Plan
Comptes Rendus

Internal Geophysics (Space Physics)
Interstellar filaments and star formation
Comptes Rendus. Géoscience, Volume 349 (2017) no. 5, pp. 187-197.

Résumé

Recent studies of the nearest star-forming clouds of the Galaxy at submillimeter wavelengths with the Herschel Space Observatory have provided us with unprecedented images of the initial conditions and early phases of the star formation process. The Herschel images reveal an intricate network of filamentary structure in every interstellar cloud. These filaments all exhibit remarkably similar widths – about a tenth of a parsec – but only the densest ones contain prestellar cores, the seeds of future stars. The Herschel results favor a scenario in which interstellar filaments and prestellar cores represent two key steps in the star formation process: first turbulence stirs up the gas, giving rise to a universal web-like structure in the interstellar medium, then gravity takes over and controls the further fragmentation of filaments into prestellar cores and ultimately protostars. This scenario provides new insight into the origin of stellar masses and the star formation efficiency in the dense molecular gas of galaxies. Despite an apparent complexity, global star formation may be governed by relatively simple universal laws from filament to galactic scales.

Métadonnées
Reçu le :
Accepté le :
Publié le :
DOI : 10.1016/j.crte.2017.07.002
Mots clés : Stars formation, ISM clouds, ISM filaments, ISM structure, Submillimeter

Philippe André 1

1 Département d'astrophysique (AIM), CEA Saclay, 91191 Gif-sur-Yvette, France
@article{CRGEOS_2017__349_5_187_0,
     author = {Philippe Andr\'e},
     title = {Interstellar filaments and star formation},
     journal = {Comptes Rendus. G\'eoscience},
     pages = {187--197},
     publisher = {Elsevier},
     volume = {349},
     number = {5},
     year = {2017},
     doi = {10.1016/j.crte.2017.07.002},
     language = {en},
}
TY  - JOUR
AU  - Philippe André
TI  - Interstellar filaments and star formation
JO  - Comptes Rendus. Géoscience
PY  - 2017
SP  - 187
EP  - 197
VL  - 349
IS  - 5
PB  - Elsevier
DO  - 10.1016/j.crte.2017.07.002
LA  - en
ID  - CRGEOS_2017__349_5_187_0
ER  - 
%0 Journal Article
%A Philippe André
%T Interstellar filaments and star formation
%J Comptes Rendus. Géoscience
%D 2017
%P 187-197
%V 349
%N 5
%I Elsevier
%R 10.1016/j.crte.2017.07.002
%G en
%F CRGEOS_2017__349_5_187_0
Philippe André. Interstellar filaments and star formation. Comptes Rendus. Géoscience, Volume 349 (2017) no. 5, pp. 187-197. doi : 10.1016/j.crte.2017.07.002. https://comptes-rendus.academie-sciences.fr/geoscience/articles/10.1016/j.crte.2017.07.002/

Version originale du texte intégral

1 Introduction

Star formation is one of the most complex processes in astrophysics, involving a subtle interplay between gravity, turbulence, magnetic fields, feedback mechanisms, heating and cooling effects (McKee and Ostriker, 2007), etc. Yet, despite this apparent complexity, the net products of the star formation process on global scales are relatively simple and robust. In particular, the distribution of stellar masses at birth or stellar initial mass function (IMF) is known to be quasi-universal (Bastian et al., 2010). Likewise, the star formation rate on both interstellar cloud and galaxy-wide scales is related to the mass of dense molecular gas available by rather well defined “star formation laws” (Lada et al., 2010, 2012; Shimajiri et al., 2017).

This paper presents an overview of recent observational results obtained with the Herschel Space Observatory (Pilbratt et al., 2010) and other facilities on the texture of nearby star-forming clouds, which suggest that it may be possible to explain, at least partly, the IMF and the global star formation efficiency in the Galaxy in terms of the quasi-universal filamentary structure of the cold interstellar medium (ISM) out of which stars form. Interestingly, the filamentary web of cold interstellar gas responsible for star formation within galaxies bears a marked resemblance to the cosmic web of intergalactic gas and dark matter leading to galaxy formation in cosmological simulations (cf. Freundlich et al., 2014; Springel et al., 2005 and Section 4 below).

2 Summary of Herschel results supporting a filamentary paradigm for star formation

Herschel imaging surveys of Galactic star-forming regions have confirmed the ubiquitousness of filaments in Galactic molecular clouds (see Fig. 1) and suggested an intimate connection between the filamentary structure of the ISM and the star formation process (André et al., 2010; Molinari et al., 2010). While molecular clouds have been known to exhibit large-scale filamentary structures for quite some time (Schneider and Elmegreen, 1979; Johnstone and Bally, 1999; Myers, 2009 and Myers, 2009), Herschel observations now demonstrate that these filaments are truly ubiquitous in the cold ISM (Hill et al., 2011; Men'shchikov et al., 2010; Wang et al., 2015), probably make up a dominant fraction of the dense gas in molecular clouds (Schisano et al., 2014; Könyves et al., 2015), present a high degree of universality in their properties (Arzoumanian et al., 2011), and are the preferred birthplaces of prestellar cores (Könyves et al., 2015; Marsh et al., 2016). This means that interstellar filaments probably play a central role in the star formation process (André et al., 2014).

Fig. 1

Herschel/SPIRE 250 μm dust continuum image of a portion of the Polaris flare translucent cloud (d ∼ 150 pc) taken as part of the HGBS survey (e.g., André et al., 2010; Miville-Deschênes et al., 2010; Ward-Thompson et al., 2010). Note the widespread filamentary structure.

2.1 A characteristic filament width

Detailed analysis of the radial column density profiles derived from Herschel submillimeter dust continuum data shows that, at least in the nearby clouds of the Gould Belt, filaments are characterized by a very narrow distribution of inner widths W with a typical FWHM value ∼ 0.1 pc (much higher than the ∼ 0.01 pc resolution provided by Herschel at the distance ∼ 140 pc of the nearest clouds) and a dispersion lower than a factor of 2 (Arzoumanian et al., 2011; Koch and Rosolowsky, 2015; Palmeirim et al., 2013). Independent measurements of filament widths have generally been consistent with this Herschel finding when performed in submillimeter continuum emission. For instance, Salji et al. (2015) found an averaged deconvolved FWHM width of 0.080.03+0.07 pc for 28 filaments detected at 850 μm with SCUBA-2. It should be pointed out, however, that significant variations around the mean inner width of ∼ 0.1 pc (by up to a factor of ∼ 2 on either side) sometimes exist along the main axis of a given filament (Juvela et al., 2012; Ysard et al., 2013). We also note that the report by Henshaw et al. (2017) of a typical width of ∼ 0.03 pc for filamentary structures observed in the 1.1-mm dust continuum with ALMA toward the infrared dark cloud G035.39–00.33 may be affected by interferometric filtering of large-scale emission.

Recently, Panopoulou et al. (2017) raised the issue of whether the existence a characteristic filament width is consistent with the scale-free nature of the power spectrum of interstellar cloud images (well described by a single power law from ∼ 0.01 pc to ∼ 50 pc – Miville-Deschênes et al., 2010, 2016). They suggested that the filament widths obtained by Gaussian fitting may be strongly correlated with the range of radii over which the filament profiles are fitted and may not indicate a genuine characteristic width. This suggestion is questionable for the following reasons. First, simulations indicate that the power spectra of synthetic cloud images including populations of simple model filaments, all 0.1 pc in diameter, remain consistent with the observed ‘scale-free’ power spectra for realistic distributions of filament contrasts over the background (A. Roy et al., in prep.). Second, in the case of dense filaments whose radial profiles often feature power-law wings, a typical inner diameter of ∼ 0.1 pc is also found using a Plummer-like model function of the form Npr=NH20/1+r/Rflat2p12, which provides a much better fit to the overall radial profiles than a simple Gaussian model (see Fig. 2 adapted from Palmeirim et al., 2013). Third, the range of radii used for the Gaussian fitting needs not be arbitrarily fixed, but can be adjusted depending on an initial estimate of the background level at each point along the crest of a filament (André et al., 2016). Finally, it is the physical inner width (in pc) that remains approximately constant from filament to filament in Herschel observations of nearby clouds, while the angular width (in arcsec) scales roughly as the inverse of the parent cloud's distance (Arzoumanian, 2012). Although the width estimates reported for low-density filaments with a low contrast over the background are clearly more uncertain, we conclude that the median FWHM filament width of 0.09 ± 0.04 pc reported by (Arzoumanian et al., 2011 – see also André et al., 2014) is most likely reflecting the presence of a genuine characteristic scale corresponding to the inner diameter of filament systems, at least in the Gould Belt.

