23.03.2013 Views

Interferometric observations of pre-main sequence disks - Caltech ...

Interferometric observations of pre-main sequence disks - Caltech ...

Interferometric observations of pre-main sequence disks - Caltech ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Università degli Studi di Milano<br />

Facoltà di Scienze Matematiche, Fisiche e Naturali<br />

Dottorato di Ricerca in Fisica, Astr<strong>of</strong>isica e<br />

Fisica applicata - XIX Ciclo<br />

FIS/05-Astronomia e Astr<strong>of</strong>isica – Matricola R05633<br />

<strong>Interferometric</strong> <strong>observations</strong> <strong>of</strong><br />

<strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> <strong>disks</strong><br />

Tutore: Pr<strong>of</strong>.ssa Antonella Natta<br />

– Andrea Isella –<br />

Coordinatore: Pr<strong>of</strong>. Gianpaolo Bellini<br />

Anno Accademico 2005-2006


A Rossana


CONTENTS<br />

I Introduction 1<br />

1 Scientific overview 3<br />

1.1 Star formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3<br />

1.2 Disk structure in the <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> phase . . . . . . . . . . . . . . 8<br />

1.3 Grain growth and planetary formation . . . . . . . . . . . . . . . . . . 15<br />

1.4 Outline and aims <strong>of</strong> the thesis . . . . . . . . . . . . . . . . . . . . . . . 19<br />

2 Introduction to interferometry 23<br />

2.1 A bit <strong>of</strong> history . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23<br />

2.2 Basic principle <strong>of</strong> astronomical interferometry . . . . . . . . . . . . . . 26<br />

2.3 Optical versus millimeter interferometry . . . . . . . . . . . . . . . . . 29<br />

2.4 Introduction to modelling techniques . . . . . . . . . . . . . . . . . . . 34<br />

II Dust in the inner disk 39<br />

3 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong> 41<br />

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

3.2 Model description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44


iv CONTENTS<br />

3.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

3.3.1 The rim shape . . . . . . . . . . . . . . . . . . . . . . . . . . . 51<br />

3.3.2 The rim SED . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

3.3.3 Rim images . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

3.3.4 Grain size distribution . . . . . . . . . . . . . . . . . . . . . . 55<br />

3.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.4.1 The 3µm bump . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

3.4.2 The rim radius . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

3.4.3 LkHa101 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.5 Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

3.6 Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

3.6.1 Evaluation <strong>of</strong> Eq. 3.1 for single grains . . . . . . . . . . . . . . 64<br />

3.6.2 SED . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

3.6.3 Scattering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64<br />

4 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong> 67<br />

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68<br />

4.2 Target stars and <strong>observations</strong> . . . . . . . . . . . . . . . . . . . . . . . 69<br />

4.3 Model description . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

4.3.1 The dust evaporation and the “puffed-up” inner rim . . . . . . . 72<br />

4.3.2 Visibility model . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

4.4 Comparison with the <strong>observations</strong> . . . . . . . . . . . . . . . . . . . . 75<br />

4.4.1 Model parameters . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

4.4.2 MWC 758 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.4.3 VV Ser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76<br />

