Articles

STAR FORMATION EFFICIENCIES AND LIFETIMES OF GIANT MOLECULAR CLOUDS IN THE MILKY WAY

Published 2011 February 18 © 2011. The American Astronomical Society. All rights reserved.
, , Citation Norman Murray 2011 ApJ 729 133 DOI 10.1088/0004-637X/729/2/133

0004-637X/729/2/133

ABSTRACT

We use a sample of the 13 most luminous WMAP Galactic free–free sources, responsible for 33% of the free–free emission of the Milky Way, to investigate star formation. The sample contains 40 star-forming complexes; we combine this sample with giant molecular cloud (GMC) catalogs in the literature to identify the host GMCs of 32 of the complexes. We estimate the star formation efficiency epsilonGMC and star formation rate per free-fall time epsilonff. We find that epsilonGMC ranges from 0.002 to 0.2, with an ionizing luminosity-weighted average 〈epsilonGMCQ = 0.08, compared to the Galactic average ≈0.005. Turning to the star formation rate per free-fall time, we find values that range up to $\epsilon _{\rm ff }\equiv \tau _{\rm ff }\cdot \dot{M}_*/M_{\rm GMC}\approx 1$. Weighting by ionizing luminosity, we find an average of 〈epsilonffQ = 0.14–0.24 depending on the estimate of the age of the system. Once again, this is much larger than the Galaxy-wide average value epsilonff = 0.006. We show that the lifetimes of GMCs at the mean mass found in our sample is 27 ± 12  Myr, a bit less than three free-fall times. The GMCs hosting the most luminous clusters are being disrupted by those clusters. Accordingly, we interpret the range in epsilonff as the result of a time-variable star formation rate; the rate of star formation increases with the age of the host molecular cloud, until the stars disrupt the cloud. These results are inconsistent with the notion that the star formation rate in Milky Way GMCs is determined by the properties of supersonic turbulence.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Giant molecular clouds (GMCs) are stellar nurseries—most stars in most galaxies are born in GMCs. However, the fraction of a GMC's gas that ends up in stars before the cloud is disrupted is rather small, with typical estimates in the Milky Way ∼2%, e.g., Evans (1991). A related question regards the rate at which stars form in GMCs. Global measurements in both the Milky Way and external galaxies (Kennicutt 1998) establish that ∼2% of the molecular gas is turned into stars over the galactic dynamical time R/vc, where R is the exponential scale length of the disk and vc is the circular velocity. Because gas disks are in hydrostatic equilibrium, the galactic dynamical time is also the disk dynamical time H/vT, where H is the disk scale height and vT is the turbulent velocity of gas in the disk. Finally, the largest GMCs have diameters that are of order the molecular gas disk scale height (Malhotra 1994) and Table 1 below, and CO linewidths ∼5 km  s−1, e.g., Solomon et al. (1987), so GMC dynamical times are again of the same order. This implies that only a few percent of the mass of a GMC is turned into stars in the GMC free-fall time.

Table 1. Star-forming Complexes and Giant Molecular Cloud Identifications

Cat WMAP l b vr D RSFR fν Ref. Cat l b vr RGMC MGMC
No. Name (deg) (deg) ( km  s−1) ( kpc) ( pc) (Jy)   No. (deg) (deg) ( km  s−1) (pc) (M)
1 G10 10.156 −0.384 15 14.5 25.0 29 1 14 10.00 −0.04 32 39.0 1.3e+05
2 G10 10.288 −0.136 10 15.2 25.0 150 1 15 10.20 −0.30 8 29.0 7.5e+05
3 G10 10.450 0.021 69 6.1 5.0 50 ... ... ... ... ... ... ...
4 G10 10.763 −0.498 −1 16.7 46.0 29 1 16 10.60 −0.40 −2 76.0 2.8e+05
5 G24 22.991 −0.345 76 10.8 49.0 124 2 97 23.00 −0.36 77 83.4 2.1e+06
6 G24 23.443 −0.237 104 9.2 5.0 482 2 100 23.39 −0.23 100 23.3 4.1e+05
7 G24 23.846 0.152 95 5.8 12.0 138 2 105 23.96 0.14 81 8.2 2.1e+04
8 G24 24.050 −0.321 85 10.3 61.0 55 2 106 24.21 −0.04 88 22.7 8.2e+04
9 G24 24.133 0.438 98 9.7 4.0 41 2 116 24.63 −0.14 84 25.8 6.9e+04
10 G24 24.911 0.134 100 6.1 43.0 179 2 118 25.18 0.16 103 23.3 1.2e+05
11 G24 25.329 −0.275 63 4.1 7.0 248 2 121 25.40 −0.24 58 52.0 5.9e+05
12 G24 25.992 0.119 106 8.8 28.0 110 2 125 25.72 0.24 111 42.7 1.5e+05
13 G30 28.827 −0.230 88 5.3 2.0 119 2 154 28.80 −0.26 88 16.3 3.2e+04
14 G30 29.926 −0.049 97 8.7 13.0 594 2 162 29.89 −0.06 99 34.9 8.3e+05
15 G30 30.456 0.443 58 3.6 3.0 36 2 165 30.41 0.46 45 5.0 1.7e+03
16 G30 30.540 0.022 44 11.9 3.0 65 2 168 30.57 −0.02 41 44.7 2.7e+05
17 G30 30.590 −0.024 99 6.3 35.0 753 2 171 30.77 −0.01 94 41.8 1.1e+06
18 G30 32.162 0.038 96 8.2 12.0 29 2 182 32.02 0.06 97 37.7 1.6e+05
19 G34 34.243 0.146 37 2.2 5.0 130 ... ... ... ... ... ... ...
20 G34 35.038 −0.490 51 3.0 3.0 130 1 200 34.70 −0.70 46 35.0 1.6e+06
21 G34 35.289 −0.073 48 11.0 62.0 25 ... ... ... ... ... ... ...
22 G37 37.481 −0.384 50 10.5 74.0 130 2 211 37.76 −0.21 63 20.8 7.7e+04
23 G37 38.292 −0.021 57 3.5 10.0 114 ... ... ... ... ... ... ...
24 G49 49.083 −0.306 68 5.6 11.0 229 1 233 49.50 −0.40 57 52.1 3.3e+06
25 G49 49.483 −0.343 60 5.7 18.0 229 2 234 49.75 −0.55 68 13.0 9.0e+04
26 G283 283.883 −0.609 0 4.0 63.0 848 3 7 283.80 0.00 −5 65.0 1.3e+06
27 G291 290.873 −0.742 −22 3.0 27.0 456 ... ... ... ... ... ... ...
28 G291 291.563 −0.569 16 7.4 67.0 232 3 17 291.60 −0.40 15 73.0 1.4e+06
29 G298 298.505 −0.522 24 9.7 69.0 313 3 26 298.80 −0.20 25 157.0 8.4e+06
30 G311 310.985 0.409 −51 3.5 3.0 56 ... ... ... ... ... ... ...
31 G311 311.513 −0.027 −55 7.4 56.0 590 4 7 311.70 0.00 −50 118.0 1.2e+06
32 G311 311.650 −0.528 34 13.6 29.0 114 3 35 311.30 −0.30 27 210.0 1.3e+07
33 G327 327.436 −0.058 −60 3.7 36.0 977 4 25 327.00 −0.40 −60 60.0 3.4e+05
34 G332 332.809 −0.132 −52 3.4 39.0 1192 4 33 332.60 0.20 −47 56.0 6.6e+05
35 G332 333.158 −0.076 −91 5.5 4.0 596 4 35 333.90 −0.10 −89 111.0 1.2e+06
36 G337 336.484 −0.219 −88 5.4 6.0 356 ... ... ... ... ... ... ...
37 G337 336.971 −0.019 −74 10.9 49.0 365 4 38 337.10 −0.10 −74 162.0 4.4e+06
38 G337 337.848 −0.205 −48 3.5 18.0 750 4 39 337.80 0.00 −56 141.0 4.0e+06
39 G337 338.412 0.120 −33 2.5 4.0 640 ... ... ... ... ... ... ...
40 G337 338.888 0.618 −63 4.4 11.0 128 4 41 339.00 0.70 −60 38.0 2.4e+05

Notes. Properties of Milky Way star-forming complexes and their host GMCs. Column 1 gives the Rahman & Murray (2010) catalog number of the star-forming complex, while Column 2 gives the associated WMAP catalog name. The next six columns give star formation complex properties: the Galactic longitude l (Column 3), latitude b (Column 4), heliocentric radial velocity vr (Column 5), heliocentric kinematic distance D (Column 6), the radius of the complex (Column 7) and free–free flux (Column 8). The ninth column gives reference to the relevant GMC catalog. Column 10 gives the catalog number of the GMC, followed by the GMC properties in the next five columns: l (Column 11), b (Column 12), radial velocity (Column 13), GMC radius (Column 14), and mass (Column 15). References: (1) Solomon et al. 1987; (2) Heyer et al. 2009; (3) Grabelsky et al. 1988; (4) Bronfman et al. 1989. The masses and radii listed are those in the original publications; in the calculations described in this paper, and in Table 2, the GMC radii have been adjusted to R0 = 8.5 kpc and the masses have been adjusted as described in the text.