Fig. 2

Mean radial column density profile observed with Herschel perpendicular to the B213/B211 filament in Taurus, for both the northern (blue curve) and the southern part (red curve) of the filament. The yellow area shows the (± 1σ) dispersion of the distribution of radial profiles along the filament. The inner solid purple curve shows the effective 18″ HPBW resolution (0.012 pc at 140 pc) of the column density map used to construct the profile. The dashed black curve shows the best-fit Plummer-like model of the form Npr=NH20/1+r/Rflat2p12, which has a power-law index p = 2.0 ± 0.4 and a diameter 2 × Rflat = 0.07 ± 0.01 pc. It matches the data very well for r ≤ 0.4 pc.

Adapted from Palmeirim et al. (2013).

The existence of this characteristic scale is challenging for numerical simulations of interstellar cloud turbulence (Smith et al., 2014; Ntormousi et al., 2016), and the origin of the common inner width of interstellar filaments is currently debated. A possible interpretation is that filaments originate from planar intersecting shock waves due to supersonic interstellar turbulence (Pudritz and Kevlahan, 2013) and that the filament width corresponds to the (magneto-)sonic scale below which the turbulence becomes subsonic in diffuse, non-star-forming molecular gases (Federrath, 2016; Padoan et al., 2001). Alternatively, a characteristic width may arise if interstellar filaments are formed as quasi-equilibrium structures in pressure balance with a typical ambient ISM pressure Pext ∼ 2–5 × 104 K·cm−3 (Fischera and Martin, 2012; S. Inutsuka, private communication). A second alternative explanation relies on the thermodynamical properties of cold ISM gas (Hocuk et al., 2016; Larson, 2005) and connects the common inner width of Herschel filaments to the thermal Jeans length at the transition density nH2104105   cm3 at which the effective equation of state of the gas P ∝ ργeff changes from γeff < 1 (below the transition density) to γeff > 1 (above the transition density). Yet another possibility is that the characteristic inner width of filaments may be set by the dissipation mechanism of magnetohydrodynamic (MHD) waves due to ion-neutral friction (Hennebelle, 2013; Hennebelle and André, 2013; Ntormousi et al., 2016). Recently, Basu (2016) and Auddy et al. (2016) pointed out that the magnetized character of dense molecular filaments (see Section 6 below) implies that they cannot be true cylinders and that their 3D configuration should be closer to magnetic ribbons, i.e. triaxial objects flattened along the direction of the large-scale magnetic field. On this basis, they developed an analytic model in which the lateral width of the ribbon-shaped filaments is independent of density, but their thickness corresponds to the Jeans scale, and a relatively flat distribution of apparent filament widths results from projection effects assuming a random set of viewing angles.

2.2 Link between dense cores and filaments

Another major result from Herschel in nearby clouds is that most (> 75%) prestellar cores are found to lie within dense filaments with column densities NH2fil7×1021 cm2 (André et al., 2010; Könyves et al., 2015; Marsh et al., 2016 – see Figs. 3 and 4). Independent, albeit more indirect, evidence that prestellar cores are closely associated with filaments is provided by recent SCUBA-2 imaging observations that show that the spatial distribution of dense cores as traced by minimal spanning trees is itself filamentary (Kirk et al., 2013; Lane et al., 2017), and highly correlated with the dense filaments detected with Herschel (Könyves et al., 2015). Quantitatively, ⪆ 80% of the candidate prestellar cores identified with Herschel lie within a radius of < 0.1 pc of the crest of their nearest filament (Bresnahan et al., 2017; Könyves et al., 2015; Ladjelate et al., in prep.). The column density threshold above which prestellar cores are found in filaments is quite pronounced, as illustrated in Fig. 5 which shows the observed core formation efficiency CFEobs(AV) = ΔMcores(AV)/ΔMcloud(AV) as a function of the “background” column density of the parent filaments in the Aquila cloud complex. As can be seen, the plot of Fig. 5 resembles a smooth step function. The combination of this threshold with the typical column density probability function or distribution of cloud/filament mass as a function of column density (Kainulainen et al., 2009; Lada et al., 2010; Schneider et al., 2013) implies that most prestellar cores form just above the column density threshold (cf. Fig. 6).

Fig. 3

(a) Herschel/SPIRE 250-μm dust continuum image of the Taurus B211/B213 filament in the Taurus molecular cloud (Marsh et al., 2016; Palmeirim et al., 2013). Note the presence of several prominent dense cores along the filament; (b) IRAM/NIKA 1.25-mm dust continuum image of the central part of the Herschel field on the left (from Bracco et al., 2017), showing at least three prominent dense cores.

Fig. 4

Spatial distribution of prestellar cores (red diamonds) overlaid on a filtered column density map of the L1495 region in the Taurus molecular cloud, derived from Herschel Gould Belt survey data. To enhance the contrast of the filamentary structure, the original column density map was filtered using the multi-scale decomposition algorithm getfilaments (Men'shchikov, 2013). (Here, for better visualization of the filaments, only transverse spatial scales up to 0.1 pc were retained). Note how prestellar cores are only seen toward dense filamentary structures with mass per unit length in excess of ∼16 Mpc−1.

From Marsh et al. (2016).

Fig. 5

Plot of the differential core formation efficiency (CFE) in the Aquila complex as a function of background column density expressed in Av units (blue histogram with error bars). The CFE was obtained by dividing the mass in the form of prestellar cores in a given column density bin by the cloud mass observed in the same column density bin. The vertical dashed line marks a fiducial threshold at AVback7. The horizontal dotted line marks the rough asymptotic value ∼ 15% of the CFE at AV > 15.

From Könyves et al. (2015).

There is a natural interpretation of the column density threshold for prestellar core formation within filaments in terms of simple theoretical expectations for the gravitational instability of nearly isothermal gas cylinders. Given the typical width Wfil ∼ 0.1 pc measured for filaments and the relation Mline ≈ Σ0 × Wfil between the central gas surface density Σ0 and the mass per unit length Mline of a filament, the threshold at AVback7 or gasback150Mpc2 corresponds to within a factor of << 2 to the critical mass per unit length Mline,crit2cs2/G16Mpc1 of isothermal long cylinders in hydrostatic equilibrium for a sound speed cs ∼ 0.2 km/s, i.e. a typical gas temperature T ∼ 10 K (Ostriker, 1964). Therefore, as pointed out by André et al. (2010), the prestellar core formation threshold approximately corresponds to the line-mass threshold above which interstellar filaments become gravitationally unstable to radial contraction and fragmentation along their length (Inutsuka and Miyama, 1992, 1997).

Some departures from the predictions of standard cylinder fragmentation theory are observed, however. Linear fragmentation models for infinitely long, isothermal equilibrium cylinders predict a characteristic core spacing of 4× the filament diameter or ∼ 0.4 pc (Inutsuka and Miyama, 1992). In contrast, recent high-resolution studies (Kainulainen et al., 2013; Kainulainen et al., 2017; Takahashi et al., 2013; Teixeira et al., 2016) provide evidence of two-level hierarchical fragmentation within thermally supercritical filaments, with two fragmentation modes:

  • • a “cylindrical” fragmentation mode corresponding to groups of cores separated by ∼ 0.3 pc;
  • • a “spherical”, Jeans-like fragmentation mode corresponding to a typical spacing ≲ 0.1 pc between cores (and within groups).

Moreover, the second mode appears to dominate since the median projected spacing observed between prestellar cores (∼0.08 pc in Aquila – Könyves et al., 2015) is consistent with the thermal Jeans length within critical filaments of 0.1 pc inner diameter at 10 K.

