<<

Whitworth et al.: The Formation of Brown Dwarfs 459

The Formation of Brown Dwarfs: Theory

Anthony Whitworth Cardiff University

Matthew R. Bate University of Exeter

Åke Nordlund University of Copenhagen

Bo Reipurth University of Hawaii

Hans Zinnecker Astrophysikalisches Institut, Potsdam

We review five mechanisms for forming brown dwarfs: (1) turbulent fragmentation of molec- ular , producing very-low- prestellar cores by shock compression; (2) collapse and fragmentation of more massive prestellar cores; (3) disk fragmentation; (4) premature ejection of protostellar embryos from their natal cores; and (5) photoerosion of pre-existing cores over- run by HII regions. These mechanisms are not mutually exclusive. Their relative importance probably depends on environment, and should be judged by their ability to reproduce the IMF, the distribution and kinematics of newly formed brown dwarfs, the binary statis- tics of brown dwarfs, the ability of brown dwarfs to retain disks, and hence their ability to sustain and outflows. This will require more sophisticated numerical modeling than is presently possible, in particular more realistic initial conditions and more realistic treatments of transport, transport, and magnetic . We discuss the mini- mum mass for brown dwarfs, and how brown dwarfs should be distinguished from .

1. INTRODUCTION form a smooth continuum with those of low-mass H-burn- ing . Understanding how brown dwarfs form is there- The existence of brown dwarfs was first proposed on the- fore the key to understanding what determines the minimum oretical grounds by Kumar (1963) and Hayashi and Nakano mass for formation. In section 3 we review the basic (1963). However, more than three decades then passed physics of , as it applies to brown dwarfs, and before brown dwarfs were observed unambiguously (Rebolo derive key analytic results, in particular the et al., 1995; Nakajima et al., 1995; Oppenheimer et al., for -limited fragmentation in various different for- 1995). Brown dwarfs are now observed routinely, and it is mation scenarios. We have deliberately assembled most of therefore appropriate to ask how brown dwarfs form, and the mathematical analysis in this one section. In sections 4 in particular to ascertain (1) whether brown dwarfs form in to 8 we consider five different mechanisms that may be in- the same way as H-burning stars, and (2) whether there is volved in the formation of brown dwarfs. Section 4 explores a clear distinction between the mechanisms that produce the possibility that turbulent fragmentation of molecular brown dwarfs and those that produce planets. clouds produces prestellar cores of such low mass that they In section 2 we argue that the mechanisms forming brown inevitably collapse to form brown dwarfs. Section 5 consid- dwarfs are no different from those forming low-mass H- ers more massive prestellar cores, and the possibility that burning stars, on the grounds that the statistical properties they fragment dynamically as they collapse, thereby spawn- of brown dwarfs (mass function, clustering properties, ki- ing with a range of . The collapse of such nematics, binary statistics, accretion rates, etc.) appear to cores ceases when the gas starts to up (due to adiabatic

459 460 Protostars and Planets V compression) and/or when becomes important. Un- dwarfs, then brown dwarfs should not be distinguished from fortunately, the interplay of dynamics and adiabatic com- stars. Many stars fail to burn , and most fail to burn pression, while likely to play a critical role (e.g., Boss et , without forfeiting the right to be called stars. The al., 2000), is hard to analyze. On the other hand, rotational reason for categorizing brown dwarfs as stars is that the sta- effects can be analyzed systematically, and therefore sec- tistical properties of brown dwarfs appear to form a contin- tion 6 explores the formation of brown dwarfs by gravita- uum with those of low-mass H-burning stars (and not with tional instabilities in disks, considering first isolated relaxed those of high-mass planets). disks, then unrelaxed disks, and finally interacting disks in clusters. Section 7 reviews the process of competitive ac- 2.1. The , Clustering Statistics, cretion, which determines how an ensemble of protostellar and Dispersion embryos evolves to populate higher masses; and then con- siders the N-body processes that (1) may eject brown dwarfs The initial mass function (IMF) is apparently continu- and low-mass stars from their natal cores, thereby termi- ous across the H-burning limit at ~0.075 M , in the Trape- nating accretion and effectively capping their masses, and zium cluster (Slesnick et al., 2004), in the σ Orionis clus- (2) may influence the binary statistics and clustering prop- ter (Bejar et al., 2001), in (Luhman, 2004), in IC 348 erties of brown dwarfs. Section 8 considers the formation of (Luhman et al., 2003), in the (Moraux et al., 2003), brown dwarfs by photoerosion of pre-existing cores that are in the field stars of the disk (Chabrier, 2003), and even pos- overrun by HII regions. Section 9 presents numerical simula- sibly in the halo (Burgasser et al., 2003b). If the IMF is tions of the birth of a whole in a large protoclus- fitted by a power law across the H-burning limit, dN /dM ∝ ter core; in these simulations, the mechanisms of section 5 M–α, estimates of α fall in the range 0.4 to 0.8 (e.g., Moraux (collapse and fragmentation), section 6 (disk fragmentation), et al., 2003). The IMF appears to extend down to a few Jupi- and section 7 (premature ejection) all occur simultaneously ter masses (e.g., Zapatero Osorio et al., 2002; McCaugh- and in tandem, and the collective properties of brown dwarfs rean et al., 2002; Lucas et al., 2005). The continuity of the formed can be extracted for comparison with observation. IMF across the H-burning limit is not surprising, since the In section 10 we summarize the principal conclusions of this processes that determine the mass of a low-mass star are pre- review. sumed to occur at relatively low (ρ < 10–8 g cm–3) Comprehensive discussions of the observational proper- and (T < 2000 K), long before the material in- ties of brown dwarfs, and how they may constrain forma- volved knows whether it will reach sufficiently high tion mechanisms, are contained in the chapters by Luhman (ρ > 1 g cm–3) to be supported in perpetuity by electron de- et al. (which deals with the entire observational picture) and generacy pressure before or after it reaches sufficiently high Burgasser et al. (which deals specifically with the issue of (T > 107 K) to burn . multiplicity). Therefore we offer only a brief summary of In the , and in Taurus, brown dwarfs the observations in section 2. Likewise, a critique of simula- appear to be homogeneously mixed with H-burning stars tions of core fragmentation is given in the chapter by Good- (Lucas and Roche, 2000; Briceño et al., 2002; Luhman, win et al.; simulations of disk fragmentation are discussed 2004), and in I their kinematics are also essen- and compared in detail in the chapter by Durisen et al.; and tially indistinguishable (Joergens, 2006b). Although they the origin of the IMF is treated in the chapter by Bonnell have been searched for — as possible signatures of for- et al., so we have limited our consideration to those aspects mation by ejection — neither a greater velocity dispersion that pertain specifically to the origin of brown dwarfs. of brown dwarfs in very young clusters, nor a diaspora of brown dwarfs around older clusters, has been found. 2. EVIDENCE THAT BROWN DWARFS FORM LIKE LOW-MASS 2.2. Binary Statistics HYDROGEN-BURNING STARS Multiple systems involving brown dwarf secondaries can We shall assume that brown dwarfs form in the same be categorized based on whether the is a -like way as H-burning stars, i.e., on a dynamical timescale, by star or another brown dwarf. gravitational instability, and with initially uniform elemen- 1. Sun-like primaries. Sun-like stars seldom have tal composition (reflecting the composition of the interstel- brown dwarf companions. At close separations (<5 AU), the lar matter out of which they form). (By implication, we dis- frequency of companions with masses in the range 0.01 to tinguish brown dwarfs from planets, a term that we will here 0.1 M is ~0.5% (Marcy and Butler, 2000); this figure is reserve for objects that form on a much longer timescale, by known rather accurately due to the numerous Doppler sur- the amalgamation of a rocky core and — if circumstances veys aimed at detecting extrasolar planets. Since the fre- allow — the subsequent accretion of a gaseous envelope. quency of companions outside this mass range (both exo- This results in planets having an initially fractionated ele- planets below, and H-burning stars above) is much higher, the mental composition with an overall deficit of elements.) paucity of brown dwarf companions is termed the “brown If this is the correct way to view the formation of brown dwarf .” However, the chapter by Luhman et al. points Whitworth et al.: The Formation of Brown Dwarfs 461