Download table as:  ASCIITypeset image

The lifetimes of GMCs have been a matter of considerable and intense debate, with estimates ranging from a single dynamical or free-fall time τff (Elmegreen 2000; Hartmann et al. 2001), to a few (Elmegreen 2007), up to 10 or more (Scoville & Hersh 1979; Scoville & Wilson 2004), although it is important to note that the last authors were careful to distinguish between the lifetime of a typical H2 molecule (which is likely in excess of 10 GMC free-fall times) and that of the GMC of which it is a part. Recent work using the NANTEN telescope has determined that GMCs in the Large Magellanic Cloud have a lifetime of 20–30 Myr (Kawamura et al. 2009; Fukui & Kawamura 2010), about 3τff.

If GMCs live much longer than 2–3 free-fall times, the question arises as to what supports them. As just noted, individual GMCs are observed to have CO linewidths (on large scales) of sufficiently high magnitude to support them against their own gravity. However, it is likely that such motions, often referred to as turbulence, decay on a dynamical time (Mac Low 1999; Ostriker et al. 2001). Hence, long-lived clouds require some energy source if their turbulence is to be maintained. Another possible way to prevent GMC collapse is to appeal to dynamically significant or dominant magnetic fields (e.g., Mouschovias 1976; Shu 1983); such fields (or weaker fields) could be the carriers of MHD turbulence (Arons & Max 1975), possibly from sources outside the cloud, which would obviate the need to maintain the turbulence from inside the cloud.

However, it is also possible that GMCs only live for 2–3 free-fall times, in which case neither magnetic fields nor turbulence are needed to support them; they could be in free fall, as suggested originally by Goldreich & Kwan (1974). These authors pointed out that the similarity between the optically thick CO emission lines and optically thin emission lines from other molecular species in the same clouds could be understood if the clouds had bulk motions with velocities in excess of the thermal linewidths. They argued that there was no clear way to support GMCs, so global collapse was to be expected, leading to large-scale bulk motions needed to explain the observed similarity in line profiles.

This suggestion was disputed by Zuckerman & Evans (1974). They give three arguments against free-fall collapse of GMCs. First, they argue that the resulting star formation rate would be too large. They arrive at this conclusion by assuming that upon collapse, all the gas in a GMC would be converted into stars. This ignores any feedback from the stars on the evolution of the cloud. As shown by Murray & Rahman (2010), Rahman & Murray (2010), and this paper, there is solid evidence that stellar feedback disrupts the host GMC when a fraction epsilonGMC ≈ 0.08 of the GMC mass is converted into stars. This eviscerates the star formation rate argument of Zuckerman & Evans (1974), a point also made by Elmegreen (2007).

The second argument of Zuckerman and Evans is that the line profiles of molecules such as H2CO, which is seen in absorption toward H ii regions, should show relative velocity shifts compared to the CO emission lines; the H2CO absorption will be blue shifted if the H ii regions are near the centers of the host GMC and the cloud is expanding, since the portion of the cloud between Earth and the H ii region is moving toward the observer, while the absorption will be red shifted if the cloud is collapsing. This argument makes the assumption that the absorption lines are formed by gas in the outer regions of the GMC, between the observer and the H ii regions, which provide the background continuum against which the H2CO lines are seen. The observations described in Rahman & Murray (2010) establish the presence of outward velocities (in the expanding bubble walls) somewhat in excess of the turbulent velocities measured by CO observations of the GMCs we examine. Despite this, neither double-peaked CO emission lines, nor redshifted (relative to the CO lines) absorption lines have been reported toward these star-forming complexes (SFCs). High-resolution molecular line observations toward individual bright polycyclic aromatic hydrocarbon emission regions might reveal both red and blue shifted CO emission, and associated molecular absorption lines.

The third argument of Zuckerman & Evans (1974) against global infall of GMCs is that there appear to be many infall centers in a given GMC. However, there is no reason to believe that there will be a single infall center in a cloud that is undergoing global collapse: given the clumpy nature of the turbulent interstellar medium (ISM), any large clump will act as a local center for collapse of gas in its immediate vicinity.

In this paper, we estimate the star formation efficiency, and efficiency per free-fall time, of Milky Way GMCs selected by their free–free luminosity. The star formation efficiency epsilonGMC is measured on a per-GMC basis; in principle it is defined as the total mass in stars, produced over the lifetime of the GMC, divided by the initial GMC mass. However, the observationally accessible quantities in the case of massive clusters are the stellar mass inferred from the ionizing flux (via free–free or Hα emission) and the current GMC mass, so that is what we use here. The star formation efficiency per free-fall time epsilonff is the fraction of a GMC (or other object) that is turned into stars per free-fall time of the GMC (Padoan 1995). Recent discussions provide both observational estimates of epsilonff (Krumholz & Tan 2007) and theoretical estimates (e.g., Krumholz & McKee 2005; Krumholz et al. 2009).

We show that both efficiencies vary widely, the latter by nearly three decades. Both have large upper limits of order ∼0.3 and ∼0.6, respectively. We also estimate the lifetimes of massive GMCs, finding that they live 1–3 free-fall times, consistent with the picture of Goldreich & Kwan (1974). In Section 2, we give our definition of the two efficiencies epsilonGMC and epsilonff. In Section 3, we describe how we select our SFCs, and how we identify their host GMCs. We estimate the lifetimes of the host GMCs in Section 4. We outline the implications of our results and discuss its relation to previous work in Section 5; in particular, we compare our results with the predictions of turbulent star formation theories in Section 5.2. We present our conclusions in the final section.

2. THE STAR FORMATION EFFICIENCY AND RATE PER FREE-FALL TIME OF MILKY WAY GIANT MOLECULAR CLOUDS

In this section, we define and compute two types of star formation efficiency. The first is the efficiency epsilonGMC of star formation in GMCs. This is nominally the fraction of a GMC that is converted into stars over the lifetime of a GMC. In fact what we actually measure in this paper is the fraction of a GMC's mass that currently resides in that GMC in the form of ionizing star clusters. This is because we find stars (and their parent GMCs) by using Wilkinson Microwave Anisotropy Probe (WMAP) to look for free–free radiation, which is powered almost exclusively by massive stars.

2.1. Star Formation Efficiency in a GMC

Star clusters emit ionizing radiation for a rather short time. Following McKee & Williams (1997), we define the ionization-weighted (or Q-weighted) stellar lifetime is

Equation (1)

In this equation

Equation (2)

where we take a minimum stellar mass of mL = 0.1 M and a maximum stellar mass of mU = 120 M. The quantity dN/dln m is the stellar initial mass function (IMF), which we take to be either that of Muench et al. (2002) modified to have a high mass slope of −1.35 rather than their −1.21, or the IMF of Chabrier (2005). With our modification to the Muench et al. slope, the two IMFs are almost identical. We take the ionizing flux q(m*) of a star of mass m* on the zero-age main sequence from Martins et al. (2005), extended to stars of 120 M using the results of Martins et al. (2008). The main sequence lifetime tms of a star of mass m* is taken from Bressan et al. (1993). The quantity 〈q〉 is the number of ionizing photons emitted per second per star, averaged over the IMF.

The integrand in Equation (1) is small for small m*, since low-mass stars produce little ionizing radiation (q(m*) is small), so the integral is not particularly sensitive to the lower limit. In other words, the integral is dominated by the high-mass end of the IMF, and by the very highest mass stars, m* ≳ 40 M in particular. The reason for the latter is that tms(m*) is a rather slowly varying function of mass for m* ≳ 40 M, while q(m*) continues to increase. We find 〈tms〉 ≈ 3.9 Myr.

It is worth noting that if a star cluster forms in time much less than 〈tms〉, then that cluster is effectively undetectable via its ionizing radiation after ∼5  Myr. This is so because the ionizing flux of such a cluster is roughly constant (for slightly less than 〈τms〉), and then plunges rapidly as the most massive stars in the cluster evolve off the main sequence; see Figure 1. In a free–free-selected population of such rapidly formed clusters, the average age will be 〈tms〉/2 ≈ 2  Myr.

Figure 1.

Figure 1. Ionizing flux of a massive cluster Q(t) vs. time. The solid line is calculated from the stellar models of Bressan et al. (1993). The dashed line is a top-hat model, with the width set to the Q-averaged main sequence lifetime of 3.9 Myr, appropriate for a cluster with a Muench et al. (2002) initial mass function, modified to have a high mass slope of −1.35.

Standard image High-resolution image

The mass of "live" stars, i.e., the mass in the clusters (of age less than 〈τms〉) containing the ionizing stars, is given by

Equation (3)

where

Equation (4)

and the IMF averaged stellar mass 〈m*〉 is defined in a manner similar to 〈q〉 in Equation (2).