4.4.4 CQ Tau . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

4.4.5 V1295 Aql . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77<br />

4.4.6 MWC 480 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

4.4.7 AB Aur . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78


CONTENTS v<br />

4.5 Comparison with <strong>pre</strong>vious analysis . . . . . . . . . . . . . . . . . . . . 79<br />

4.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

4.6.1 Presence <strong>of</strong> large grains . . . . . . . . . . . . . . . . . . . . . 81<br />

4.6.2 Inclination and position angle . . . . . . . . . . . . . . . . . . 83<br />

4.6.3 Improving the model constrains . . . . . . . . . . . . . . . . . 84<br />

4.7 Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . 85<br />

5 More on the “puffed–up” inner rim 95<br />

5.1 Fuzzy and sharp rims . . . . . . . . . . . . . . . . . . . . . . . . . . . 97<br />

5.2 New shapes for the rim . . . . . . . . . . . . . . . . . . . . . . . . . . 99<br />

5.3 New observational results . . . . . . . . . . . . . . . . . . . . . . . . . 101<br />

III Gas in the inner disk 105<br />

6 AMBER/VLTI spectroscopy <strong>of</strong> HD 104237 107<br />

6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108<br />

6.2 Observations and data reduction . . . . . . . . . . . . . . . . . . . . . 110<br />

6.3 Results and discussions . . . . . . . . . . . . . . . . . . . . . . . . . . 112<br />

6.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115<br />

IV The structure <strong>of</strong> the outer disk 117<br />

7 The keplerian disk <strong>of</strong> HD 163296 119<br />

7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

7.2 Observations and data reduction . . . . . . . . . . . . . . . . . . . . . 121<br />

7.2.1 PBI <strong>observations</strong> . . . . . . . . . . . . . . . . . . . . . . . . . 121<br />

7.2.2 SMA <strong>observations</strong> . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

7.2.3 VLA <strong>observations</strong> . . . . . . . . . . . . . . . . . . . . . . . . 122<br />

7.3 Observational results . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

7.3.1 Disk morphology and apparent size . . . . . . . . . . . . . . . 123


vi CONTENTS<br />

7.3.2 Disk kinematic . . . . . . . . . . . . . . . . . . . . . . . . . . 124<br />

7.3.3 Free-free contribution and spectral index . . . . . . . . . . . . 128<br />

7.3.4 Disk mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130<br />

7.4 Disk Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131<br />

7.4.1 Continuum emission . . . . . . . . . . . . . . . . . . . . . . . 132<br />

7.4.2 CO emission . . . . . . . . . . . . . . . . . . . . . . . . . . . 133<br />

7.5 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

7.5.1 Method <strong>of</strong> analysis . . . . . . . . . . . . . . . . . . . . . . . . 134<br />

7.5.2 Continuum emission . . . . . . . . . . . . . . . . . . . . . . . 135<br />

7.5.3 CO emission . . . . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

7.6 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138<br />

7.6.1 The disk outer radius . . . . . . . . . . . . . . . . . . . . . . . 139<br />

7.6.2 The gaseous disk . . . . . . . . . . . . . . . . . . . . . . . . . 141<br />

7.7 Summary and conclusions . . . . . . . . . . . . . . . . . . . . . . . . 142<br />

V Conclusions and future prospects 149<br />

8 Summary and Conclusions 151<br />

8.1 The inner disk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

8.2 The outer disk . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154<br />

8.3 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155<br />

9 Future developments 157<br />

9.1 Gas in the inner disk <strong>of</strong> Herbig Ae stars . . . . . . . . . . . . . . . . . 157<br />

9.2 Probing planet formation through interferometric <strong>observations</strong> . . . . . 161<br />

Published Papers 169<br />

Bibliography 171


Part I<br />

Introduction


CHAPTER 1<br />

1.1 Star formation<br />

Scientific overview<br />

Star formation in our galaxy is a continuous process that occurs inside large concentra-<br />

tions <strong>of</strong> cold gas and dust which contain more than 50% <strong>of</strong> the interstellar matter. Char-<br />

acterized by temperatures <strong>of</strong> about 10 K, extents <strong>of</strong> tens <strong>of</strong> parsec and masses between<br />

10 4 and 10 6 M⊙, the Giant Molecular Clouds are sustained against the gravitational<br />

collapse by supersonic turbulent motions (Zuckerman and Evans, 1974) which lead to<br />

strong density inhomogeneity inside the cloud. When the gravity on these smaller scale<br />

condensations wins the gas <strong>pre</strong>ssure, the material collapses and the stellar formation<br />

starts. During the contraction, clouds break in smaller parts following a fragmentation<br />

process that ends with the formation <strong>of</strong> cores when the collapse becomes adiabatic, i.e.,<br />

the gas com<strong>pre</strong>ssion corresponds to an increase in temperature. Before that, the col-<br />

lapse has been isothermal since the gas density was too low to <strong>pre</strong>vent the cooling <strong>of</strong> the<br />

cloud.<br />

Cores generally have extent <strong>of</strong> about 0.1 pc; the gas density is between 10 4 and<br />

10 5 molecules cm −3 and the temperature is similar to the temperature <strong>of</strong> the original<br />

molecular cloud. Given these properties, the typical core mass is <strong>of</strong> few M⊙, higher than