out that the paucity of close brown dwarf companions to 2.3. Disks, Accretion, and Outflows Sun-like stars may simply reflect the overall paucity of brown dwarfs. At larger separations (>100 AU), only about Young brown dwarfs are observed to have ex- 20 brown dwarf companions to Sun-like stars have been cesses indicative of circumstellar disks, like young H-burn- found to date, indicating that these systems are also rare ing stars (Muench et al., 2001; Natta and Testi, 2001; Jaya- (frequency ~1 to 2%), although this estimate is based on wardhana et al., 2003; Mohanty et al., 2004). Brown-dwarf more limited statistics (e.g., Gizis et al., 2001; McCarthy disk lifetimes are estimated to be 3–10 m.y., again like H- and Zuckerman, 2004). burning stars. From their Hα emission-line profiles, there is 2. Brown-dwarf primaries. BD-BD binary systems evidence for ongoing magnetospheric accretion onto brown (and binary systems with very-low-mass H-burning prima- dwarfs (Scholz and Eislöffel, 2004), and the inferred accre- ries) have an observed multiplicity of ~10–20% for sepa- tion rates form a continuous distribution with those for H- rations greater than ~2 AU (Bouy et al., 2003; Burgasser burning stars, fitted approximately by M ~ 10–8 M yr –1 (M/ et al., 2003a; Close et al., 2003; Gizis et al., 2003), and all M )2 (Muzerolle et al., 2003, 2005). To sustain these esti- but 5 of the ~75 known BD-BD binaries have separations mated accretion rates, brown dwarfs only require rather low- less than ~20 AU (see Table 1 in the chapter by Burgasser mass disks ~10–4 M . Finally, the spectra of brown dwarfs et al.). Below ~2 AU, the multiplicity is unknown because also show forbidden emission lines suggestive of outflows most surveys to date have been imaging surveys that cannot like those from H-burning stars (Fernández and Comerón, resolve close systems. Only a few spectroscopic BD-BD 2001; Natta et al., 2004), and recently an outflow from a binaries have been discovered (Basri and Martin, 1999; brown dwarf has been resolved spatially (Whelan et al., Stassun et al., 2006; Kenyon et al., 2005). Some authors 2005). Thus, in the details of their circumstellar disks, ac- speculate that the overall multiplicity might be ~30–50%, cretion rates, and outflows, young brown dwarfs appear to based on the positions of brown dwarfs in color- mimic H-burning stars very closely, and to differ signifi- diagrams (Pinfield et al., 2003) and the statistics of radial cantly only in scale. velocity variations (Maxted and Jeffries, 2005). However, Joergens (2006a) has examined the radial of 10 2.4. Rotation and X-Rays BDs and very-low-mass H-burning stars, and finds no bina- ries with separations less than ~0.1 AU, and only two ob- The rotational properties of brown dwarfs also appear jects with variability on timescales greater than 100 days that to connect smoothly with those of very-low-mass H-burn- might indicate companions at greater than ~0.2 AU. Thus, ing stars (e.g., Joergens et al., 2003). There is a decrease in the peak in the separation distribution for BD-BD binaries the amplitude of periodic photometric variations with de- is likely to be at ~1 to 4 AU. In contrast, the distribution of creasing mass, presumably because the decreasing separations in binaries with Sun-like primaries peaks at temperature leads to weaker coupling between the gas and ~30 AU (Duquennoy and Mayor, 1991). The mass ratio dis- the and hence smaller spots. As with H-burn- tribution also seems to be dependent on primary mass, with ing stars, brown dwarfs show evidence for braking by their BD-BD binaries having a distribution that peaks toward accretion disks; those with strong accretion signatures α equal masses (qPEAK ~ 1), while binaries with Sun-like pri- (broad H emission) are exclusively slow rotators. X-ray maries have qPEAK ~ 0.3. The implication is that, as the pri- emission is detected from M-type brown dwarfs, with prop- mary mass decreases, (1) the multiplicity decreases, (2) the erties very similar to the X-ray emission from very-low- distribution of semimajor axes shifts to smaller separations mass H-burning stars (Preibisch et al., 2005), but none is and becomes narrower (logarithmically), and (3) the distri- detected from L-type brown dwarfs, which are too cool to bution of mass ratios shifts toward unity — with these trends support surface magnetic activity. all continuing across the divide between H-burning stars and brown dwarfs. 2.5. Synopsis 3. Exotica. Finally, we note that some brown dwarfs are components in more complex systems. There are at least Given the continuity of statistical properties between six systems currently known in which an H-burning star is brown dwarfs and H-burning stars, it is probably unhelp- orbited at large distance (>50 AU) by a BD-BD binary. ful to distinguish the formation of brown dwarfs from the Indeed, preliminary results suggest that wide brown dwarf formation of stars, and in the rest of this review we will companions to H-burning stars are 2–3× more likely to be only use the H-burning limit at ~0.075 M as a reference in a close BD-BD binary than are field brown dwarfs (Bur- point in the range of stellar masses. The D-burning limit at gasser et al., 2005). Cases also exist in which a close bi- ~0.012 M falls in the same category. We will then define nary with H-burning components is orbited by a wide brown a star as any object forming on a dynamical timescale by dwarf, while Bouy et al. (2005) have recently reported the gravitational instability. With this definition, there may well discovery of what is likely to be a triple brown dwarf sys- be a small overlap between the mass range of stars and that tem. However, the statistics of these exotic systems are cur- of planets. Given that in the immediate future we are un- rently too small to interpret quantitatively. likely to know too much more than the masses and radii of 462 Protostars and Planets V the lowest-mass objects, and certainly not their internal also assume that the is approximately , and composition, we will simply have to accept that there is a that the Rosseland- and Planck-mean opacities due to κ– κ– κ gray area in the range 0.001–0.01 M that may harbor both are (to order of magnitude) the same, R(T) = P(T) = 1(T/ β κ –3 2 –1 β stars and planets, and possibly even hybrids. K) with 1 = 10 cm g and emissivity index = 2 in It follows that understanding how brown dwarfs form is the far-infrared and submillimeter. With this assumption, our important, not just for its own sake, but because it is a key estimates will again not apply to the hot gas (T > 103 K) element in understanding why most stars have masses in where dust sublimates and the opacity falls abruptly with the range 0.01 M –100 M — and hence why there are lots increasing temperature (before picking up at even higher of hospitable stars like the Sun with long-lived habitable temperatures, due to the H– ). In sections 3.4 to 3.6, we zones, and enough heavy elements to produce rocky planets relax these assumptions. and . The high-mass cutoff is probably due to the fact that makes it hard to form the highest- 3.1. Three-Dimensional Collapse and Hierarchical mass stars. The low-mass cutoff is probably due to the opac- Fragmentation ity limit. By studying brown dwarf formation we seek to confirm and quantify the low-mass cutoff. In a uniform three-dimensional medium, an approxi- mately spherical fragment of mass MF3 will only condense 3. STAR-FORMATION THERMODYNAMICS out if it is sufficiently massive

In this section we review the basic thermodynamics of π5a6 1/2 and fragmentation. The first three sub- M > M = (1) F3 J3 3ρ sections deal — respectively — with three-dimensional col- 36G lapse and hierarchical fragmentation (section 3.1); two-di- mensional one-shot fragmentation of a shock compressed or equivalently, if it is sufficiently small and dense layer (section 3.2); and fragmentation of a disk (section 3.3). We describe and contrast these different environments, and π5a6 in each case estimate the minimum mass for star formation. ρ > ρ = (2) F3 J3 3 2 In section 3.3 we conclude that brown dwarf formation by 36G MF3 fragmentation of disks around Sun-like stars is more likely in the cooler outer parts (R > 100 AU), and this may ex- (Subscript F is for fragment, J is for Jeans, and 3 is for three- plain why brown dwarf companions to Sun-like stars almost dimensional.) The timescale on which the fragment con- always have large separations. In section 3.4 we explain denses out is why impulsive compression does not promote cooling of an optically thick fragment. In section 3.5 we suggest that close 3π 1/2M 2/3 –1/2 BD-BD binaries may be formed by secondary fragmenta- t = 1 – J3 (3) F3 ρ tion promoted by the softening of the equation of state when 32G MF3 H2 dissociates, and enhanced cooling due to the opacity gap. In section 3.6 we speculate that close BD-BD binaries may The molecular- gas from which stars are forming form in the outer parts of disks around Sun-like stars. today in the is expected to be approximately We caution that analytic estimates cannot capture all the isothermal, with T ~ 10 K, as long as it can radiate effi- nonlinear effects that are likely to occur, and are probably ciently via molecular lines and dust continuum. Therefore it important, in a process as chaotic as star formation; they has been argued, following Hoyle (1953), that star forma- are therefore only indicative. A full understanding of any tion proceeds in molecular clouds by a process of hierar- mode of star formation requires detailed simulations with chical fragmentation in which an initially massive low-den- all potentially influential physical effects included. How- sity cloud (destined to form a protocluster of stars) satisfies ever, as long as converged, robust simulations with all this condition (1) and starts to contract. Once its density has 2 –1 physics properly included remain beyond the compass of increased by a factor of f , MJ3 is reduced by a factor of f , current supercomputers, analytic estimates provide useful and hence parts of the cloud can condense out indepen- insights into the trends to be expected. dently, thereby breaking the cloud up into < f subclouds. In sections 3.1 to 3.3, we will be mainly concerned with Moreover, as long as the gas remains approximately isother- contemporary star formation in the disk of the Milky Way, mal, the process can repeat itself recursively, breaking the and therefore with molecular hydrogen at temperatures T < cloud up into ever smaller “sub-sub...subclouds.” 100 K where the rotational levels are not strongly excited. The process ends when the smallest sub-sub...subclouds In this regime the adiabatic exponent is γ = 5/3 and the iso- are so optically thick, and/or collapsing so fast, that the PdV sound is a = 0.06 km s–1 (T/K)1/2. With these work being done on them cannot be radiated away fast assumptions, our estimates will not apply to the hot gas (T > enough and they heat up. This process is presumed to de- 103 K) where the equation of state is softened by effects like termine the minimum mass for star formation (e.g., Rees, H2 dissociation (e.g., the inner regions of disks). We will 1976; Low and Lynden-Bell, 1976), and is usually referred to Whitworth et al.: The Formation of Brown Dwarfs 463

as the opacity limit, but see Masunaga and Inutsuka (1999) will subsequently increase its mass by a large factor through for a more accurate discussion. To estimate MMIN3 we first accretion before its condensation becomes nonlinear. Fi- formulate the PdV heating rate for a spherical fragment, nally, individual fragments will be backwarmed by the am- neglecting the background radiation field, bient radiation field from other cooling fragments, which in principle fill a significant fraction of the celestial , and again this will tend to increase M . dV 3M a2 dR 3M a2 GM 1/2 MIN3 H ≡ –P = –F3 F3 ~ F3 F3 (4) dt RF3 dt RF3 RF3 3.2. Two-Dimensional One-Shot Fragmentation of a Shocked Layer 1/2 where in putting dRF3/dt ~ –(GMF3/RF3) we are assuming the collapse is dynamical. By comparison, the maximum In fact, three-dimensional hierarchical fragmentation may radiative of a spherical fragment is be an inappropriate paradigm for star formation in molecu- lar clouds. There is growing evidence that, once a is assembled, star formation proceeds very rapidly, 4πR2 σ T4 L = F3 SB (5) essentially “in a crossing ” (Elmegreen, 2000). In this τ τ–1 ( R(T) + P (T)) scenario, star formation occurs in molecular clouds only where two or more turbulent flows of sufficient density col- τ– τ– where the optical depths are given by R(T) = P(T) = lide with sufficient ram pressure to produce a shock-com- κ 2 π 2 3MF3 1(T/K) /4 RF3. pressed layer or filament that then fragments to produce pre- If we follow Rees (1976) and assume that the fragment is stellar cores; in cases where flows converge simultaneously τ– τ––1 marginally optically thick, we can put ( R(T) + P (T)) = 2, from several directions, isolated cores may even form (see and the requirement that L > H then reduces to Padoan and Nordlund, 2002, 2004) (see also section 4). A basic model of this scenario can be constructed with relatively few parameters by considering two flows having π29 m24a36 1/7 ρ < ρ = (6) preshock density ρ, colliding at relative speed v, to produce F3 C3 2 5 6 3 2 12 18 2 3 5 G MFSc h a shock-compressed layer. If the effective isothermal sound speed in the resulting layer is a, the density is ~ρ(v/a)2. The ρ ρ Conditions (2) and (6) require J3 < C3 and hence layer is initially contained by the ram pressure of the inflow- ing gas, and until it fragments it has a rather flat density pro- file. It fragments at time t , while it is still accumulating, 2π2 1/4 3 1/2 F2 5 mPL a M > M = (7) and the fastest-growing fragments have mass MF2, radius F3 MIN3 2433 m2 c RF2 (in the plane of the layer), and half-thickness ZF2 (per- pendicular to the plane of the layer) given by 1/2 "5 Here mPL = (hc/G) = 5.5 × 10 g is the Planck mass, and π ρ 1/2 so MMIN3 is essentially the Chandrasekhar mass a fac- tF2 ~ (2a/ G v) (8) tor of (a/c)1/2 ~ 10–3. We also note the relatively weak de- ∝ 1/4 7 π 3ρ 1/2 pendence of MMIN3 on T ( T ) and the relatively strong MF2 ~ (8a / G v) (9) dependence on m (∝m–9/4). For contemporary local star for- –24 4 3 π ρ 1/2 mation we substitute m = 4.0 × 10 g and a = 1.8 × 10 cm RF2 ~ (2a / G v) (10) –1 s to obtain MMIN3 ~ 0.004 M . 5 π ρ 3 1/2 In general, the limiting fragment will not necessarily be ZF2 ~ (8a / G v ) (11) marginally optically thick, but it is trivial to substitute for τ– τ– R(T) and P(T), and it turns out that, for contemporary star (see Whitworth et al., 1994a,b). We note that (1) this mode formation, the value of MMIN3 is unchanged. This is be- of fragmentation is “two-dimensional” because the cause — coincidentally — the limiting fragment in contem- that assemble a fragment out of the shock-compressed layer porary star formation is marginally optically thick. are initially largely in the plane of the layer; (2) it is “one- There are, however, some serious problems with three- shot” in the sense of not being hierarchical; (3) the frag- dimensional hierarchical fragmentation. There is no conclu- ments are initially flattened objects (RF2/ZF2 ~ v/2a >> 1); sive evidence that it operates in , and nor does it seem (4) MF2 is not simply the standard three-dimensional Jeans to occur in numerical simulations of star formation. This is mass (MJ3, equation (1)) evaluated at the postshock density because a protofragment inevitably condenses out more and velocity dispersion — it is larger by a factor of ~(v/ slowly than the larger structure of which it is part, by virtue a)1/2; and (5) in reality the colliding flows will contain den- of the fact that it is, at every stage, less Jeans unstable than sity substructure that acts as seeds for fragmentation and that larger structure (see equation (3)). Therefore the proto- gives rise to a range of MF2 values. fragment is very likely to be merged with other nearby frag- From equation (9), we see that the fragment mass de- ments before its condensation becomes nonlinear. In addi- creases monotonically with the mass-flux in the colliding ρ tion, even if a protofragment starts off with mass ~MJ3, it flows, v. As in hierarchical fragmentation, a fragment in 464 Protostars and Planets V a shock-compressed layer will only condense out if it is able and the dissipated in the accretion shock. They find to remain approximately isothermal by radiating efficiently. that for T ~ 10 K, the minimum mass is 0.0027 M . This The PdV heating rate for a flattened fragment in a layer is fragment starts with mass ~0.0011 M , but continues to grow by accretion as it condenses out. dV pv2 2πR2 Z 2a5 H ≡ –P F2 ==F2 F2 (12) 3.3. Fragmentation of a Circumstellar Disk dt 4 tF2 G Another scenario that may be more relevant to contem- The radiative cooling rate of the fragment is porary star formation than three-dimensional hierarchi- cal fragmentation is fragmentation of a circumstellar disk. There are three critical issues here: (1) Under what circum- 2πR2 σ T4 8π5m4a11/15c2h3Gρv L = F2 SB = (13) stances does a disk fragment gravitationally? (2) Can the τ τ–1 τ τ–1 ( R(T) + P (T)) ( R(T) + P (T)) resulting fragments cool fast enough to condense out? (3) Can the resulting fragments lose angular momentum fast τ– τ– where the optical depths are now given by R(T) = P(T) = enough to condense out? The last two issues are critical, ρ π 1/2κ 2 (2a v/ G) 1(T/K) . The requirement that L > H then because if a fragment cannot cool and lose angular momen- reduces to a limit on the mass flux in the colliding flows tum fast enough, it is likely to bounce and be sheared apart. For simplicity we consider an equilibrium disk. 1. The condition for an isolated disk to fragment gravi- 4π5m4a6/15c2h3 ρv < (14) tationally is that the surface density, Σ, be sufficiently large τ τ–1 ( R(T) + P (T)) αε If we assume the fragment is marginally optically thick, and Σ > Σ = (16) T π τ– τ––1 G set ( R(T) + P (T)) = 2, we obtain

