THE DISRUPTION OF GIANT MOLECULAR CLOUDS BY RADIATION PRESSURE & THE EFFICIENCY OF STAR FORMATION IN GALAXIES

, , and

Published 2009 December 29 © 2010. The American Astronomical Society. All rights reserved.
, , Citation Norman Murray et al 2010 ApJ 709 191 DOI 10.1088/0004-637X/709/1/191

0004-637X/709/1/191

ABSTRACT

Star formation is slow in the sense that the gas consumption time is much longer than the dynamical time. It is also inefficient; star formation in local galaxies takes place in giant molecular clouds (GMCs), but the fraction of a GMC converted to stars is very small, epsilonGMC ∼ 5%. In luminous starbursts, the GMC lifetime is shorter than the main-sequence lifetime of even the most massive stars, so that supernovae can play no role in GMC disruption. We investigate the disruption of GMCs across a wide range of galaxies from normal spirals to the densest starbursts; we take into account the effects of H ii gas pressure, shocked stellar winds, protostellar jets, and radiation pressure produced by the absorption and scattering of starlight on dust grains. In the Milky Way, a combination of three mechanisms—jets, H ii gas pressure, and radiation pressure—disrupts the clouds. In more rapidly star-forming galaxies such as "clump" galaxies at high-redshift, ultra-luminous infrared galaxies (ULIRGs), and submillimeter galaxies, radiation pressure dominates natal cloud disruption. We predict the presence of ∼10–20 clusters with masses ∼107M in local ULIRGs such as Arp 220 and a similar number of clusters with M* ∼ 108M in high redshift clump galaxies; submillimeter galaxies will have even more massive clusters. We find that epsilonGMC = πGΣGMCc/(2(L/M*)) for GMCs that are optically thin to far-infrared radiation, where ΣGMC is the GMC gas surface density. The efficiency in optically thick systems continues to increase with ΣGMC, but more slowly, reaching ∼35% in the most luminous starbursts. The disruption of bubbles by radiation pressure stirs the interstellar medium (ISM) to velocities of ∼10 km  s−1 in normal galaxies and to ∼100 km  s−1 in ULIRGs like Arp 220, consistent with observations. Thus, radiation pressure may play a dominant dynamical role in the ISM of star-forming galaxies.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

The Kennicutt law (Kennicutt 1998)

Equation (1)

relates the star formation surface density $\dot{\Sigma }_*$ to the gas surface density Σg and the local dynamical time Ω ≈ vc/R in disk galaxies, where vc is the circular velocity of the galaxy and R is the distance from the galactic center. The dimensionless constant η ≈ 0.017 is surprisingly small, a finding that is interpreted as showing that star formation is a slow process. Star formation is similarly slow on smaller scales within galaxies (Kennicutt et al. 2007; Bigiel et al. 2008; Leroy et al. 2008; Krumholz & Tan 2007).

Remarkably, Equation (1) holds for galaxies like the Milky Way, with rather modest star formation rates of order a solar mass per year, for starburst galaxies with star formation rates of order tens of solar masses per year, for ultra-luminous infrared galaxies (ULIRGs) with star formation rates around one hundred solar masses per year, and for submillimeter galaxies with star formation rates in excess of one thousand solar masses per year. There are indications, however, that the star formation efficiency η may be larger in ULIRGs and submillimeter galaxies, with η ∼ 0.1 (Bouché et al. 2007).

The large range of galaxies that obey Equation (1) suggests that whatever process sets the efficiency of star formation operates in galaxies with very different conditions in their interstellar media (ISMs). For example, the gas surface density in the Milky Way at 8 kpc is Σg ≈ 2 × 10−3 g cm−2 (Boulares & Cox 1990), while that in the two 100 pc star-forming disks of the ULIRG Arp 220 is Σg ≈ 7 g cm−2; the mean gas densities of the two galaxy's star-forming disks also differ by a factor of ≳104. Although the range in turbulent velocities in the ISM is not so dramatic, from ∼6 km  s−1 in the Milky Way to ∼60–80 km  s−1 in Arp 220 (Downes & Solomon 1998), the turbulent pressure in Arp 220 exceeds that in the Galaxy by a factor of ∼106.

Another signature of inefficient star formation relates to individual giant molecular clouds (GMCs). In the Milky Way, almost all stars are believed to form in such clouds. However, the fraction epsilonGMC of a GMC that is turned into stars is quite low, around 5% in the Milky Way (Williams & McKee 1997; Evans et al. 2009), with a similar value inferred from the global star formation efficiencies in other galaxies (see, e.g., McKee & Ostriker 2007).

One class of explanation for this low star formation efficiency is that gas in the ISM is prevented from collapsing by, for example, magnetic fields, cosmic rays, or by externally driven turbulence (Parker 1969; Sellwood & Balbus 1999; Ostriker et al. 2001). A second class of explanation is known by the name of "feedback": the injection of energy and momentum into the ISM by stellar processes so that star formation alters the ISM conditions and limits the rate at which gas turns into stars. The form the feedback takes is not currently agreed upon. Suggested mechanisms include supernova heating, deposition of momentum by supernovae, heating by photoionizing radiation from massive stars, deposition of momentum by expanding bubbles of photoionized H ii region gas, deposition of momentum by the shocked winds from massive stars, and jets from protostars (e.g., McKee & Ostriker 1977; Wada & Norman 2001; Matzner 2002; Li & Nakamura 2006; Cunningham et al. 2008).

In this paper, we study these feedback processes and assess the role that they play in disrupting GMCs across a wide range of star-forming galaxies. In addition, we focus on a somewhat less well-studied form of feedback: deposition of momentum by the absorption and scattering of starlight by dust grains (O'Dell et al. 1967; Chiao & Wickramasinghe 1972; Elmegreen 1983; Ferrara 1993; Scoville et al. 2001; Scoville 2003; Thompson et al. 2005). Although the magnetic fields in starburst galaxies can be large (∼ few mG for Arp 220; Thompson et al. 2006; Robishaw et al. 2008), we neglect them throughout this paper in order to focus on the competition between various processes that contribute to disrupting GMCs.

It is important to distinguish between two arenas in which galactic feedback likely operates: galactic disks in the large, and in the main units of star formation, GMCs. While there is rather sharp debate in the literature, we will assume that GMCs are at least marginally gravitationally bound objects, and hence that they are unlikely to be supported by feedback acting on the scale of galactic disks as a whole. As noted above, observations in our own and nearby galaxies establish that only ∼5% of the gas in a GMC ends up in stars; the rest of the gas is dispersed back into the ISM. Something is disrupting GMCs, but it is unlikely to be large-scale turbulence in the galaxy as a whole. Instead, GMCs must be disrupted by the stars that form in them. A number of authors have argued that galactic GMCs are disrupted by expanding H ii regions (e.g., Matzner 2002; Krumholz et al. 2006); this mechanism cannot, however, work in luminous starbursts (Matzner 2002). The fact that these galaxies nonetheless have roughly similar star formation efficiencies suggests that another disruption mechanism must be competitive with expanding H ii regions.

In this paper, we argue that the radiation pressure produced by the largest few star clusters in a GMC, acting on dust grains in the gas, is the primary mechanism by which GMCs are disrupted in more luminous starbursts and massive GMCs (see also Scoville et al. 2001; Harper-Clark & Murray 2009, and Pellegrini et al. 2007, 2009). Protostellar jets also provide an important contribution, particularly in the early stages of the evolution. In spirals like the Milky Way, both expanding H ii regions and radiation pressure are comparably important, depending on the size and mass of the cluster, and supernovae also play an important role in the latest stages of disruption.

1.1. Is Feedback Really Necessary?

A key thesis of this work is that stellar feedback is crucial for understanding the low observed values of the star formation efficiency in galaxies. In contrast, Krumholz & McKee (2005) present an explanation of the Kennicutt-Schmidt law (Equation (1)) that does not invoke an explicit form of feedback. Their argument is that turbulent motions prevent the collapse of the bulk of the gas in a GMC (or in other bound objects); only if the density is above a critical density, which depends on the Mach number of the flow, do stars actually form. The fraction of gas in a turbulent flow that lies above this critical density is small, leading to the low observed star formation efficiency per dynamical time.

We find this argument to be compelling, as far as it goes. As long as turbulence is maintained, only a small fraction of gas will collapse into stars per dynamical time. However, the assumption of a constant level of turbulence is essential to the Krumholz & McKee (2005) argument. A key, and yet unanswered, question is thus what maintains the turbulence? If the turbulence in a GMC is not maintained, then the GMC will contract, leading to an increase in the mean density and a decrease in the dynamical time. Indeed, simulations find that turbulence decays on ∼1 crossing time (Mac Low 1999; Ostriker et al. 2001), so that a continued source of energy is needed to maintain the turbulent support of the cloud. It is possible, in principle, that gravitational contraction of a GMC can sustain the turbulence, maintaining approximate virial equilibrium and a slow contraction of the cloud. We argue, however, that an independent internal source of turbulence, provided by stars, is crucial to maintaining the slow rate of star formation.

As an example, we apply this argument to Arp 220. The ISM of Arp 220 has a turbulent Mach number ${\cal M}\approx 100$. The fraction of a GMC (or any bit of molecular gas) that is sufficiently dense to be converted into stars in a free-fall time is then ≃0.013–0.05 for GMC's with a virial parameter αvir = 0.1–1 (see Figure 3 of Krumholz & McKee 2005).7 According to this argument, a GMC will convert half its gas into stars in 10–40 free-fall times, reasonably consistent with Equation (1) for any αvir. However, this assumes that the cloud does not collapse and reduce its free-fall time. In reality, if turbulence can only maintain αvir ∼ 0.1, a cloud is likely to collapse, leading to a rapid increase in density and a decrease in the free-fall time. If the star formation rate per free-fall time remains roughly constant, then the actual star formation rate will increase rapidly with time, and most of the gas in the cloud will be turned into stars in roughly one initial (large-scale) free-fall time. Thus, the model of Krumholz & McKee (2005) for the low star formation efficiency in galaxies relies critically on maintaining sufficient levels of turbulence so that αvir ∼ 1. On larger scales—above the characteristic GMC size—the equivalent argument is that the galactic disk must have Toomre Q ∼ 1, as we discuss below.

1.2. This Paper

The remainder of this paper is organized as follows. In Section 2, we collect a number of relevant astrophysical results used in our modeling. In Section 3, we describe a simple one-dimensional model for the disruption of GMCs which includes the effects of H ii gas pressure, protostellar jets, radiation pressure, gas pressure associated with shocked stellar winds, and wind shock generated cosmic rays (many of the details of how we model these forces are given in Appendix A). In Section 4, we present the results of our numerical modeling of GMC disruption in star-forming galaxies. To explore the wide range of conditions seen in galaxies across the Kennicutt-Schmidt law, we consider models for GMCs in the Milky Way, M82, Arp 220, and the z ∼ 2 galaxy Q2346-BX 482. In Section 5, we discuss the implications of our results, the origin of turbulence in galaxies, and explain physically why radiation pressure is the only source of stellar feedback in principle capable of disrupting GMCs across the huge dynamic range in ISM conditions from normal galaxies to the densest starbursts.

