in Table2),and where the critical density,n cr,is given by1n cr,H +x H2(reaction75),we use a slightly different expression,taken from Dove et al.(1987):log k He=log k He,h−(log k He,h−log k He,l)|∇·v|,(4) where n(M)is the local number density of species M.The Neufeld&Kaufman(1993) and Neufeld,Lepp&Melnick(1995)treatments assume that only collisions with H2are important in determining the CO or H2O rotational cooling rates,as is appropriate in a fully molecular gas.However,in metal-poor gas,the H2abundance will often be much smaller than the abundances of atomic hydrogen or atomic helium.We therefore use an effective number density n eff,given by(Pavlovski et al.2002;Smith&Rosen2003)n eff=n H2+√2.2.Initial ConditionsWe model the initial dark matter distribution,following Bromm et al.(2001),as anisolated and roughly spherical overdensity described by the top-hat approximation with alow level of small-scalefluctuation added.We consider a halo virializing at z vir=30,whichcorresponds to a roughly3σpeak with a total mass of2×106M⊙,corresponding to2×105M⊙in baryons.The overdense region has a proper radius of150pc.For more details on theset-up of the dark matter and the gas distribution see Bromm et al.(2001).We carry outfive numerical experiments with different metallicities ranging from Z=0to Z=0.1Z⊙.Aside from the difference in metallicity,these calculations have identicalinitial conditions.The simulations were initialized at a redshift z=100,with an initialgas temperature T gas,i=200K.We performed a number of low resolution simulations,withN DM=13400dark matter particles and N gas=65536gas particles.In these simulations,our numerical resolution was the same as in Bromm et al.(2001),i.e.M res=150M⊙.Wealso performed three high resolution simulations with N DM=107177and N gas=8×105;inthese simulations,the mass resolution was M res=12.5M⊙.In all of our simulations,boththe dark matter and the gas particles were endowed with the same Hubble expansion andwere set in rigid rotation with a spin parameter of0.05,just as in Bromm et al.(2001).Wesummarize the properties of all of the runs in Table1.3.ResultsIn each of our calculations the halo contracts rapidly and the gas density increases byseveral orders of magnitude.A disk-like structure builds up in the very center of each halowith a density of n≈103cm−3and a diameter of10to20pc,as illustrated in Figure1.Thedisk mass at the time illustrated is of order M disk∼2–3×104M⊙,and so it is well-resolved even in the low resolution simulations.The disk is supported by a combination of thermalpressure and rotation and its growth is fed by material falling in alongfilaments and sheets.This complex disk provides the background for the small-scale initial densityfluctuations togrow,some of which become gravitationally unstable and collapse to form stars.In contrastto the results of Bromm et al.(2001),wefind that fragmentation of the disk occurs in allof our simulations,and that the degree of fragmentation is independent of the metallicity ofthe gas.This result can be understood as follows.Regardless of the metal content,H2line coolingmaintains the gas temperature at densities above103cm−3close to the CMB temperature(T CMB=105K at the redshift z=37.5considered in Figure2);even in the primordial case,the gas temperature is within a factor of two of this value.The resulting equation of stateis therefore approximately isothermal(at high densities)or softer than isothermal(at low densities).The thermal behavior of the gas is depicted in the left column of Figure2.Massive,isothermal disks are known to be prone to gravitational fragmentation(Bodenheimer 1995;Saigo et al.2000)and so it is unsurprising that our disks fragment.The physical conditions where the equation of state changes from softer than isothermalto isothermal or stiffer,i.e.from a regime where the gas cools with increasing density to one where the temperature remains constant or rises again,imprint a characteristic mass inthe fragmentation process(Jappsen et al.2005;Larson2005).Because this fragmentation occurs at approximately the same density and temperature in each simulation,the mass distribution