where a is the local sound speed and ε is the local epicyclic (30)1/2 m3 a 1/2 M > M = PL (15) frequency (Toomre, 1964). Condition (16) is also the condi- F2 MIN2 π3 2 m c tion for spiral modes to develop in the disk, and these will have the effect of redistributing angular momentum. As a and for contemporary local star formation, this gives MMIN2 ~ result, the inner parts of the disk may simply accrete onto 0.001 M . Once again, if we treat the completely general the central primary star and the outer parts may disperse case by including the optical-depth terms, we obtain essen- without fragmenting (Laughlin and Bodenheimer, 1994; tially the same value for MMIN2, because — purely coinci- Nelson et al., 1998). Thus, if fragmentation is to occur, it dentally — the limiting mass for contemporary local star for- must occur on a dynamical timescale. mation is again marginally optically thick. 2. The condition for a fragment to cool fast enough to Although one-shot two-dimensional fragmentation of a condense out is therefore that the fragment can radiate away, shock-compressed layer and three-dimensional hierarchical on a dynamical timescale, the being deliv- fragmentation give essentially the same expression for the ered by compression. Gammie (2001) has shown that for a minimum mass (MMIN2 ~ MMIN3), they give very different Keplerian disk, this condition can be written as a constraint ≠ expressions for the Jeans mass (MJ2 MJ3 ), and there are on the cooling time other important differences. In particular, two-dimensional one-shot fragmentation bypasses all the problems associated 3 with three-dimensional hierarchical fragmentation. There is t < (17) COOL Ω no backwarming because there are no other local fragments filling the part of the celestial sphere into which a fragment radiates (i.e., perpendicular to the layer). More importantly, where Ω is the local orbital angular speed. there is little likelihood of fragments merging, since frag- We assume that, as a disk forms, Σ increases sufficiently Σ ments with MF2 ~ MJ2 condense out faster than any other slowly that it does not greatly exceed T when the disk be- structures in a layer [whereas fragments in a three-dimen- comes unstable. It then follows that the radius, growth time, sional medium with MF3 ~ MJ3 condense out slower than all and mass of the fastest growing fragment are larger structures in that medium (e.g., Larson, 1985)]. Fi- nally, because condensation in a layer is so fast, growth by a R = (18) accretion is limited. FD ε Boyd and Whitworth (2004) have analyzed the radiative cooling of a fragmenting layer, taking into account not only 1 the PdV heating of a condensing fragment, but also the on- t = (19) FD ε going accretion (as matter continues to flow into the layer) Whitworth et al.: The Formation of Brown Dwarfs 465

Therefore the condition for a fragment to lose angular mo- a3 M = (20) mentum fast enough to condense out has to be evaluated FD ε G by numerical simulation. The chapter by Durisen et al. deals with this problem. The compressional heating rate for a fragment is thus 3.4. Nonlinear Thermodynamics dV 3M a2 3a5 H = P ==FD (21) Impulsive triggers that produce rapid compression will dt 2tFD 2G always help to amplify self-, because the freefall time ∝ ρ–1/2 varies as tFF . However, there is only a very restrict- and the radiative cooling rate is ed temperature range within which rapid compression will help an optically thick protofragment to cool more rapidly (and thereby avoid the likelihood of its bouncing and be- 2πR2 σ T4 4π6m4a10/15c2h3ε2 L = FD SB = (22) ing sheared apart). Suppose that the Rosseland-mean opac- τ τ–1 τ τ–1 ( R(T) + (T)) ( R(T) + (T)) κ– ∝ ρα β P P ity is given by R T , and that the gas has a ratio of specific γ. τ– τ– εκ– π where now R(T) = P(T) = a (T)/ G. Consequently the re- Then if a fragment condenses out quasistatically, its cool- ∝ ρα + (1 + β)/3 quirement that L > H reduces to ing time varies as tCOOL , so tCOOL only de- β α creases as fast as tFF, with increasing density, if < –(6 + 5)/2. This condition is only likely to be satisfied in the tem- ε2 4π6Gm4/45c2h3 < (23) perature range where refractory dust sublimates (1500– 5 τ τ–1 a ( R(T) + P (T)) 3000 K). If instead the fragment is compressed impulsively — and ∝ To illustrate the discussion we consider the specific case therefore adiabatically — its cooling time varies as tCOOL ρα + 4/3 + (β – 3)(γ – 1) of a Keplerian disk around a Sun-like star, with , so now tCOOL only decreases as fast as β α γ tFF, with increasing density, if < 3 – (6 + 11)/6( – 1). 1/2 –3/2 For T < 100 K, we have γ = 5/3 and so enhanced cooling M★ D ε(D) = 2 × 10–7 s–1 (24) requires β < –1/4, which is unlikely. For 100 K < T < M AU O 1000 K, γ = 7/5 and enhanced cooling requires β < –19/12, which is even less likely. For 1000 K < T < 3000 K, H2 dis- 1/4 –1/2 sociation gives γ ~ 1.1 and enhanced cooling requires β < L★ D T(D) = 300 K (25) –15, which may occur during sublimation, but then neces- L AU O sarily only over a small temperature range. For 3000 K < T < 10,000 K, γ ~ 5/3 and H– opacity gives α ~ 1/2 and β ~ –1 1/8 –1/4 ∝ ρ7/2 and hence a(D) = 1 km s (L★/L ) (D/AU) , where D 4, so tCOOL . At even higher temperatures where Kra- ∝ ρ–2 is distance from the Sun-like star. The fastest-growing frag- mers opacity dominates tCOOL , and where electron ∝ ρ–2/3 ment then has mass scattering opacity dominates tCOOL , but by this stage a fragment is very opaque and very strongly bound. –1/2 3/8 3/4 M★ L★ D M = 3 × 10–5 M (26) 3.5. Forming Close Brown Dwarf–Brown Dwarf FD M L AU O O Binaries

and condition (23) is only satisfied for It is possible that there is a secondary fragmentation regime at T ~ 2000 K, due to the softening of the equation 3/7 –1/7 of state caused by H dissociation and the enhanced cooling M★ L★ 2 D > 90 AU (27) that occurs in the opacity gap between dust sublimation and M L O O H– opacity. A spherical cloud of mass M in hydrostatic equi- librium at this temperature has radius –5/28 15/56 M★ L★ M > 0.003 M (28) FD M L M O O R ~ 4 AU (29) 0.1 M 3. Angular momentum is removed from a condensing fragment by gravitational torques, in the same way that If the cloud fragments into a binary and virializes, the binary angular momentum is redistributed in the disk as a whole should have a separation of the same order. Equation (29) Σ Σ when > T. This is a nonlinear and stochastic process, and is actually a good mean fit to the values plotted on Fig. 6 there is no analytic estimate of the rate at which it occurs. of the chapter by Burgasser et al., suggesting that BD-BD 466 Protostars and Planets V binaries may be produced by secondary fragmentation facil- ber of stars being formed from each core (Goodwin and itated by H2 dissociation and/or the opacity gap. Kroupa, 2005).