2. ASTROPHYSICAL ELEMENTS

In this section, we collect several pieces of phenomenology and physics that we believe are relevant to star formation in galactic disks and GMCs. We order these items according to the amount of support they enjoy, from substantial to slim. The key conclusion below is that a significant fraction of all stars are formed in compact (∼ few pc radius) massive star clusters that in turn reside in GMCs. Given the importance of a few star clusters that are small compared to the GMC as a whole, a one-dimensional model for GMC disruption is a reasonable first approximation; this is presented in Section 3.

2.1. Marginally Stable Disks (Q ≈ 1)

Quirk (1972) showed that normal galaxies have gas disks with Q ≈ 1. Kennicutt (1989) refined this to the statement that within the star-forming part of normal galactic disks, 1/4 ≲ Q ≲ 0.6. At large radii, he found Q>1 and a lack of star formation. More recently, Leroy et al. (2008) studied star formation in detail in 23 nearby galaxies. They found that if they accounted for only the gas surface density, as done above, their disks were stable, with Q ≈ 3–4; using the total (gas plus stars) surface density resulted in Q ≈ 2 with a slight variation in radius. Unlike Kennicutt (1989), they find that star formation occurs at large radii, beyond Kennicutt's Q>1 radius, albeit at reduced rates.

These studies were restricted to normal galaxies and employed a fixed sound speed as the estimate for the random velocity. However, there is evidence that disks with turbulent velocity vT ≫  cs also satisfy Q ∼ 1 when vT is used in evaluating Q (e.g., Thompson et al. 2005; and our discussion of the starbursts M82, Arp 220, and Q2346-BX 482 in Section 4). Motivated by these observations and by theoretical considerations, we will assume that all star-forming galaxies have Q ≈ 1.

2.2. The Toomre Mass and Giant Molecular Clouds

We assume that galactic disks initially fragment on the disk scale height H ≈ (vT/vc)r. The fragments will form gravitationally bound structures with a mass given by the Toomre mass, MT ≃ πH2Σg. Near the location of the Sun, the gas surface density Σg ≈ 2 × 10−3 g cm−2 and H ∼ 300 pc, giving MT ≈ 2 × 106M.

This scenario is consistent with observations of GMCs in our Galaxy; in the Milky Way, half the gas is in molecular form in giant molecular clouds with a characteristic mass of order 5 × 105M (Solomon et al. 1987), but with a rather wide range of masses. The number of clouds N(m) of mass m is given by

Equation (2)

with an exponent αG ∼ 1.8 (Solomon et al. 1987) or 1.6 (Williams & McKee 1997), so that most of the mass is in the largest clouds. As a cautionary note, we note that Engargiola et al. (2003) find αG = 2.6 ± 0.3 in M33, which suggests that lower mass clouds contribute a significant fraction of the total mass. In the Milky Way, the largest GMCs have masses of order ∼3 × 106M (Solomon et al. 1987), roughly consistent with the Toomre mass. In the Milky Way, and possibly in other galaxies, molecular clouds are surrounded by atomic gas with a similar or slightly smaller mass.

The clouds appear to be somewhat centrally concentrated. We will often employ a Larson-law density distribution,

Equation (3)

where r is the distance from the center of the GMC. We also explored isothermal models ρ(r) ∼ 1/r2; we find that such clouds are slightly easier to disrupt than the less centrally concentrated Larson-law clouds in the optically thick limit.

2.2.1. But are there Molecular Clouds in ULIRGs?

GMCs are observed in the Milky Way and in nearby star-forming galaxies such as M82. We see clumps of gas in "chain" or "clump" star-forming galaxies at z = 2, such as Q2346-BX 482 discussed below. These have been interpreted as self-gravitating, i.e., as GMCs (e.g., Genzel et al. 2008). However, we do not know of any direct evidence for Toomre-mass self-gravitating objects in ULIRGs. There is some evidence against such objects: since the clouds are self-gravitating, they will have a slightly higher velocity dispersion than that of the disk out of which they form. Increasing the velocity dispersion will alter the inferred gas mass (see Downes & Solomon 1998). Putting too much gas in gravitationally bound objects will increase the apparent gas mass, possibly making it larger than the dynamical mass.

On the other hand, ongoing star formation is clearly seen in Arp 220. Star formation probably requires densities exceeding 106 cm−3 to proceed. Thus, there is evidence that some gas is gravitationally bound. Moreover, there are numerous massive compact star clusters observed in Arp 220 (Wilson et al. 2006), indicating that massive, bound, and relatively compact accumulations of gas existed in the recent past. Motivated by these considerations, we will assume that Toomre-mass self-gravitating objects exist in all star-forming galaxies, including ULIRGs.

2.3. Gas Clump and Stellar Cluster Mass Distributions

Most of the gas in Milky Way GMCs is diffuse (n ≲ 3 × 102 cm−3), but a fraction of order 10% is in the form of dense gas clumps, with sizes around 1 pc (Lada & Lada 2003) and masses from a few tens to a few thousand solar masses. The clumps have a mass distribution similar to that for clouds (Equation (2)), with an exponent αc ≈ 1.7 (Lada & Lada 2003).

In both the Milky Way (Elmegreen & Efremov 1997; van den Bergh & Lafontaine 1984) and in nearby galaxies (McKee & Williams 1997; Kennicutt et al. 1989; Murray & Rahman 2010;), the number of stellar clusters of mass m* is given by

Equation (4)

with αcl ≈ 1.8. In other words, most stars form in massive clusters; in the Milky Way, at least, these clusters are made from gas in massive gas clumps, inside of massive GMCs.

2.4. The Sizes of Star Clusters

Star clusters are observed to have sizes ranging from rcl ≈ 0.1 pc (for Mcl ≈ 10 M) to rcl ≈ 10 pc (Mcl ≈ 108M). There are hints that clusters with masses Mcl ≲ 104M have a mass–radius relation of the form

Equation (5)

with m0 = 104 M and β ≈ 0.4 (Lada & Lada 2003), but it is entirely possible that this is a selection effect. Intermediate mass clusters, those with 104MMcl ≲ 3 × 106M have rcl ≃ 2 pc independent of mass (β = 0), albeit with substantial scatter. This characteristic size is seen for young (≲30 Myr) and old (≳30 Myr) clusters in M51 with masses in the range 103–106M (Scheepmaker et al. 2007), for superstar clusters in M82 with Mcl = 105–4 × 106M (McCrady et al. 2003; McCrady & Graham 2007), and in globular clusters with masses ∼105–106M (Harris 1996). Finally, high-mass clusters with Mcl ≳ 106M have β = 0.6 and m0 = 106 M in Equation (5) (Walcher et al. 2005; Evstigneeva et al. 2007; Barmby et al. 2007; Rejkuba et al. 2007; Murray 2009). When using the radii of stellar clusters in our GMC models, we will be guided by these observed mass–radius relations.

3. A MODEL OF CLUSTER & GMC DISRUPTION

We employ a one-dimensional model for the GMC, and place the star cluster at the center. The cluster blows a bubble in the GMC; we model this using a thin shell approximation, where the swept-up mass resides in a shell with a thickness much less than the radius of the bubble. The size and mass of the GMC are taken from observed GMCs in the Milky Way, GMCs in M82, and the large clump (which probably contains mostly ionized gas) in Q236-BX 482. For example, in the Milky Way we model the specific cases of G298.4-0.3, which has RGMC ≈ 150 pc and MGMC = 3 × 106M (Grabelsky et al. 1988), and W49, with RGMC = 22 pc and MGMC ≈ 7 × 105M, where we use the mean of the CO and virial mass estimates (Simon et al. 2001).

We also need the properties of the disk in which the GMC lives; the effective disk radius Rd, the disk dynamical time tdynRd/Vc, the turbulent velocity vT, the disk mass surface density Σd, and metallicity Z/Z (relative to solar). We choose the values of these parameters to match those of four galaxies to which we compare our models: the Milky Way, M82, Q2346-BX 482, and Arp 220. In that sense these are not free parameters. Table 1 summarizes the observed input parameters for the galactic disks in the systems we model.

Table 1. Observed Galaxy Parameters

Galaxy Rd tdyn Σ vT Z/Z $\dot{M}_*$ (obs)
   kpc  yr g  cm-2  km  s−1   M  yr−1
Milky Way 8.0 3.6 × 107 2 × 10−3 6 1 1.3
M82  0.35 3.0 × 106 0.1 15 1.5 4
BX482 7.0 2.9 × 107 4 × 10−2 53 1 140
Arp 220 0.1 3.3 × 105 7 61 3 120

Notes. Observed galaxy properties. Column 1 gives the name of the model. The next five columns give model input parameters: the disk radius (Column 2), dynamical time Rd/vc (Column 3), gas surface density (Column 4), turbulent velocity vT (Column 5; recall that H = [vT/vc]Rd), and metallicity in solar units (Column 6). The metallicity is not that well constrained in Arp 220. Column 7 gives the observed star formation rate for the galaxies.

Download table as:  ASCIITypeset image

Table 2 gives the inferred or assumed input parameters related to the GMC and its central star cluster: the mass MGMC and radius RGMC of the GMCs and the stellar mass and radius of the star clusters. Table 2 also lists the GMC's star formation efficiency epsilonGMC and the shell velocity when the GMC is disrupted—we interpret the latter as the turbulent velocity induced in the ISM, vT.

Table 2. GMC and Star Cluster Properties

Galaxy RGMC MGMC rcl M* epsilonGMC vT
   pc M  pc M    km  s−1
MW G298.4-0.3 150 3 × 106 2 105 0.03 10
MW W49 22 7.5 × 105   2 4.3 × 104 0.06 9
M82 30 3 × 106 1.5   7 × 105 0.24 10
BX482 925 109 13 2.7 × 108 0.27 50
Arp 220 5 4 × 107 3.5 1.4 × 107 0.38 50

Notes. Columns 2–5 give the assumed GMC and star cluster properties: the radius of the GMC RGMC (Column 2), the mass of the GMC MGMC (Column 3), the star cluster radius rcl (Column 4), and the stellar mass of the star cluster M* (Column 5). Columns 6 and 7 give the predictions of our model for the star formation efficiency in the GMC epsilonGMC and the shell velocity when the GMC is disrupted, which we also interpret as the turbulent velocity vT induced in the ISM of the Galaxy.