Figure 1 also illustrates a weakness associated with using measurements of the ionizing flux to estimate the stellar mass. Suppose a cluster forms in a period of time short compared to 〈τms〉, and suppose further that we measure Q for this cluster when it has an age of 4  Myr. Since Q(4 Myr) ≈ 0.5Q(0), we will infer that the cluster mass is about half its true mass. More generally, the masses of clusters with ages 3 Myr < τ < 5 Myr will be underestimated by a factor ranging from ∼1–5. On the other hand, as was just noted, clusters older than ≈5  Myr have Q(t) ≪ Q(0), and so will be difficult to detect at all using any measure related to Q.

Since we can measure the mass M* in live stars, we define the efficiency of star formation as

Equation (5)

As just noted, this stellar mass estimate M* refers only to stars younger than ∼4 Myr. If a star cluster takes longer than 〈τms〉 to form, then the stellar mass associated with that cluster will be larger than that given by Equation (3). Most GMCs have probably given birth to stars older than 〈τms〉, so M* as defined here is a lower limit on the total mass of stars formed in any particular GMC.

2.2. Star Formation Efficiency per Free-fall Time

The second efficiency we calculate is called the star formation efficiency per free-fall time, epsilonff. For star formation in objects (such as GMCs) of class X, this is defined as

Equation (6)

where τff-X is the free-fall time of objects of class X, MX is the total gas mass in objects of class X, and $\dot{M}_{*X}$ is the star formation rate in such objects (Krumholz & Tan 2007). We restrict our attention in this paper to GMCs.

We start by computing the Galaxy-averaged value of epsilonff. The total molecular gas mass in the Milky Way is Mtot ≈ 109M, e.g., Dame (1993). The total ionizing luminosity of the Milky Way is 3.2 × 1053  s−1 (see, e.g., Murray & Rahman 2010, and references therein).

The star formation rate is given by (Smith et al. 1978)

Equation (7)

Using Equation (4) and Q = 3.2 × 1053  s−1, Murray & Rahman (2010) found $\dot{M}_*=1.3\,M_\odot \,\,{\rm yr }^{-1}$.

As noted above, essentially all stars form in GMCs, and essentially all the molecular gas is in GMCs, so we use MX = Mg = 109M. The free-fall time for a GMC varies with the GMC mass, but for now we use the value assumed by Krumholz & Tan (2007), τff = 4.4 Myr. The Galaxy wide average star formation efficiency per free-fall time is then

Equation (8)

This is roughly a factor of three smaller than found by Krumholz & Tan (2007). They used $\dot{M}_*=3\,M_\odot \,\,{\rm yr }^{-1}$, asserting that this was the value found by McKee & Williams (1997) (although the latter authors actually found $\dot{M}_*=4.0\,M_\odot \,\,{\rm yr }^{-1}$). The reason for the larger star formation rate is that McKee & Williams (1997) used an IMF with a large mass to light ratio, that of Scalo (1986), which is no longer believed to be a good representation of the actual IMF in the Milky Way.

3. STAR-FORMING GIANT MOLECULAR CLOUDS

In this section we measure the the efficiency, epsilonGMC, of star formation in individual Milky Way GMCs, and the efficiency of star formation per free-fall time, epsilonff. We select the most rapidly star-forming GMCs in the Galaxy: the 32 GMCs we select are responsible for 31% of the star formation in the Galaxy.

We start with the WMAP free–free thermal radio fluxes of sources found by Murray & Rahman (2010). Examination of Spitzer and MSX images in the direction of the WMAP sources reveals one to several "star-forming complexes" in the direction of each of the WMAP sources (Rahman & Murray 2010). A SFC is identified on the basis of the morphology in the 8 μm Spitzer and MSX images, combined with measurements of radio recombination line (or molecular line) radial velocities taken from the literature. The radial velocity data are given in Table 4 of Rahman & Murray (2010). The result is a list of 40 SFCs with Galactic coordinates and radial velocities, (l, b, vr). These complexes have characteristic sizes of ∼25 pc, ranging between 2 pc and 70 pc (see Table 1). The spread in vr is ≲15–20 km  s−1.

Once all the SFCs are identified, we divide the free–free flux in each WMAP source between the SFCs contained in that source. In two cases, G283 and G327, there is only a single SFC in the WMAP source. We assign the entire free–free flux of the WMAP source to that SFC; in both cases the size of the SFC (determined by the extent of the region outlined by the radio recombination line velocity measurements) is similar to that of the WMAP source, supporting the view that only that source is responsible for the free–free emission seen by WMAP in that direction.

In four of the WMAP sources there are exactly two SFCs; we split the flux between the SFCs based on the radio recombination line fluxes associated with each source. Such a division is a rather crude approximation; ground-based radio continuum fluxes toward each of these complexes are systematically lower than the continuum fluxes measured by WMAP, but the flux ratios vary unpredictably from object to object. In its favor, this recombination line ratio split was consistent with the relative amount of 8 μm flux associated with the SFCs (accounting for the fact that the total 8 μm flux appears to scale as the square of the free–free flux, Rahman & Murray (2010)).

Rahman & Murray (2010) identified four SFCs in G10, three in G34 and G311, and five in G337. In all these cases the SFCs are well separated in (l, b, vr) space, as was the 8 μm emission in (l, b). The division of the free–free emission amongst the sources was made in the same manner as for the previous cases.

The two WMAP sources G24 and G30 are very confused, with eight and six SFCs, respectively; the corresponding catalog numbers are in the range 5–18 inclusive. We again used the recombination line ratios to assign the free–free flux to each source, but we have less faith in the results; to indicate this, we plot these points using open polygons, as opposed to the filled polygons plotted for the less confused WMAP sources.

Having assigned fluxes to each SFC, we then use the kinematic distance D to that region to calculate the free–free luminosity Lν = 4πD2fν emitted by that region, and the rate Q = 1.33 × 1026Lν s−1 (Murray & Rahman 2010) of ionizing photons emitted by each source per second required to power the observed free–free luminosity. In cases where the distances listed in Rahman & Murray (2010) and the relevant GMC catalog do not agree, we use the distance of the former. In the case of Rahman & Murray sources 5, 7, 9, 12, 15, and 35, where those authors do not resolve the distance ambiguity, we use the distance that is closer to that of the relevant GMC catalog distance.

Following McKee & Williams (1997) we then increase the rate of ionizing photons by a factor 1.37 to account for the fact that some of those photons are absorbed by dust, and hence do not contribute to the free–free emission detected by WMAP.

The live stellar mass was then calculated using Equation (3); a given GMC almost certainly has given birth to stars older than 〈tms〉, so M* as defined here is a lower limit on the total mass of stars formed in the GMCs we examine.

The next step in the calculation of epsilonGMC and epsilonff is to identify the host GMCs of our SFCs. We search the GMC catalogs of Solomon et al. (1987), Grabelsky et al. (1988), Bronfman et al. (1989), and Heyer et al. (2009) for objects having centers at the same Galactic longitude within ±0.4 deg, Galactic latitude within ±0.4 deg (∼50  pc at D = 10  kpc, comparable to the radius of a typical cloud), and the same heliocentric radial velocity within ±20 km  s−1.

We find matches for 32 out of our 40 SFCs; results are given in Table 1. A number of GMCs are listed in both Solomon et al. (1987) and Heyer et al. (2009), in which case we use the values in the latter. We corrected the GMC radii to account for the difference between our assumed value for the distance to the Galactic center R0 = 8.5  kpc and that used in the GMC surveys. We have followed Williams & McKee (1997, their Table 1) in correcting the masses of the GMCs in the first three surveys. Heyer et al. use the same distance to the Galactic center we do (8.5  kpc); we use MGMC = 2 ×MLTE, as those authors recommend (corresponding to $X_{\rm CO}=1.9\times 10^{20}\,{\rm cm }^{-2} (\rm{K }\,{\rm km \,\, s}^{-1})^{-1}$).

Two of the matches are to objects in Grabelsky et al. (1988) that those authors identify as "cloud complexes," assigning them catalog numbers 26 and 35 (corresponding to our numbers 29 and 32). The complexes contain more than one GMC, but Grabelsky et al. (1988) do not give separate masses or radii for the component GMCs. The smaller of the two complexes has a mass nearly twice that of the most massive GMC, and is well separated from the GMCs in the plots below. In the figures we plot these complexes as open circles, but connect them to filled squares; the latter represents an attempt to break up the complexes, assuming that the most massive GMC contains half the mass of the entire complex. However, in calculating the various averages presented below, we have simply used the total mass contained in the complex.

The average star-forming GMC mass in our ionizing luminosity-selected sample is 1.5 × 106M. The ionizing luminosity-weighted GMC mass is 2.3 × 106M. The average and Q-weighted GMC radii are 61  pc and 82  pc, respectively. Similarly, the average and Q-weighted free-fall times are 9  Myr and 10  Myr.

These averages include the confused GMCs associated with SFCs 5 to 18. These GMCs have been investigated by both Solomon et al. (1987) and Heyer et al. (2009); the masses of these GMCs differ between the two studies (as noted above, we use the results of Heyer et al. 2009 where possible). While the average properties of the confused sources (and their host GMCs) are similar to those of the unconfused sources, the measured properties of individual confused sources may be unreliable.

