• Keine Ergebnisse gefunden

The dependence of accretion rates with the mass of the central star

1.3 Accretion as a tracer of protoplanetary disk evolution

1.3.3 The dependence of accretion rates with the mass of the central star

The dependence ofM˙acconMwas initially not explored in detail from a theoretical point of view. The basic reason was that in a viscous disk where viscosity is driven by MRI turbulence there is no dependence between the two quantities according to theory. For example, using the layered accretion disk theory, Gammie (1996) showed thatM˙acc in the inner disk is:

M˙acc = 1.8×108 ( α

102 )2(

Σa 100g·cm2

)3

M yr1, (1.25) where Σa is the surface density in the active layer of the disk, and α the usual param-eter to describe the disk viscosity. In this result there is no explicit dependence upon M, and no further theoretical arguments were available at that time to predict any de-pendence between these two observables. As I describe in Sect. 1.4.1 in detail, the first observational works aimed at studying accretion in cTTs were all focused on objects with M 0.5 1 M (e.g., Valenti et al., 1993; Kenyon & Hartmann, 1995; Hartmann et al., 1998). In the next years more efforts were done to study lower mass objects, and, in particular, it became clear that brown dwarfs (BDs) posses circumstellar disks like cTTs, with infrared excess (e.g., Natta & Testi, 2001; Jayawardhana et al., 2003a) and accretion signatures (e.g., Jayawardhana et al., 2003b; Natta et al., 2004; Mohanty et al., 2005). The initial idea was that, if BD formation is just a scaled-down version of star formation, M˙acc should scale proportionally to M down to the BD regime. However, thanks to the suffi-ciently large interval ofMaccessible to observations (roughly 2 orders of magnitude when including BDs) theM˙acc-M relation could be observationally investigated, and, strikingly, this dependence was found to beM˙acc∝M2, with a considerably large spread in the values of M˙acc at any M (e.g., Muzerolle et al., 2003; Natta et al., 2004; Mohanty et al., 2005;

Natta et al., 2006). These results called for an explanation, and different possibilities have been explored in the next years. In particular, it became clear that, given thatM changes negligibly during the Class II phase, theM˙acc-Mplane is a diagnostic tool for the evolution of accretion during this phase, yielding information on the relative amount of time spent in various regions of the plane by the objects (Clarke & Pringle, 2006). So far, no final solution of this problem has been found. Here I discuss the possible scenarios introduced to explain the M˙acc-M dependence, while I discuss the observational results and the still open issues on this relation at the beginning of this Thesis in Sect. 1.4.1.

1. Introduction

Bondi-Hoyle accretion

An early explanation for the observed M˙acc-M relation was suggested by Padoan et al.

(2005), who explained this as a consequence of Bondi-Hoyle accretion. They suggested that, given that the origin of disk turbulence driving the viscous evolution is not yet clear, the only possibility to compare viscous evolution models with observation is to assume ``ad hoc'' values for the viscosity. Moreover, all YSOs are still associated with their parental cloud, so the assumption that disks evolve in isolation could be incorrect. Their proposed evolution of YSOs is divided in two phases. In the first one the core collapses and there is spherical accretion which ends with the formation of a central PMS object surrounded by a disk. This very massive disk is gravitationally unstable and rapidly accretes onto the central object in less than 1 Myr. In the second phase the system accretes material from the large-scale gas distribution in the parent cloud, and the disk acts as a short-term buffer between the large-scale structure and the central PMS star. The accretion of material from the parent cloud is driven by Bondi-Hoyle accretion, whose rate is:

M˙BH= 4πG2ρ

(c2+v2)3/2M2, (1.26) where ρis the gas density of the parental cloud,cits sound speed,v the gas velocity relative to the star, and all these quantities depend on the environment, therefore are in-dependent on the properties of each individual YSO. If the values ofM˙acc are assumed to be the same as expected from Bondi-Hoyle accretion, than the order-of-magnitude of the observation is matched and the dependence ofM˙accwithMis a power-law with exponent 2, as expected.

However, there are several caveats to this description. First, the assumption thatM˙acc M˙BH means that all the mass which is accreted on the star+disk system steady flows in the disk to the central PMS star. In fact, the actual rate of accretion on the star could be lower (e.g., Mohanty et al., 2005). Then, this description implies that M˙acc itself should depend strongly on the environment in which the YSO is located. This was, however, never observed, as described by, e.g., Mohanty et al. (2005), Hartmann et al. (2006), and Alexander & Armitage (2006). Finally, explaining the evolution of protoplanetary disks without angular momentum transport is a questionable assumption that could not explain, for example, the fact that TTauri stars posses large disks (e.g., Hartmann et al., 2006) and that disks in the early phases are very small (e.g., Miotello et al., 2014), so they have to expand afterwards. Therefore, this explanation is not thought to be descriptive of the phenomenon.