Download table as:  ASCIITypeset image

This paper argues that two forces dominate the dynamics of GMCs, self-gravity, and radiation pressure from stars. It follows that the mass and radius of the GMC have a strong effect on the dynamics. If, for example, GMCs in the Milky Way are smaller (at a fixed mass) than those employed in our models, while the stellar luminosity is held fixed, then it might be that a compact GMC will not be disrupted. For example, in Section 4.1, we model the Milky Way star-forming region G298.4-0.3 using a GMC with MGMC = 3 × 106M and RGMC = 150 pc. If we reduce the radius of the GMC by a factor of 2.5 to RGMC ≈ 60 pc, the bubble blown by the cluster expands to ∼30 pc, but then stalls and collapses.

We would argue that the collapse of the GMC would lead to further star formation, which would then likely disrupt the cloud. However, we have modeled about a dozen Milky Way star-forming regions using GMC masses and radii, and cluster luminosities, taken from the literature; in all cases, our models predict that the GMCs are disrupted.

In the model for M82, we use a mass near but slightly below the maximum mass in Keto et al. (2005); for the radius, we use the mass–radius relation for M82 in the observations just cited (E. Keto 2008, private communication), i.e., MGMC = 3 × 106M and RGMC = 30 pc. We note that the mean particle number density in these clouds is n ∼ 1000 cm−3, much larger than for GMCs of similar mass in the Milky Way. More importantly, the surface density ΣG ≈ 0.2 g cm−2, 10 times larger than the Milky Way value 2 × 10−2 g cm−2. We argue in this paper that this factor of 10 explains why M82 has several clusters with Mcl ≳ 106M, while the most massive cluster in the Milky way barely reaches 105M.

For Q2346-BX 482, we also use the observed mass and radius, MGMC = 109M and RGMC ≈ 925 pc, of the giant clump (R. Genzel 2008, private communication). We note that this clump may or may not contain any molecular gas; there are two ways to estimated the mass, one based on the ionizing flux and the radius of the clump, and the second based on the estimated extinction and an assumed gas-to-dust ratio.

For Arp 220, where little is known about the size of GMCs, we estimate H/RGMC ≈ 5 leading to RGMC = 5 pc; if other galaxies are any guide, and it is not clear that they are since conditions in Arp 220 are so extreme, this is an underestimate of RGMC. This makes the model GMC more difficult to disrupt than may be the case for real GMCs in Arp 220. The mass of the GMC is taken to be the Toomre mass, with a Larson-like (ρ ∼ 1/r) internal density profile.

For the purposes of our simplified modeling, the stars are assumed to lie in a single massive cluster of total mass (gas plus stars) Mcl, which forms a mass of stars M* = epsilonclMcl with luminosity L and radius rcl, surrounded by the remnants of the gas out of which it formed, with mass Mg = (1 − epsiloncl)Mcl. We use a Muench et al. (2002) stellar initial mass function (IMF) to relate the cluster luminosity to its mass. The quantity epsiloncl characterizes the efficiency with which cluster gas is turned into stars. In our Galaxy, epsiloncl ≈ 0.3–0.5 (Lada & Lada 2003). In our models, we fix epsiloncl ≃ 0.5 and adjust the cluster mass Mcl (or equivalently, the cluster stellar mass M*) to find under what conditions the central star cluster can disrupt its host GMC. Physically, we expect that star formation will self-adjust to form a cluster of approximately this mass. For the Milky Way and M82, the star cluster masses we infer by this method are comparable to those observed.

We assume that all the stars in the cluster are formed at the same time, which we set to be t = 0.

We model the impact of the central star cluster on the surrounding GMC using the thin shell approximation. As the star cluster evolves, driving winds, jets, and radiation into the overlying gas, a shell forms; this is seen directly in massive star-forming regions in the Milky Way (e.g., Churchwell et al. 2006; Murray & Rahman 2010).

We calculate the shell's dynamics as it sweeps up mass and disrupts the GMC. The relation between shell radius r, shell velocity v(r), shell mass Msh(r), and shell momentum Psh(r) is given by

Equation (6)

where Msh(r) is the mass of the shell, which increases as the shell radius r increases, and Psh = Msh(r)v(r) is the momentum of the shell. For a Larson-like density profile of the GMC, ρ ∝ r−1, the mass of the shell is given by

Equation (7)

for r < RGMC and by

Equation (8)

for RGMC < r < H. The quantity ρdisk ≡ Σg/(2H) is the mean density of the galactic disk.

We have experimented with both isothermal (MGMC(r) ∼ r) and Larson (MGMC(r) ∼ r2) GMCs and find qualitatively similar results.

The momentum equation for the shell is

Equation (9)

The inward forces on the right-hand side of Equation (9) are the self-gravity Fgrav of the swept-up shell and the mutual gravity of the stars in the largest cluster and the shell, and the force Fturb exerted by the turbulent motions of the gas in the GMC on the shell. The outward forces on the right-hand side of Equation (9) acting to disrupt the GMC include the force Fjet associated with momentum deposited by jets from star formation, the force FH ii due to the thermal pressure from ionized gas (H ii regions) and from shocked stellar winds, and the force Frad associated with radiation pressure. In some models, we include the effects of hot gas Fhot and cosmic rays Fcr associated with shocked stellar wind. A more detailed description of how we implement these forces is given in Appendix A.

We assume that the shell captures essentially all the ultraviolet radiation from the cluster; if the shell has holes, as real shells in the Milky Way certainly do, photons moving in the direction of the holes will escape, but then there is no shell to push on in that direction.

The situation is different for hot gas. The extraordinary effectiveness of hot gas in driving a shell relies on the gas being confined (Castor et al. 1975). Shocked but unconfined stellar winds add a force comparable to that of radiation pressure, a fact which follow from the line-driven nature of hot star winds, combined with the fairly high effective continuum opacity of the winds. As just mentioned, bubble shells in the Milky Way are far from complete.

It is not so clear that bubbles in galaxies with high surface densities, such as Arp 220, will also have holes. Observations of the distribution of column density in the Milky Way are consistent with a log-normal distribution of column density, but with a rather low dispersion (Wong et al. 2008; Goodman et al. 2009). We will assume that bubbles in M82 and Arp 220 also have a log-normal column density distributions, with a mean value much higher than that in the Milky Way. We show explicitly in this paper that even in for perfect confinement, shocked stellar winds are not important for GMC disruption in ULIRGs. We discuss the effect of a log-normal column density distribution on the optical depth below.

Much of the literature on bubbles around massive stars and massive star clusters assume that shocked stellar winds dominate the dynamics (e.g., Castor et al. 1975; Weaver et al. 1977; Chu & Mac Low 1990). However, observations of H ii regions in the Milky Way suggest that the pressure in such hot gas is equal to that of the associated H ii (104 K) gas (Dorland & Montmerle 1987; McKee et al. 1984; Harper-Clark & Murray 2009). The most likely interpretation of these results is that neither hot gas nor cosmic rays are confined inside bubbles in the Milky Way or the LMC, but rather escape (Harper-Clark & Murray 2009). Accordingly, we neglect the last two terms on the right-hand side of Equation (9) in our Milky Way models and in our M82 models; calculations which include these pressures predict bubble sizes that are far too large compared to observations in the Milky Way. Accordingly, with the exception of Section 4.4, we ignore the pressure associated with shocked stellar winds. We do include stellar-wind and cosmic-ray pressure in our Arp 220 models, but there they make little difference.

In Section 1, we made a distinction between feedback in the disk as a whole and feedback in GMCs. This distinction is important for a number of reasons. In particular, we believe that supernova (SN) explosions largely contribute to the former, but not the latter. In nearby ULIRGs, where the bulk of the star formation takes place in ∼100 pc disks, the dynamical time is tdyn = R/vc ≈ 5 × 105 yr, much less than the main-sequence lifetime of even the most massive stars. Hence, a gravitationally bound object in a ULIRG cannot be disrupted on a dynamical time by SNe resulting from stars formed in that object. In the Milky Way, even though the dynamical time is longer than the main-sequence lifetime of massive stars, it is clear that GMCs are in the process of being disrupted well before SNe occur (Murray & Rahman 2010). Although SNe may be important during late stages of GMC disruption, and for stirring up the galactic disk as a whole, they cannot be the main agent that disrupts GMCs, either in Milky Way-like galaxies or in ULIRGs.

In the densest starbursts, which have mean gas densities ∼103−4 cm−3, SNe rapidly lose the majority of their energy to radiative losses (e.g., Thornton et al. 1998). Under these conditions, the primary role of SNe is to stir up the bulk of the ISM via the momentum they supply, rather than to heat up the gas and/or create a hot phase of the ISM in pressure equilibrium with the rest of the mass (Thompson et al. 2005). To see that the latter is untenable, we estimate the density nh of hot gas required for a virialized ISM at ∼107T7 K to be in pressure equilibrium with the bulk of the gas, i.e., for ph = nhkT ≃ πGΣ2g. Given this density, we find that there is a critical surface density Σc above which thermal X-ray emission from the hot gas would exceed the observed correlation between X-ray and FIR emission (LX ≃ 10−4LFIR; Ranalli et al. 2003): Σc ≃ 0.04(Hd/100 pc)−2/5T3/57. Galaxies with Σd ≳ Σc, which includes the majority of luminous star-forming galaxies, cannot have a dynamically important hot ISM. Instead, the dominant role of SNe is to stir up the dense gas via the momentum imparted in the snowplow phase. This may even be true in galaxies with Σc ≲ Σg, because the hot gas can vent via galactic winds or fountains. For example, in the Milky Way, which has Σg ≃ 2 × 10−3 g cm−2 ≪ Σc, the hot ISM is believed to contribute only ∼10% of the total pressure (Boulares & Cox 1990). For this reason, we will not consider the pressure due to the hot ISM in this paper, although the momentum supplied by SNe is important in the late stages of star cluster and GMC evolution.

Having summarized the basic elements of our model, we now describe its application to galaxies ranging from the Milky Way to the most luminous starbursts (Section 4). We then discuss the implications of these results (Section 5).

4. RESULTS

4.1. The Milky Way: G298.4-0.3 and W49

Recent work has revealed that the Milky Way harbors a number of very young massive star clusters, with masses ranging up to 105M (Figer et al. 1999; Brandner et al. 2008; Figer et al. 2006; Murray & Rahman 2010). For example, consider the case of G298.4-0.3 in the Carina arm. Murray & Rahman (2010) show that the free–free flux emerging from this region implies a total ionizing flux Q = 7.7 × 1051 s−1 and suggest that ∼60% of this comes from a single massive cluster residing in the prominent ∼50 pc bubble revealed by Spitzer GLIMPSE images. Most of the remaining flux comes from two clusters associated with the two giant H ii regions G298.2-0.3 and G298.9-0.4; both sources appear to lie in the rim of the bubble seen in the GLIMPSE images. Grabelsky et al. (1988) find two massive GMCs in this region, numbers 24 and 26, both with MGMC ≈ 3 × 106M and RGMC ≈ 100 pc. Their radial velocities are 22 km  s−1 and 24 km  s−1, in good agreement with the range of radio recombination line, i.e., H ii region, radial velocities in this direction, which range from +16 km  s−1 to 30.3 km  s−1, with a mean ∼+23 km  s−1.