If we confine our attention to unconfused GMCs, the average and Q-weighted radii are 85  pc and 110  pc. Similarly, the average and Q-weighted free-fall times for the unconfused GMCs are 13  Myr and 15  Myr, while the corresponding masses are 2.1 × 106M and 2.9 × 106M.

Having found the mass of the host GMCs, we can estimate the star formation efficiency. The results are plotted in Figure 2.

Figure 2.

Figure 2. Ratio epsilonGMCM*/(MGMC + M*) of the mass of young (t < 〈τms〉) stars in star-forming Milky Way GMCs of mass MGMC. Filled triangles are for GMCs in Solomon et al. (1987) and unconfused GMCs in Heyer et al. (2009), open triangles correspond to the 14 confused GMCs in SFCs G24 and G30 (Heyer et al. 2009), filled squares are for GMCs in Grabelsky et al. (1988), and filled pentagons are for GMCs in Bronfman et al. (1989). The two open circles are sources 29 and 32; as discussed in the text, the corresponding Grabelsky et al. sources are GMC cloud complexes rather than single GMCs. The lines join them to solid squares, assuming that the free–free sources reside in the largest GMC in the complex, and assuming that cloud has a mass that ranges between the total mass and half the total mass in the complex. The stellar masses are found from WMAP free–free fluxes as described in the text. The dashed line represents the luminosity limit Q>1051  s−1 (Equation (9); see the text). We find 〈epsilonGMC〉 = 0.09, the Q-weighted average is 〈epsilonGMCQ = 0.08, and the GMC mass-weighted average is $\langle \epsilon _{\rm GMC }\rangle _{M_{\rm GMC}}=0.03$.

Standard image High-resolution image

The sample average 〈epsilonGMC〉 = 0.08 for our 32 SFCs. We stress again that epsilonGMC as defined here is a lower limit to the total star formation efficiency of the host GMC. Any stars in clusters older than 〈τms〉 will not be detected by WMAP and hence are not included in our estimate of M*.

This average is a bit higher than that usually quoted (typical estimates are more like 0.02), but we have selected the most active SFCs in the Galaxy, rather than a Galaxy wide average.

The Q-weighted efficiency is defined as 〈epsilonGMCQ ≡ ΣiQiepsilonG,i/(ΣiQi); we find 〈epsilonGMCQ = 0.07. Finally, if we weight by the mass of the host GMC, we find $\langle \epsilon _{\rm GMC }\rangle _{M_{\rm GMC}}\equiv \Sigma _i M_{\rm GMC,i} \epsilon _{G,i}/(\Sigma _i M_{\rm GMC,i})=0.03$.

3.0.1. Is the Apparent Trend of epsilonGMC with MGMC + M*MGMC Real?

There is an apparent trend of epsilonGMC with GMC mass in Figure 2; low-mass GMCs have higher star formation efficiencies than do high-mass GMCs. Is this trend a physical effect, or is it due to the way the observations are carried out?

We argue that the trend is largely explained by two effects, the first an observational selection effect, namely our luminosity limit. The second effect is due to the fact that the quantity on the y-axis (essentially M*/MGMC) involves the quantity on the x-axis (MGMC). If there is some scatter in MGMC, e.g., driven by a scatter in RGMC, it will generate a trend of epsilonGMC with MGMC.

The first effect arises because we have selected our sources on the basis of ionizing flux; essentially, we require Q>1051  s−1. It is expected on general grounds (and we find it to be true of our sample) that the total stellar mass in a GMC tends to increase with GMC mass. We have selected against low-mass clusters, and since low-mass GMCs tend to have low mass clusters, we will not find SFCs in low-mass GMCs with low or even moderate epsilonGMC.

To be more quantitative, M* = 1.6 × 104(Q/1051  s−1) M, from Equation (3). We can only find clusters more massive than this, or

Equation (9)

This is shown as the dashed line in Figure 2.

There are four systems with clusters smaller below the line, i.e., with M* < 104M: numbers 15, with (MGMC, epsilonGMC) = (4.3 × 103M, 0.273), 20 (1.7 × 106M, 0.002), 18 (3.2 × 105M, 0.02), and 40 (2 × 105M, 0.04), in order of increasing stellar mass. Numbers 15 and 18 are in the highly confused WMAP SFC 30; they are in this sample not because they are luminous sources, but because they reside near luminous sources on the sky. Source 40 is also in a highly confused WMAP region, G337, which contains five sources, so it has entered the sample for the same reason that numbers 15 and 18 did.

We can check that the apparent lower limit of epsilonGMC with MGMC is due to this selection effect. Mooney & Solomon (1988) compiled far-infrared (FIR) luminosities and masses for a sample of GMCs. They point out that FIR luminosity is a proxy for bolometric luminosity, and for the star formation rate, particularly in SFCs. Since Q (or free–free luminosity) is a proxy for bolometric luminosity, we can combine the two samples.

To calibrate the two measures of luminosity, we note that there are four regions in Mooney & Solomon (1988) for which Rahman & Murray (2010) give Q, M17A, W49, W43 (source 17 in the current paper), and W51. Mooney & Solomon (1988) calculate LFIR/MGMC, whereas we can calculate Lbol/MGMC, using the conversion from Q to Lbol in Murray & Rahman (2010). Doing so, we find that the ratio of LFIR/MGMC to Lbol/MGMC is 2.5, 2.3, 5.6, and 6 for the four common sources.

We plot the light-to-mass ratios for both samples in Figure 3. For the four common sources we use the respective ratios of Lbol/LFIR to convert the FIR fluxes; for the remainder of the Mooney & Solomon (1988) sources we use the average ratio Lbol/LFIR = 4 to make the conversion. We note that this rather large ratio is consistent with the fact that the bulk of the ionizing photons make it out of their host GMCs.

Figure 3.

Figure 3. Ratio Lbol/MGMC of the mass for GMCs in Mooney & Solomon (1988) (open squares) and in the present paper (filled triangles). The dashed line again represents the luminosity limit Q>1051  s−1, using the conversion Lbol/LFIR = 4. It can be seen that the trend of epsilonGMC with MGMC seen in Figure 2 is to a large extent driven by this luminosity selection. The remaining component of that trend, the upper envelope, is consistent with errors in the measurement of RGMC, particularly in confused regions (see the text).

Standard image High-resolution image

From this plot we can see that part of the apparent trend in epsilonGMC with MGMC in Figure 2, the lower boundary, is in fact due to selection.

The second effect arises from the fact that we are plotting M*/MGMC (ignoring the ≲30% contribution to the denominator from M*) versus MGMC, combined with the difficulty in measuring GMC radii in crowded fields, such as those examined by Solomon et al. (1987) and Heyer et al. (2009). We use the virial mass for the GMCs, which is proportional to RGMC, and the latter is subject to large uncertainties, or order 2 or so; this is apparent from a comparison of the date in these two papers compared to that in the papers from the Columbia group. This scatter (or bias) in RGMC leads to a scatter in MGMC, and in M*/MGMC, which spreads the data points out along a line with slope −1 in the plot; this is of course the slope of the dashed line in Figure 2. Note that the objects in crowded fields tend to have both lower masses and higher values of epsilonGMC, as would be expected if RGMC were systematically underestimated.

Finally there is an effect similar to cosmic variance, call it Galactic variance, that will result in a decrease in the upper envelope of epsilonGMC as MGMC increases. The upper envelope of the points in Figure 2 is roughly flat for MGMC < 106M, but drops abruptly for larger masses. In section 4 we estimate that GMCs live ∼27 Myr, while the free–free emission from the star clusters we see is on for only 4 Myr, about 1/7th as long. If the star formation rate is variable for any reason, it follows that a single cloud is unlikely to be found near its maximal value of epsilonGMC. In a large collection of clouds, the upper envelope of the distribution of epsilonGMC will be well sampled; in a smaller collection, the distribution will not be so well sampled, and the upper limit is likely to be below the maximal value of epsilonGMC reached by any individual cloud.

The most massive GMC in the Milky Way has a mass ∼6 × 106M; there are roughly 200 GMCs with masses above 106M. If the star formation rate varies rapidly, we would expect to see only (1/7) × 200 ≈ 30 GMCs near their maximum luminosity; we find 12 in the sample reported on here (we note that there are a significant number of luminous free–free emitting cluster residing in WMAP sources somewhat less luminous than those in the current sample, so at least some of the expected 20 remaining clusters may be in those sources). Thus, the distribution is reasonably well sampled if we include clouds down to 106M. However, as we move to larger GMC masses, the distribution will become less and less well sampled, so we should expect, and we do find, that the maximum value of epsilonGMC decreases with increasing MGMC.

This suggests that the apparent trend of decreasing epsilonGMC with increasing MGMC is due largely to selection and sampling effects.

3.1. The Star Formation Rate per Free-fall Time

Since we know both the mass and radius of the host GMC, we can calculate the GMC free-fall time $\tau _{\rm ff }\equiv \sqrt{3\pi /(32 G\bar{\rho })}$, where $\bar{\rho }\equiv 3M_{\rm GMC}/(4\pi R_{\rm GMC}^3)$. We can then find the star formation efficiency per free-fall time. However, we cannot use Equation (8) directly, since the star formation in GMCs is very likely not steady state: the super-virial expansion rate of the bubble walls, combined with the fact that the radiation force often exceeds the self-gravity of the GMC (see Column 9 in Table 2) indicates that the host GMCs will shortly be disrupted.