These findings have stimulated recent theoretical work to try and account for two fragmentation modes based on modifications to the standard cylinder fragmentation theory. Clarke et al. (2016) explored the properties of non-equilibrium, accreting filaments and Gritschneder et al. (2017) investigated the role of geometrical effects (i.e. the fact that real filaments are not perfectly straight, but often exhibit some bends). While further work is still needed to get a perfect match with observations, it seems likely that a relatively simple model of filament fragmentation (including, for instance, the effects of low levels of turbulence – Clarke et al., 2017) may be able to explain the observed distribution of core spacings along filaments.

Overall, the observational results summarized in this section support a filamentary paradigm for solar-type star formation, in two main steps (André et al., 2014): first, large-scale compression of interstellar material in supersonic MHD flows generates a cobweb of filaments ∼ 0.1 pc in width in the ISM; second, the densest filaments fragment into prestellar cores (and subsequently protostars) by gravitational instability above Mline,crit. This paradigm differs from the classical “gravo-turbulent” fragmentation picture (Hennebelle and Chabrier, 2008; Mac Low and Klessen, 2004; Padoan and Nordlund, 2002) in that it relies on the unique properties of filamentary geometry, such as the existence of a critical mass per unit length for nearly isothermal filaments.

3 Merits of the filamentary paradigm of star formation

The filamentary paradigm of star formation sketched at the end of Section 2 has several merits. One of them is that it provides a simple explanation for the origin of the lognormal “base” of the stellar initial mass function (IMF), for stellar masses around ∼0.3 M (Bastian et al., 2010; Chabrier, 2005).

Indeed, as most prestellar cores appear to form in filaments just above the threshold for cylindrical gravitational instability, one expects that there should be a characteristic prestellar core mass corresponding to the local Jeans (or Bonnor–Ebert) mass in marginally critical filaments. The thermal Bonnor–Ebert mass within ∼ 0.1-pc-wide critical filaments at ∼ 10 K with Mline ≈ Mline,crit ∼ 16Mpc−1 and surface densities clgascrit160Mpc2 is MBE1.3cs4/G2cl0.5M. This is very close to the peak of the prestellar core mass function (CMF) at ∼0.6 M as observed in the Aquila cloud complex (Könyves et al., 2015 – cf. Fig. 7). Moreover, the observed prestellar CMF closely resembles the IMF, apart from a global shift in mass scale corresponding to an efficiency factor ɛcore ∼ 30–50% (Alves et al., 2007; Könyves et al., 2015), which can be attributed to mass loss due to the effects of outflows during the protostellar phase (Machida and Matsumoto, 2012; Matzner and McKee, 2000). The Herschel results therefore support the view that gravitational fragmentation of filaments is the dominant physical mechanism generating prestellar cores and stars around the peak of the CMF/IMF.

Fig. 6

Mass in the form of prestellar cores in the Aquila molecular cloud as a function of background column density in Av units. The peak seen at Av ∼ 10 implies that most prestellar cores form just above the column density threshold AVback10, in marginally supercritical filaments.

Adapted from Könyves et al. (2015).

Naively, one would expect gravitational fragmentation to result in a narrow prestellar CMF sharply peaked at the median thermal Jeans mass. However, a Salpeter-like power-law tail can quickly develop if turbulence has generated a field of initial density fluctuations within the filaments in the first place (Inutsuka, 2001). To investigate this possibility, Roy et al. (2015) carried out a statistical analysis of the line-mass fluctuations observed with Herschel along a sample of 80 subcritical or marginally supercritical filaments in three clouds of the Gould Belt. They found that the line-mass fluctuations are in the linear regime and that their power spectrum is well fit by a power law [P(k) ∝ kα with α = –1.6 ± 0.3], which is consistent with the 1D power spectrum generated by subsonic Kolmogorov turbulence (α  = −5/3). Starting from such an initial power spectrum of line-mass fluctuations, the theoretical analysis by Inutsuka (2001) shows that the density perturbations quickly evolve – in about two free-fall times or ∼ 0.5 Myr for a critical 0.1-pc-wide filament – from a mass distribution similar to that of CO clumps (Kramer et al., 1998) to a population of protostellar cores whose mass distribution approaches the Salpeter power law at the high-mass end. This occurs because small-scale perturbations (i.e. small core masses) grow faster than large-scale perturbations, resulting in a steepening of the mass distribution with time.

4 Formation of interstellar filaments: observational constraints

Broadly speaking, three classes of models have been proposed to explain the origin of interstellar filaments depending on whether gravity, supersonic turbulence, or magnetic fields play the primary role in generating the filamentary structure. When gravity dominates, e.g., in strongly Jeans-unstable clouds, it is known to amplify initial anisotropies, leading to large-scale collapse first along the shortest axis, creating sheets or pancakes, then along the second dimension producing filaments, and finally ending up with spheroidal clumps/cores (Lin et al., 1965; Zel’dovich, 1970). This is the favored mechanism to explain the growth of large-scale structure in the Universe and the formation of filaments in the cosmic web (Bond et al., 1996; Shandarin and Zeldovich, 1989), and it has been proposed to operate in molecular clouds as well (Burkert and Hartmann, 2004; Gómez and Vázquez-Semadeni, 2014). Observationally, there is little doubt that gravity is indeed the main player in strongly self-gravitating “hub-filament” systems, where a cluster-forming hub is observed at the center of a converging network of filaments (Myers, 2009). Such systems are quite spectacular in high-mass star-forming regions (e.g., MonR2: Didelon et al., 2015; Pokhrel et al., 2016; SDC335: Peretto et al., 2013), but also exist in clouds forming mostly (or only) low- to intermediate-mass stars (e.g., B59: Peretto et al., 2012; L1688: Ladjelate et al., in prep.). Clearly, however, filaments are already widespread in gravitationally unbound clouds, such as atomic (HI) clouds (Clark et al., 2014; Kalberla et al., 2016; McClure-Griffiths et al., 2006) or translucent molecular clouds (e.g., Polaris: see Fig. 1 – Men'shchikov et al., 2010; Miville-Deschênes et al., 2010), where the dominant role of gravity is difficult to invoke. At least in these clouds, the filamentary structure must arise from other mechanisms, and the role of large-scale supersonic flows (turbulent or not) is likely. There is growing evidence of the presence of large-scale supersonic streams in the nearby ISM, possibly resulting from the action of stellar winds (Bouy and Alves, 2015). Each of these large-scale supersonic flows generates a locally planar shock wave, which compresses interstellar matter. A filament is naturally produced at the intersection of two such planar shock waves (Pudritz and Kevlahan, 2013), or as a result of the focusing effect generated by a single, but deformed MHD shock wave (Inoue and Fukui, 2013). Direct evidence of the role of large-scale compressive flows has been found with Herschel in the Pipe Nebula in the form of filaments with asymmetric column density profiles, which most likely result from the compression by the winds of the neighboring Sco OB2 association (Peretto et al., 2012).

Recent molecular line studies of the projected velocity field observed toward several nearby filament systems set interesting constraints on the origin of dense star-forming filaments. Both velocity gradients along and perpendicular to the major axis of molecular filaments have been detected (Kirk et al., 2013; Palmeirim et al., 2013; Peretto et al., 2006), but the gradients across filaments often dominate (Fernández-López et al., 2014; Mundy et al., in prep.). Such transverse velocity gradients, sometimes seen consistently along the entire length of the filamentary structure (see Fig. 8), would not be observed if the filaments were embedded in (and accreting from) a spherical or cylindrical ambient cloud. They are most readily explained if the filaments are forming and growing inside sheet-like structures. Numerical MHD simulations of filament formation within shock-compressed layers generated by large-scale supersonic flows can reproduce this kinematic pattern (Chen and Ostriker, 2014; Inutsuka et al., 2015).

Fig. 7

Observed core mass function (dN/dlogM) for the 446 candidate prestellar cores (upper blue histogram and triangles), and 292 robust prestellar cores (lower blue histogram and triangles) identified with Herschel in the Aquila cloud (Könyves et al., 2015). The shaded area in light blue reflects the uncertainties in the prestellar CMF. Lognormal fits (red curves) and a power-law fit to the high-mass end of the CMF (black solid line) are superimposed.

Adapted from Könyves et al. (2015).