Accordingly, our model for G298.4-0.3 consists of a GMC with RGMC = 150 pc and MGMC = 3 × 106M. In the spirit of our simplified one-dimensional modeling, we lump all of the star clusters together into a central star cluster with L ≈ 7 × 107L and initial cluster radius rcl = 1.5 pc. Half the Galactic star formation takes place in 17 complexes, with a minimum Q = 3 × 1051 (Murray & Rahman 2010), so G298.4-0.3 is representative of star-forming clusters in the Milky Way.

The astute reader may note that the density of the GMC associated with G298.4-0.3 is lower than that of typical GMCs. The large star clusters (including G298) mentioned in the first paragraph of this section are all associated with very massive GMCs, in fact, with the most massive GMCs in the Milky Way. The surface densities of Milky Way GMCs are roughly constant (e.g., Solomon et al. 1987). It follows that the volume densities of massive GMCs are lower than the volume densities of typical (less massive) GMCs.

However, there do exist fairly massive GMCs with smaller radii, for example the GMC associated with the W49 star formation region. Simon et al. (2001) find a GMC with a radius of 22 pc and a CO luminosity derived mass of ≈5 × 105M, and a viral mass ≈106M. Using a mass 7.5 × 105M, the mean density of the W49 GMC is n ≈ 700 cm−3. The ionizing flux associated with W49 is Q ≈ 3 × 1051 s−1 when corrected for dust absorption (Murray & Rahman 2010); this is about half the ionizing flux of G298.4-0.3. The corresponding luminosity is L = 2.5 × 1041 erg  s−1 and the corresponding stellar mass Mcl = 4 × 104M. We present models of both G298.4-0.3 and W49.

The top two panels of Figure 1 show the radius of the shell surrounding the central cluster in G298.4-0.3 as a function of time (upper left panel) and the shell velocity as a function of shell radius (upper right panel). The shell starts at our putative initial radius for the cluster of ∼1.5 pc, and reaches r ∼ 80 pc at about 6.5 Myr (dashed line), at which point the star cluster luminosity has dropped by a factor of 3. The most massive stars begin to explode after about 3.6 Myr (vertical dotted line), while the last O stars explode after about 1.3 × 107 yr. The solid vertical line marks the dynamical time R/vc for the Milky Way at R = 8 kpc. Note that the dynamical time for the GMC is somewhat shorter, ∼6 Myr.

Figure 1.

Figure 1. Shell radius r as a function of time, and velocity, forces, and momentum (total and broken down by source) as a function of r(t) in our model for G298.4-0.3 in the Milky Way. The shell is sheared apart when it reaches the Hill radius (∼200 pc), where we end our integration. Upper left: radius vs. time. The vertical dotted, dashed, and solid lines mark when the first cluster SNe explode, the central cluster luminosity drops to 1/3 of its initial value, and t reaches the Milky Way dynamical timescale, respectively. Upper right: velocity of the shell as a function of r(t). Note that the asymptotic velocity is comparable to the turbulent velocity of the Milky Way disk. Lower left: the upper-most solid line is the total outward force, consisting of the momentum supplied by protostellar jets (labeled Fjet), by H ii gas pressure (FH ii) and radiation pressure on dust grains (Frad). The dotted line is the total inward force (Fin), dominated by ram pressure from the overlying gas at small shell radii, and by the self-gravity of the shell at large radii. All forces act only at the shell radius r(t), i.e., this is not a plot of the run of force with radius at a fixed time. Lower right: momentum of the swept-up shell (labeled Pjet), together with the momentum deposited by radiation (Prad), gas pressure ($P_{\rm H\,{\mathsc {ii}}}$), and protostellar jets (Pjet). The bulk of the momentum is supplied by radiation and gas pressure.

Standard image High-resolution image

The lower left panel of Figure 1 shows the forces acting on the shell as a function of shell radius in this model; recall that in our thin shell approximation all outward forces are contact forces, so they only act at the radius of the shell. Since the velocity only varies by a factor ∼2 for t ≳ 300, 000 yr, the plot can also be used to infer (roughly) the force acting on the shell as a function of time. Note that the radiation and H ii gas pressures drop after 3.6 Myr, when the bubble has r ∼ 80 pc, because the luminosity and ionizing flux of the cluster drop rapidly. Despite this, the bubble continues to expand at very nearly the same rate. The evolution after 3.8 Myr, i.e., radii larger than ∼80 pc, may underestimate the rate of expansion somewhat; the bubble may expand slightly more rapidly after several megayears due to energy input by SNe. We say the rate of expansion may be underestimated since the hot gas from the SNe is likely to escape the bubble as easily as the hot gas from shocked winds apparently does. In addition to SNe, other unmodeled effects also become important at late times and large radii. For example, because the inner radius of the shell exceeds the outer radius of the initial GMC, the surface density of the gas decreases to AV ≈ 1, so that ionizing photons from the Galactic radiation field can penetrate and ionize the shell.

We halt the integration in our model when the radius of the expanding bubble exceeds the Hill radius rHill of the GMC, i.e., when the tidal shear from the Galaxy exceeds the self-gravity of the GMC: rHill ≈ (MGMC/2M(r))1/3a, where M(r) = v2ca/G is the enclosed dynamical mass of the galaxy at the galactocentric radius a of the GMC. After this time, the remaining molecular gas will be dispersed (although not necessarily converted to atomic gas).

Figure 1 shows that the central cluster in G298.4-0.3 should disrupt its natal GMC. What force is responsible for this disruption? At the current radius, ∼55 pc, the radiation force and the gas pressure force are within a factor of 2 of each other, and will remain so until most of the O stars explode; the force from protostellar jets is substantially smaller. However, at early times, the jet force was as much as a factor of 2 larger than the radiation pressure force, and the gas pressure force was negligible.

Finally, the lower right panel of Figure 1 plots the momentum of the shell as a function of time. At the current radius of the bubble in G298.4-0.3, r ∼ 55 pc, the radiation has deposited about twice the momentum supplied by the H ii gas pressure. The stellar jets are not active at this time, but over the time they were active (corresponding to radii below ∼15 pc) they deposited a momentum comparable to that of the radiation pressure (at those early times). In these models, a combination of protostellar jets and radiation pressure disrupts the natal cluster, while a combination of gas and radiation pressure disrupts the GMC. The shell velocity at late times is of the order of the turbulent velocity seen in the ISM of the Galaxy (upper right panel), demonstrating that, even in the absence of supernovae, massive star formation can generate turbulent motions on large (50 pc or larger) scales comparable to those observed.

Figure 2 shows the results for W49; again, we find that the GMC will be disrupted by the stars that power the observed free–free radio emission.

Figure 2.

Figure 2. Shell radius r as a function of time, and velocity, forces, and momentum (total and broken down by source) as a function of r(t) in our model for W49 in the Milky Way. Upper left: radius vs. time. Upper right: the velocity of the shell as a function function of r(t). Once again, the asymptotic velocity is comparable to the turbulent velocity of the Milky Way disk. Lower left: the upper-most solid line is the total outward force, consisting of the momentum supplied by protostellar jets (labeled Fjet), by H ii gas pressure (FH ii) and radiation pressure on dust grains (Frad). The dotted line is the total inward force (Fin). Lower right: momentum of the swept-up shell (labeled Pjet), together with the momentum deposited by radiation (Prad), gas pressure ($P_{\rm H\,{\mathsc {ii}}}$), and protostellar jets (Pjet). As in the model for G298.4-0.3, the bulk of the momentum is supplied by radiation and gas pressure.

Standard image High-resolution image

We argue below that the fraction epsilonGMC is proportional to ΣGMC based on a simple force balance. The GMC in W49 has a significantly higher surface density ΣGMC ≈ 0.1 g cm−2 compared to ≈0.01 g cm−2 for G298.4-0.3. The observed epsilonGMC for the two objects are different, although not by a factor of 10.

Because star formation is a complex process, it would be surprising if it were not somewhat stochastic, so that, for example, the star formation efficiency varied from GMC to GMC even when RGMC and MGMC are held fixed. Our estimate of the efficiency is then more in the nature of a lower bound.

We have run models in which we artificially lowered the stellar mass (and hence luminosity) in our models for G298.4-0.3 and W49. We can reduce the stellar mass by a factor of ∼6 and still see the GMC in G298.4-0.3 disrupted. Lowering the mass further results in a failed bubble, i.e., the GMC is not disrupted.

In contrast, we find that if we use a lower efficiency for W49 by more than 40%, our models predict that the GMC will not be disrupted.

We note that W49 is younger than G298.4-0.3, based on the size of the observed bubble in each object. Star formation will continue as the bubble in W49 expands, so that the final epsilonGMC will be larger than that currently observed. However, it seems unlikely that the stellar mass will increase by more than a factor of 2, so that the final stellar mass will remain within a factor ∼2–3 of the critical mass required to disrupt the GMC.

4.2. The Starburst M82

M82 is one of the nearest (D = 3.6 Mpc; Freedman et al. 1994) starburst galaxies, with an infrared luminosity LIR = 5.8 × 1010L (Sanders et al. 2003). The galaxy is small compared to the Milky Way, with a circular velocity vc ≈ 110 km  s−1 (Young & Scoville 1984), and a CO inferred gas mass 2 × 108M inside r = 350 pc (Weiß et al. 2001) (adjusted to our assumed distance), yielding a gas surface density Σg ≈ 0.1 g cm−2 and a gas fraction fg ≈ 0.2. The metallicity is 1.2–2.0 times solar (Smith et al. 2006).

The radius and mass of the most massive star clusters in M82 are well established; there are about 200 clusters with M>104M (Melo et al. 2005) and about ∼20 well-studied superstar clusters (Mcl>105M). With 1 arcsec corresponding to a spatial scale of 17.5 pc, a number of superstar clusters are resolved by Hubble Space Telescope (HST; Smith & Gallagher 2001; McCrady et al. 2003; McCrady & Graham 2007). Typical half-light-projected radii for these massive objects are ∼0farcs08 or 1.4 pc. McCrady et al. (2003) list 20 such clusters. The total mass of the 15 clusters for which they measure viral masses is ∼1.4 × 107M. Their largest cluster, "L," is a monster, with a mass of 4 × 106M and a half-light radius of 1.5 pc; more typical masses are ∼5 × 105M. A rough fit of the form (Equation (4)) gives αcl ≈ 1.9 (McCrady & Graham 2007).