3.6. Forming Close Brown Dwarf–Brown Dwarf 4.1. Core Structure and Low-Mass Core Formation Binaries in Disks Observed prestellar cores generally have Bonnor-Ebert Burgasser et al. (2005) have noted that — modulo the (BE)-like density profiles, with relatively sharp outer edges small-number statistics involved — a brown dwarf in a wide (Bacmann et al., 2000; Motte and André, 2001; Kirk et al., (>200 AU) about a Sun-like star is apparently more 2005). Starless cores have flat BE-like profiles, while cores likely to be in a close BD-BD (<20 AU) than with detected Class 0 or Class I protostars have more cen- a brown dwarf in the field. If this trend is confirmed, it trally peaked density profiles, presumably as a result of their suggests that BD-BD binaries are formed in disks, and then deeper potential wells. Since well-defined cores with sharp may be ejected. For a close BD-BD binary in a wide orbit edges define finite mass reservoirs for the stars that form around a Sun-like star, the internal of the inside them, the core observations reinforce the notion that BD-BD binary is typically comparable to or larger than the the masses of stars are, at least statistically, strongly influ- binding energy of the BD-BD binary to the Sun-like star, enced by the masses of the cores within which they form and therefore some BD-BD binaries should be able to sur- (Padoan and Nordlund, 2002, 2004). vive ejection. It is therefore important to consider whether it is possi- Moreover, we know from the Toomre criterion that for ble to form brown dwarfs directly, from correspondingly low- fragments condensing out of disks, losing angular momen- mass cores. The standard argument against brown dwarfs tum is a critical issue. The smaller a condensation becomes, being formed directly is that the density needs to be very the slower the rate at which angular momentum can be lost high for a fragment with brown dwarf mass to collapse grav- by gravitational torques. Therefore such a condensation may itationally. If one considers low-amplitude perturbations on be strongly disposed to binary fragmentation when its ther- top of a typical mean density of 104 to 105 cm–3, he will mal support is weakened at T ~ 2000 K. Since only a sub- conclude that only fluctuations on the order of the Jeans set of brown dwarfs is in close BD-BD binaries, this sug- mass (1–3 M ; equation (1)) are able to collapse in a typical gestion does not require that all brown dwarfs are formed molecular cloud. However, many cores actually have very in disks. high density contrast relative to their surroundings, and so it is inappropriate to use the Jeans mass at the mean density 4. FORMING VERY-LOW-MASS as an estimate of the resulting protostellar mass. PRESTELLAR CORES IN It is also important to realize that cores having suffi- TURBULENT CLOUDS ciently high density to form brown dwarfs directly need not — indeed must not — be very common, and therefore A number of surveys (Testi and Sargent, 1998; Peng et we can appeal to exceptional circumstances to generate al., 1998; Motte et al., 1998, 2001; Motte and André, 2001; them. As a measure of how exceptional the circumstances Johnstone et al., 2000; Sandell and Knee, 2001; Onishi et must be, consider the IMF (in the form dN /d ln M). At high al., 2002; Reid and Wilson, 2005) have noted the close simi- masses, the IMF falls off with the Salpeter (1955) expo- larity between the core mass function (CMF) and the stel- nent –1.35, and if this power-law IMF continued unbroken lar initial mass function (IMF). Cores with brown dwarf down to the brown dwarf regime, 104 × 1.35 = 250,000 brown masses have been observed in several of these surveys, but dwarfs with M ~ 0.01 M would be formed for every mas- since these low-mass cores are usually below the complete- sive star with M ~ 100 M . Instead, the IMF peaks near ness limit, their statistics are unreliable. However, the simi- ~0.3 M and falls off at low masses; adopting the Chabrier larity between the CMF and the IMF suggests that each core (2003) IMF for the Milky Way, we find that the number of gives rise to a similar low number of stars, whose final mas- brown dwarfs formed with M ~ 0.01 M is actually about ses are heavily influenced by the mass of the core. Indeed, equal to the number of massive stars formed with M ~ protostars are generally observed inside relatively well-de- 100 M , as predicted theoretically by Zinnecker (1984). fined envelopes, as implied by the definition of Class 0 and In order for a high-density core to form, its mass must be Class I objects (André et al., 1993, 2000; Tachihara et al., concentrated into a small . If is the fluid velocity, 2002). This is not to say that the final masses of stars are accumulation of mass is measured by the quantity –∇ · u = the same as the masses of the prestellar cores in which they D ln ρ/Dt (the co-moving time derivative of log density), form. Protostars are associated with outflows, which can and so individual cores form at local maxima of –∇ · u. remove a significant fraction of the mass of the protostellar The flow toward a convergence point is generally super- envelope, and cores may split into more than one star (see sonic, and so standoff shocks develop, separating the un- also Hosking and Whitworth, 2004; Stamatellos et al., 2005). shocked upstream gas from the shocked and nearly stagnant However, low-mass cores appear typically to harbor only downstream gas. The standoff shocks correspond to density one or two protostars (e.g., Tachihara et al., 2002), and the jumps ~M 2 if the shocked gas is dominated by gas pres- overall binary statistics are inconsistent with a larger num- sure; and ~MA (the Alfvénic Mach number) if the shocked Whitworth et al.: The Formation of Brown Dwarfs 467

gas is dominated by magnetic pressure. As matter accumu- As an example, consider the formation of a core with lates around the convergence point, a deepening gravita- temperature ~10 K and central density ~107 cm–3, suffi- tional potential well develops, and the accumulated and ciently high to form a star with M ~ 0.1 M . If the average stagnant gas forms a growing Bonnor-Ebert like core, stabi- cloud density is ~104 cm–3 and the Mach number is 5, f only lized by the external ram pressure, ρ|u|2. needs to be ~3; i.e., the upstream flow needs to be focused If the density at the center of the core increases to more onto a 3× smaller standoff shock surface. This must be rather than about 14× the boundary density (just inside the shock common, and indeed at M ~ 0.1 M the IMF has hardly surface), the core becomes gravitationally unstable and col- dropped below its maximum. lapses. However, even before it becomes unstable, such a Repeating this estimate for a core with M ~ 0.01 M one core has — by virtue of the shock jump at its boundary — finds that the flow needs to be much better focused. The a total density contrast (between its center and the surround- standoff shock surface now needs to be ~300× smaller than ing inflowing unshocked gas) that can greatly exceed a fac- the upstream source area, and so the linear size of the core tor of 14. Moreover, if the inflowing gas runs out before needs to be ~17× smaller than the scale of the upstream the core has become sufficiently massive and dense to col- flow. This does require a rather exceptional focusing of the lapse, then the core will expand and disperse due to the de- upstream flow, but then brown dwarfs with M ~ 0.01 M crease in ram pressure at its boundary. actually are very rare. The fraction of brown dwarfs that In the last column of Table 3 in Bacmann et al. (2000), form in this manner can only be ascertained quantitatively at least 8 of the 9 cores listed have density contrasts exceed- by performing very-high-resolution numerical simulations. ing 14, but no strong indication of collapse. Bacmann et Figure 1 illustrates an example intermediate between al. conclude that the structures are best fitted with Bonnor- these two extremes, i.e., a core with final mass ~0.03 M . Ebert , but since they cannot explain how the cores can be stable at these high-density contrasts, they go on to 4.2. Envelope Breakout and Continued Accretion discuss other density profiles and models, which do not fit the data as well as BE-spheres. Similar core profiles have The sharp density transitions surrounding accreting cores been observed in many other surveys. There is thus direct produce more well-defined cores than would exist in a sub- observational evidence for the creation of prestellar cores sonic medium. Nevertheless, there will be some continued with large density contrasts between their centers and the accretion after a collapsed object is formed, until the up- surrounding medium, as expected when cores are produced stream supply of gas is exhausted. Except toward the end as stagnant structures by supersonic flows. Indeed, cores of the process, such cores correspond to Class 0 objects, formed in numerical simulations of molecular cloud turbu- which by definition have not yet accreted half their final lence have internal velocity dispersions and rotation veloci- mass (André et al., 2000). ties that are entirely consistent with those of observed cores (cf. Figs. 8 and 9 in Nordlund and Padoan, 2003). With this scenario in mind, we now investigate semi- quantitatively the range of densities that is possible, and the circumstances under which the densities attained would be high enough to lead to the collapse of brown dwarf mass cores. Any particular converging flow may be broadly char- ρ acterized by three parameters: the upstream density 0, the Mach number M of the upstream flow, and the of focusing of the upstream flow toward a three-dimensional convergence point. As a measure of the latter we take the ratio f of the surface area that encloses the upstream flow to the surface area of the standoff shock that surrounds the central BE-like structure. Since the upstream flow is super- sonic, its is essentially inertial (i.e., constant veloc- ity), until it encounters the standoff shock. Thus, by mass ρ ρ conservation, its density increases from 0 to f 0. At the standoff shock the density increases by an additional fac- Fig. 1. Density and velocity profiles, through a core forming 2 from a spherically symmetric convergent flow. The thin line rep- tor M for hydroshocks (or MA for MHD shocks). Finally, from just inside the shock to the center of the BE-like core resents an early stage when the ~0.015 M core approximates to a stable and stagnant BE-like configuration; the boundary shock the density increases by a further factor of <14 (~14, in the is at ~300 AU. The thick line represents a later stage after the core marginally stable case). The total density increase from the has become unstable and started to collapse; the boundary shock is 2 upstream source to the core center is ~14 fM . This factor now at ~350 AU, because the core is more massive (~0.025 M ). has the right dependence to account for the formation of Outside the shock is the convergent inertial flow. Note the very brown dwarf mass cores, in that it can be arbitrarily large, large density contrast, even between the gas at the edge of the core albeit with decreasing probability. and the ambient medium at ~1500 AU. 468 Protostars and Planets V