5 Evolution of molecular filaments: observational constraints

Locally at least, i.e. close to the filament axis, it appears that the transverse velocity gradients observed toward supercritical filaments are driven by gravity. Their magnitude is indeed roughly consistent with what is expected from the accretion of background cloud material given the gravitational potential of the central filaments (e.g., Mundy et al., in prep.; Palmeirim et al., 2013; Shimajiri et al., in prep.). In the case of the Taurus B211/B213 filament, the estimated accretion rate of 27 − 50 M/pc/Myr is such that it would take ∼ 1–2 Myr for the filament to double its mass (Palmeirim et al., 2013). Another indication that thermally supercritical filaments are accreting background cloud material is the detection in dust and/or CO emission of low-density striations or sub-filaments perpendicular to the main filaments, and apparently feeding them from the side. Examples include the Musca filament (Cox et al., 2016), the Taurus B211/B213 filament in Taurus (Goldsmith et al., 2008; Palmeirim et al., 2013), the Serpens-South filament in Aquila (e.g., Kirk et al., 2013), and the DR21 ridge in Cygnus X (Hennemann et al., 2012; Schneider et al., 2010). The conclusion that supercritical filaments are generally accreting mass from their parent cloud is independently supported by the fact that their internal velocity dispersion is observed to increase with central column density, which is suggestive of an increase in virial mass per unit length with time (Arzoumanian et al., 2013).

This accretion process likely plays a key role in effectively preventing dense filaments from collapsing to spindles and allowing them to maintain approximately constant inner widths while evolving. Accretion supplies gravitational energy to supercritical filaments, which is then converted into turbulent kinetic energy (Heitsch et al., 2009; Klessen and Hennebelle, 2010) and may explain the increase in velocity dispersion (σtot) with column density observed by Arzoumanian et al. (2013). The central diameter of such accreting filaments is expected to be on the order the effective Jeans length DJ,eff2σtot2/GΣ0, which Arzoumanian et al. (2013) have shown to remain close to ∼ 0.1 pc. Toy models of this process are discussed by Heitsch (2013) and Hennebelle and André (2013). Given the importance of magnetic fields in the formation and evolution of filaments (see Section 6 below), the “turbulence” driven in the interior of dense filaments by the accretion of background matter is likely to be MHD in nature rather than purely hydrodynamic. It is therefore expected to dissipate due to ion-neutral friction. As shown by Hennebelle and André (2013), a dynamical equilibrium can then be established between the supply and the dissipation of kinetic energy inside the filament, resulting in an almost constant inner width ∼ 0.1 pc, in good agreement with observations (see Fig. 9).

Fig. 8

Transverse velocity gradient observed in 12CO(1–0) across the Taurus B211/B213 filament in Taurus (Palmeirim et al., 2013), based on CO data from Goldsmith et al. (2008). Note how the redshifted CO emission (in red) is mostly seen to the north-east and the blueshifted CO emission (in blue) is mostly seen to the south-west of the filament (displayed in green).

A possible manifestation of accretion-driven MHD turbulence may have been detected in the form of substructure and velocity-coherent “fibers” in several star-forming filaments, such as Taurus B211/B213 (Hacar et al., 2013; Henshaw et al., 2016; Tafalla and Hacar, 2015 – see Fig. 10). Not all filaments consist of multiple fiber-like structures, however. The Musca filament, for instance, appears to be a 6-pc-long velocity-coherent sonic filament with much less substructure than the Taurus B211/B213 filament and no evidence for multiple velocity-coherent fibers (Cox et al., 2016; Hacar et al., 2016). In the context of the accretion-driven turbulence picture introduced earlier, Cox et al. (2016) proposed that the Musca cloud may represent an earlier evolutionary stage that the Taurus B211/B213 filament in which the main filament has not yet accreted sufficient mass and energy to develop a multiple system of filamentary sub-components. A thermally supercritical filament may indeed be expected to develop a more complex system of intertwined fibers as the filament system grows in mass per unit length and its internal velocity dispersion increases.

Fig. 9

Plot of filament width against filament central column density in the accretion-driven MHD turbulence model proposed by Hennebelle and André (2013) (blue curve). The data points with error bars correspond to width measurements for 278 nearby filaments based on Herschel observations (Arzoumanian et al., 2011). The black straight line running from top left to bottom right shows the thermal Jeans length as a function of column density and represents the expected dependence of filament width on density for self-gravitating filaments in the absence of accretion.

6 Role of magnetic fields

All-sky dust polarization observations with the Planck satellite at 850 μm have revealed very organized magnetic fields on large scales (> 1–10 pc) in interstellar clouds (Planck Collaboration Int. XIX, 2015 – see Fig. 11). Low-density filamentary structures tend to be aligned with the magnetic field, while dense star-forming filaments tend to be perpendicular to the magnetic field (Planck Collab. Int. XXXII, 2016; Planck Collab. Int. XXXV, 2016), a trend also seen in optical and near-IR polarization observations (Chapman et al., 2011; Palmeirim et al., 2013; Panopoulou et al., 2016; Soler et al., 2016). This trend has been quantified by the Planck Collaboration using the Histogram of Relative Orientations (HRO) and the unprecedented statistics provided by Planck dust polarization observations in nearby molecular clouds (Planck Collab. Int. XXXV, 2016). The HRO (Soler et al., 2013) is a statistical tool that measures the relative orientation between the plane-of-sky magnetic field B and the local column density gradient. Using this technique, Soler et al. (2013) were able to show that a switch occurs at log10NH/cm2=21.9±0.2 between filamentary structures preferentially parallel to the magnetic field (low NH) and filamentary structures preferentially perpendicular to the magnetic field (high NH). Expressed in terms of H2 column densities, and taking into account the fact that Planck underestimates the central column density of narrow ∼ 0.1 pc filaments due to beam dilution, the switch occurs at NH27×1021cm2, which corresponds to the critical mass per unit length for 10 K gas filaments of ∼ 0.1 pc width.

Fig. 10

Fine (column) density structure of the B211/B213 filament based on a filtered version of the Herschel 250 μm image of Palmeirim et al. (2013) using the algorithm getfilaments (Men'shchikov, 2013). In this view, all transverse angular scales larger than 72′′ (or ∼ 0.05 pc) were filtered out to enhance the contrast of the small-scale structure. The color scale on the right is in MJy/sr at 250 μm. The colored curves display the velocity-coherent fibers identified by Hacar et al. (2013) using C18O and N2H+ line observations.

There is also a hint from Planck polarization observations of the nearest clouds that the direction of the magnetic field may change within dense filaments from nearly perpendicular in the ambient cloud to more parallel in the filament interior (Planck Collab. Int. XXXIII, 2016).

All of these findings suggest that magnetic fields play a crucial role in the formation and evolution of star-forming filaments. In particular, they support the view that dense filaments form by accumulation of interstellar matter along field lines. Comparison with MHD simulations of molecular cloud evolution also shows that the Planck results are only consistent with models in which the turbulence is sub-Alfvénic or at most Alfvénic (Planck Collab. Int. XXXV, 2016). Note that a similar conclusion had been reached by Li et al. (2009) based on a comparison of the magnetic field orientation inferred on large (∼ 10–100 pc) scales by optical polarimetry with the magnetic field orientation inferred on small (<< 1 pc) scales by ground-based submillimeter polarimetry.

The low resolution of Planck polarization data (5′–10′ at best or 0.2–0.4 pc in nearby clouds) is however insufficient to probe the organization of field lines in the ∼ 0.1 pc interior of filaments. Consequently, the detailed role of magnetic fields in the evolution of star-forming filaments and their fragmentation into prestellar cores remains a key open issue for the future.

7 Conclusions and prospects

The observational results discussed in this paper are quite encouraging as they point to a unified picture for star formation in the molecular clouds of galaxies. They suggest that both the IMF (see Section 3) and the star formation efficiency in dense molecular gas (André et al., 2014; Shimajiri et al., 2017) are largely set by the “microphysics” of core formation along filaments. Many open issues remain, however, especially concerning the detailed role of magnetic fields (see Section 6). High-resolution, high-dynamic range dust polarization observations at far-IR/submillimeter wavelengths with future space observatories (possibly SPICA) should be able to test the hypothesis, tentatively suggested by Planck polarization results, that the B field becomes nearly parallel to the long axis of star-forming filaments in their dense interiors on scales < 0.1 pc, due to, e.g., gravitational compression. Should this prove to be the case, it could explain both how dense filaments maintain a roughly constant ∼ 0.1 pc width while evolving and why the observed core spacing along filaments is significantly shorter than the characteristic fragmentation scale of 4× the filament diameter expected in the case of non-magnetized nearly isothermal gas cylinders (see Section 2.2). This would have profound implications for our understanding of core and star formation in the Universe.