The masses and radii of the GMCs in M82 are also known; the distribution is well fitted by Equation (2), with αG ≈ 1.5 ± 0.1, and a maximum mass of MGMC ≈ 3 × 106M (Keto et al. 2005); the radii of these 3 × 106M GMCs are RGMC ∼ 30 pc (E. Keto 2008, private communication). The Toomre mass is ≈7 × 106M. Both are comparable to the mass of the two largest superstar clusters given by McCrady & Graham (2007). Either there were more massive GMCs in M82 in the past, or epsilonGMC ≈ 1 for the GMCs out of which these two clusters formed.

Tables 1 and 2 summarize our assumed galaxy, GMC, and star cluster properties in M82. These are all motivated by, and reasonably consistent with, the observations summarized above. Our results for the disruption of GMCs are summarized in Figure 3. The top two panels show the shell radius as a function of time and the velocity as a function of radius. The main-sequence lifetime of a 120 M star (the dashed vertical line) is comparable to the disk dynamical time (Rd/vc, the solid vertical line). The GMC is disrupted (reaches the Hill radius) about one disk dynamical time after the cluster forms. The velocity of the shell reaches higher values than those found in our Milky Way model because of the much larger cluster masses in M82, combined with the fact that the star cluster radii in the two galaxies are nearly the same. Initially, the shell velocity is comparable to the escape velocity from the cluster. The velocity begins to slow once the swept-up mass is similar to the mass in the cluster (at r ∼ 4 pc).

Figure 3.

Figure 3. As in Figure 1, but for a massive star cluster in the starburst M82. Upper left: bubble radius as a function of time. The dashed line is when the star cluster luminosity drops to 1/3 of its initial value, while the vertical solid and partially obscured dotted lines mark disk dynamical time and the (nearly equal) lifetime of a 120 M star. As in Figure 1, we halt the integration when the shell reaches the Hill radius; this occurs before the first supernovae explode. Upper right: velocity of the swept-up shell in the M82 model as a function of shell radius. Lower left: forces in our M82 model, plotted against the radius of the swept up shell. Line styles are the same as in Figure 1. Lower right: momentum of the swept-up shell in the M82 model, with the same line styles as in Figure 1. The bulk of the momentum is supplied by radiation; the contribution from gas pressure is negligible. The contribution of the protostellar jets is about 20% that of radiation pressure.

Standard image High-resolution image

What is responsible for disrupting the GMC in our M82 model? The lower panels of Figure 3 show that the cluster gas on small scales is expelled by a combination of protostellar jets and radiation pressure, while the overlying GMC is disrupted primarily by radiation pressure. The contribution from the gas pressure of the H ii region is negligible over most of the evolution: the lower right panel shows that the H ii gas pressure contribution to the momentum is ∼10% of that contributed by radiation when the shell radius reaches the original size of the GMC.

4.3. The Clump Galaxy Q2346-BX 482

Q2346-BX 482 is a redshift z = 2.26 disk galaxy with a disk radius of Rd ≈ 7 kpc and a gas mass, as estimated from inverting the Schmidt Law in Kennicutt (1998), of Mg ≈ 3 × 1010M (Genzel et al. 2008). We interpret the clumps in rapidly star-forming redshift z ∼ 2 galaxies as Toomre-mass GMCs, with radii RGMC ≈ 1 kpc, and we model the giant clump in BX 482, as one of the most extreme examples of this phenomenon. With turbulent and circular velocities of vT ≈ 55 km  s−1 and vc ≈ 235 km  s−1, respectively, we infer a disk scale height of H ≈ 1.6 kpc, and using ϕG = 3 a GMC size of order 500 pc. Our model uses the observed GMC radius, around 1 kpc (R. Genzel 2008, private communication).

The mass of the central star cluster is set so that the luminosity matches that observed, roughly L ≈ 4 × 1011L (R. Genzel, private communication). We take the radius of the star cluster to be ∼7 pc (see Section 2.4). In reality, there will likely be many star clusters, with a distribution given by Equation (4), and with a spread in age of several to ten megayears. We assume solar metallicity, consistent with the observations. Finally, the gas mass of the clump is not known, but we assume it is roughly the Toomre mass, ∼109M. This is consistent with the mass of ionized gas for the observed luminosity of the clump, at the observed size RGMC ≈ 1 kpc, and for a stellar population less than 4 Myr old (Equations (A9) and (A11)). The clump is seen to emit Hα, with an estimated Av ≈ 2, suggesting that the mass cannot be much larger than 109M. If Av is larger, then both the luminosity and potentially the GMC mass could be larger.

Figure 4 shows the results of our model for the giant GMC in Q2346-BX 482. The right panel shows that radiation pressure dominates the evolution of the GMC at nearly all times. The GMC is disrupted in about 15 Myr, half the disk dynamical timescale. The shell velocity, shown in the middle panel, is ∼60–80 km  s−1 when the radius is ∼1 kpc, in reasonable agreement with the observed velocity dispersion of the galaxy.

Figure 4.

Figure 4. Upper left: radius of the shell as a function of time in the model for the giant clump in the z = 2.26 galaxy Q2346-BX 482. The dotted, dashed, and solid lines mark when the first cluster SNe explode, the central cluster luminosity drops to 1/3 of its initial value, and t reaches the disk's dynamical timescale, respectively. Upper right: the velocity of the swept-up shell. For the vast majority of the evolution, the shell velocity is comparable to the observed velocity dispersion of the gas, ∼55 km  s−1 (Genzel et al. 2008). Lower left: forces in our model for the giant clump in BX482, with line styles as in Figure 1. Lower right: momentum of the swept-up shell, and the momentum supplied by H ii gas pressure (short dash line), protostellar jets (solid line labeled Pjet) and radiation pressure (long dash line). Radiation pressure provides about 90% of the momentum of the shell.

Standard image High-resolution image

Our conclusion that radiation pressure is disrupting the massive clump in BX482 is directly supported by observations, independent of the specific assumptions in our model: the self-gravity of the clump is

Equation (10)

where we have scaled M to the Toomre mass. This can be compared directly with the radiation pressure force,

Equation (11)

The clump should thus be expanding.

4.4. The Ultra-luminous Infrared Galaxy Arp 220

Arp 220, at ∼77 Mpc, is the prototypical ULIRG in the local universe. The gas mass of each of the rd ≈ 100 pc star-forming disks in Arp 220 is 109M, the circular velocity vc ≈ 300 km  s−1, vT ≈ 80 km  s−1, and the disk scale height H = (vT/vc)rd ≈ 23 pc (Downes & Solomon 1998; Sakamoto et al. 1999). The mean surface density is Σg ≈ 7 g cm−2, about 100 times larger than that of M82 and more than a thousand times higher than in the Milky Way. We estimate that Arp 220 has GMC masses of ≈5 × 107M, RGMC ≈ 5 pc and a turbulent velocity in each GMC of ∼170 km  s−1, about twice that measured for the disk as a whole. Although the metallicity is uncertain, we take a fiducial metallicity of 3 times solar; this increase in metallicity is important because it increases the dust optical depth and hence the overall importance of radiation pressure.

Wilson et al. (2006) found ∼40 young superstar clusters in and around Arp 220; they estimate masses for about a dozen, with a number having Mcl ∼ (2–4) × 106M; the largest has Mcl ≈ 107M. Given the huge extinction toward the twin disks, this is likely to be a rather conservative lower limit on the mass of the most massive cluster in the system. The clusters are unresolved in the HST images (which have a resolution of order 15 pc at the distance of Arp 220), except possibly their brightest cluster, with a half-light radius rcl ≈ 20 pc. Wilson et al. (2006) do not obtain either a velocity dispersion or a half-light radius for their clusters, so they cannot calculate a dynamical mass. Rather, they use a Salpeter (1955) IMF and Bruzual & Charlot (1993) stellar synthesis models combined with their photometry.

We find that for Mcl = 1.4 × 107M (L = 3 × 1010L), even a Toomre-mass GMC (4 × 107M) in Arp 220 would be disrupted (see Figure 5). The disruption of the GMC occurs on the dynamical time of the disk, well before any supernovae explode in the GMC's central star clusters.

Figure 5.

Figure 5. Upper left: bubble shell radius rsh as a function of time in a model for the disruption of a GMC by a star cluster in the ULIRG Arp 220. The dotted, dashed, and solid lines mark when the first cluster SNe explode, the central cluster luminosity drops to 1/3 of its initial value, and t reaches the disk's dynamical timescale, respectively. Note that in Arp 220, the disk and GMC dynamical times are short compared to the main-sequence lifetime of massive stars, unlike in our other models (see also Table 1). Upper right: velocity of the swept-up shell as a function of shell radius. Lower left: the forces as a function of rsh. The uppermost solid line is the total outward force Fout, consisting of five components: the force exerted by protostellar jets (solid line; Fjet), the force exerted by H ii gas pressure multiplied by 100 (short dash line; FH ii) the force exerted by radiation pressure on dust grains (long dash line; Frad), the force exerted by shocked stellar winds (Fhot; solid), and the force exerted by cosmic rays produced in stellar wind shocks (Fcr; dot-dashed). Radiation pressure dominates the outward force at nearly all times. The dotted line is the total inward force Fin, dominated by the self-gravity of the shell. Lower right: momentum of the shell (solid line labeled Psh), and the contribution from radiation pressure (long dash line), protostellar jets (solid line labeled Pjet), and shocked stellar wind gas (short dash line labeled Phot).

Standard image High-resolution image

Our estimated GMC mass in Arp 220 (5 × 107M) is a factor of 10 higher than the largest GMC mass seen in M82; this is a result of the much larger surface density in Arp 220 compared to M82. In contrast, the star cluster masses found so far in Arp 220 are only a factor 2–3 times larger than the masses of the largest clusters observed in M82, the latter being around (2–4) × 106M (McCrady & Graham 2007). Given that our predicted GMC star formation efficiency is not that different in the two cases (Table 2), we expect that more massive clusters are lurking in Arp 220.

The right panel of Figure 5 shows the forces acting on the shell of swept-up mass in our model of Arp 220. As in Figures 1 and 3, the force due to protostellar jets is initially similar to that due to radiation pressure. This situation lasts only while the shell accelerates from the initial clump radius of about 4 pc, until the shell reaches a little less than 6 pc. For the rest of the evolution radiation pressure provides the only significant outward force. The outward force supplied by ionized gas is completely negligible; the short dash line in the figure is the gas pressure multiplied by 100 (Figure 5, lower left panel). Both the hot gas and cosmic rays produced by shocked stellar winds are dynamically unimportant, even though we have assumed complete trapping of the shocked stellar wind material (an assumption that fails in the Milky Way; Harper-Clark & Murray 2009).