of the fragments does not vary greatly from simulation to simulation.In eachof our low resolution simulations,between15and20fragments are formed,with massesin the range102M⊙≤M≤103M⊙.The resulting mass spectra are shown as the solid lines in the right-hand column of Figure2.For comparison,the thermal Jeans mass for gas with temperature T gas=T CMB is approximately2.3×103M⊙at n=103cm−3,the density above which H2line cooling maintains the gas as approximately isothermal,and 230M⊙at n=105cm−3,the density above which we form sinks.To quantify whetherthe mass spectra differ significantly,we can use the two-sided Kolmogorov-Smirnov test to determine for each of our low resolution,metal-enriched simulations the probability thatthe mass spectrum produced is drawn from the same underlying distribution as in the Z=0 simulation.Wefind probabilities of0.16,0.05,0.49and0.63for runs Z-4,Z-3,Z-2and Z-1 respectively,indicating that the differences between the fragment mass spectra are in generalnot statistically significant.The one exception is the Z=10−3Z⊙case,and even then the significance is marginal.In our high resolution simulations,wefind a greater degree of fragmentation and a broader fragment mass distribution.However,just as in our low resolution simulations,the mass distribution is dominated by fragments with masses M>100M⊙.Wefind veryfew objects with masses close to our resolution limit of12.5M⊙,and there is no indication from our current results that we are missing a substantial population of objects with masses below this limit.As before,we can compare the fragment mass distributions resulting fromour three high-resolution simulations with the K-S test.Wefind that the probabilities thatthe fragment mass distributions in runs Z-3-BIG and Z-1-BIG are drawn from the same underlying distribution as in run Z0-BIG are0.78and0.34respectively,indicating that again the differences are not statistically significant.One quantity does change significantly as the metallicity of the gas increases,the peak temperature reached during the evolution.In the zero-metallicity case,the peak temperatureT peak reached due to compressional heating during the initial contraction of the gas is∼103K.At higher metallicities,however,T peak gradually declines,until it hardly exceeds the CMB temperature in the run with Z=0.1Z⊙.This decrease in T peak reflects the growing importance of metal-line cooling in cold gas as the metallicity increases.From Figure2it is apparent that the metals begin to significantly affect T peak once the metallicity exceeds Z=10−3Z⊙.This is a lower threshold than found in Jappsen et al.(2007a),but the difference is easily understood as a consequence of the fact that in these simulations we start with cold gas,and that the H2cooling rate of the cold,low-density gas in these simulations is much lower than H2cooling rate in the warm,low-density gas studied in Jappsen et al.(2007a).Our threshold of Z∼10−3Z⊙is in reasonable agreement with the Bromm et al.(2001)threshold of Z∼10−3.5Z⊙;however,the key difference here is that at lower metallicities,the H2alone is sufficient to cool the gas,and the outcome of the fragmentation process does not appear to depend on whether the cooling comes from the H2 or from the metals.Finally,as noted above,we performed two runs using the simplified chemical model of Paper I(runs Z-3S and Z-1S)in order to investigate the importance of treating the high density metal chemistry accurately.The results of these runs are illustrated in Figure3, along with the results of the corresponding runs utilizing the full chemical network(runs Z-3and Z-1).Comparison of the phase diagrams of these runs shows very little difference in the outcome.This result is easy to understand if,as argued in Paper I,CO and H2O formation and cooling are unimportant at low densities.At high densities,we might naively expect to see a greater difference between the simulations,but in practice,all of the gas at those densities is located at or very close to the CMB temperature,and the identity of the coolant or coolants responsible for keeping it there does not appear to be of great importance.That said,our