Class I and II objects are observed to accrete from low- n that reproduce the observed ratio of brown dwarfs to H- mass envelopes (Motte and André, 2001) and have much- burning stars. reduced accretion rates, <10–6 M yr–1. Once the natal en- velope is consumed there is no longer a strong coupling 5. COLLAPSE AND FRAGMENTATION OF between the collapsed object (moving under the N-body LARGE PRESTELLAR CORES influence of neighboring objects and gas) and the surround- ing medium, and one expects the collapsed object to pick While a very-low-mass prestellar core (<0.1 M ) must up speed relative to the surrounding gas, and to then accrete collapse to form either a single brown dwarf or a multiple in a manner similar to Bondi-Hoyle accretion (Padoan et brown dwarf system, larger prestellar cores (>1 M ) are ex- al., 2005; Krumholz et al., 2005b). In a turbulent medium pected to form clusters of stars having a range of masses. the problem is more complicated (Krumholz et al., 2005a), We can identify five mechanisms that may play a role in de- but the scaling is more or less as for Bondi-Hoyle accretion, termining the final stellar masses: (1) During the approxi- with a prefactor Φ ~ 1 to 5 that accounts for the complica- mately isothermal initial collapse phase, as the pressure be- tions. Quantitative estimates indicate that the accretion after comes increasingly unimportant and the collapse approaches breakout from the natal core is insignificant. freefall, self-gravity will amplify any existing density sub- structure. (2) Then, when the density reaches n ~ 1011 cm–3, H2 4.3. The Viability of Turbulent Fragmentation the gas becomes optically thick and switches rather sud- denly from approximate isothermality to approximate adia- Although turbulent fragmentation generates a mass func- baticity. At this juncture, a network of shock waves devel- tion for prestellar cores, matching broadly the observed ops to slow the collapse down, and nonlinear interactions stellar IMF, there are two caveats that should be born in between these shock waves produce and amplify further mind when considering the formation of brown dwarfs. substructure (sheets, filaments, and isolated prestellar cores). First, in turbulent fragmentation (Padoan and Nordlund, (3) Some of these structures will have sufficient angular mo- 2002), brown dwarfs are produced by a small subset of low- mentum to form disks, and these may then fragment due to mass cores, i.e., those that are sufficiently dense to collapse rotational instability (see section 6). Finally, once a proto- gravitationally. At its low-mass end, the overall (i.e., fully stellar embryo (i.e., an object that is sufficiently well bound sampled) CMF should be dominated by a much larger num- to be treated dynamically as a single entity) has condensed ber of transient (i.e., nonprestellar) cores. Moreover, many out of a fragment, its subsequent evolution and final mass of these transient cores will be only slightly less dense than will be determined by (4) competitive accretion and (5) dy- the ones that spawn brown dwarfs. They will therefore be namical interaction (see section 7). Here we concentrate on detectable in submillimeter continuum observations, and mechanisms (1) and (2), since these are the ones that distin- they should also be somewhat longer lived. Thus, even al- guish the evolution of a high-mass core from the low-mass lowing for selection effects, the observed CMF should fall cores considered in section 4. much less steeply with decreasing mass than the stellar IMF. High-mass prestellar cores (>10 M ) invariably display Recent estimates of the CMF (Nutter and Ward-Thompson, nonlinear internal density structure, and at the typical den- in preparation), which probe to lower masses by using longer sity (n ~ 105 cm–3) and temperature (T ~ 10 K) in a pre- H2 exposures, actually suggest the opposite, but completeness , the Jeans mass is MJ3 ~ 0.8 M (equation (1)). remains a concern (e.g., Kirk et al., 2006). Therefore, in the absence of a significant magnetic field, Second, turbulent fragmentation predicts a ratio of brown they are very likely to fragment during collapse. dwarfs to H-burning stars, R , which is exponentially sensi- Even the smallest cores are usually far from spherical, tive to the Alfvénic Mach number on the largest scales (MA) and have been modeled as being either prolate or oblate and to the mean cloud density (n). Regions with smaller (Myers et al., 1991; Ryden, 1996; Jones et al., 2001; Good- values of MA and/or n should generate stellar populations win et al., 2002; Curry, 2002; Myers, 2005), with prolate with significantly fewer brown dwarfs, and regions with models being favored statistically, although in reality cores larger values should generate stellar populations with signi- are probably triaxial — or even more complicated — in ficantly more brown dwarfs. However, in nature R appears their full three-dimensional structure (Jones et al., 2001; to vary little over a wide range of local star-forming envi- Goodwin et al., 2002). Gravity works to enhance anisotro- ronments (see chapter by Luhman et al.). pies in collapsing objects (Lin et al., 1965) with collapse This problem can be overcome if nature selects a narrow occurring fastest along the shortest axis to form sheets and range of MA and n. For example, Whitworth (2005) has noted filaments, which then subsequently fragment (Bastien, 1983; that contemporary star formation may only proceed rapidly Inutsuka and Miyama, 1992). Thus, it is unsurprising that if the gas couples thermally to the dust (so that it can avail hydrodynamical simulations of both oblate and prolate cores itself of broadband — as distinct from molecular-line — are prone to fragmentation (e.g., Bastien et al., 1991; Boss, cooling). This requires that the ram pressure in shocks pro- 1996). ducing prestellar cores exceeds a critical value, which con- Prestellar cores also tend to have complex internal ve- 2 5 –3 verts into the constraint nTMA > PCRIT ~ 10 cm K. This locity fields (Larson, 1981; Myers and Benson, 1983; Ar- constraint may help to select the combinations of MA and quilla and Goldsmith, 1985), but since prestellar cores can Whitworth et al.: The Formation of Brown Dwarfs 469 only be observed from a single direction, the interpretation Rice et al. (2003) have corroborated equation (17) nu- of these velocities is difficult. If cores are assumed to be in merically by performing SPH simulations of disk fragmen- solid-body rotation, the ratio of rotational to gravitational tation with a parameterized cooling law of the form du/dt = energy is typically β ≡ R /|Ω| ~ 0.03, with some cases as –uΩ/β (where u is the specific internal energy and Ω the high as β ~ 0.1 (Goodman et al., 1993; Barranco and Good- orbital angular speed). Endemic fragmentation occurs with man, 1998). However, the observed velocities are more β = 3 but not with β = 5. However, because this cooling law likely to be turbulent in nature, i.e., less well ordered than results in indefinite cooling, whereas there is a limit to the solid-body rotation (Myers and Gammie, 1999; Burkert and supply of rotational energy that can be tapped through shock Bodenheimer, 2000). Quite low levels of turbulence (e.g., heating, by the time fragmentation occurs the temperatures Goodwin et al., 2004a), and/or global rotation (e.g., Boss, have dropped to rather low values, which may be hard to 1986; Bonnell and Bate, 1994a; Hennebelle et al., 2004; realize in nature. Cha and Whitworth, 2003) are sufficient to make a collaps- Rafikov (2005) argues that the opacity of gravitationally ing core fragment into a small ensemble of protostellar unstable disks is so high, and the cooling times are there- embryos. fore so long, that fragments can only condense out on a dy- Unfortunately, collapse and fragmentation can only be namical timescale in the outer, cooler regions of massive explored by means of numerical simulations, and there is disks. (Although the treatment is somewhat different, his a huge and poorly constrained range of admissible initial conclusion resonates with the simple analysis we have pre- conditions, which makes the extraction of robust theorems sented in section 3.3 and carries the same caveats.) very hard (see, e.g., Hennebelle et al., 2004). Moreover, Johnson and Gammie (2003) point out that the effects almost all simulations to date use barotropic equations of of opacity may be mitigated in temperature regimes where state [i.e., P = P(ρ)]. These barotropic equations of state are dust sublimates and hence the opacity decreases abruptly designed to mimic the expected thermal behavior of proto- with increasing temperature. However, dust sublimation stellar gas, but they do not capture the thermal inertia effects effects are confined to the temperature range T > 200 K, and that become important at the juncture when the gas starts to the most critical effects occur at T > 2000 K, so they are heat up due to adiabatic compression, and it appears (Boss probably only relevant to fragmentation close to the cen- et al., 2000) that these thermal inertia effects play a criti- tral star (D < 10 AU). cal, deterministic role in gravitational fragmentation. Proper Cai et al. (2006) report simulations of disk evolution tak- treatment of the energy equation and the associated radia- ing radiation transport into account, and conclude that — tive transport (e.g., Whitehouse and Bate, 2006) is needed even at low — the opacity in the hot inner parts to make these simulations more realistic. of a disk is too high to allow fragmentation on a dynamical timescale, in agreement with Rafikov (2005). 6. DISK FRAGMENTATION In contrast, Boss (2004) presents simulations of disk evo- lution taking radiation transport into account and concludes We organize our discussion of disk fragmentation un- that fragments of do condense out — or at der the headings (1) isolated, relaxed disks; (2) unrelaxed least that gravitationally bound fragments form that would disks; and (3) interacting disks. subsequently condense out. He argues that the protofrag- ments in his simulations bypass the effects of high opacity 6.1. Isolated Relaxed Disks by transporting energy convectively. However, it is not clear how can cool a fragment that is condensing out The dynamical fragmentation of isolated relaxed disks on a dynamical timescale. The coherent small-scale motions is discussed in detail in the chapter by Durisen et al. Al- he attributes to convection may actually be manifestations though the emphasis there is on the of planets, the of a fragment that is unable to cool and is bouncing — or same issues pertain to the formation of brown dwarfs, viz., will soon bounce — prior to being sheared apart. This needs under what circumstances do disks become unstable against to be investigated further, but the numerical complexities fragmentation? Is equation (16) a sufficient condition for are considerable. gravitational instability (in which case, what is the precise Σ Σ value of T) or is it also necessary for to increase rap- 6.2. Unrelaxed Disks idly (in order to avoid the disk simply being accreted and dispersed by torques due to spiral density waves)? Does In numerical simulations of the collapse and fragmen- equation (17) determine whether fragments can cool fast tation of intermediate- and high-mass prestellar cores, the enough to condense out? What role is played by thermo- first single protostars to form usually quickly acquire mas- dynamic effects like H2 dissociation, dust sublimation, and sive circumstellar disks, and secondary protostars then con- convection? What are the properties of dust in disks, and dense out of these disks. This pattern is common for rotat- how well mixed are the gas and dust? Does the survival of a ing cores (e.g., Bonnell, 1994; Bonnell and Bate, 1994a,b; fragment depend on its ability to lose angular momentum Turner et al., 1995; Whitworth et al., 1995; Burkert and rapidly? Some of these issues are only beginning to be in- Bodenheimer, 1996; Burkert et al., 1997), for turbulent cores vestigated. (e.g., Bate et al., 2002b, 2003; Goodwin et al., 2004b), and 470 Protostars and Planets V for cores that are subjected to a sudden increase in exter- 7. PREMATURE EJECTION OF nal pressure (e.g., Hennebelle et al., 2004). The disks thus PROTOSTELLAR EMBRYOS formed fragment before they have time to relax to an equi- librium state. Indeed, the material accreting onto the disk The scenario where brown dwarfs form by premature is often quite lumpy, and this helps to seed fragmentation. ejection is closely linked to — but ultimately independent Under this circumstance, gravitational fragmentation is more of — the notion of competitive accretion. In the ejection likely simply because protofragments are launched directly hypothesis, brown dwarfs are simply protostellar embryos into the nonlinear regime of gravitational instability, rather that get separated from their reservoir of accretable mate- than having first to grow through the linear phase. rial at an early stage, and so in the context of brown dwarf Even so, a protofragment still has to be able to cool and formation it is irrelevant whether there are other stars com- lose angular momentum on a dynamical timescale, if it is to peting for this same material. The importance of competi- condense out, rather than bouncing and then being sheared tive accretion for intermediate- and high-mass stars is ar- apart. Fragmentation of unrelaxed disks can only be studied gued in the chapter by Bonnell et al. The contentious issue by means of numerical simulations, and to date no simula- is whether protostellar embryos — i.e., the first, very-low- tions of the process have been performed that treat prop- mass (~0.003 M ) high-density star-like objects — exist for erly the energy equation and the associated transport of - long enough to do much competing. At one extreme, it is diation. Since young protostars have relatively high lumi- argued that a protostellar embryo forms following the non- nosities, an important consideration will be irradiation of homologous collapse of a much larger gravitationally un- the disk by the primary at its center and any other stable core, and therefore it accretes mainly from its own nearby protostars. co-moving placenta. At the other extreme, it is argued that a protostellar embryo quickly becomes sufficiently decoupled 6.3. Interacting Disks from the ambient gas that it can roam around competing with other embryos for the same reservoir of accretable gas. Another way in which disk fragmentation can be trig- For the purpose of this section, we assume that nature gered is by an impulsive interaction with another disk, or cleaves to the second extreme. with a naked star. In the dense protocluster environment where most stars are presumed to form, such interactions 7.1. Competitive Accretion in Gas-rich Protoclusters must be quite frequent, since many very young protostel- lar disks have diameters >300 AU and the mean separa- Once an ensemble of protostellar embryos has formed, tion between neighboring protostars in a typical cluster is still deeply embedded in its parental prestellar core and/or <3000 AU. Indeed, ~50% of solar-type stars end up in bi- parental disk, the individual embryos evolve by accreting nary systems with semimajor axes a < 1000 AU, so the no- gas, and by interacting dynamically with one another. The tion of a disk evolving in the gravitational field of a single, accretion histories of individual embryos differ due to their isolated protostar is probably rather artificial. varying circumstances, leading to a of protostellar Boffin et al. (1998) and Watkins et al. (1998a,b) have masses, extending from high masses down to below the H- simulated parabolic interactions between two protostellar burning limit. Those that spend a long time moving slowly disks, and between a single protostellar disk and a naked through the dense gas near the center of the core can grow to protostar. All possible mutual orientations of spin and orbit high mass. Conversely, those that spend most of their time are sampled, and the gas is assumed to behave isothermally, moving rapidly through the diffuse gas in the outer reaches which is probably a reasonable assumption, since the disks of the core do not grow much. This process of “competi- are large (initial radius 1000 AU) and most of the second- tive accretion” may be a major element in the origin of the ary protostars form at large distances (periastra > 100 AU). IMF, as first pointed out by Zinnecker (1982). The critical parameter turns out to be the effective shear Over the past decade, many numerical simulations of the viscosity in the disk. If the Shakura-Sunyaev parameter is formation of star clusters by fragmentation and competitive α –3 low, SS ~ 10 , most of the secondary protostars have mas- accretion have been performed (e.g., Chapman et al., 1992; α ses in the range 0.001 M to 0.01 M . Conversely, if SS Turner et al., 1995; Whitworth et al., 1995; Bonnell et al., α –2 is larger, SS ~ 10 , most of the secondary protostars have 1997, 2001a,b; Klessen et al., 1998; Klessen and Burkert, masses in the range 0.01 M to 0.1 M . The formation of 2000, 2001; Bate et al., 2002a,b, 2003; Goodwin et al., low-mass companions is most efficient for interactions in 2004b, 2004c). Bonnell et al. (1997, 2001a,b) have shown which the orbital and spin angular momenta are aligned; through numerical simulations and analytical arguments that on average 2.4 low-mass companions are formed per inter- competitive accretion in large-N protoclusters can reproduce action in this case. If the orbital and spin angular momenta the general form of the IMF. are randomly orientated, then on average 1.2 companions are formed per interaction. It is important that such simu- 7.2. Unstable Multiple Systems lations be repeated, with a proper treatment of the energy equation and the associated energy transport, to check their Multiplicity studies of main-sequence and evolved stars fidelity, and to establish whether low-mass companions can have revealed that 15–25% of all stars, when studied in suf- form at closer periastra. ficient detail, are triple or higher-order multiples (e.g., Toko- Whitworth et al.: The Formation of Brown Dwarfs 471