The middle panel of Figure 5 shows that radiation pressure will stir the ISM of Arp 220 to ∼50 km  s−1, somewhat less than the escape velocity from the cluster and similar to the velocity dispersion seen in CO observations.

5. DISCUSSION

5.1. The Importance of Radiation Pressure

Using four examples that cover conditions ranging from Milky Way-like spirals to the densest starbursts (see Tables 1 and 2), we have explored the physical processes that can disrupt GMCs, one of the basic building blocks of star formation. We find that radiation pressure produced by the absorption and scattering of starlight by dust grains can contribute significantly to disrupting GMCs in nearly all types of galaxies. By contrast, protostellar jets are important only at early times during GMC disruption, while the thermal gas pressure in H ii regions is important for GMC dispersal in spiral galaxies like the Milky Way, but not in more luminous starbursts. For the Milky Way and M82, where the observations are particularly detailed, our results demonstrate that observed massive star clusters have precisely the luminosities and structural properties required to disrupt Toomre-mass GMCs via radiation pressure.

The results presented here support Thompson et al.'s (2005) model of radiation pressure supported star-forming galaxies. In that paper, we focused on the large-scale properties of star formation in galaxies and the fueling of massive black holes in galactic nuclei. Here we have extended that model by taking into account the fact that star formation is not smooth and homogeneous; rather, most stars form in massive star clusters inside massive GMCs (Section 2). Our conclusion that Toomre-mass GMCs can be disrupted by radiation pressure is qualitatively and quantitatively similar to Thompson et al.'s conclusion that radiation pressure can regulate star formation in galactic disks to have Toomre's Q ≈ 1.

It is useful to consider simple scaling arguments in order to understand why, over the range of surface densities probed by the observed Schmidt law (10−3  g  cm-2 ≲ Σg ≲ 10  g cm−2), radiation pressure is the most viable mechanism for GMC disruption. To rough approximation, the self-gravity of the gas in a GMC is

Equation (12)

which varies by a more than a factor of ∼106 from normal galaxies to starbursts.

We can compare this force directly to the radiation pressure force. In the optically thin limit8

Equation (13)

where Ψ is the light-to-mass ratio in cgs units. Thus,

Equation (14)

where we have scaled to values appropriate to the Galaxy. We see that if epsilonGMC increases with gas surface density, as our calculations indicate (e.g., Table 2), then radiation pressure provides a mechanism for GMC disruption in both spiral galaxies like the Milky Way and in somewhat denser and more luminous galaxies.

Another way to express this is to note that the minimum efficiency

Equation (15)

For galaxies with sufficiently large surface densities, Σg ≳ 0.5 g cm−2, GMCs will be opaque to the emission by dust in the far-infrared (FIR). This increases the radiation pressure force so that (in the thin shell approximation used here)

Equation (16)

where τrad = κFIRΣsh/2, κFIR is the Rosseland mean opacity of the GMC in the FIR, and Σsh is the surface density of material in the shell. Comparing the optically thick radiation pressure force with that due to gravity, we find that

Equation (17)

where we have scaled to a relatively high value for the Rosseland-mean dust opacity (see below). Note that the ratio in Equation (17) does not explicitly depend on stellar/gas mass, because both Fsh and Frad are ∝M2. Using the scalings in Appendix A, it is easy to see that no other stellar feedback process has this property. Indeed, most of the previously suggested support mechanisms scale as FMβ* with β ⩽ 1, namely, H ii gas pressure, stellar winds, and pressures associated with shocked stellar winds. For this reason, although many feedback mechanisms are competitive with gravity in normal spirals, the self-gravity of the disk quickly overwhelms the forces due to stellar feedback in starburst galaxies. In contrast to these other feedback processes, radiation pressure in optically thick gas scales as FradM*Mg, so that it is at least, in principle, possible that radiation pressure can disrupt GMCs even in the densest, most gas-rich environments (e.g., ULIRGs and z ∼ 2 galaxies). Radiation pressure is, to our knowledge, the only stellar feedback process that has this property.

Equation (17) shows that the efficiency of star formation in a GMC at very high densities is sensitive to the metallicity and dust composition, which influence the FIR opacity κFIR, and to the stellar IMF, which determines the light-to-mass ratio of the stellar population Ψ. In very dense environments, there are some reasons for suspecting that the IMF may be top heavy (e.g., Murray 2009), as appears to be the case in regions of massive star formation more generally (Krumholz et al. 2007). If this is indeed the case, it would increase Ψ and thus decrease the epsilonGMC required for GMC disruption.9 For a relatively normal IMF, however, the star formation efficiency in GMCs must be appreciable at high densities, with epsilonGMC ∼ 0.25 or perhaps even larger.

5.2. The FIR Optical Depth

A key part of our argument for the importance of radiation pressure is the fact that star clusters have very high surface densities, which can trap the FIR radiation produced by dust, thus enhancing the radiation force by a factor of ∼τrad in the optically thick limit (Equation (17)). Figure 6 shows the Rosseland mean dust optical depth (τrad) through the shell as a function of shell radius rsh(t) in the M82 and Arp 220 models (compare with Figures 3 and 5). In the M82 model, τrad>1 for r ≲ 20 pc, while for Arp 220 τrad ≫ 1 at all radii. In our Milky Way models, by contrast, we find that the GMC is essentially always in the optically thin limit, i.e., opaque to the UV but not to the FIR. Figure 6 also shows the effective optical depth in both the M82 and Arp 220 models (dashed lines), which we define as

Equation (18)

The effective optical depth quantifies the enhanced coupling of photons emitted by a cluster in the center of a GMC compared to photons originating at the mid-plane of a uniform density disk. In M82, the effective optical depth at the end of the bubble evolution is about equal to the mean optical depth at the mid-plane. By contrast, in our Arp 220 models, the momentum deposited per photon in the bubble shell is a factor of ∼3 larger than would deposited by a photon traversing a uniform density disk. This shows that the effect of the radiation pressure may be 3 times larger than calculated using the mean mid-plane optical depth, as was done in Thompson et al. (2005). This effect increases the importance of radiation pressure in the densest galaxies, where it is needed most, effectively decreasing the normalization in Equation (17). Note also that because most of the star formation—and thus radiation—occurs in a few massive star clusters in the most massive GMCs (Section  2), there is very little "cancellation" due to different radiation sources driving the gas in different directions (as was suggested by Socrates et al. 2008); the distribution of radiation sources in galaxies is not well approximated as infinite and homogeneous (Section  2).

Figure 6.

Figure 6. Rosseland mean optical depth (1 + τrad) (solid lines) and the effective optical depth τeff (Equation (18)) (dashed lines) as a function of radius in the M82 (left) and Arp 220 (right) models, respectively. Note that τeff is a factor ∼3 larger than the mean optical depth of the disk in Arp 220 (the horizontal dotted line), enhancing the magnitude of the radiation pressure force.

Standard image High-resolution image

Given the turbulent and clumpy nature of the ISM in GMCs, one may question whether or not the photon coupling efficiency is as large as τrad or τeff, since these expressions assume uniform shells of matter. We have, after all, used the argument that GMCs are porous to argue that hot gas from shocked stellar winds escapes rapidly from GMCs in the Galaxy. The optical depth τrad is measured from the center of the protocluster outward and is proportional to the column density of overlying gas. The optical depth through a GMC has been observationally measured by a number of authors (e.g., Wong et al. 2008; Goodman et al. 2009) and is consistent with a log-normal distribution. Numerical simulations also find log-normal surface density distributions (Ostriker et al. 2001).10 Using the convention that $x = ln (N_{\rm H}/\bar{N}_{\rm H})$, Goodman et al. (2009) find a dispersion in the logarithm of the column density of 0.25 < σ < 0.53, and a range of mean logarithmic column densities −0.14 < μ < −0.02, where μ ≡ 〈x〉. The relation between μ and σ and their numerical values agree well with the high Mach number turbulence simulations of Ostriker et al. (2001).

In Appendix B, we estimate the reduction in the effective FIR optical depth, using an expression given in McKee & Tan (2008), for a log-normal distribution of column density. We find that the correction is of order 10% (see Figure 7). We apply this correction in our models for M82 and Arp 220; in the Milky Way and Q2346-BX 482, the optical depth to the FIR is so small that the correction is negligible.

Figure 7.

Figure 7. Ratio of the effective optical depth of a patch medium to the mean optical depth $\tau _{\rm patchy}/\bar{\tau }$, as given by Equation (B3). The lines correspond to $\bar{\tau }=10$, 30, and 50 coming down from the top of the figure. The reduction in the optical depth expected from the lognormal density distribution is rather modest, amounting to at most 30% for plausible values of μ.

Standard image High-resolution image

This correction is based on observations that do not probe, and simulations that do not include, some of the physics relevant to our models (e.g., the Rayleigh–Taylor and photon-bubble instabilities; Blaes & Socrates 2003; Turner et al. 2007; Thompson 2008; Krumholz et al. 2009). Nonetheless, these results provide some support for our argument that the dense ISM will be opaque to FIR radiation, thus significantly enhancing the radiation pressure force in the optically thick limit. In the absence of very large changes in the IMF (Ψ) or the average dust-to-gas ratio (κFIR), this enhancement is required for radiation pressure to disrupt GMCs in dense starbursts. For Milky Way-like galaxies, our conclusions are not sensitive to the radiation force in the optically thick limit, since the GMCs/star clusters are not optically thick except on very small scales.

McKee & Tan (2008) also give an expression for the reduction in optical depth when there is a literal hole in the shell wall, i.e., no absorption at all, so that photons stream out freely. If this were to occur in Arp 220 or M82, even for holes that pierce as little as 2% of the shell, our calculations show that essentially all the gas in the natal GMC must be in stars for radiation pressure to disrupt (what little is left of) the GMC.

Is it possible that the bubble walls in very dense regions do not have even such small holes? One observational hint comes from our own Galaxy. Observations of mature H ii regions clearly show breaches in bubble walls. Examples include the very extended free–free emission associated with giant H ii regions, and similarly extended 8 μm emission associated with massive star-forming regions (e.g., Murray & Rahman 2010).