Table 2. Star-forming Complex Efficiencies

Catalog Lν Q M* MGMC RGMC τff epsilonGMC epsilonff Frad/FGrav
Number ( erg  s−1 Hz−1) ( s−1) (dust adjusted) (M) (M) (pc) ( Myr)      
1 7.27e+24 1.32e+51 2.10e+04 4.77e+05 102.3 24.71 0.042 0.267 5.79
2 4.13e+25 7.53e+51 1.19e+05 2.09e+06 57.6 5.00 0.054 0.069 0.55
4 9.64e+24 1.76e+51 2.79e+04 8.56e+05 166.0 38.14 0.032 0.308 6.28
5 1.72e+25 3.14e+51 4.99e+04 4.12e+06 82.6 6.11 0.012 0.019 0.12
6 4.86e+25 8.86e+51 1.41e+05 8.11e+05 23.0 2.03 0.148 0.077 0.68
7 5.53e+24 1.01e+51 1.60e+04 4.87e+04 9.5 2.19 0.247 0.139 3.66
8 6.96e+24 1.27e+51 2.01e+04 1.67e+05 23.1 4.50 0.107 0.124 2.32
9 4.60e+24 8.38e+50 1.33e+04 1.30e+05 24.3 5.49 0.093 0.131 2.79
10 7.94e+24 1.45e+51 2.30e+04 1.61e+05 15.8 2.58 0.125 0.082 1.32
11 4.97e+24 9.06e+50 1.44e+04 4.35e+05 19.0 2.08 0.032 0.017 0.17
12 1.02e+25 1.85e+51 2.94e+04 3.11e+05 44.2 8.70 0.086 0.193 3.57
13 3.98e+24 7.26e+50 1.15e+04 6.30e+04 16.3 4.33 0.155 0.172 4.62
14 5.36e+25 9.77e+51 1.55e+05 1.69e+06 35.7 2.71 0.084 0.058 0.41
15 5.56e+23 1.01e+50 1.61e+03 4.29e+03 6.4 4.10 0.273 0.287 21.61
16 1.10e+25 2.00e+51 3.17e+04 5.38e+05 44.7 6.72 0.056 0.096 1.31
17 3.56e+25 6.49e+51 1.03e+05 2.43e+06 46.2 3.32 0.041 0.035 0.22
18 2.32e+24 4.24e+50 6.72e+03 3.22e+05 38.6 6.99 0.020 0.037 0.58
20 1.39e+24 2.54e+50 4.03e+03 1.73e+06 27.0 1.76 0.002 0.001 0.01
22 1.71e+25 3.11e+51 4.94e+04 1.65e+05 22.3 4.27 0.230 0.252 5.39
24 8.56e+24 1.56e+51 2.48e+04 3.33e+06 37.6 2.08 0.007 0.004 0.02
25 8.87e+24 1.62e+51 2.57e+04 1.55e+05 11.2 1.57 0.142 0.057 0.80
26 1.62e+25 2.95e+51 4.68e+04 1.12e+06 55.2 6.41 0.040 0.066 0.68
28 1.51e+25 2.76e+51 4.38e+04 1.20e+06 62.0 7.35 0.035 0.066 0.70
29 3.51e+25 6.40e+51 1.02e+05 7.22e+06 133.4 9.47 0.014 0.034 0.21
31 3.85e+25 7.02e+51 1.11e+05 1.71e+06 207.9 37.80 0.061 0.592 9.84
32 2.51e+25 4.58e+51 7.27e+04 1.12e+07 178.5 11.77 0.006 0.019 0.11
33 1.59e+25 2.91e+51 4.61e+04 2.61e+05 56.9 13.87 0.150 0.533 13.12
34 1.64e+25 2.99e+51 4.75e+04 5.51e+05 57.7 9.74 0.079 0.198 3.12
35 2.15e+25 3.92e+51 6.22e+04 1.03e+06 117.4 20.70 0.057 0.303 4.86
37 5.17e+25 9.42e+51 1.50e+05 3.53e+06 160.5 17.86 0.041 0.186 1.85
38 1.10e+25 2.00e+51 3.17e+04 9.61e+05 41.8 4.55 0.032 0.037 0.36
40 2.95e+24 5.38e+50 8.54e+03 2.04e+05 39.8 9.18 0.040 0.095 1.96

Notes. Efficiencies of Milky Way star-forming complexes and their host GMCs. Column 1 gives the Rahman & Murray (2010) catalog number of the star-forming complex. The rest of the columns give star formation complex properties; free–free luminosity Lu (Column 2), ionizing photons emitted per second Q (Column 3), stellar mass (Column 4), GMC mass adjusted from the original as described in the text (Column 5), GMC radius adjusted to R0 = 8.5  kpc, and using the distance from Column 6 of Table 1 (Column 6), τff (Column 7), GMC efficiency (Column 8), epsilonff (Column 9), and radiation force divided by the force of gravity (Column 10).

Download table as:  ASCIITypeset image

In fact, most of the star formation takes place on a timescale shorter than 〈tms〉. We are clearly detecting many main-sequence 40–50 M stars (via their ionizing flux), but the clusters are no longer embedded in their natal gas clumps; instead, they have blown rather large bubbles in the surrounding ISM.

Rather than using Equation (8) directly, we combine Equations (7) and (8):

Equation (10)

Note that τff = 1.65 × 104R3/2pc/M1/26 yr.

The results using this estimate are given in Figure 4, which shows that epsilonff ranges from ≈10−3 to 0.59, with a mean 〈epsilonff〉 = 0.14.

Figure 4.

Figure 4. Star formation rate per free-fall time, epsilonff ≡ [M*/ (MGMC + M*)](τff/〈τms〉) = epsilonGMCff/〈τms〉) of free–free selected GMCs of mass MGMC. Symbols as in Figure 1. The sample average 〈epsilonff〉 = 0.18, the Q-weighted 〈epsilonffQ = 0.17, and the GMC-mass weighted average, $\langle \epsilon _{\rm ff }\rangle _{M_{\rm GMC}}=0.08$. These should be compared to the Milky Way average value of epsilonff = 0.013, shown by the solid line. The two dashed lines show the range around epsilonff ≈ 0.02 discussed by Krumholz & McKee (2005) and Krumholz & Tan (2007).

Standard image High-resolution image

The Q-weighted efficiency per free-fall time is

Equation (11)

similar to the sample average value. The GMC-weighted average is 〈epsilonffGMC = 0.08.

These estimates range from a factor 0.08/0.0057 = 14 to a factor of 26 times the value obtained by averaging the total star formation rate of the Milky Way over all the molecular gas inside the solar circle. We conclude that at least 30% of the star formation in GMCs occurs in very rapid bursts, 10 times higher than the Galaxy-averaged star formation rate per free-fall time.

Once again, these are lower limits for epsilonff, since not all the ionizing clusters are 3.9 Myr old. An arguably more realistic but still simple estimate is to take a mean cluster age of 〈tms〉/2 (this assumes that the clusters form in a time short compared to 3.9 Myr) leading to an estimate of 〈epsilonff〉 = 0.3.

A third estimate uses the smaller of the dynamical time τdyn = R/v and 〈tms〉. We argue that our dynamical times are likely to be overestimates, since the radial velocity spread, given in Table 4 of Rahman & Murray (2010), is a lower limit to the true bubble expansion velocity. We use the smaller of 〈τms〉 and the dynamical time

Equation (12)

Using this estimate we find

Equation (13)

for a Q-weighted average.

4. GMC LIFETIMES

The result that ∼30% of the star formation in the Milky Way occurs in 32 GMCs is truly remarkable. This becomes apparent when we compare the total molecular gas mass in the Milky Way, Mtot ≈ 109M (Dame 1993), to the mass in the 32 star-forming GMCs, with total gas mass M ≈ 5.9 × 107M; 30% of the star formation takes place in clouds that contain only 6% of the molecular gas mass. For context, the number of GMCs with masses above 105M ≈ 1100, while the number with masses above 106M ≈ 250. Thus, the bulk of massive GMCs in the Milky Way do not house WMAP sources.

All the SFCs we examine contain expanding bubbles, with typical expansion velocities of ∼10 km  s−1 and radii Rb ranging from 3  pc to 100  pc (Rahman & Murray 2010). In the more vigorously star-forming GMCs, the outward force exerted by radiation from the stars exceeds the inward force of gravity acting on the GMC; we use

Equation (14)

for the force due to radiation, where ξ = 8 × 10−11 erg  s−1 ·  s, appropriate for our choice of IMF. The outward force due to the pressure of the ionized gas is estimated as Fgas = 4πR2bP, where P is the gas pressure. We use the expression P = nkT, where k is Boltzmann's constant, the temperature of the ionized gas is T = 7000 K, and

Equation (15)

where αrec is the recombination coefficient.