Acknowledgments

I am grateful to my colleagues D. Arzoumanian, V. Könyves, A. Men'shchikov, P. Palmeirim, Y. Shimajiri, A. Bracco for their key contributions and to S. Inutsuka and P. Hennebelle for insightful discussions. This work was supported by the European Research Council under the European Union's Seventh Framework Programme (ERC Grant Agreement No. 291294 – ‘ORISTARS’).


Bibliographie

[Alves et al., 2007] L. Alves; J. Lada Alves; M. Lombardi; C.J. Lada The mass function of dense molecular cores and the origin of the IMF, Astron. Astrophys., Volume 462 (2007), p. L17

[André et al., 2014] P. André; J. Di Francesco; D. Ward-Thompson; S.I. Inutsuka; R.E. Pudritz; J.E. Pineda From filamentary networks to dense cores in molecular clouds: toward a new paradigm for star formation (H. Beuther; R.S. Klessen; C.P. Dullemond; T. Henning, eds.), Protostars and Planets VI, 2014, pp. 27-51

[André et al., 2010] P. André; A. Men'shchikov; S. Bontemps et al. From filamentary clouds to prestellar cores to the stellar IMF: Initial highlights from the Herschel Gould Belt Survey, Astron. Astrophys., Volume 518 (2010), p. L102

[André et al., 2016] P. André; V. Revéret; V. Könyves; D. Arzoumanian; J. Tige; P. Gallais; H. Roussel; J. Le Pennec; L. Rodriguez; E. Doumayrou; D. Dubreuil; M. Lortholary; J. Martignac; M. Talvard; C. Delisle; F. Visticot; L. Dumaye; C. De Breuck; Y. Shimajiri; F. Motte; S. Bontemps; M. Hennemann; A. Zavagno; D. Russeil; N. Schneider; P. Palmeirim; N. Peretto; T. Hill; V. Minier; A. Roy; K.L.J. Rygl Characterizing filaments in regions of high-mass star formation: high-resolution submilimeter imaging of the massive star-forming complex NGC 6334 with ArTeMiS, Astron. Astrophys., Volume 592 (2016), p. A54

[Arzoumanian, 2012] D. Arzoumanian Characterizing interstellar filaments as revealed by the Herschel Gould Belt survey: insights into the initial conditions for star formation, Université Paris-7, 2012 (PhD thesis)

[Arzoumanian et al., 2011] D. Arzoumanian; P. André; P. Didelon; V. Könyves; N. Schneider; A. Men'shchikov; T. Sousbie; A. Zavagno; S. Bontemps; J. di Francesco; M. Griffin; M. Hennemann; T. Hill; J. Kirk; P. Martin; V. Minier; S. Molinari; F. Motte; N. Peretto; S. Pezzuto; L. Spinoglio; D. Ward-Thompson; G. White; C.D. Wilson Characterizing interstellar filaments with Herschel in IC 5146, Astron. Astrophys., Volume 529 (2011), p. L6

[Arzoumanian et al., 2013] D. Arzoumanian; P. André; N. Peretto; V. Könyves Formation and evolution of interstellar filaments. Hints from velocity dispersion measurements, Astron. Astrophys., Volume 553 (2013), p. A119

[Auddy et al., 2016] S. Auddy; S. Basu; T. Kudoh A magnetic ribbon model for star-forming filaments, Astrophys. J., Volume 831 (2016), p. 46

[Bastian et al., 2010] N. Bastian; K.R. Covey; M.R. Meyer Universal stellar initial mass function? A critical look at variations, Annu. Rev. Astron. Astrophys., Volume 48 (2010), pp. 339-389

[Basu, 2016] S. Basu Molecular cloud fragmentation and core collapse, IAU Symposium. Vol. 315, 2016, pp. 85-90

[Bond et al., 1996] J.R. Bond; L. Kofman; D. Pogosyan, How filaments of galaxies are woven into the cosmic web, 380 (1996), pp. 603-606

[Bouy and Alves, 2015] H. Bouy; J. Alves Cosmography of OB stars in the solar neighbourhood, Astron. Astrophys., Volume 584 (2015), p. A26

[Bracco et al., 2017] A. Bracco; P. Palmeirim; P. André et al. Probing changes of dust properties along a chain of solar-type prestellar and protostellar cores in Taurus with NIKA, 2017 (Astron. Astrophys. 604, A52)

[Burkert and Hartmann, 2004] A. Burkert; L. Hartmann Collapse and fragmentation in finite sheets, Astrophys. J., Volume 616 (2004), pp. 288-300

[Chabrier, 2005] G. Chabrier The initial mass function: from Salpeter (E. Corbelli; F. Palla; H. Zinnecker, eds.), The initial mass function 50 years later, Vol. 327, Astrophysics and Space Science Library, 2005

[Chapman et al., 2011] N.L. Chapman; P.F. Goldsmith; J.L. Pineda et al. The magnetic field in taurus probed by infrared polarization, Astrophys. J., Volume 741 (2011), p. 21

[Chen and Ostriker, 2014] C.-Y. Chen; E.C. Ostriker Formation of magnetized prestellar cores with ambipolar diffusion and turbulence, Astrophys. J., Volume 785 (2014), p. 69

[Clark et al., 2014] S.E. Clark; J.E.G. Peek; M.E. Putman Magnetically aligned H I fibers and the rolling hough transform, Astrophys. J., Volume 789 (2014), p. 82

[Clarke et al., 2017] S.D. Clarke; A.P. Whitworth; A. Duarte-Cabral; D.A. Hubber Filamentary fragmentation in a turbulent medium, Mon. Not. Roy. Astron. Soc., Volume 468 (2017), p. 2489

[Clarke et al., 2016] S.D. Clarke; A.P. Whitworth; D.A. Hubber Perturbation growth in accreting filaments, Mon. Not. Roy. Astron. Soc., Volume 458 (2016), p. 319

[Cox et al., 2016] N.L.J. Cox; D. Arzoumanian; P. André; K.L.J. Rygl; T. Prusti; A. Men'shchikov; P. Royer; A. Kospal; P. Palmeirim; A. Ribas; V. Könyves; J.-P. Bernard; N. Schneider; S. Bontemps; B. Merin; R. Vavrek; C. Alves de Oliveira; P. Didelon; G.L. Pilbratt; C. Waelkens Filamentary structure and magnetic field orientation in Musca, Astron. Astrophys., Volume 590 (2016), p. A110

[Didelon et al., 2015] P. Didelon; F. Motte; P. Tremblin et al. From forced collapse to H ii region expansion in Mon R2: envelope density structure and age determination with Herschel, Astron. Astrophys., Volume 584 (2015), p. A4

[Federrath, 2016] C. Federrath On the universality of interstellar filaments: theory meets simulations and observations, Mon. Not. Roy. Astron. Soc., Volume 457 (2016), p. 375

[Fischera and Martin, 2012] J. Fischera; P.G. Martin Physical properties of interstellar filaments, Astron. Astrophys., Volume 542 (2012), p. A77

[Freundlich et al., 2014] J. Freundlich; C.J. Jog; F. Combes Local stability of a gravitating filament: a dispersion relation, Astron. Astrophys., Volume 564 (2014), p. A7

[Goldsmith et al., 2008] P.F. Goldsmith; M. Heyer; G. Narayanan; R. Snell; D. Li; C. Brunt Large-scale structure of the molecular gas in Taurus revealed by high linear dynamic range spectral line mapping, Astrophys. J., Volume 680 (2008), pp. 428-445

[Gómez and Vázquez-Semadeni, 2014] G.C. Gómez; E. Vázquez-Semadeni Filaments in simulations of molecular cloud formation, Astrophys. J., Volume 791 (2014), p. 124

[Gritschneder et al., 2017] M. Gritschneder; S. Heigl; A. Burkert Oscillating filaments. I. Oscillation and geometrical fragmentation, Astrophys. J., Volume 834 (2017), p. 202

[Hacar et al., 2016] A. Hacar; J. Kainulainen; M. Tafalla; H. Beuther; J. Alves The Musca cloud: a 6 pc-long velocity-coherent, sonic filament, Astron. Astrophys., Volume 587 (2016), p. A97