Initial conditions and basic viscous evolution

Alexander & Armitage (2006) firstly proposed an explanation for the M˙acc M2 relation as a consequence of properties of the disks at formation, followed by classical viscous evolution. From the solution of the surface density evolution in the context of similarity

1.3 Accretion as a tracer of protoplanetary disk evolution

solution (ν ∝Rγ, Eq. 1.13), by expressing the constantC, that is:

C = 3Md(0)(2−γ)ν1

2R21 , (1.27)

they derive:

Σ(R, t) = Md(0)(2−γ)

2πR21rγ T(5/2γ)/(2γ)exp [

−r(2γ) T

]

, (1.28)

where Md(0) is the initial disk mass, r = R/R1 is a dimensionless radius, R1 is the scale radius that sets the initial disk size, and T = t/tν + 1 is the dimensionless time. As the viscous timescale is:

tν = R21

3(2−γ)2ν1, (1.29)

they derive thatM˙acc, i.e. the accretion rate onto the star (r 0) is:

M˙acc = Md(0)

2(2−γ)tνT(5/2γ)/(2γ), (1.30)

which is the same as in Eq. (1.17), but with the proportionality constant expressed. Their assumption is that the M˙acc M2 relation is valid already for the initial accretion rates.

Given that observations show that in BDs the disk-to-star mass ratio is comparable to that of cTTs, they assume that Md(0) M, thus their assumption is fullfilled if the viscous timescale depends onM in the following way:

tν 1

M. (1.31)

This has strong implications on possible observables. First of all, Eq. (1.31) implies that the viscous timescale is longer for BDs than for TTauri stars, with typicallytν 106 yr for BDs if tν 104 105 yr for cTTs (e.g., Hartmann et al., 1998). Lower mass objects will thus evolve more slowly with time than higher mass ones. As a consequence, the spread in the values ofM˙accon theM˙acc-M plane should be smaller at lowerM. Then, from the expression of the viscous timescale and the viscosity itself they derive that, typically, the initial disk radius for BDs are larger than for cTTs, with scale radii R0 for BDs 50-100 AU. However, the fact that the viscous timescale is shorter for higher mass stars implies that disks around these objects will spread for viscous evolution faster than those around BDs, so there would be no observable differences at later stages. Implicit in their discussion there is the suggestion that the observed spread of values of M˙acc at anyM is due to age differences.

A similar approach of suggesting that the observed spread is due to the imprint of the initial conditions at formation has been used by Dullemond et al. (2006) to explain the M˙acc-M relation. They model the collapse of various rotating molecular cloud cores and the subsequent viscous evolution of the disk which forms. By assuming that cores with significantly different masses have similar rotation rates to breakup rate ratios they find

1. Introduction

a trend of M˙acc M1.8±0.2. The observed spread can be then interpreted as a spread in the initial core rotation rates. This simple but descriptive solution has some observable consequences, as well. First, the process of star formation is similar for higher mass stars and BDs. Then, the mass of the disk should depend strongly on M, roughly Md M2, and, in particular, should be a tightly correlated withM˙acc.

Assumptions on disk viscosity

A possible explanation could be that the viscosity of the disk, which drives accretion, depends on M. If MRI is driving the disk turbulence, Mohanty et al. (2005) suggest that the amount of ionization, and thus the turbulence, is lower for cooler, i.e. lower-mass, objects. This qualitative argument has been later on somehow developed by Hartmann et al. (2006), who suggest that M˙acc depends on M if irradiation by the central star is taken into account in a layer accretion model (Gammie, 1996). Basically,M˙accis given by the accretion rate of the layered model at a critical radius (RC) where the temperature is high enough to enhance MRI in the disk. They assume this critical radius to be the inner edge of the dusty disk, which is frontally illuminated by the central star. By fixing the dust destruction temperature they derive that RC L1/2 . Assuming a classical alpha-viscosity description of the disk, i.e. ν =αcs/Ω, they obtain that:

M˙ ∝νΣ∝αc2s1Σa∝αΣaL3/4 M1/2, (1.32) where also in this caseΣais the surface density of the active layer of the disk. In PMS stars, at least forM ≲2M, the dependence of the stellar luminosity with the mass isL ∝M2, so, if the disk is purely internally irradiated, the accretion rate atRC would be:

M˙ ∝αΣaM. (1.33)

With this description the mass accretion rate depends on the stellar mass, but the slope of this dependence is not as steep as observed. However, this description could represent the upper envelope of accretion rates observed (Hartmann et al., 2006). The suggestion would be that the upper envelope is composed by objects whose disk is described by the layered accretion disk model with irradiation from the central star, while the other objects falling well below this upper envelope in the M˙acc-M plane (mostly BDs and low-mass TTauri stars) have a small and magnetically active disk that substantially evolved viscously and has now a lower rate of accretion. This implies that lower-mass objects evolve faster than solar-mass cTTs, so the evolution of M˙acc with time should be different for the two classes of objects, withM˙acc slowly decreasing for higher mass cTTs with respect to BDs and low-mass stars.

Self-gravitation

Some efforts have been made in the literature to explain the evolution of protoplanetary disks under self-regulated disks which experience gravitational instabilities driving the disk

1.3 Accretion as a tracer of protoplanetary disk evolution turbulence and the subsequent accretion process. In particular, Vorobyov & Basu (2008) discuss howM˙acc would depend onM in systems where gravitational instabilities are the only driver of disk turbulence, and hence accretion. By carrying out several simulations on the first 3 Myr of evolution of low- and intermediate-mass TTauri stars (0.25M< M <

3.0M), they derive aM˙acc-Mrelation with a power-law of exponent 1.7. At the same time, they predict that M˙acc should correlate almost linearly with the disk mass. Finally, they discuss that there could be a bimodality in the observed M˙acc-M relation, with different behaviors in the very-low-mass end and in the M > 0.2M range. Following this work, Vorobyov & Basu (2009) developed a set of models to describe the evolution of both objects withM >0.2Mand of lower M. In order to be able to describe both classes of objects they include in their description also an effective turbulentα-viscosity withα=0.01 to allow for accretion in the lower mass objects, where the disk is not massive enough to drive substantial gravitational instabilities. Also in this case the values ofM˙accpredicted by their models are similar to those observed. The M˙acc-M relation is clearly reproduced, with an exponent of the power-law which is 1.8 when including all the time-averaged models at all M. They also suggest that the large spread of M˙acc at any given M is partly due to intrinsic variability during the evolution and partly to object-to-object variations due to different initial conditions. Finally, they predict with their models a strong bimodality in the M˙acc-M relation, with values of the exponent of the power-law varying from 2.9 for YSOs withM <0.2Mto 1.5 for objects with 0.2M ≤M <3.0M.

Photoevaporation

As I discussed in Sect. 1.3.2, protoplanetary disks are subject to photoevaporation from the central star. Under this description, Clarke & Pringle (2006) suggested that the lowest accretion rate measurable at a given mass is set by the onset of photoevaporation. As the decrease of M˙acc with time is a power-law decay, young stars spend most of their time at the lowest possible accretion rates, and this is the situation at which it is more plausible to observe them. Therefore, theM˙acc-M relation could be driven in large samples by these low-accretors. Following on this path, if the lowest accretion is set by photoevaporation (see Sect. 1.3.2), then the most plausible accretion rate of a star isM˙acc=M˙wind, withM˙wind being the mass-loss rate due to photoevaporation. Clarke & Pringle (2006) compute the ex-pectedM˙acc-Mrelation in the case in which UV radiation dominates the photoevaporative flow, which results in the following mass-loss rate dependence ofM:

M˙wind (Mϕ)1/2, (1.34)

where ϕis the ionizing flux. In the UV photoevaporation caseϕsimply scales with stellar luminosity, implying that M˙wind M1.35. This power-law exponent, however, is much lower than what is observed.

On the other hand, steeper dependence onM is found when considering X-ray driven photoevaporation, as recently shown by Ercolano et al. (2014). In the case of X-ray

photo-1. Introduction

evaporation, Owen et al. (2012) derived that the mass-loss rate is described as:

M˙wind= 6.25×109

( LX 1030 ergs−1

)1.14

M yr1, (1.35) where LX is the X-ray luminosity, and the very weak (∼M0.068) dependence on the stel-lar mass is ignored. With this description the M˙acc-M relation is driven by the M-LX

one, which can be constrained observationally. Observations of various objects in Orion (Preibisch et al., 2005) and in Taurus (Güdel et al., 2007) suggest that LX M1.51.7 for objects withM <1M, which results in a dependenceM˙acc∼M1.51.7 in agreement with most of the observations, as I discuss in the next section.