vinin and Smekhov, 2002; Tokovinin, 2004). It follows that thereby create protostellar embryos that at the outset are the formation of multiple stars is an important element of inevitably low-mass (~0.005 M ), very prone to dynamical the star-formation process. interaction (being effectively point masses, albeit with grav- The multiplicity fraction among PMS stars is poorly ity softening), and unable to merge. Thus, although forma- known, partly because of the difficulty in studying stars in tion by ejection seems inevitable, it may be less efficient the embedded phase, but it appears to be at least as high than these simulations suggest. As discussed by Goodwin as — and probably much higher than — that for more and Kroupa (2005) and Hubber and Whitworth (2005), the evolved stars (e.g., Reipurth, 2000; Looney et al., 2000; observed binary statistics are incompatible with too many Koresko, 2002; Reipurth et al., 2002, 2004; Haisch et al., ejections, and the number of protostellar embryos under- 2004) (see also chapters by Duchêne et al., Goodwin et al., going dynamical interactions within a single prestellar core and Burgasser et al.). Most stars are formed in embedded should be relatively small (N < 4). (2) They do not include clusters, and there is increasing evidence that the primary magnetic fields. Boss (2002) argues that a magnetic field building blocks of clusters are small subclusters of <20 stars may promote fragmentation of a collapsing core by inhib- that quickly dissolve and merge to form the more extended iting the formation of a central density peak; however, his cluster (e.g., Teixeira et al., 2006). code does not capture all the possible MHD effects, in par- Nonhierarchical multiple systems, in which the time- ticular the anisotropy of magnetic pressure and the torques averaged distances between components are comparable, exerted by a twisted field. In contrast, Hosking and Whit- are inherently unstable (e.g., van Albada, 1968). Within worth (2004) have simulated the collapse of a rotating core about 100 crossing times a triple system is likely to have with imperfect MHD, and find that fragmentation is inhib- ejected one member, most likely the least-massive compo- ited. More work is needed. nent, since the ejection probability scales approximately as In their original paper, Reipurth and Clarke (2001) con- the inverse third power of the mass (e.g., Anosova, 1986; jectured that brown dwarfs formed by ejection could have a Mikkola and Valtonen, 1986). Although most nonhierarchi- higher-velocity dispersion than more massive H-burning cal systems disintegrate in this way, the existence of nu- stars, and that this might be detected either in observations merous stable hierarchical triple systems shows that this is of the velocity dispersions of young star clusters or as a not always the case. Ejected members leave with a velocity diaspora of brown dwarfs around more evolved clusters. that, to first order, is comparable to the velocity attained at They also pointed out that violent ejections would result in pericenter in the close triple encounter, and depends on the smaller accretion disks, and therefore a shortened quasi- geometry of the encounter, the energy and angular momen- T Tauri phase. Numerical simulations of star-cluster forma- tum of the system, and the masses of the components (e.g., tion in highly turbulent massive cores (e.g., Bate et al., Standish, 1972; Monaghan, 1976; Sterzik and Durisen, 2003; Bate and Bonnell, 2005) find that the brown dwarfs 1995, 1998; Armitage and Clarke, 1997). Ejection of brown do indeed have quite a high-velocity dispersion (~2 to 4 km dwarfs from systems that are marginally hierarchical is also s–1), but they are difficult to distinguish from the low-mass possible (Jiang et al., 2004). H-burning stars, because both are frequently ejected from The disintegration of a small multiple system is most small dynamically unstable groups. Simulations of low-den- likely to occur during the deeply embedded Class 0 phase, sity star-forming regions by Goodwin et al. (2005) find that while massive accretion from a surrounding envelope is still the velocity dispersions are typically somewhat lower (~0.5 taking place (Reipurth, 2000). If the ejection leads to an to 1 km s–1), and again the velocity dispersions of brown escape, then the accretion halts and the final mass of the dwarfs and H-burning stars are hard to distinguish, because object is capped (Klessen et al., 1998). they are partially masked by the velocity dispersion between the different cores in which small subgroups (<5) of stars 7.3. Dynamical Ejection as a Source of are born. For the same reasons, segregation of brown dwarfs Brown Dwarfs from H-burning stars may be difficult to detect. Moreover, given their low ejection , some brown dwarfs formed If a protostellar embryo is ejected from its natal core with by ejection retain significant disks (see Fig. 2), and are there- a mass below the H-burning limit, then it becomes a brown fore presumably well able to sustain accretion and outflows, dwarf (Reipurth and Clarke, 2001). All that is needed for as observed (see section 2.3 and chapter by Luhman et al.). this to happen is for a prestellar core to spawn more than Thus kinematical and spatial information on brown dwarfs, two protostellar embryos; for at least one of them to be less and the signatures of accretion and outflow, provide less- massive than 0.075 M at the outset; and for one of them to powerful constraints on brown dwarf formation than were stay less than 0.075 M long enough to be ejected by dy- initially surmised by Reipurth and Clarke (2001). namical interaction with the other embryos. Simulations suggest that forming more than two proto- 7.4. Dynamical Interactions in Early stellar embryos in a collapsing core is routine, as is the ejec- tion of brown dwarfs and very-low-mass stars (e.g., Bate et al., 2002a; Delgado-Donate et al., 2003, 2004; Goodwin The binary statistics of brown dwarfs provide another po- et al., 2004a,b). However, these simulations may be mis- tential constraint on formation mechanisms (see section 2.2 leading: (1) They use sink particles (Bate et al., 1995), and and the chapters by Burgasser et al., Duchêne et al., Good- 472 Protostars and Planets V

Reipurth (2000) notes that the different sizes and separa- tions of shocks in Herbig-Haro flows can be understood as a fossil record of the dynamical evolution of a newly formed binary. He also notes that dynamical interactions may on occasion lead to a departure from the standard evolution- ary picture of a star passing smoothly from the Class 0 stage through the Class III stage. Instead, stochastic dynamical interactions may lead to the sudden ejection of an object from the Class 0 or I stage, resulting in its abrupt appear- ance as a Class II or even a Class III object. The infrequent young binaries with infrared companions may be related to such events. Finally, the FU Orionis eruptions may be re- lated to the formation of a close binary, and could result from the viscous interactions of circumstellar material as the two components in a newly formed binary spiral together (Bon- nell and Bastien, 1992; Reipurth and Aspin, 2004).

8. PHOTOEROSION OF PRE-EXISTING CORES

A fifth — and somewhat separate — mechanism for Fig. 2. Four stars formed in a low-mass core (~5 M ) with a low forming brown dwarfs is to start with a pre-existing core of initial level of turbulence (ETURB ~ 0.05 |EGRAV|) (Goodwin et al., standard mass (i.e.,

Fig. 3. Two sequences illustrating the brown dwarf formation mechanisms that occur in the simulations of Bate et al. (2002a,b, 2003). The upper sequence shows two brown dwarfs (square and triangle) forming in a circumbinary disk, while the lower sequence shows two brown dwarfs (square and triangle) forming in a filament. In both cases, these objects remain as brown dwarfs because they are dynamically ejected before they can accrete sufficient mass to ignite H-burning. In the upper case, they are ejected from a multiple system formed by disk fragmentation. In the lower case, the two brown dwarfs form in separate filaments, but then fall into, and are ejected from, the multiple system that exists at the intersection of the two filaments. Each panel is 600 AU across. 474 Protostars and Planets V support the ejection hypothesis for the origin of brown by stars and planets. In hotter environments and at earlier dwarfs, with disk fragmentation also helping to generate dy- epochs, MMIN was probably larger, and brown dwarfs were namically unstable systems. therefore less common. We also suggest that the thermody- The brown dwarfs formed in these simulations do not namics of disks make it easier for proto brown dwarfs to frequently have companions. With three individual calcu- condense out at large radii >100 AU, and that these proto lations now published, the overall frequency of BD-BD and brown dwarfs may fragment to produce close BD-BD bi- very-low-mass binaries is ~5%. Most of these systems are naries, due to the dissociation of H2 and the opacity gap at closer than ~20 AU. Sun-like stars with brown dwarf com- T ~ 2000 K. panions at separations less than 10 AU are also very rare, We have discussed five possible mechanisms for form- at ~2%. Wider brown dwarf companions at hundreds or ing brown dwarfs. Turbulent fragmentation of molecular thousands of AU are much more common, although none clouds may deliver prestellar cores of such low mass that of these systems have reached dynamic stability when the their subsequent collapse (and possible fragmentation) can simulations are terminated. Most of these results are con- only yield brown dwarfs. The collapse and fragmentation sistent with current observations, with the exception of the of more massive cores is likely to deliver protostellar em- BD-BD binary frequency, where the value from the simu- bryos with a wide range of masses, many of which are too lations is about three times lower than the observed value low to support hydrogen burning. Similarly, disk fragmen- (~15%). There are at least two possible reasons for this. tation is likely to deliver low-mass protostellar embryos. First, most observed and calculated BD-BD binaries have These protostellar embryos may undergo competitive ac- separations <20 AU, but the simulations are unable to re- cretion (as a result of which some of them evolve to higher solve disks <10 AU, so the discrepancy might be solved mass) and dynamical interactions (as a result of which some with better resolution. If this is not the case, missing physics of them are ejected before they reach H-burning masses). may be the problem (e.g., the effects of radiative transfer Lastly, cores that find themselves overrun by an HII region on disk fragmentation). Finally, only one brown dwarf that may be photoeroded by the resulting ionization front and was ejected during any of the simulations had a resolved end up spawning brown dwarfs. None of these mechanisms disk (radius R > 10 AU). This implies that brown dwarfs is mutually exclusive, and in the most advanced simulations formed from highly turbulent cores should only have small of cluster formation, collapse and fragmentation, disk frag- disks. In contrast, the simulations of cores with low levels mentation, competitive accretion, and dynamical ejection of turbulence performed by Goodwin et al. (2004a,b) pro- all occur concurrently. However, these simulations do not duce brown dwarfs with somewhat larger disks (see Fig. 2). capture all the deterministic physics. It requires a fully radi- This indicates that disk size may be a function of birth en- ative, effectively inviscid, three-dimensional magnetohydro- vironment. Although many brown dwarfs are observed to dynamical simulation to evaluate properly the thermal ef- have disks from their spectral energy distributions, the dis- fects that influence the minimum mass for star formation, tribution of their sizes is not currently known. the angular momentum transport processes that influence the binary statistics, and the N-body dynamics that influence 10. CONCLUSIONS the clustering properties of brown dwarfs. Work to develop and validate such codes is ongoing. Since the statistical properties of brown dwarfs appear We believe that all the proposed mechanisms operate in to form a continuum with those of low-mass H-burning nature, and that once they have been properly modeled, the stars, we have argued that brown dwarfs form like stars, i.e., ultimate task will be to determine their relative contribu- on a dynamical timescale and by gravitational instability, tions to the overall brown dwarf population. These relative with an homogeneous initial elemental composition, the contributions may depend on environment, metallicity, and same as the from which they form. In , and may therefore lead to local and/or cosmological this regard brown dwarfs are distinct from planets, which variations in the ratio of brown dwarfs to H-burning stars, we define to be objects that form on a longer timescale, by and in the binary and accretion statistics of brown dwarfs. the accumulation of a rocky core and — if circumstances allow — the subsequent acquisition of a gaseous envelope, REFERENCES leading to an initially fractionated elemental composition and a deficit of light elements. André Ph., Ward-Thompson D., and Barsony M. (1993) Astrophys. J., We have evaluated the minimum mass for a brown dwarf 406, 122–141. by considering the cooling required for a very-low-mass André Ph., Ward-Thompson D., and Barsony M. (2000) In Protostars and Planets IV (V. Mannings et al., eds.), pp. 59–96. Univ. of Arizona, prestellar core to condense out on a dynamical timescale. Tucson. We have treated several different scenarios: hierarchical Anosova J. P. (1986) Astrophys. Space Sci., 124, 217–241. three-dimensional fragmentation; one-shot two-dimensional Armitage P. J. and Clarke C. J. (1997) Mon. Not. R. Astron. Soc., 285, fragmentation of a shock-compressed layer; and fragmenta- 540–546. tion of a Toomre-unstable disk. All three cases yield values Arquilla R. and Goldsmith P. F. (1985) Astrophys. J., 297, 436–454. Bacmann A., André Ph., Puget J.-L., Abergel A., Bontemps S., and Ward- of MMIN in the range 0.001–0.004 M for contemporary Thompson D. (2000) Astron. Astrophys., 361, 555–580. star formation in the solar vicinity. This suggests that there Barranco J. A. and Goodman A. A. (1998) Astrophys. J., 504, 207–222. may be some overlap between the range of masses occupied Basri G. and Martín E. L. (1999) Astron. J., 118, 2460–2465. Whitworth et al.: The Formation of Brown Dwarfs 475