choice of chemical model does have an impact on the fragment mass distribution,as illustrated in the right-hand column of Figure3.In runs Z-3S and Z-1S,we obtain fragments that are systematically smaller than in the corresponding runs Z-3and Z-1.Further investigation suggests that this difference is a result of the fact that the fragments form at a slightly later time in the runs performed using the simplified chemical model than in the runs performed using the full model,possibly indicating that the timing of the fragmentation is sensitive to the cooling time of the gas.4.Summary and ConclusionThe current study shows that gas in simulations of the type considered here,with low initial temperature,moderate initial rotation,and a top-hat dark-matter overdensity,willreadily fragment into multiple objects,regardless of metallicity,provided that enough H2 is present to cool the gas.Rotation leads to the build-up of massive disk-like structures in these simulations which allow smaller-scalefluctuations to grow and become gravitationally unstable to form protostars.The resulting mass spectrum of fragments peaks at a few hundred solar masses,roughly corresponding to the thermal Jeans mass in the disk-like structure.Our major conclusion is that the metallicity threshold at Z=10−3.5Z⊙reported by Bromm et al.(2001)does not represent a critical metallicity above which gas fragments and below which it does not.Rather,this threshold simply marks the point at which metal-line cooling becomes more important than H2cooling at the gas densities and temperatures relevant for fragmentation in the disk.This metallicity threshold only represents a criti-cal metallicity for fragmentation in these simulations if H2formation is strongly suppressed or not considered(e.g.Bromm et al.2001).This scenario may be relevant if the extra-galactic UV background is strong(Haiman et al.2000),although recent work suggests that even in this case,suppression of H2formation is less important than previously supposed (Wise&Abel2007;O’Shea&Norman2007b).We alsofind no indication that the mass spectrum of fragments formed in these simu-lations depends on the metallicity.In our low-resolution simulations,many of the fragments are only marginally resolved,but we obtain very similar results from our high-resolution simulations,in which the fragments are well-resolved,suggesting that this lack of influence is real.We further note that simulations performed by Smith&Sigurdsson(2007)that follow the evolution of cold,metal-enriched gas from cosmological initial conditions until collapse dofind tentative evidence for a metallicity threshold at around Z=10−3Z⊙that must be surmounted before fragmentation will occur.On the other hand,the simulations presented in Paper V that follow the collapse of hot,initially ionized gas into isolated NFW halosfind no evidence for any fragmentation at or below this metallicity.We conclude that the question of whether there is a critical metallicity below which the formation of low-mass stars is impossible remains unresolved.Our current results show no evidence for such a critical metallicity:fragmentation occurs even in our Z=0simulation, due only to efficient H2cooling,and the mass spectrum of the fragments formed does not appear to differ significantly from those obtained from our higher metallicity simulations. However,the degree of fragmentation occurring in these simulations appears to be a con-sequence of the initial conditions chosen:simulations utilizing substantially different initial conditionsfind very different results(Jappsen et al.2007b;Smith&Sigurdsson2007).To make further progress in understanding the role(if any)that metal-line cooling plays inpromoting fragmentation,we need to develop a much better understanding of how metals are initially dispersed into the high-redshift interstellar and intergalactic medium,in order to be able to select the most appropriate initial conditions for our simulations.Our results call attention to an alternative scenario for the transition to the present-day low-mass IMF.Omukai et al.