For the force of gravity we estimate

Equation (16)

The ratio Frad/Fgrav is given in Column 9 of Table 2, and the ratio (Frad + Fgas)/Fgrav is plotted in Figure 5. More than half of the unconfused sources have a total outward force larger than the force of self-gravity. This is consistent with our interpretation of the radial velocity spreads as being due to expansion of the bubble walls, and strongly suggests that the star clusters are disrupting their host GMCs. We note that in almost all our sources, Frad>Fgas, as expected for such massive clusters (Murray et al. 2010).

Figure 5.

Figure 5. Ratio of outward (radiation and gas pressure) forces to the self-gravity force of the GMCs, (Frad + Fgas)/Fgrav, as a function of the host GMC mass. Twelve of the eighteen unconfused clusters (shown as filled polygons) have ratios larger than unity, indicating that the star clusters should be blowing expanding bubbles, consistent with the 8 micron bubble morphology seen in Spitzer or MSX images of the star-forming complexes. In almost all cases the radiation pressure is larger than the gas pressure, with the average ratio Frad/Fgas ∼ 2.

Standard image High-resolution image

The largest bubbles appear likely to disrupt the host GMC; the bubble walls have enough momentum to sweep up the rest of the gas in the host cloud, driving it to r ∼ 200 pc, at which point tidal shear will complete the disruption. We see these clouds in their death throes.

We can estimate the lifetimes of massive (MGMC ≳ 106M) Milky Way GMCs against disruption by the effects of the star clusters that form in them. To do so we assume that the massive GMCs harboring the massive clusters we examine here are drawn from the massive GMC population as a whole. This assumption implies that all massive GMCs will eventually form massive star clusters, but leaves open the possibility that less massive GMCs (say with masses below ∼105M) do not necessarily form many stars. For example, cloud fragments from the objects we have found, which may well have masses as large as 105M, may not be self-gravitating. If they are not, they may avoid any substantial star formation until the next time they find themselves inside a (larger) gravitationally bound object.

To proceed with our estimate of massive GMC lifetimes, we need to estimate the number of GMCs in the parent population of our star-forming objects. We do so in two different ways. First, we estimate the number of GMCs in the Milky Way that are required to produce a fraction f* ≈ 0.31 of the observed total star formation rate, $\hbox{\it dM}_*/dt=1.3\,M_\odot \,\,{\rm yr }^{-1}$. This requires an estimate of the GMC distribution function, which is known to follow a power-law dNGMC/dmm−α truncated at an upper mass MU (see the Appendix). The most massive GMC known in the Milky Way is object number 24 in Grabelsky et al. (1988, which does not correspond to any of the bright WMAP sources considered in this paper). We follow McKee & Williams (1997) and adopt MU = 6 × 106M, but we also report the result of using MU = 1.1 × 107M, which we take as a very firm upper limit. The power-law index α ≈ 1.5–1.6 (Casoli et al. 1984; Dame et al. 1986; Digel et al. 1996; Rathborne et al. 2009). We truncate the power law at the lower end as well, denoting the lower limit to the GMC mass distribution by ML; our results depend only weakly on ML.

The second method of estimating GMC lifetimes uses the mass function of GMCs integrated from MU down to a critical GMC mass Mcrit. This critical GMC mass can be defined in multiple ways. We define it as that mass which results in a mean GMC mass equal to the Q-weighted mean GMC mass in our sample.

We proceed to the first estimate. To find the star formation rate in clouds with masses greater than some mass MGMC, we use the Galaxy-wide average star formation rate per free-fall time, and integrate from MU down toward the least massive clouds, with mass ML. Let f* be the fraction of the Galactic star formation rate that occurs in clouds with mass Mf* or larger. We show in the Appendix that the critical mass (denoted by $M_{f_*}$ here since it is defined by the fraction of the total star formation rate produced by clouds at or above that mass) is

Equation (17)

where δ = (7/4) − α. Figure 6 shows the cumulative fraction of the star formation rate produced in GMCs with masses greater than M, for two values of MU, our best estimate 6 × 106M (Williams & McKee 1997), and a firm upper limit MU = 1.1 × 107M, and two values of α (1.5 and 1.6). Taking the two extreme cases, the value of $M_{f_*=0.31}=1.05\times 10^6\,M_\odot$ (MU = 6 × 106M, α = 1.6) and $M_{f_*=0.31}=2.9\times 10^6\,M_\odot$ (MU = 1.1 × 107M, α = 1.5).

Figure 6.

Figure 6. Cumulative fraction f*(>M) of the Galactic star formation rate produced in GMCs with MGMC>M. We show four cases, with MU and α as follows: MU = 1.1 × 107M and α = 1.5 (solid line), MU = 1.1 × 107M and α = 1.6 (dashed line), MU = 6 × 106M, and α = 1.5 (dotted line), and finally MU = 6 × 106M and α = 1.6 (dot-dash line). The vertical solid and dashed lines denote the GMC mass at which f* = 0.31 is reached for the first and last case; the corresponding GMC masses are 2.9 × 106M and 1.05 × 106M.

Standard image High-resolution image

Figure 7 shows the cumulative number N(>M) of GMCs with masses MGMC>M. The number of Milky Way GMCs with masses larger than M(f* = 0.31) = 1.05 × 106M is 87 (left panel, dashed vertical line, and for M(f* = 0.31) = 2.9 × 106M there are 210 GMCs (right panel, dashed vertical line). We use the relation

Equation (18)

where we assume that the average lifetime of the star clusters in our sample is 〈τms〉 = 3.9  Myr. Note that the luminosity of a cluster decays much more slowly after 3.9  Myr; the luminosity decays to half the luminosity at 〈τms〉 only when the cluster age is ≈8  Myr (about a dynamical time for the massive GMCs we consider here). Thus, the radiation pressure from the cluster acts for about 2〈τms〉 (while the H ii gas pressure essentially vanishes after 〈τms〉).

Figure 7.

Figure 7. Cumulative number of GMCs in the Milky Way, plotted as a function of GMC mass MGMC. The left panel shows N(>M) for MU = 1.1 × 107M and α = 1.5, while the right panel assumes MU = 6 × 106M and α = 1.6. These two pairs bracket the observationally supported parameters. The total number of clouds is N ≈ 10,000. The vertical dashed line is at MGMC = 2.9 × 106M (left) and MGMC = 1.05 × 106M (right), corresponding to the mass of GMCs that produce 31% of the star formation in the Milky Way; there are ∼87 (left) or 210 (right) such clouds. The vertical solid lines at MGMC = 4.9 × 105M (left) and MGMC = 9.5 × 105M (right) correspond to the critical mass that yields a mean GMC mass equal to the Q-weighted average GMC mass in our sample; there are 343 (left) or 231 (right) GMCs in the Milky Way having masses this larger or larger.

Standard image High-resolution image

We find that the lifetime of the parent GMCs is ∼11  Myr and ∼26  Myr for the two extreme cases.

The Q-weighted average free-fall time of these GMCs is ≈9  Myr, and 10  Myr for unconfused GMCs, so our first estimate is that massive GMCs live 1.1–2.9 free-fall times before they are disrupted by expanding bubbles produced by the star clusters they contain.

Now for the second estimate for the lifetimes of massive GMCs, which uses the Q-weighted mean GMC mass for the parent GMCs in our ionizing luminosity-selected sample of SFCs. We use the GMC mass function, $\hbox{\it dN}_{\rm GMC}/dm$, to calculate the mean GMC mass as a function of the lower mass cutoff (denoted Mcrit). We then adjust Mcrit until the mean GMC mass is equal to the Q-weighted mean GMC mass in our sample. Having found the appropriate Mcrit, we then calculate the number of GMCs having masses above Mcrit, and compare that to the number of GMCs in our sample.

The Q-weighted mean GMC mass of our sample is 2.3 × 106M. The Mcrit that gives the same mean mass, assuming MU = 6 × 106M and α = 1.6, is 9.5 × 105M, corresponding to N(>9.5 × 105M) = 231; see Figure 7. The estimated GMC lifetime is ∼28  Myr, or ∼3τff.

If we use MU = 1.1 × 107M and α = 1.5, we find N(>4.9 × 105M) = 343, a lifetime of 42  Myr, or ∼4.2τff. These lifetimes are somewhat larger than our first estimate; averaging we take the lifetime of massive GMCs in the Milky Way to be 27 ± 12  Myr, or ∼2.7 ± 1τff.

The ionizing luminosity of some of our SFCs is rather low, while our original WMAP sample was chosen for its high luminosity. The confusion of the WMAP sources results in the inclusion of low luminosity SFCs, which in turn leads to our overcounting the number of GMCs in a true luminosity-selected sample. We can try to correct for this by making a cut in Q, at a value we will denote Qcut. When we do so, we find that the average GMC mass increases as a function of Qcut. So while it is true that increasing Qcut decreases the number of star-forming GMCs in our sample, which tends to increase the estimated lifetimes of the GMCs, at the same time 〈MGMC〉(Qcut) also increases, which tends to decrease the estimated lifetimes of the GMCs. The net result is that the ratio NGMC/Ncut increases slowly, from 4.6 to 6.1, and the estimated lifetime increases by 30% by the time Qcut = 5 × 1051  s−1; at this value of Qcut, there are only seven sources left in the sample, and the mean mass is no longer well defined, i.e., the mean GMC mass begins to drop with increasing Qcut beyond this value.