[Hacar et al., 2013] A. Hacar; M. Tafalla; J. Kauffmann; A. Kovács Cores, filaments, and bundles: hierarchical core formation in the L1495/B213 Taurus region, Astron. Astrophys., Volume 554 (2013), p. A55

[Heitsch et al., 2009] F. Heitsch; J. Ballesteros-Paredes; L. Hartmann Gravitational collapse and filament formation: comparison with the Pipe Nebula, Astrophys. J., Volume 704 (2009), pp. 1735-1742

[Hennebelle, 2013] P. Hennebelle On the origin of non-self-gravitating filaments in the ISM, Astron. Astrophys., Volume 556 (2013), p. A153

[Hennebelle and André, 2013] P. Hennebelle; P. André Ion-neutral friction and accretion-driven turbulence in self-gravitating filaments, Astron. Astrophys., Volume 560 (2013), p. A68

[Hennebelle and Chabrier, 2008] P. Hennebelle; G. Chabrier Analytical theory for the initial mass function: CO clumps and prestellar cores, Astrophys. J., Volume 684 (2008), pp. 395-410

[Hennemann et al., 2012] M. Hennemann; F. Motte; N. Schneider; P. Didelon; T. Hill; D. Arzoumanian; S. Bontemps; T. Csengeri; P. André; V. Konyves; F. Louvet; A. Marston; A. Men'shchikov; V. Minier; Q. Nguyen Luong; P. Palmeirim; N. Peretto; M. Sauvage; A. Zavagno; L.D. Anderson; J.-P. Bernard; J. Di Francesco; D. Elia; J.Z. Li; P.G. Martin; S. Molinari; S. Pezzuto; D. Russeil; K.L.J. Rygl; E. Schisano; L. Spinoglio; T. Sousbie; D. Ward-Thompson; G.J. White The spine of the swan: a Herschel study of the DR21 ridge and filaments in Cygnus X, Astron. Astrophys., Volume 543 (2012), p. L3

[Henshaw et al., 2016] J.D. Henshaw; P. Caselli; F. Fontani; I. Jimenez-Serra; J.C. Tan; S.N. Longmore; J.E. Pineda; R.J. Parker; A.T. Barnes Investigating the structure and fragmentation of a highly filamentary IRDC, Mon. Not. Roy. Astron. Soc., Volume 463 (2016), pp. 146-169

[Henshaw et al., 2017] J.D. Henshaw; I. Jiménez-Serra; S.N. Longmore; P. Caselli; J.E. Pineda; A. Avison; A.T. Barnes; J.C. Tan; F. Fontani Unveiling the early-stage anatomy of a protocluster hub with ALMA, Mon. Not. Roy. Astron. Soc., Volume 464 (2017), p. L31-L35

[Hill et al., 2011] T. Hill; F. Motte; P. Didelon; S. Bontemps; V. Minier; M. Hennemann; N. Schneider; P. André; A. Men'shchikov; L.D. Anderson; D. Arzoumanian; J.-P. Bernard; J. di Francesco; D. Elia; T. Giannini; M.J. Griffin; V. Konyves; J. Kirk; A.P. Marston; P.G. Martin; S. Molinari; Q. Nguyen Luong; N. Peretto; S. Pezzuto; H. Roussel; M. Sauvage; T. Sousbie; L. Testi; D. Ward-Thompson; G.J. White; C.D. Wilson; A. Zavagno Filaments and ridges in Vela C revealed by Herschel: from low-mass to high-mass star-forming sites, Astron. Astrophys., Volume 533 (2011), p. A94

[Inutsuka and Miyama, 1992] S.-I. Inutsuka; S.M. Miyama Self-similar solutions and the stability of collapsing isothermal filaments, Astrophys. J., Volume 388 (1992), pp. 392-399

[Inutsuka and Miyama, 1997] S.-I. Inutsuka; S.M. Miyama A production mechanism for clusters of dense cores, Astrophys. J., Volume 480 (1997), p. 681

[Johnstone and Bally, 1999] D. Johnstone; J. Bally JCMT/SCUBA submillimeter wavelength imaging of the integral-shaped filament in Orion, Astrophys. J. Lett., Volume 510 (1999), p. L49-L53

[Juvela et al., 2012] M. Juvela; I. Ristorcelli; L. Pagani; Y. Doi; V.M. Pelkonen; D.J. Marshall; J.-P. Bernard; E. Falgarone; J. Malinen; G. Marton; P. McGehee; L.A. Montier; F. Motte; R. Paladini; L.V. Toth; N. Ysard; S. Zahorecz; A. Zavagno Galactic cold cores. III. General cloud properties, Astron. Astrophys., Volume 541 (2012), p. A12

[Kainulainen et al., 2009] J. Kainulainen; H. Beuther; T. Henning; R. Plume Probing the evolution of molecularcloud structure. From quiescence to birth, Astron. Astrophys., Volume 508 (2009), p. L35-L38

[Kainulainen et al., 2013] J. Kainulainen; S.E. Ragan; T. Henning; A. Stutz High-fidelity view of the structure and fragmentation of the high-mass, filamentary IRDC G11, Astron. Astrophys., Volume 557 (2013), p. A120

[Kainulainen et al., 2017] J. Kainulainen; A.M. Stutz; T. Stanke; J. Abreu-Vicente; H. Beuther; T. Henning; K.G. Johnston; S.T. Megeath Resolving the fragmentation of high line-mass filaments with ALMA: the integral shaped filament in Orion A, Astron. Astrophys., Volume 600 (2017), p. A141

[Kalberla et al., 2016] P.M.W. Kalberla; J. Kerp; U. Haud; B. Winkel; N. Ben Bekhti; L. Flöer; D. Lenz Cold Milky Way HI Gas in Filaments. 2016, Astrophys. J., Volume 821 (2016), p. 117

[Kirk et al., 2013] H. Kirk; P.C. Myers; T.L. Bourke; R.A. Gutermuth; A. Hedden; G.W. Wilson Filamentary accretion flows in the embedded serpens South Protocluster, Astrophys. J., Volume 766 (2013), p. 115

[Klessen and Hennebelle, 2010] R.S. Klessen; P. Hennebelle Accretion-driven turbulence as universal process: galaxies, molecular clouds, and protostellar disks, Astron. Astrophys., Volume 520 (2010), p. A17

[Koch and Rosolowsky, 2015] E.W. Koch; E.W. Rosolowsky Filament identification through mathematical morphology, Mon. Not. Roy. Astron. Soc., Volume 452 (2015), pp. 3435-3450

[Könyves et al., 2015] V. Könyves; P. André; A. Men'shchikov; P. Palmeirim; D. Arzoumanian; N. Schneider; A. Roy; P. Didelon; A. Maury; Y. Shimajiri; J. Di Francesco; S. Bontemps; N. Peretto; M. Benedettini; J.-P. Bernard; D. Elia; M.J. Griffin; T. Hill; J. Kirk; B. Ladjelate; K. Marsh; P.G. Martin; F. Motte; Q. Nguyên Luong; S. Pezzuto; H. Roussel; K.L.J. Rygl; S.I. Sadavoy; E. Schisano; L. Spinoglio; D. Ward-Thompson; G.J. White A census of dense cores in the Aquila cloud complex: SPIRE/PACS observations from the Herschel Gould Belt survey, Astron. Astrophys., Volume 584 (2015), p. A91

[Kramer et al., 1998] C. Kramer; J. Stutzki; R. Rohrig; U. Corneliussen Clump mass spectra of molecular clouds, Astron. Astrophys., Volume 329 (1998), pp. 249-264

[Lada et al., 2012] C.J. Lada; J. Forbrich; M. Lombardi; J.F. Alves Star formation rates in molecular clouds and the nature of the extragalactic scaling relations, Astrophys. J., Volume 745 (2012), p. 190

[Lada et al., 2010] C.J. Lada; M. Lombardi; J.F. Alves On the star formation rates in molecular clouds, Astrophys. J., Volume 724 (2010), pp. 687-693

[Larson, 2005] R.B. Larson Thermal physics, cloud geometry and the stellar initial mass function, Mon. Not. Roy. Astron. Soc., Volume 359 (2005), pp. 211-222