Ultracompact H ii regions, on the other hand, furnish an example of almost completely confined UV and optical radiation fields. This is clear evidence against holes with a cover fraction of more than a few percent, since many of the central stars are luminous enough to power detectable free–free emission if 1% of UV light escapes. De Pree et al. (2000) give examples of objects in W49 with ionizing fluxes Q ∼ 2 × 1049 s−1; if 1% of this flux were to free stream out of the confined region, it would produce a free–free luminosity Lν ∼ 1.5 × 1021 erg s−1 Hz−1; at the distance of W49 (11 kpc) this is 10 mJy, easily detected in their maps. The gas densities in their objects reach n ∼ 106 cm−3, which makes them similar to our estimates of GMC densities in Arp 220. The surface densities of ultracompact H ii regions are also high, but two orders of magnitude or more below our estimates for Arp 220 GMCs, suggesting that radiation trapping in the latter case may be yet more complete. Clearly, more work is needed to address this question with any certainty.

5.3. The Origin of Large-scale Turbulent Motions

In all of our calculations, the shell velocity at rHd is similar to the turbulent velocity required to maintain Toomre's Q ∼ 1 in the ambient disk, thus staving off self-gravity on the largest scales. This is not a coincidence: disrupting the GMC—which we have explicitly taken to be the Toomre mass—requires a force comparable to that needed to stir up the gas in the disk proper to Q ∼ 1. We have interpreted the shell velocity at large radii as a turbulent velocity because the disruption of the shell by shear, tidal forces, and instabilities (e.g., Rayleigh–Taylor) will inevitably convert this relatively ordered kinetic energy into random fluid motions. More sophisticated models are clearly called for, but it is also clear that such models must include the effects of radiation pressure, particularly in very optically thick systems like Arp 220, but also in models of Milky Way-like galaxies.

5.4. The Global Star Formation Efficiency in Galaxies

A key question we are unable to fully address is how the efficiency of star formation in a GMC (epsilonGMC) relates to the global star formation rate in the galaxy as a whole, as encapsulated by, e.g., the parameter η in the Kennicutt law (Equation (1)) or the observed Schmidt law. We briefly discuss some of the relevant issues here, but leave a more detailed analysis of this problem to future work.

In our models of GMCs in Milky Way-like spirals, the time for a star cluster to disrupt the natal GMC is 1/3 to 1/2 the local dynamical time. When we add to this the time for the GMC to contract to its present size, the total time involved will be ∼2 longer than the local dynamical time of the disk. If the subsequent supernovae do not greatly prolong the process of reincorporating the bulk of the gas back into the disk, which we suspect is correct, the time averaged star formation rate per unit area will be

Equation (19)

with η given roughly by η ≈ epsilonGMC/2 ≈ 0.02, in good agreement with the observations.

The number of clusters capable of disrupting a Toomre-mass GMC, in our feedback model, should be proportional to the number of Toomre masses, ∼(R/H)2 ∼ 700. The corresponding number of giant H ii regions observable at any time is (R/H)2tMS/(2tdyn) ≈ 30, where tMS ≈ 3.6 × 106 years is the lifetime of an early O star, which is also the time over which a star cluster will emit a large luminosity of ionizing photons. This estimate is consistent with the fact that half of the free–free emission in the Galaxy is produced by about 17 giant H ii regions (Murray & Rahman 2010).

The GMC star formation efficiency is epsilonGMC = 0.38 for our fiducial Arp 220 model (Table 2); more generally, it is epsilonGMC ∼ 0.25 in the optically thick limit for typical IMFs and dust-to-gas ratios (Equation (17)). This suggests that η is also ≈0.25, higher than implied by the Kennicutt relation (Equation (1)), although reasonably consistent with the conclusions of Bouché et al. (2007). However, this estimate assumes that after the dispersal of the GMC, gas falls back into the disk and forms a new GMC in a single dynamical time; as the left panel of Figure 5 shows, the timescale for cluster disruption is in fact ∼1/10 of the timescale for the cluster luminosity to drop significantly. Thus, the star cluster luminosity can drive motions in the remaining gas ∼vT ∼ (H/R)vc, sufficient for hydrostatic equilibrium over many (∼10) dynamical timescales. Because the photon diffusion timescale is rapid compared to the dynamical timescale, the medium cannot be supported stably (Thompson et al. 2005; Thompson 2008). Hydrostatic balance will only be maintained in a statistical sense within a volume ∼4H3 of the star cluster. After the cluster's luminosity decreases on a timescale ∼tMS, the gas will recollapse to form a new GMC and the process will repeat until gas exhaustion. We suspect that this may reduce η by a factor of ∼tdyn/tMS, but clearly more work is needed.

We thank Reinhard Genzel and Chris McKee for useful conversations. N.M. is supported in part by the Canada Research Chair program and by NSERC. E.Q. is supported in part by NASA grant NNG06GI68G and the David and Lucile Packard Foundation. T.A.T. is supported in part by an Alfred P. Sloan Fellowship.

APPENDIX A: FORCES INCLUDED IN THE MODELS

In this section, we describe the forces acting on gas surrounding massive star clusters embedded in giant molecular clouds.

A.1. Inward Forces

A.1.1. Gravity

We assume that the gravitational force acting on the bubble shell consists of three components, Fgrav = Fstars + Fshell + Fdisk. The central star cluster exerts a force on the bubble shell given by

Equation (A1)

the shell self-gravity is

Equation (A2)

while the mass in the galactic disk exerts a force

Equation (A3)

The last force is that exerted on the part of the shell rising vertically away from the disk; gas in the plane of the disk will not feel this force, but we ignore this complication, just as we ignore Coriolis forces.

A.1.2. Turbulent Pressure

The pressure in the ISM and in the GMC provides an inward force on the shell. We refer to these pressures as turbulent pressure, although there may be other components such as magnetic or cosmic-ray pressure. We approximate them as

Equation (A4)

Here we have assumed that the GMC has a Larson-like mass distribution. A similar expression holds for clouds with ρ(r) ∼ 1/r2. The pressure of the ISM is given by

Equation (A5)

A.2. Outward Forces

A.2.1. Gas Pressure Forces

The large luminosity Q (number per second) of ionizing photons produced by the massive stars in a young star cluster will photoionize and heat gas in the vicinity of the cluster, raising the gas pressure above that in the neutral gas. This hot gas will exert an outward force on the bubble wall, with a magnitude

Equation (A6)

The gas pressure is taken to be

Equation (A7)

where T = 8000 K (Gardner et al. 1970) and

Equation (A8)

The volume V = 4πr3/3 and the recombination coefficient αrec ≈ 4 × 10−13. In models for clusters in the Milky Way, the shell velocity is subsonic, so the gas pressure is roughly constant throughout the bubble. If there is any hot gas or cosmic rays due to, for example, shocked stellar winds, they will also be roughly isobaric, and in pressure equilibrium with the H ii gas. If the hot gas or cosmic-ray pressure is in excess of the estimate given here, the H ii gas will be confined to a fraction of the bubble volume. However, observations of Carina and massive Milky Way and LMC clusters suggest that the pressure is well approximated by (A7) and (A8) (Harper-Clark & Murray 2009).

For large enough stellar clusters, those with at least one 35 M star, Q is proportional to L,

Equation (A9)

or about 3.6 Rydbergs (1Ryd ≈ 2.2 × 10−11 erg is the energy required to ionize hydrogen). The H ii gas pressure inside the bubble is then

Equation (A10)

assuming a unit filling factor for the H ii gas, i.e., ignoring the hot shocked winds.

It is helpful to have an estimate for n in terms of the luminosity. The number density n(L, r) is given by

Equation (A11)

If the bubble is breached, so that H ii gas leaks out, the pressure in the bubble will be lower than the estimate given here, but the force experienced by the bubble wall will actually be larger. The reason is that fewer ionizing photons are absorbed in the bubble cavity, leaving more to heat the gas on the interior of the bubble wall. The heated gas escapes away from the wall into a partial vacuum, exerting a force

Equation (A12)

on the wall. The mass loss rate is given by $4\pi r^2 m_p n c_{\rm H\,\mathsc{ii}}$. In our numerical work, we have assumed that the H ii gas leaks out of the bubble, since the rapidly expanding ionized gas at the ionization front will exert a larger force, and we want an upper limit on the efficacy of H ii gas pressure.

In models for ULIRGs like Arp 220, the bubble expansion velocity is larger than the sound speed of 8000 K gas, so the H ii pressure may be twice that in the subsonic case even if no gas escapes from the bubble. However, the pressure associated with H ii gas is negligible in ULIRGs, as we show in the main text.

A.2.2. Forces Associated With Shocked Stellar Winds

Hot gas from shocked stellar winds is often thought to be important in the formation of bubbles around massive stars or star clusters. We argue here that it is not; see also Harper-Clark & Murray (2009). We further argue that cosmic-ray pressure cannot be important in Milky Way bubbles; if they were, the bubbles would expand more rapidly than is observed. We show in this paper that neither hot gas nor cosmic-ray pressure is relevant in ULIRGs.

A.2.3. The Force Due to Hot, Shocked Stellar Wind Gas: X-ray Constraints

O stars emit high velocity massive winds; in clusters these winds are seen to shock and emit diffuse X-rays at ∼ KeV energies (Seward et al. 1979; Oey 1996; Smith et al. 2005). If the stellar winds are confined to the bubble interior, the associated pressure can be far larger than the ram pressure of the wind (Castor et al. 1975). The force is given by

Equation (A13)

where Ph = 2Eh/(3V). The energy equation for the hot gas is

Equation (A14)

recall that V is the volume of a bubble of radius r, and Λ is the cooling function. The wind luminosity Lw = (1/2)η(v/c)Lbol, (η ≈ 0.5 is the fraction of stellar luminosity scattered in the wind) and Lbol is the bolometric luminosity of the cluster.

We follow Castor et al. (1975) to find the number density nh of hot gas, i.e., we assume that heat conduction (at the Spitzer rate CSpT5/2h) drives a mass flow into the bubble interior (possibly supplemented from cold gas clouds in the interior of the bubble) at a rate

Equation (A15)

Given Eh and the radius of the bubble we find the hot gas pressure; from Equation (A15), we find nh, and hence the temperature. With nh and Th, we can find the cooling rate (the third term on the right-hand side of the energy Equation (A14).

The solar-mass stars in clusters also emit X-rays; in low resolution (non-Chandra) observations, these stars appear as diffuse emission, so care must be taken to distinguish the two. Assuming that one can extract the stellar emission, the diffuse X-ray flux can be used to constrain the pressure of any hot gas component. Harper-Clark & Murray (2009) use this to show that for clusters in the Milky Way and the LMC, the X-ray gas is in approximate pressure equilibrium with the H ii gas, and that the H ii gas has a filling factor near unity (greater than ∼0.1). The implication is that the shocked winds either radiate away the bulk of their energy, or else leak out of the bubble wall. In either case, they do not play any role in the dynamics of the bubble, a conclusion also reached by McKee et al. (1984).

Harper-Clark & Murray (2009) show that observations of ultracompact H ii regions (Rauw et al. 2002; Tsujimoto et al. 2006) reveal a similar story.