5. DISCUSSION

A number of authors, e.g., Mooney & Solomon (1988), Mead et al. (1990), and Evans (1991), have examined the ratio LFIR/LCO (a proxy for epsilonGMC) in Milky Way GMCs. They found that this ratio varies by nearly three orders of magnitude; this is consistent with the findings in this paper, and is illustrated in Figure 3.

Similar result have been found recently by Schruba et al. (2010), who examine the ratio of Hα to CO luminosity in nearby galaxies. They find that selecting for H ii regions leads to very high ratios on small scales, whereas selecting for CO peaks leads to very low ratios on small scales. Again, this suggests large variations in epsilonGMC on small scales (although their smallest scales tend to be larger than individual GMCs in the relevant host galaxies).

Along the same lines, Lada et al. (2010) examine nearby SFCs in the Milky Way and argue that the star formation rate per free-fall time varies strongly from cloud to cloud.

The massive clusters in the sample examined here are capable of ionizing a substantial fraction of the mass in their host GMCs. For a cluster emitting Q ionizing photons per second, all of which are absorbed by hydrogen in the host GMC, we have

Equation (19)

where αrec ≈ 3.57 × 10−13 cm3 s−1 is the recombination coefficient.

This expression assumes that all the ionizing photons emitted by the stars in a GMC are absorbed in that GMC. However, the WMAP results show that a substantial fraction, and in many cases a majority, of the ionizing photons escape the host GMC. For example, source 29, at l = 298°, has a radius in excess of a degree (about twice the WMAP beam). This corresponds to R ≈ 200 pc, compared to RGMC = 137 pc; the bulk of the free–free emission comes from outside the host GMC in this object. We estimate that on average, half or more of the ionizing photons escape from the host GMC. In addition, it is estimated that about 35% of the ionizing photons emitted by hot stars are absorbed by dust (McKee & Williams 1997).

Finally, if all the ionizing photons are absorbed before they reach RGMC, the ionized mass will again be smaller than the estimate in Equation (19).

If we account only for the presence of dust, we find that, on average, ∼50% of the GMC gas could be ionized. This is consistent with the fact that the CO-inferred masses correlate well with the virial masses in the GMC catalogs we use; the average ratio of CO-inferred mass to virial mass is also 0.5. (This paper uses the virial GMC masses throughout.) However, it is likely that much of the non-molecular gas contributing to the viral mass is atomic rather than ionized. Indeed, if the sizes of the WMAP sources are any indication, only a small fraction of the non-molecular gas can be ionized; most of the photons leak out of the host GMC, and so cannot ionize as much gas inside RGMC.

We note that we are using the currently measured virial mass, along with the current, less than 4  Myr old (ionizing) stellar mass. The mass of the cloud cannot change on a timescale much less than the dynamical time (the bubble wall expansion velocity and the ionized gas sound speed are both comparable to the reported GMC linewidths), either by feedback driven loss of mass or by accretion. Since the star clusters we find are less than half a GMC free-fall time old, their masses reflect the current GMC mass rather than any previous (larger or smaller) mass associated with their host GMC. It follows that the high values of epsilonGMC and epsilonff are not the result of recent mass loss from the host GMCs.

Figure 5 shows that the sum of radiation pressure and gas pressure forces exceeds the self gravity in 19 of our 32 GMCs, indicating that these GMCs should be in the process of being disrupted. The presence of expanding bubbles in these systems is consistent with this notion, and with the predictions of Murray et al. (2010). This strongly suggests that these GMCs are at the end of their lives, either as converging flows or as gravitationally bound objects.

Both the observations and the theory of Murray et al. (2010) suggest that the process of disruption takes about a free-fall time, since the expansion velocity of the bubbles is v ∼ 15  km  s−1, similar to the observed CO linewidths. The time for turbulence to decay (assuming the clouds are initially turbulence supported) is about one free-fall time, so the time for an unsupported GMC to collapse is ∼1–2τff. The total life time of an unsupported GMC, from formation through collapse and star formation to disruption, would then be 1.5–2.5τff, similar to the value we measure. In other words, the fact that massive GMCs in the Milky Way appear to live only about two free-fall times is consistent with the suggestion of Goldreich & Kwan (1974) that all GMCs are in the process of collapse; the short lifetimes of massive GMCs obviate the need for any means of support, including driven turbulence or magnetic support.

As noted in the introduction, the argument of Zuckerman & Evans (1974), that rapid collapse of GMCs would lead to a star formation rate in the Milky Way that was much higher than the observed rate, does not apply if the stars that are formed disrupt the host GMC. We have seen that the GMCs we examine which have epsilonGMC ∼ 0.08 appear to be in the process of disruption. This suggests that the star formation rate is much less than MGMCff because only a small fraction of MGMC is converted into stars before the cloud is disrupted by radiation pressure.

While our results are consistent with free-fall collapse of GMCs, they do not require such a collapse. It may be, for example, that the bulk of the material in GMCs is held up by magnetic fields, but that over one dynamical time (∼10 Myr), some 10%–15% of the GMC accretes onto one or two massive clumps. Roughly half of this material will form stars, with the rest dispersed back in to the ISM, resulting in the observed 〈epsilonGMCQ ∼ 0.08. It may also be that the GMCs are not gravitationally bound to begin with; measurements of the virial parameters of massive clouds are near unity.

The last statement is consistent with the short lifetimes we have inferred for massive GMCs—dissipating sufficient energy to make a strongly bound GMC would require more than a single free-fall time. One could also imagine that some supply of energy could maintain a cloud with a virial parameter near unity. However, since the clouds only live for one or two free-fall times, it is difficult to envision a steady state arising, in which decay of turbulent energy is balanced by some source of energy.

Adding up the star formation from all the GMCs, ∑epsilonGMCMGMC/(2τff) (the factor two comes from our estimate of the GMC lifetime) the star formation rate is only a factor of ∼2 larger than the observed Milky Way star formation rate. This implies that the fragments of the disrupted GMCs must reassemble in ∼2τff; if they did not, the star formation rate would be lower than observed. In addition, if the fragments of the disrupted 106M GMCs do not reassemble after a short time, more of the molecular gas would reside in smaller clouds, contrary to observations. Both these observations strongly suggest that the cloud fragments reassemble in less than an orbital period.

We note that the lifetimes of the majority of massive GMCs cannot be much larger than a few τdyn, for two reasons: first, we would see massive GMCs between spiral arms, something for which there is little observational evidence. Second, the star formation rate would be much less than that observed. In other words, the majority of the most massive GMCs in the Milky Way cannot be supported against collapse by any mechanism that suppresses star formation for more than two or three GMC free-fall times.

5.1. Dynamical Star Formation

Hartmann et al. (2001) and Ballesteros-Paredes & Hartmann (2007) argue that GMCs and stars form rapidly, in a dynamical time or less, based on observations of SFCs within 1  kpc of the Sun. The observational argument is simple: most of the GMCs in the solar vicinity have substantial star formation, so that the age of the star clusters and that of the GMCs must be similar. Note that their "substantial star formation" is not detectable by WMAP, since there is little ionizing radiation associated with most of their sources. Since the average age of their star clusters is of order 1–3  Myr, they conclude that the typical GMC lifetime is also ∼1–3  Myr.

The different (short) lifetimes for the host GMCs found by these authors, compared to the lifetimes found here, arise from several causes. First is the observational fact that they find a ratio of total to star-containing GMCs of 1.5, compared to our ratio of free–free dim to free–free luminous objects of ∼4, for a total to luminous ratio of ∼5. By itself, this difference accounts for a ratio of more than three in the estimated GMC lifetime.

It may well be the case that most of the Milky Way clouds with MGMC>1.5 × 106M (the parent population of our sample) host clusters with a total Q that lies below our selection criteria. If we counted such clouds then we might find a ratio similar to that found by Ballesteros-Paredes & Hartmann (2007). However, it would not be correct to conclude that the GMC lifetime was then 1.5 times 〈τms〉, unless all the GMCs were in the process of being disrupted (by some mechanism not related to radiation pressure or H ii gas pressure, since either would not be adequate in such dim sources).

A second source for the difference in estimated GMC lifetime arises because Ballesteros-Paredes & Hartmann (2007) employ an average age for their stars of ∼2  Myr, with an upper limit of 5  Myr, while we use a total lifetime of 〈τms〉 = 3.9  Myr. We believe that the correct procedure is to use the total (embedded) lifetime of the stellar tracers, rather than the average age. We add the qualifier embedded since the host GMC may be disrupted before the life of the stellar tracers employed is reached. The total embedded lifetime is in general longer than the average embedded lifetime, so using the latter will underestimate the GMC lifetime. Using their upper limit of 5  Myr (given in their Section 3) as the total embedded lifetime would result in an increase in their estimated GMC lifetime by a factor of 2.5, to about 5  Myr. If we then multiply by the ratio of total to star-containing GMCs (1.5), the GMC lifetime in their sample is ∼7.5  Myr, somewhat larger than a third of the ∼19  Myr lifetime found in our sample.