[Li et al., 2009] H.-b. Li; C.D. Dowell; A. Goodman; R. Hildebrand; G. Novak Anchoring magnetic field in turbulent molecular clouds, Astrophys. J., Volume 704 (2009), pp. 891-897

[Lin et al., 1965] C.C. Lin; L. Mestel; F.H. Shu Gravitational collapse of a uniform spheroid, Astrophys. J., Volume 142 (1965), p. 1431

[Mac Low and Klessen, 2004] M.-M. Mac Low; R.S. Klessen Control of star formation by supersonic turbulence, Rev. Mod. Phys., Volume 76 (2004), pp. 125-194

[Machida and Matsumoto, 2012] M.N. Machida; T. Matsumoto Impact of protostellar outflow on star formation: effects of the initial cloud mass, Mon. Not. Roy. Astron. Soc., Volume 421 (2012), pp. 588-607

[Marsh et al., 2016] K.A. Marsh; J.M. Kirk; P. André; M.J. Griffin; V. Konyves; P. Palmeirim; A. Men'shchikov; D. Ward-Thompson; M. Benedettini; D.W. Bresnahan; J.D. Francesco; D. Elia; F. Motte; N. Peretto; S. Pezzuto; A. Roy; S. Sadavoy; N. Schneider; L. Spinoglio; G.J. White A census of dense cores in the Taurus L1495 cloud from the Herschel, Mon. Not. Roy. Astron. Soc., Volume 459 (2016), pp. 342-356

[Matzner and McKee, 2000] C.D. Matzner; C.F. McKee Efficiencies of low-mass star and star cluster formation, Astrophys. J., Volume 545 (2000), pp. 364-378

[McClure-Griffiths et al., 2006] N.M. McClure-Griffiths; J.M. Dickey; B.M. Gaensler et al. Magnetically dominated strands of cold hydrogen in the Riegel–Crutcher Cloud, Astrophys. J., Volume 652 (2006), pp. 1339-1347

[McKee and Ostriker, 2007] C.F. McKee; E.C. Ostriker Efficiencies of star formation, Annu. Rev. Astron. Astrophys., Volume 45 (2007), pp. 565-687

[Men'shchikov, 2013] A. Men'shchikov A multi-scale filament extraction method: getfilaments, Astron. Astrophys., Volume 560 (2013), p. A63

[Men'shchikov et al., 2010] A. Men'shchikov; P. André; P. Didelon; V. Könyves; N. Schneider; F. Motte; S. Bontemps; D. Arzoumanian; M. Attard; A. Abergel; J.-P. Baluteau; J.-P. Bernard; L. Cambresy; P. Cox; J. di Francesco; A.M. di Giorgio; M. Griffin; P. Hargrave; M. Huang; J. Kirk; J.Z. Li; P. Martin; V. Minier; M.A. Miville-Deschênes; S. Molinari; G. Olofsson; S. Pezzuto; H. Roussel; D. Russeil; P. Saraceno; M. Sauvage; B. Sibthorpe; L. Spinoglio; L. Testi; D. Ward-Thompson; G. White; C.D. Wilson; A. Woodcraft; A. Zavagno Filamentary structures and compact objects in the Aquila and Polaris clouds observed by Herschel, Astron. Astrophys., Volume 518 (2010), p. L103

[Miville-Deschênes et al., 2016] M.-A. Miville-Deschênes; P.-A. Duc; F. Marleau; J.-C. Cuillandre; P. Didelon; S. Gwyn; E. Karabal Probing interstellar turbulence in cirrus with deep optical imaging: no sign of energy dissipation at 0.01 pc scale, Astron. Astrophys., Volume 593 (2016), p. A4

[Miville-Deschênes et al., 2010] M.-A. Miville-Deschênes; P.G. Martin; A. Abergel; J.-P. Bernard; F. Boulanger; G. Lagache; L.D. Anderson; P. André; H. Arab; J.-P. Baluteau; K. Blagrave; S. Bontemps; M. Cohen; M. Compiegne; P. Cox; E. Dartois; G. Davis; R. Emery; T. Fulton; C. Gry; E. Habart; M. Huang; C. Joblin; S.C. Jones; J. Kirk; T. Lim; S. Madden; G. Makiwa; A. Menshchikov; S. Molinari; H. Moseley; F. Motte; D.A. Naylor; K. Okumura; D. Pinheiro Goncalves; E. Polehampton; J.A. Rodon; D. Russeil; P. Saraceno; N. Schneider; S. Sidher; L. Spencer; B. Swinyard; D. Ward-Thompson; G.J. White; A. Zavagno Herschel–SPIRE observations of the Polaris flare: Structure of the diffuse interstellar medium at the sub-parsec scale, Astron. Astrophys., Volume 518 (2010), p. L104

[Molinari et al., 2010] S. Molinari; B. Swinyard; J. Bally et al. Clouds, filaments, and protostars: the Herschel Hi-GAL Milky Way, Astron. Astrophys., Volume 518 (2010), p. L100

[Myers, 2009] P.C. Myers Filamentary structure of star-forming complexes, Astrophys. J., Volume 700 (2009), pp. 1609-1625

[Ntormousi et al., 2016] E. Ntormousi; P. Hennebelle; P. André; J. Masson The effect of ambipolar diffusion on low-density molecular ISM filaments, Astron. Astrophys., Volume 589 (2016), p. A24

[Ostriker, 1964] J. Ostriker The equilibrium of polytropic and isothermal cylinders, Astrophys. J., Volume 140 (1964), p. 1056

[Padoan et al., 2001] P. Padoan; M. Juvela; A.A. Goodman; Å. Nordlund The turbulent shock origin of proto-stellar cores, Astrophys. J., Volume 553 (2001), pp. 227-234

[Padoan and Nordlund, 2002] P. Padoan; Å. Nordlund The stellar initial mass function from turbulent fragmentation, Astrophys. J., Volume 576 (2002), pp. 870-879

[Palmeirim et al., 2013] P. Palmeirim; P. André; J. Kirk; D. Ward-Thompson; D. Arzoumanian; V. Konyves; P. Didelon; N. Schneider; M. Benedettini; S. Bontemps; J. Di Francesco; D. Elia; M. Griffin; M. Hennemann; T. Hill; P.G. Martin; A. Men'shchikov; S. Molinari; F. Motte; Q. Nguyen Luong; D. Nutter; N. Peretto; S. Pezzuto; A. Roy; K.L.J. Rygl; L. Spinoglio; G.L. White Herschel view of the Taurus B211/3 filament and striations: evidence of filamentary growth?, Astron. Astrophys., Volume 550 (2013), p. A38

[Panopoulou et al., 2017] G.V. Panopoulou; I. Psaradaki; R. Skalidis; K. Tassis; J.J. Andrews A closer look at the ‘characteristic’ width of molecular cloud filaments, Mon. Not. Roy. Astron. Soc., Volume 466 (2017), pp. 2529-2541

[Panopoulou et al., 2016] G.V. Panopoulou; I. Psaradaki; K. Tassis The magnetic field and dust filaments in the Polaris Flare, Mon. Not. Roy. Astron. Soc., Volume 462 (2016), pp. 1517-1529

[Peretto et al., 2006] N. Peretto; P. André; A. Belloche Probing the formation of intermediate- to high-mass stars in protoclusters. A detailed millimeter study of the NGC 2264 clumps, Astron. Astrophys., Volume 445 (2006), pp. 979-998

[Peretto et al., 2012] N. Peretto; P. André; V. Könyves; N. Schneider; D. Arzoumanian; P. Palmeirim; P. Didelon; M. Attard; J.-P. Bernard; J. Di Francesco; D. Elia; M. Hennemann; T. Hill; J. Kirk; A. Men'shchikov; F. Motte; Q. Nguyen Luong; H. Roussel; T. Sousbie; L. Testi; D. Ward-Thompson; G.J. White; A. Zavagno The Pipe Nebula as seen with Herschel: formation of filamentary structures by large-scale compression?, Astron. Astrophys., Volume 541 (2012), p. A63

[Peretto et al., 2013] N. Peretto; G.A. Fuller; A. Duarte-Cabral; A. Avison; P. Hennebelle; J.E. Pineda; P. André; S. Bontemps; F. Motte; N. Schneider; S. Molinari Global collapse of molecular clouds as a formation mechanism for the most massive stars, Astron. Astrophys., Volume 555 (2013), p. A112