(2005)stress the importance of dust-induced cooling at high densities which may become important at metallicities as low as Z≈10−6–10−5Z⊙depending on the adopted dust model(Schneider et al.2006).Recent numerical simulations by Clark et al.(2007)support this point of view and predict the existence of low-mass stars and even brown dwarfs with metal abundances of∼10−5the solar value(see also Tsuribe&Omukai2006).We acknowledge useful discussions with Tom Abel,Volker Bromm,Paul C.Clark, Kazuyuki Omukai,and Naoki Yoshida.This research was supported in part by the US National Science Foundation under Grants No.PHY05-51164and AST03-07854,and by the Emmy Noether Program of the German Science Foundation under Grant No.KL1358/1. Computations were performed at the McKenzie cluster at the Canadian Institute for The-oretical Astrophysics.RSK and SCOG thank the staffand scholars of the Kavli Institute for Theoretical Physics for their hospitality during thefinal stages of the preparation of this paper,while M-MML thanks the Deutscher Akademischer Austausch Dienst and the Max-Planck-Gesellschaft for research stipends that helped support this work.REFERENCESAbel,T.,Bryan,G.L.,&Norman,M.L.2002,Science,295,93Abrahamsson,E.,Krems,R.V.,&Dalgarno,A.2007,ApJ,654,1171Adams,N.G.,&Smith,D.1976a,Int.J.Mass Spectrom.Ion.Phys.,21,349Adams,N.G.,&Smith,D.1976b,J.Phys.B.,9,1439Adams,N.G.,Smith,D.,&Grief,D.1978,Int.J.Mass Spectrom.Ion.Phys.,26,405 Adams,N.G.,Smith,D.,&Millar,T.J.,1984,MNRAS,211,857Adams,N.G.,Smith,D.,&Paulson,J.F.1980,J.Chem.Phys.,72,288Aldrovandi,S.M.V.&Pequignot,D.1973,A&A,25,137Alge,E.,Adams,N.G.,&Smith,D.1983,J.Phys.B,16,1433Andreazza,C.M.,&Singh,P.D.1997,MNRAS,287,287Anicich,V.G.,Futrell,J.H.,Huntress,W.T.,&Kim,J.K.1975,Int.J.Mass Spectrom.Ion.Phys.,18,63Anicich,V.G.,Huntress,W.T.,&Futrell,J.H.1976,Chem.Phys.Lett.,40,233 Azatyan,V.V.,Aleksandrov,E.N.,&Troshin A.F.1975,Kinet.Catal.,16,306 Barinovs,G.,&van Hemert,M.C.2006,ApJ,636,923Barlow,S.G.1984,PhD thesis,Univ.ColoradoBate,M.R.,Bonnell,I.A.,&Price,N.M.1995,MNRAS,277,362Baulch D.L.,Cobos,C.J.,Cox,R.A.,Esser,C.,Frank,P.,Just,Th.,Kerr,J.A.,Pilling, M.J.,Troe,J.,Walker,R.W.,&Warnatz,J.1992,J.Phys.Chem.Ref.Data,21, 411Beers,T.C.,&Christlieb,N.2005,ARA&A,43,531Benz,W.1990,in Numerical Modelling of Nonlinear Stellar Pulsations Problems and Prospects,ed.J.R.Buchler(Dordrecht:Kluwer),269Bodenheimer,P.1995,ARA&A,33,199Bray,I.,Burgess,A.,Fursa,D.V.,&Tully,J.A.2000,A&AS,146,481Bromm,V.,Larson,R.B.2004,ARA&A,42,79Bromm,V.&Loeb,A.2003,Nature,425,812Bromm,V.,Ferrara,A.,Coppi,P.S.,&Larson,R.B.2001,MNRAS,328,969 Bromm,V.,Coppi,P.S.,&Larson,R.B.2002,ApJ,564,23Carty,D.,Goddard,A.,K¨o hler,S.P.K.,Sims,I.R.,&Smith,I.W.M.2006,J.Phys.Chem.A,110,3101Cen,R.1992,ApJS,78,341Chabrier,G.2003,PASP,115,763Christlieb,N.,Bessell,M.S.,Beers,T.C.,Gustafsson,B.,Korn,A.,Barklem,P.S.,Karls-son,T.,Mizuno-Wiedner,M.,Rossi,S.2002,Nature,419,904Clark,P.C.,Glover,S.C.O.,Klessen,R.S.2007,ApJ,submitted;arXiv:0706.0613 Cohen,N.,&Westberg,K.R.1979,J.Phys.Chem.,83,46Cohen,N.,&Westberg,K.R.1983,J.Phys.Chem.Ref.Data,12,531Dalgarno,A.,Du,M.L.,&You,J.H.1990,ApJ,349,675Dean,A.J.,Davidson,D.F.,&Hanson,R.K.1991,J.Phys.Chem.,95,183Dove,J.E.,Rusk,A.C.M.,Cribb,P.H.,&Martin,P.G.1987,ApJ,318,379 Fairbairn,A.R.1969,Proc.R.Soc.London A,312,207Federer,W.,Villinger,H.,Howorka,F.,Lindinger,W.,Tosi,P.,Bassi,D.,&Ferguson,E.1984a,Phys.Rev.Lett.,52,2084Federer,W.,Ferguson,E.,Tosi,P.,Villinger,H.,Bassi,D.,Howorka,F.,Lindinger,W., 1984b,Contrib.-Symp.At.Surf.Phys.,eds.Howorka,F.,Lindinger,W.,Maerk, T.D.,(Innsbruck:Austria),p.141Fehsenfeld,F.C.1976,ApJ,209,638Ferland,G.J.,Peterson,B.M.,Horne,K.,Welsh,W.F.,&Nahar,S.N.1992,ApJ,387, 95Field,D.,Adams,N.G.,&Smith,D.1980,MNRAS,192,1Frank,P.1986,Proc.Int.Symp.Rarefied Gas Dyn.,2,422Frank,P.,Bhaskaran,K.A.,&Just,Th.1988,bust.Proc.,21,885 Frank,P,&Just,Th.1984,Proc.Int.Symp.Shock Tubes Waves,14,706Frebel,A.,Johnson,J.L.,&Bromm,V.2007,MNRAS,380,40Geppert,W.D.,et al.,2005,J.Phys.:Conf.Ser.,4,26Gerlich, D.1982,in‘Symposium on Atomic and Surface Physics’,eds.Lindinger,W., Howorka,F.,M¨a rk,T.D.,(Dordrecht:Kluwer),p.304Gerlich,D.,&Horning,S.1992,Chem.Rev.,92,1509Glover,S.C.O.2005,Space Sci.Reviews,117,445Glover,S.C.O.2007,in preparation.(Paper III)。