There is also a physical effect that may contribute to the difference in estimated GMC lifetimes found by Ballesteros-Paredes & Hartmann (2007) and those found here. Table 1 of Ballesteros-Paredes & Hartmann (2007) lists 21 small GMCs within 1  kpc of the Sun, containing a total of 4.2 × 106M of gas, of which 14 show some signs of star formation. The average GMC mass in their sample is 2 × 105M.

The GMCs considered here are much more massive (average MGMC = 15 × 105 versus 2 × 105M) in our sample, much larger, and more diffuse, so the GMCs in our sample have longer free-fall times than the GMCs in Ballesteros-Paredes & Hartmann (2007). The more massive clouds will naturally take longer to form and collapse than less massive clouds. If the less massive clouds are destroyed by the same mechanism (a combination of radiation and gas pressure), then the larger clouds will live a factor M1/4 ∼ 1.6 longer, or ∼12  Myr, (although they may survive for the same number of free-fall times).

We conclude that the GMC lifetimes, measured in terms of free-fall times, found by Ballesteros-Paredes & Hartmann (2007) are consistent with those we find, given the different GMC masses considered and the uncertainties involved.

5.2. Turbulence and Star Formation

The high fraction of gas turned into stars in a free-fall time in star-forming Milky Way GMCs (∼16%–25%) is surprising in light of recent theories invoking turbulence to regulate the rate of star formation, e.g., Padoan (1995); Krumholz & McKee (2005). In these theories, the rate of star formation is related to the amount of gas above a critical density, set by the requirement that the local Jeans length be less than the sonic length (the length at which the velocity of turbulent motions equals the sound speed of the gas). The critical density depends on the Mach number $ {\cal M}\equiv \sigma /c_s$ and the virial parameter

Equation (20)

where σ is the one-dimensional line-of-sight velocity dispersion.

We have calculated both $ {\cal M}$ and αvir for the host GMCs in our sample, and in the parent GMC population, finding values clustering about $ {\cal M}\approx 10\hbox{--}15$ and αvir ≈ 0.3–2, with the lower values for αvir seen in the more massive GMCs, those with MGMC>106M.2 The average virial parameter for the Heyer et al. sample objects with MGMC>106M is α = 0.6 ± 0.3; the subsample with high free–free luminosities has an average α = 0.9. While not highly significant, this is consistent with the notion that the host GMCs in the latter objects are being disrupted by their daughter star clusters.

In calculating $ {\cal M}$ we adopted T = 15 K, and a mean molecular weight of 3.9 × 10−24 g. We then used these values in Equation (30) of Krumholz & McKee (2005) to find the predicted value of epsilonff ≈ 0.02. This is roughly a factor of 10 lower than the average epsilonff in our luminosity selected sample.

Krumholz & Tan (2007) have estimated epsilonff in different environments in the Milky Way, finding values ranging from 0.013 for GMCs and other diffuse objects, to 0.2 (their CS(5-4) point). As noted above, we find a similar value for the Milky Way average GMC but a much higher value for actively star-forming GMCs, which we infer to be at the end of their lives. One natural interpretation is that the star formation rate per free-fall time varies with time in a given GMC, increasing rapidly just before the cloud is disrupted. Since there is no indication that the Mach number or virial parameters of our rapidly star-forming GMCs differ dramatically from the sample mean values (and the Mach number dependence of epsilonff in the turbulence picture depends only weakly on $ {\cal M}$ in any case), this suggests that turbulence is not, by itself, controlling the rate of star formation in Milky Way GMCs.

This point is worth stressing. Krumholz & McKee (2005) estimate the upper and lower limits to the turbulence limited star formation rate, finding values a factor 10 below their best estimate to 6 times above. They state that a more realistic estimate is a factor of three. We have measured epsilonff to be a factor of 8–12 larger than their best estimate in more than a dozen GMCs in the Milky Way. Furthermore, these same GMCs seen to be in the process of disruption, so that they are at the end of their lifetimes, suggesting that every massive GMC will experience a similar burst of rapid star formation. Since none of the massive GMCs in the catalogs we have considered have very low virial parameters (or very high Mach numbers), there is no indication that these objects have extreme values of either αvir or $ {\cal M}$; we conclude that supersonic turbulence is not the dominant determinant of the star formation rate in Milky Way GMCs.

Rapid star formation rates (per free-fall time) have implications for star formation in external galaxies. Murray et al. (2010) modeled star formation in clump galaxies, which contain SFCs with R ∼ 0.2–1  kpc and M ∼ 108–109M ("clumps"). Murray et al. (2010) found that the clumps should be disrupted by radiation pressure, with epsilonGMC ∼ 0.3. Since the free-fall time of the clump is several times longer than 〈τms〉, the implication was that epsilonff ≈ 0.3 or larger. Krumholz & Dekel (2010) state that this is "an extraordinarily high value of epsilonff." We have seen that rate of star formation is achieved by a number of GMCs in the Milky Way, which also have free-fall times several times longer than 〈τms〉. Since both the Milky Way and the clump forming galaxies appear to have marginally stable disks, and to have star formation rates consistent with the Kennicutt value (although in the latter case the evidence is rather slim), high values of epsilonff might be expected in external galaxies. A concomitant result is that the star formation efficiency of clump galaxies may well saturate at epsilonGMC ∼ 0.3, rather than proceeding to epsilonGMC ≈ 1, as envisioned by Krumholz & Dekel (2010).

6. CONCLUSIONS

We have presented four major observational results in this paper. First, we have estimated a lower limit to the fraction of gas in a massive Milky Way GMC that will be converted into stars over the lifetime of that GMC, finding epsilonGMC ≈ 0.08. Second, we have measured the star formation rate per free-fall time, epsilonff ≈ 0.16, for our ionizing luminosity-selected sample of GMCs. Third, we have estimated the lifetime of massive (MGMC ≈ 106M Milky Way GMCs, finding that they live two to three free-fall times. Finally, we have shown that many if not most of these clouds are in the process of being disrupted by the radiation pressure of the star clusters they have given birth to.

The strong implication of these results is that the rate of star formation in Milky Way GMCs, and by extension, in the Galaxy as a whole, is not set by the properties of supersonic turbulence.

This research has made use of NASA's Astrophysics Data System. The author is supported in part by the Canada Research Chair program and by NSERC of Canada.

APPENDIX: THE GMC MASS DISTRIBUTION FUNCTION

The GMC mass distribution function is generally fit by a simple power law

Equation (A1)

with α ≈ 1.5 (Casoli et al. 1984; Dame et al. 1986; Digel et al. 1996; Rathborne et al. 2009). These papers show that Milky Way GMCs have masses ML < m < MU with ML ≈ 103M and MU = 6 × 106M; we adopt this value as our best estimate following Williams & McKee (1997). There are two objects in Table 2 that appear to have GMC masses above this limit, numbers 29 and 32 (Grabelsky et al. catalog numbers 26 and 35) but both are described by those authors as "cloud complexes." They list three other objects with masses in excess of 6 ×  106M, none of which they consider to be a single GMC. We take MU = 1.1 × 107M as a firm upper limit on the mass of the most massive GMC in the Galaxy.

The total mass in molecular gas is Mg = 109M (Dame 1993); we assume that it is all in GMCs.

It is useful to employ the natural logarithm of the mass as the independent variable

Equation (A2)

where β = α − 1. The quantity NU is approximately the number of clouds in the mass range from MU/2 to MU.

Using this notation, the total mass in GMCs is related to the upper and lower mass cutoff, and the power-law index, by

Equation (A3)

Equation (A4)

The observed quantities are MTot, MU, ML, and the power-law slope α, so we solve for NU in terms of these four:

Equation (A5)

where γ ≡ (2 − α). Using the numerical values given above, we find

Equation (A6)

The number of GMCs with mass larger than m is

Equation (A7)

Performing the integral we find

Equation (A8)

From Equation (A5),

Equation (A9)

The total number of clouds is $N_{\rm Tot}\approx \rm{13{,}000}$.

Next we calculate the star formation rate of the Galaxy under the assumption that

Equation (A10)

with a fixed epsilonff, i.e., the same value of epsilonff is used for all GMCs. The total star formation rate is

Equation (A11)

To evaluate τff we will use the fact that the surface density Σ0MGMC/(πR2GMC) of GMCs is constant, independent of MGMC, e.g., Solomon et al. (1987). Then τff(MGMC) = τff(MU)(MGMC/MU)1/4 and

Equation (A12)

Letting δ = (7/4) − α, the fraction of star formation that takes place in GMCs with masses greater than m is

Equation (A13)

It follows that the GMC mass $M_{f_*}$ at which the cumulative star formation rate reaches a fraction f* of the total star formation rate is

Equation (A14)

Footnotes

  • We note that the sample in Bronfman et al. (1989) has anomalously high values of αvir, apparently driven by the high linewidths reported there. Despite this, the values of epsilonGMC and epsilonff are similar to those found for the other populations we consider.

Please wait… references are loading.
10.1088/0004-637X/729/2/133