[Pilbratt et al., 2010] G.L. Pilbratt; J.R. Riedinger; T. Passvogel; G. Crone; D. Doyle; U. Gageur; A.M. Heras; C. Jewell; L. Metcalfe; S. Ott; M. Schmidt Herschel Space Observatory. An ESA facility for far-infrared and submillimetre astronomy, Astron. Astrophys., Volume 518 (2010), p. L1

[Planck Collab. Int. XXXII, 2016] Planck Collab. Int. XXXII Planck intermediate results. XXXII. The relative orientation between the magnetic field and structures traced by interstellar dust, Astron. Astrophys., Volume 586 (2016), p. A135

[Planck Collab. Int. XXXIII, 2016] Planck Collab. Int. XXXIII Planck intermediate results. XXXIII. Signature of the magnetic field geometry of interstellar filaments in dust polarization maps, Astron. Astrophys., Volume 586 (2016), p. A136

[Planck Collab. Int. XXXV, 2016] Planck Collab. Int. XXXV Planck intermediate results. XXXV. Probing the role of the magnetic field in the formation of structure in molecular clouds, Astron. Astrophys., Volume 586 (2016), p. A138

[Planck Collaboration Int. XIX, 2015] Planck Collaboration Int. XIX Planck intermediate results. XIX. An overview of the polarized thermal emission from Galactic dust, Astron. Astrophys., Volume 576 (2015), p. A104

[Pokhrel et al., 2016] R. Pokhrel; R. Gutermuth; B. Ali; T. Megeath; J. Pipher; P. Myers; W.J. Fischer; T. Henning; S.J. Wolk; L. Allen; J.J. Tobin A Herschel–SPIRE survey of the Mon R2 giant molecular cloud: analysis of the gas column density probability density function, Mon. Not. Roy. Astron. Soc., Volume 461 (2016), pp. 22-35

[Pudritz and Kevlahan, 2013] R.E. Pudritz; N.K.-R. Kevlahan Shock interactions, turbulence and the origin of the stellar mass spectrum, Philos. Trans. R. Soc. Ser. A, Volume 371 (2013), p. 20248

[Roy et al., 2015] A. Roy; P. André; D. Arzoumanian; N. Peretto; P. Palmeirim; V. Könyves; N. Schneider; M. Benedettini; J. Di Francesco; D. Elia; T. Hill; B. Ladjelate; F. Louvet; F. Motte; S. Pezzuto; E. Schisano; Y. Shimajiri; L. Spinoglio; D. Ward-Thompson; G. White Possible link between the power spectrum of interstellar filaments and the origin of the prestellar core mass function. 2015, Astron. Astrophys., Volume 584 (2015), p. A111

[Salji et al., 2015] C.J. Salji; J.S. Richer; J.V. Buckle; J.D. Francesco; J. Hatchell; M. Hogerheijde; D. Johnstone; H. Kirk; D. Ward-Thompson; JCMT GBS Consortium The JCMT Gould Belt Survey: properties of star-forming filaments in Orion A North, Mon. Not. Roy. Astron. Soc., Volume 449 (2015), pp. 1782-1796

[Schisano et al., 2014] E. Schisano; K.L.J. Rygl; S. Molinari; G. Busquet; D. Elia; M. Pestalozzi; D. Polychroni; N. Billot; S. Carey; R. Paladini; A. Noriega-Crespo; T.J.T. Moore; R. Plume; S.C.O. Glover; E. Vazquez -Semadeni The identification of filaments on far-infrared and submillimiter images: morphology, physical conditions and relation with star formation of filamentary structure, Astrophys. J., Volume 791 (2014), p. 27

[Schneider et al., 2013] N. Schneider; P. André; V. Könyves; V. Bontemps; S. Motte; F. Federrath; C. Ward-Thompson; D. Arzoumanian; D. Benedettini; M. Bressert; E. Didelon; P. Di Francesco; J. Griffin; M. Hennemann; M. Hill; T. Palmeirim; P. Pezzuto; S. Peretto; N. Roy; A. Rygl; K.L.J. Spinoglio; L.G. White What determines the density structure of molecular clouds? A case study of Orion B with Herschel, Astrophys. J. Lett., Volume 766 (2013), p. L17

[Schneider et al., 2010] N. Schneider; T. Csengeri; S. Bontemps; F. Motte; R. Simon; P. Hennebelle; C. Federrath; R. Klessen Dynamic star formation in the massive DR21 filament, Astron. Astrophys., Volume 520 (2010), p. A49

[Schneider and Elmegreen, 1979] S. Schneider; B.G. Elmegreen A catalog of dark globular filaments, Astrophys. J. Suppl. Ser., Volume 41 (1979), pp. 87-95

[Shandarin and Zeldovich, 1989] S.F. Shandarin; Y.B. Zeldovich The large-scale structure of the universe: turbulence, intermittency, structures in a self-gravitating medium, Rev. Mod. Phys., Volume 61 (1989), pp. 185-220

[Shimajiri et al., 2017] Y. Shimajiri; P. André; J. Braine; V. Könyves; N. Schneider; S. Bontemps; B. Ladjelate; A. Roy; Y. Gao; H. Chen Testing the universality of the star formation efficiency in dense molecular gas, 2017 (Astron. Astrophys. 604, A74)

[Smith et al., 2014] R.J. Smith; S.C.O. Glover; R.S. Klessen On the nature of star-forming filaments – I. Filament morphologies, Mon. Not. Roy. Astron. Soc., Volume 445 (2014), pp. 2900-2917

[Soler et al., 2016] J.D. Soler; F. Alves; F. Boulanger; A. Bracco; E. Falgarone; G.A.P. Franco; V. Guillet; P. Hennebelle; F. Levrier; P.G. Martin; M.A. Miville-Deschênes Magnetic field morphology in nearby molecular clouds as revealed by starlight and submillimetre polarization, Astron. Astrophys., Volume 596 (2016), p. A93

[Soler et al., 2013] J.D. Soler; P. Hennebelle; P.G. Martin; M.A. Miville-Deschênes; C.B. Netterfield; L.M. Fissel An imprint of molecular cloud magnetization in the morphology of the dust polarized emission, Astrophys. J., Volume 774 (2013), p. 128

[Springel et al., 2005] V. Springel; S.D.M. White; A. Jenkins; C.S. Frenk; N. Yoshida; L. Gao; J. Navarro; R. Thacker; D. Croton; J. Helly; J.A. Peacock; S. Cole; P. Thomas; H. Couchman; A. Évrard; J. Colberg; F. Pearce, Simulations of the formation, evolution and clustering of galaxies and quasars, 435 (2005), pp. 629-636

[Tafalla and Hacar, 2015] M. Tafalla; A. Hacar Chains of dense cores in the Taurus L1495/B213 complex, Astron. Astrophys., Volume 574 (2015), p. A104

[Takahashi et al., 2013] S. Takahashi; P.T.P. Ho; P.S. Teixeira; L.A. Zapata; Y.-N. Su Hierarchical fragmentation of the Orion molecular filaments, Astrophys. J., Volume 763 (2013), p. 57

[Teixeira et al., 2016] P.S. Teixeira; S. Takahashi; L.A. Zapata; P.T.P. Ho Two-level hierarchical fragmentation in the northern filament of the Orion Molecular Cloud 1, Astron. Astrophys., Volume 587 (2016), p. A47

[Wang et al., 2015] K. Wang; L. Testi; A. Ginsburg; C.M. Walmsley; S. Molinari; E. Schisano Large-scale filaments associated with Milky Way spiral arms, Mon. Not. Roy. Astron. Soc., Volume 450 (2015), pp. 4043-4049

[Ysard et al., 2013] N. Ysard; A. Abergel; I. Ristorcelli; M. Juvela; L. Pagani; V. Könyves; L. Spencer; G. White; A. Zavagno Variation in dust properties in a dense filament of the Taurus molecular complex (L1506), Astron. Astrophys., Volume 559 (2013), p. A133

[Zel’dovich, 1970] Y.B. Zel’dovich Gravitational instability: an approximate theory for large density perturbations, Astron. Astrophys., Volume 5 (1970), pp. 84-89


Commentaires - Politique