Bastien P. (1983) Astron. Astrophys., 119, 109–116. Duquennoy A. and Mayor M. (1991) Astron. Astrophys., 248, 485–524. Bastien P., Arcoragi J.-P., Benz W., Bonnell I. A., and Martel H. (1991) Elmegreen B. G. (2000) Astrophys. J., 530, 277–281. Astrophys. J., 378, 255–265. Fernández M. and Comerón F. (2001) Astron. Astrophys., 380, 264–276. Bate M. R. (2005) Mon. Not. R. Astron. Soc., 363, 363–378. Gammie C. F. (2001) Astrophys. J., 553, 174–183. Bate M. R. and Bonnell I. A. (2005) Mon. Not. R. Astron. Soc., 356, Gizis J. E., Kirkpatrick J. D., Burgasser A., Reid I. N., Monet D. G., et al. 1201–1221. (2001) Astrophys. J., 551, L163–L166. Bate M. R., Bonnell I. A., and Price N. M. (1995) Mon. Not. R. Astron. Gizis J. E., Reid I. N., Knapp G. R., Liebert J., Kirkpatrick J. D., et al. Soc., 277, 362–376. (2003) Astron. J., 125, 3302–3310. Bate M. R., Bonnell I. A., and Bromm V. (2002a) Mon. Not. R. Astron. Goodman A. A., Benson P. J., Fuller G. A., and Myers P. C. (1993) Astro- Soc., 332, L65–L68. phys. J., 406, 528–547. Bate M. R., Bonnell I. A., and Bromm V. (2002b) Mon. Not. R. Astron. Goodwin S. P. and Kroupa P. (2005) Astron. Astrophys., 439, 565–569. Soc., 336, 705–713. Goodwin S. P., Ward-Thompson D., and Whitworth A. P. (2002) Mon. Bate M. R., Bonnell I. A., and Bromm V. (2003) Mon. Not. R. Astron. Not. R. Astron. Soc., 330, 769–771. Soc., 339, 577–599. Goodwin S. P., Whitworth A. P., and Ward-Thompson D. (2004a) Astron. Béjar V. J. S., Martín E. L., Zapatero Osorio M. R., Rebolo R., Barrado Astrophys., 414, 633–650. y Navacués D., et al. (2001) Astrophys. J., 556, 830–836. Goodwin S. P., Whitworth A. P., and Ward-Thompson D. (2004b) Astron. Boffin H. M. J.,Watkins S. J., Bhattal A. S., Francis N., and Whitworth Astrophys., 419, 543–547. A. P. (1998) Mon. Not. R. Astron. Soc., 300, 1189–1204. Goodwin S. P., Whitworth A. P., and Ward-Thompson D. (2004c) Astron. Bonnell I. A. (1994) Mon. Not. R. Astron. Soc., 269, 837–848. Astrophys., 423, 169–182. Bonnell I. A. and Bastien P. (1992) Astrophys. J., 401, L31–L34. Goodwin S. P., Hubber D. A., Moraux E., and Whitworth A. P. (2005) Bonnell I. A. and Bate M. R. (1994a) Mon. Not. R. Astron. Soc., 269, Astron. Nachr., 326, 1040–1043. L45–L48. Haisch K. E., Greene T. P., Barsony M., and Stahler S.W. (2004) Astron. Bonnell I. A. and Bate M. R. (1994b) Mon. Not. R. Astron. Soc., 271, J., 127, 1747–1754. 999–1004. Hayashi C. and Nakano T. (1963) Prog. Theor. Phys., 30, 460–474. Bonnell I. A., Bate M. R., Clarke C. J., and Pringle J. E. (1997) Mon. Not. Hennebelle P., Whitworth A. P., Cha S.-H., and Goodwin S. P. (2004) R. Astron. Soc., 285, 201–208. Mon. Not. R. Astron. Soc., 348, 687–701. Bonnell I. A., Bate M. R., Clarke C. J., and Pringle J. E. (2001a) Mon. Hester J. J., Scowen P. A., Sankrit R., Lauer T. R., Ajhar E. A., et al. Not. R. Astron. Soc., 323, 785–794. (1996) Astron. J., 111, 2349–2360. Bonnell I. A., Clarke C. J., Bate M. R., and Pringle J. E. (2001b) Mon. Hosking J. G. and Whitworth A. P. (2004) Mon. Not. R. Astron. Soc., 347, Not. R. Astron. Soc., 324, 573–579. 1001–1010. Boss A. P. (1986) Astrophys. J. Suppl., 62, 519–552. Hoyle F. (1953) Astrophys. J., 118, 513–528. Boss A. P. (1996) Astrophys. J., 468, 231–240. Hubber D. A. and Whitworth A. P. (2005) Astron. Astrophys., 437, 113– Boss A. P. (2002) Astrophys. J., 568, 743–753. 125. Boss A. P. (2004) Astrophys. J., 610, 456–463. Inutsuka S.-I. and Miyama S. M. (1992) Astrophys. J., 388, 392–399. Boss A. P., Fisher R. T., Klein R. I., and McKee C. F. (2000) Astrophys. Jayawardhana R., Ardila D. R., Stelzer B., and Haisch K. E. Jr. (2003) J., 528, 325–335. Astron. J., 125, 1515–1521. Bouy H., Brandner W., Martín E. L., Delfosse X., Allard F., and Basri G. Jiang I.-G., Laughlin G., and Lin D. N. C. (2004) Astron. J., 127, 455– (2003) Astron. J., 126, 1526–1554. 459. Bouy H., Martín E. L., Brandner W., and Bouvier J. (2005) Astron. J., Joergens V. (2006a) Astron. Astrophys., 446, 1165–1176. 129, 511–517. Joergens V. (2006b) Astron. Astrophys., 448, 655–663. Boyd D. F. A. and Whitworth A. P. (2004) Astron. Astrophys., 430, 1059– Joergens V., Fernández M., Carpenter J. M., and Neuhäuser R. (2003) 1066. Astrophys. J., 594, 971–981. Briceño C., Luhman K. L., Hartmann L., Stauffer J. R., and Kirkpatrick Johnson B. M. and Gammie C. F. (2003) Astrophys. J., 597, 131–141. J. D. (2002) Astrophys. J., 580, 317–335. Johnstone D., Wilson C. D., Moriarty-Schieven G., Giannakopoulou- Burgasser A. J., Kirkpatrick J. D., Reid I. N., Brown M. E., Miskey C. L., Creighton J., and Gregersen E. (2000) Astrophys. J. Suppl., 131, 505– and Gizis J. E. (2003a) Astrophys. J., 586, 512–526. 518. Burgasser A. J., Kirkpatrick J. D., Burrows A., Liebert J., Reid I. N., et al. Jones C. E., Basu S., and Dubinski J. (2001) Astrophys. J., 551, 387– (2003b) Astrophys. J., 592, 1186–1192. 393. Burgasser A. J., Kirkpatrick J. D., and Lowrance P. J. (2005) Astron. J., Kenyon M. J., Jeffries R. D., Naylor T., Oliveira J. M., and Maxted P. F. L. 129, 2849–2855. (2005) Mon. Not. R. Astron. Soc., 356, 89–106. Burkert A. and Bodenheimer P. (1996) Mon. Not. R. Astron. Soc., 280, Kirk H., Johnstone D., and Di Francesco J. (2006) Astrophys. J., 646, 1190–1200. 1009–1023. Burkert A. and Bodenheimer P. (2000) Astrophys. J., 543, 822–830. Kirk J. M., Ward-Thompson D., and André Ph. (2005) Mon. Not. R. Burkert A., Bate M. R., and Bodenheimer P. (1997) Mon. Not. R. Astron. Astron. Soc., 360, 1506–1526. Soc., 289, 497–504. Klessen R. S. and Burkert A. (2000) Astrophys. J. Suppl., 128, 287–319. Cai K., Durisen R. H., S., Boley A. C., Mejía A. C., Pickett M. K., Klessen R. S. and Burkert A. (2001) Astrophys. J., 549, 386–401. and D’Alessio P. (2006) Astrophys. J., 636, L149–L152. Klessen R. S., Burkert A., and Bate M. R. (1998) Astrophys. J., 501, Cha S.-H. and Whitworth A. P. (2003) Mon. Not. R. Astron. Soc., 340, L205–L208. 91–104. Koresko C. D. (2002) Astron. J., 124, 1082–1088. Chabrier G. (2003) Publ. Astron. Soc. Pac., 115, 763–795. Kroupa P. and Bouvier J. (2003) Mon. Not. R. Astron. Soc., 346, 369– Chapman S. J., Pongracic H., Disney M. J., Nelson A. H., Turner J. A., 380. and Whitworth A. P. (1992) Nature, 359, 207–210. Krumholz M. R., McKee C. F., and Klein R. I. (2005a) Nature, 438, 332– Close L. M., Siegler N., Freed M., and Biller B. (2003) Astrophys. J., 334. 587, 407–422. Krumholz M. R., McKee C. F., and Klein R. I. (2005b) Astrophys. J., 618, Curry C. L. (2002) Astrophys. J., 576, 849–859. 757–768. Delgado-Donate E. J., Clarke C. J., and Bate M. R. (2003) Mon. Not. R. Kumar S. S. (1963) Astrophys. J., 137, 1121–1125. Astron. Soc., 342, 926–938. Larson R. B. (1981) Mon. Not. R. Astron. Soc., 194, 809–826. Delgado-Donate E. J., Clarke C. J., and Bate M. R. (2004) Mon. Not. R. Larson R. B. (1985) Mon. Not. R. Astron. Soc., 214, 379–398. Astron. Soc., 347, 759–770. Laughlin G. and Bodenheimer P. (1994) Astrophys. J., 436, 335–354. 476 Protostars and Planets V