If the bubbles that form around star clusters in ULIRGs are also leaky, the shocked wind pressure would be negligible, as in the Milky Way. However, even if we assume that the winds are perfectly trapped, the high ambient pressures in ULIRGs ensure that the shocked gas is not dynamically important.

To see this, note that we expect a total number of young (ionizing) clusters in a ULIRG to be

Equation (A16)

If there are ∼20 clusters powering Arp 220, each must have Lcl ≈ 5 × 1010L, and a mass Mcl ≈ 2 × 107M assuming a normal (Muench et al. 2002) IMF; this cluster luminosity is also that needed to disrupt a GMC in Arp 220, as we showed in Section 4.4. The wind luminosity of a single such cluster is Lw ≈ 7 × 1041 erg  s−1. We note that Lcl ≈ 5 × 1010L is about 3–4 times the luminosities of known young star clusters in Arp 220 (Wilson et al. 2006).

From Figure 5, the bubble disrupts the GMC (RGMC ≈ 10 pc) after a time t ∼ 5 × 1012 s. Assuming no losses, the wind energy accumulated in the bubble is Eh = 2.5 × 1054 erg. The pressure is 2Eh/3V = 10−5 dynes cm−2, about equal to the ambient pressure in the disk but well below the radiation pressure, τF/c ≈ 10−4 dynes cm−2.

The force we have estimated from the hot gas is 1.6 × 1035 dynes; in the right panel of Figure 5 the force from the hot gas is a factor of ∼5 below this estimate. The reason is that the hot gas suffers radiative losses, so that we have overestimated Eh. We note that in our numerical work we have assumed that the hot gas fills the volume of the bubble; if it does not, the cooling rate will be higher than we have calculated.

A.3. The Force Due to Wind Generated Cosmic Rays

In addition to producing hot gas, wind shocks will generate cosmic rays. Perhaps a third of the wind energy may be expected to be deposited into cosmic rays. If, as observed in the Milky Way, the hot gas is advected out of the shell, it will probably take the cosmic rays with it, so that both the hot gas and the cosmic rays are not dynamically relevant.

We note that the pressure due to cosmic rays, if perfectly confined, would be 1/6 that of similarly confined shocked winds. As noted in the introduction, the dynamics of bubbles in the Milky Way and the LMC are not consistent with the high pressures associated with trapped hot gas; it follows that cosmic rays are probably also not confined to the interiors of such bubbles.

If the shocked winds in ULIRGs also flow through the swept-up shell, cosmic rays are unlikely to be dynamically important. Even if the cosmic rays are confined, the cosmic-ray pressure is still negligible. The cluster bolometric luminosity is 2 × 1044 erg  s−1, so the wind luminosity is 5 × 1041 erg  s−1. Assuming 1/3 of this is deposited in cosmic rays, the cosmic-ray luminosity is Lcr ≈ 1.6 × 1041 erg  s−1. The cosmic-ray pressure when the bubble radius is 10 pc is

Equation (A17)

This is about a half of the mean dynamical pressure in the galaxy, but only several percent of the dynamical pressure in the bubble. This is still likely to be an overestimate of the cosmic-ray pressure, as cosmic rays will quickly diffuse out of the bubble. The cosmic-ray mean free path in the Milky Way is 0.1 pc < λcr < 1 pc (Schlickeiser 2002); scaling to the lower value, the time for a cosmic ray to diffuse out of a GMC in Arp 220 is

Equation (A18)

where Dcr = λcrc/3 is the cosmic-ray diffusion coefficient. This diffusion time is ∼10–20 times shorter than the bubble dynamical time, so even if the mean free path for cosmic rays in Arp 220 is substantially smaller than in the Milky Way, cosmic rays will diffuse out of the bubbles in the ULIRG. Figure 5 shows that the cosmic-ray pressure is a little less than 1% of the radiation pressure; the lower cosmic-ray pressure is the result of taking Pcr = Lcrtcr/V, which accounts for diffusive losses, instead of Equation (A17). Note that we have optimistically ignored cosmic-ray losses due to inelastic proton–proton collisions, which cool the cosmic-ray population with energies ≳GeV on a timescale tpp ≈ 5 × 103(104cm-3/n) yr, where we have scaled to the average volumetric gas density of Arp 220.

A.3.1. Protostellar Jets

While stars are actively accreting in the protocluster, we assume that they will emit high velocity outflows or jets. Once the cluster disrupts, the accretion halts, and the jets turn off. We assume that this happens over a time given by ϕff times tff of the cluster; in addition, we allow for an extra factor of 2 to account for the fact that the accretion disks will not vanish once the cluster gas is dissipated, but instead will accrete onto the star in a disk viscous time:

Equation (A19)

We take ϕff = 1. The jets deposit momentum into the surrounding gas at a rate

Equation (A20)

The jets expel mass at a fraction epsilonjet ≈ 0.1–0.3 of the mass accretion rate (Matzner & McKee 2000), at a velocity vjet that depends on the mass of the star. We estimate vjet by calculating the escape velocity from the surface of a star of mass m, using the radius of a star of that mass found from the main-sequence radius as given by the Padova models; this is a slight over estimate of the escape velocity, since accreting stars are larger than main-sequence stars, but we compensate for this by using the escape velocity rather than the (larger) terminal velocity of a wind or jet. We then average over the IMF to yield vjet. For a Muench et al. IMF, we find vjet ≈ 3 × 107 cm s−1, but for top heavy IMFs this can increase to 108 cm s−1. We varied epsilonjet over the range 0.01–0.3 giving an effective velocity in the range 5–90 km s−1, bracketing that found by, e.g., Matzner & McKee (2000) and Nakamura & Li (2007). The results did not depend strongly on this parameter, but because the clusters we consider are so dense, we expect a substantial amount of momentum cancellation (from jets colliding nearly head on) we used low values for epsilonjet, typically 0.03 in the massive cluster models, i.e., not the Milky Way model, presented here.

A.3.2. Radiation Force

The radiation force is given by

Equation (A21)

where τrad is the Rosseland mean opacity through the shell. The gas in the shell has a temperature that is of order 100 K, so for most clusters in the Milky Way τrad < 1. We calculate the optical depth as follows: we assume that the shell has a thickness 1/10 of its radius, and divide it into 100 pieces. We assume that the density is constant through the shell. We calculate Teff from L and the radius r of the shell. Given T and ρ, we use the opacity tables of Semenov et al. (2003) to find the Rosseland mean opacity in the outermost slice of the shell. Subsequently, we find the optical depth, then integrate the radiative transfer equations inward through the shell. The factor of unity in Equation (A21) accounts for the fact that the temperature of the radiation field is of order 30,000 K at the inside edge of the shell, before the photons have encountered any dust grains; upon striking a dust grain, the photons provide an impulse per unit time given by L/c.

APPENDIX B: THE EFFECT OF A LOG-NORMAL DENSITY DISTRIBUTION

In calculating the radiation pressure when the bubble wall is optically thick to FIR radiation, a first estimate of the optical depth is the product of the opacity and the mean value of the column density through the shell,

Equation (B1)

where $\bar{N}_{\rm H}\equiv \Sigma {sh}/m_p$ is the mean column density through the shell.

However, the material in the shell is swept-up interstellar material, and it is known that the column density through the ISM is path dependent. There is good evidence from numerical simulations that the column density distribution is log normal, i.e., that

Equation (B2)

We have already noted that the shell is prone to instabilities, but the growth time of the instabilities will be little faster than the dynamical time. Since the shell radius grows on the dynamical time, the swept-up gas is likely to retain its initial column density distribution. It follows that Equation (B2) should be a better characterization of the shell column than simply taking the mean swept-up column.

Another point regards the effects of stellar winds. In low surface density galaxies like the Milky Way, the pressure of shocked stellar wind arising from a cluster of massive stars is far above the hydrostatic pressure at the disk mid-plane. Hence, we should have expected that bubbles will be punctured by such winds, and this is what is seen in the Milky Way. However, in a galaxy in which the surface density is high, such as Arp 220, the hot gas pressure that shocked winds can maintain is below the hydrostatic pressure of the overlying gas; the winds cool, as we saw in Section 4.4. In that case, the bubbles are likely to retain the column density distribution they inherit from the ISM.

Since the column density varies from one direction to another, one might wonder if the "holes" corresponding to low columns will allow most of the FIR radiation that is trapped inside the bubble (when the mean optical depth is large) to leak out, reducing the force on the shell.

We show that this is not the case; the force on the shell is reduced by no more than ∼30%, using current models of $f(\ln (N_{\rm H}/\bar{N}_{\rm H}))$. The fundamental reason behind this result is that, even though the column density distribution is log normal, it is still not that broad, meaning that there are effectively few holes in the bubble wall when the column density is large.

The relevant characterization of a "hole" in this context is the effect of a hole on the trapping of FIR radiation. In a galaxy like Arp 220, the dominant opacity is dust opacity. However, since the effective optical depth in the bubble wall is large, an assertion we are about to check, the radiation field is in rough local thermodynamic equilibrium. Crudely speaking, when a photon is absorbed by a dust grain, it is replaced by another photon of about the same energy (moving in a random direction) thermally emitted by that dust grain. We then make use of an analogy to Lyα photons in a scattering dominated regime, and appropriate the results of McKee & Tan (2008). These authors present an expression for the effective optical depth τpatchy of a patchy medium,

Equation (B3)

In this expression, we denote the mean optical depth by $\bar{\tau }$. The term in the denominator involving $4/3\bar{\tau }$ accounts for the presence of very low column density holes.

Figure 7 shows the ratio of the effective optical depth to the mean optical depth, $\tau _{\rm patchy}/\bar{\tau }$, as a function of the mean of the logarithm of the column density, $\mu \equiv \langle \ln N_{\rm H}/\bar{N}_{\rm H}\rangle$, for the density distribution given in Equation (B2).

We discussed the rather weak constraints on the value of μ in Section 5.2.

Footnotes

  • A viral parameter αvir = 1 corresponds to a cloud that is just gravitationally bound, while the smaller value of αvir = 0.1 corresponds to a cloud that is approaching free-fall conditions.

  • By this, we mean that the GMC is optically thick to the UV, but optically thin to the re-radiated FIR emission.

  • Note that as IMF becomes arbitrarily top-heavy Ψ → 4πGcT, where κT is the Thomson opacity. This sets a minimum on the ratio Fsh/Frad for any stellar population: Fsh/Frad|min → (κTFIR)epsilon−1GMC ∼ 10−2(30 cm2  g-1FIR)epsilon−1GMC.

  • 10 

    The observations measure τrad along the line of sight from the Earth through the cloud rather than from the center of the cloud outward, but the two surface density distributions should not differ dramatically.

Please wait… references are loading.
10.1088/0004-637X/709/1/191