Lin C. C., Mestel L., and Shu F. H. (1965) Astrophys. J., 142, 1431–1446. Preibisch T., McCaughrean M. J., Grosso N., Feigelson E. D., Flaccomio Looney L. W., Mundy L. G., and Welch W. J. (2000) Astrophys. J., 529, E., et al. (2005) Astrophys. J. Suppl., 160, 582–593. 477–498. Rafikov R. R. (2005) Astrophys. J., 621, L69–L72. Low C. and Lynden-Bell D. (1976) Mon. Not. R. Astron. Soc., 176, 367– Rebolo R., Zapatero Osorio M. R., and Martín E. L. (1995) Nature, 377, 390. 129–131. Lucas P. W. and Roche P. F. (2000) Mon. Not. R. Astron. Soc., 314, 858– Rees M. J. (1976) Mon. Not. R. Astron. Soc., 176, 483–486. 864. Reid M. A. and Wilson C. D. (2005) Astrophys. J., 625, 891–905. Lucas P. W., Roche P. F., and Tamura M. (2005) Mon. Not. R. Astron. Reipurth B. (2000) Astron. J., 120, 3177–3191. Soc., 361, 211–232. Reipurth B. and Aspin C. (2004) Astrophys. J., 608, L65–L68. Luhman K. L. (2004) Astrophys. J., 617, 1216–1232. Reipurth B. and Clarke C. J. (2001) Astron. J., 122, 432–439. Luhman K. L., Stauffer J. R., Muench A. A., Rieke G. H., Lada E. A., Reipurth B., Rodríguez L. F., Anglada G., and Bally J. (2002) Astron. J., Bouvier J., and Lada C. J. (2003) Astrophys. J., 593, 1093–1115. 124, 1045–1053. Marcy G. W. and Butler R. P. (2000) Publ. Astron. Soc. Pac., 112, 137– Reipurth B., Rodríguez L. F., Anglada G., and Bally J. (2004) Astron. J., 140. 127, 1736–1746. Masunaga H. and Inutsuka S. (1999) Astrophys. J., 510, 822–827. Rice W. K. M., Armitage P. J., Bate M. R., and Bonnell I. A. (2003) Mon. Maxted P. and Jeffries R. (2005) Mon. Not. R. Astron. Soc., 362, L45– Not. R. Astron. Soc., 339, 1025–1030. L49. Ryden B. S. (1996) Astrophys. J., 471, 822–831. McCarthy C. and Zuckerman B. (2004) Astron. J., 127, 2871–2884. Salpeter E. E. (1955) Astrophys. J., 121, 161–167. McCaughrean M. J. and Andersen M. (2002) Astron. Astrophys., 389, Sandell G. and Knee L. B. G. (2001) Astrophys. J., 546, L49–L52. 513–518. Scholz A. and Eislöffel J. (2004) Astron. Astrophys., 419, 249–267. McCaughrean M. J., Zinnecker H., Anderson M., Meeus G., and Lodieu Slesnick C. L., Hillenbrand L. A., and Carpenter J. M. (2004) Astrophys. N. (2002) The Messenger, 109, 28–36. J., 610, 1045–1063. Mikkola S. and Valtonen M. J. (1986) Mon. Not. R. Astron. Soc., 223, Stamatellos D., Whitworth A. P., Boyd D. F. A., and Goodwin S. P. (2005) 269–278. Astron. Astrophys., 439, 159–169. Mohanty S., Jayawardhana R., Natta A., Fujiyoshi T., Tamura M., and Standish E. M. (1972) Astron. Astrophys., 21, 185–191. Barrado y Navascués D. (2004) Astrophys. J., 609, L33–L36. Stassun K. G., Mathieu R. D., and Valenti J. A. (2006) Nature, 440, 311. Monaghan J. J. (1976) Mon. Not. R. Astron. Soc., 176, 63–72. Sterzik M. F. and Durisen R. H. (1995) Astron. Astrophys., 304, L9–L12. Moraux E., Bouvier J., Stauffer J. R., and Cuillandre J.-C. (2003) Astron. Sterzik M. F. and Durisen R. H. (1998) Astron. Astrophys., 339, 95–112. Astrophys., 400, 891–902. Sterzik M. F. and Durisen R. H. (2003) Astron. Astrophys., 400, 1031– Motte F. and André Ph. (2001) Astron. Astrophys., 365, 440–464. 1042. Motte F., André Ph., and Neri R. (1998) Astron. Astrophys., 336, 150– Tachihara K., Onishi T., Mizuno A., and Fukui Y. (2002) Astron. Astro- 172. phys., 385, 909–920. Motte F., André Ph., Ward-Thompson D., and Bontemps S. (2001) Astron. Teixeira P. S., Lada C. J., Young E. T., Marengo M., Muench A., et al. Astrophys., 372, L41–L44. (2006) Astrophys. J., 636, L45–L48. Muench A. A., Alves J., Lada C. J., and Lada E. A. (2001) Astrophys. J., Testi L. and Sargent A. I. (1998) Astrophys. J., 508, L91–L94. 558, L51–L54. Tokovinin A. A. (2004) In The Environment and Evolution of Double and Muzerolle J., Hillenbrand L., Calvet N., Briceño C., and Hartmann L. Multiple Stars (C. Allen and C. Scarfe, eds.), pp. 7–14. IAU Collo- (2003) Astrophys. J., 592, 266–281. quium 191, Rev. Mex. Astron. Astrofis. Ser. Conf. 21. Muzerolle J., Luhman K. L., Briceño C., Hartman L., and Calvet N. Tokovinin A. A. and Smekhov M. G. (2002) Astron. Astrophys., 382, 118– (2005) Astrophys. J., 625, 906–912. 123. Myers P. C. (2005) Astrophys. J., 623, 280–290. Toomre A. (1964) Astrophys. J., 139, 1217–1238. Myers P. C. and Benson P. J. (1983) Astrophys. J., 266, 309–320. Turner J. A., Chapman S. J., Bhattal A. S., Disney M. J., Pongracic H., Myers P. C. and Gammie C. F. (1999) Astrophys. J., 522, L141–L144. and Whitworth A. P. (1995) Mon. Not. R. Astron. Soc., 277, 705–726. Myers P. C., Fuller G. A., Goodman A. A., and Benson P. J. (1991) Astro- Umbreit S., Burkert A., Henning T., Mikkola S., and Spurzem R. (2005) phys. J., 376, 561–572. Astrophys. J., 623, 940–951. Nakajima T., Oppenheimer B. R., Kulkarni S. R., Golimowski D. A., Mat- van Albada T. S. (1968) Bull. Astron. Inst. Netherlands, 19, 479–499. thew K., and Durrance S. T. (1995) Nature, 378, 463–465. Watkins S. J., Bhattal A. S., Boffin H. M. J., Francis N., and Whitworth Natta A. and Testi L. (2001) Astron. Astrophys., 376, L22–L25. A. P. (1998a) Mon. Not. R. Astron. Soc., 300, 1205–1213. Natta A., Testi L., Muzerolle J., Randich S., Comerón F., and Persi P. Watkins S. J., Bhattal A. S., Boffin H. M. J., Francis N., and Whitworth (2004) Astron. Astrophys., 424, 603–612. A. P. (1998b) Mon. Not. R. Astron. Soc., 300, 1214–1224. Nelson A. F., Benz W., Adams F. C., and Arnett D. (1998) Astrophys. J., Whelan E. T., Ray T. P., Bacciotti F., Natta A., Testi L., and Randich S. 502, 342–371. (2005) Nature, 435, 652–654. Nordlund Å. and Padoan P. (2003) In Turbulence and Magnetic Fields in Whitehouse S. C. and Bate M. R. (2006) Mon. Not. R. Astron. Soc., 367, (E. Falgarone and T. Passot, eds.), pp. 271–298. LNP 32. Vol. 614, Springer, Berlin. Whitworth A. P. (2005) In Cores to Clusters (M. S. N. Kumar et al., eds.), Onishi T., Mizuno A., Kawamura A., Tachihara K., and Fukui Y. (2002) pp. 15–29. Astrophys. Space Sci. Library Ser. 324, Kluwer, Dordrecht. Astrophys. J., 575, 950–973. Whitworth A. P. and Zinnecker H. (2004) Astron. Astrophys., 427, 299– Oppenheimer B. R., Kulkarni S. R., Matthews K., and Nakajima T. (1995) 306. , 270, 1478–1479. Whitworth A. P., Bhattal A. S., Chapman S. J., Disney M. J., and Turner Padoan P. and Nordlund Å. (2002) Astrophys. J., 576, 870–879. J. A. (1994a) Mon. Not. R. Astron. Soc., 268, 291–298. Padoan P. and Nordlund Å. (2004) Astrophys. J., 617, 559–564. Whitworth A. P., Bhattal A. S., Chapman S. J., Disney M. J., and Turner Padoan P., Kritsuk A., Norman M. L., and Nordlund Å. (2005) Astrophys. J. A. (1994b) Astron. Astrophys., 290, 421–427. J., 622, L61–L64. Whitworth A. P., Chapman S. J., Bhattal A. S., Disney M. J., Pongracic H., Peng R., Langer W. D., Velusamy W. T., Kuiper T. B. H., and Levin S. and Turner J. A. (1995) Mon. Not. R. Astron. Soc., 277, 727–746. (1998) Astrophys. J., 497, 842–849. Zapatero Osorio M. R., Béjar V. J. S., Martiín E. L., Rebolo R., Barrado y Pinfield D. J., Dobbie P. D., Jameson R. F., Steele I. A., Jones H. R. A., Navasqués D., et al. (2002) Astrophys. J., 578, 536–542. and Katsiyannis A. C. (2003) Mon. Not. R. Astron. Soc., 342, 1241– Zinnecker H. (1982) N.Y. Acad. Sci. Ann., 395, 226–235. 1259. Zinnecker H. (1984) Mon. Not. R. Astron. Soc., 210, 43–56.