4 Scientific overview<br />

the critical Jeans mass for the gravitational instability (∼1M⊙). The gravity overwhelms<br />

the thermal <strong>pre</strong>ssure and the core collapses in a free-fall time scale <strong>of</strong> few 10 5 years.<br />

While the core rotation has been found to be negligible compared to gravity (Ar-<br />

quilla and Goldsmith, 1986; Caselli et al., 2002), magnetic field and turbulence could<br />

play an important role in the core stability delaying or <strong>pre</strong>venting the star formation.<br />

Many authors think that molecular core are indeed supported against gravity by mag-<br />

netic field (Ostriker et al., 1999; Li et al., 2004); they can collapse only after the mag-<br />

netic field has been dissipated by ambipolar diffusion (Spitzer, 1978) on a time scale<br />

<strong>of</strong> 10 6 yr, an order <strong>of</strong> magnitude higher than the typical free-fall time. Balancing grav-<br />

ity with magnetic energy, critical masses comparable with core masses can be obtained<br />

with a magnetic field <strong>of</strong> about 50µG. Recent measurements <strong>of</strong> magnetic fields in few<br />

dense cores (Crutcher et al., 2006) suggest that most cores are closed to equilibrium,<br />

but with large uncertainties. Moreover there are authors who strongly disagree with this<br />

scheme, and believe that magnetic field has no role in the core collapse (Nakano, 1998).<br />

Turbulent motions can also <strong>pre</strong>vent cores from collapsing. In many cases, the velocities<br />

measured from the width <strong>of</strong> molecular lines are indeed large enough to sustain the core<br />

against the gravity. On the other hand even turbulent motions dissipate in a time com-<br />

parable to the free-fall time (Nakano, 1998) and the core collapse can be only delayed.<br />

If the core rotation can not <strong>pre</strong>vent the collapse, the specific angular momentum<br />

determines the stellar multiplicity <strong>of</strong> the final system and has a strong influence on the<br />

structure <strong>of</strong> the circumstellar environment. Hydrodynamic models showed indeed that<br />

the fragmentation <strong>of</strong> an isothermal core depends strongly on its initial angular momen-<br />

tum (Matsumoto and Hanawa, 2003; Boss and Myhill, 1995) and that that process is<br />

likely to be the dominant mode <strong>of</strong> binary formation (Bodenheimer et al., 2000). For<br />

an exhaustive description <strong>of</strong> the stellar multiplicity see the review <strong>of</strong> Goodwin et al.<br />

(2006).<br />

In the case <strong>of</strong> a slowly rotating sphere <strong>of</strong> gas <strong>of</strong> constant angular velocityΩ, the<br />

collapse is monolithic and leads to a structure composed by four concentric regions<br />

(Terebey, Shu and Cassen, 1984):


1.1 Star formation 5<br />

• in the center <strong>of</strong> the system an accreting core forms with radius R⋆ and increasing<br />

mass M⋆= ˙Macct. Since the collapse is adiabatic the gas <strong>pre</strong>ssure re-establishes<br />

itself trying to reach the hydrostatic equilibrium. Both the gravitational energy<br />

produced during the contraction and the accretion luminosity Lacc= GM⋆ ˙Macc/R⋆<br />

are released by the system that become a protostar;<br />

• an accreting disk forms between R⋆ < R < Rc. The disk outer radius, Rc =<br />

Ω 2 cst 3 /16, is the centrifugal radius defined as the place where the rotational ve-<br />

locity vrot=ΩR 2<br />

in f /R is equal to the free-fall velocity v f f∼ (GM⋆) 1/2 R −1/2 . The<br />

centrifugal radius corresponds to the maximum distance at which the infalling<br />

material impact the disk;<br />

• a spherical envelope in free-fall between Rc


6 Scientific overview<br />

Log νFν<br />

Log νFν<br />

Log νFν<br />

Log νFν<br />

-7<br />

-8<br />

-9<br />

-7<br />

-8<br />

-9<br />

-7<br />

-8<br />

-9<br />

-7<br />

-8<br />

-9<br />

Core<br />

Active<br />

disk<br />

CLASS II<br />

Passive<br />

disk<br />

CLASS III<br />

Remnant<br />

disk<br />

CLASS 0<br />

CLASS I<br />

Protostar<br />

Star<br />

Star<br />

11 12 13 14 15<br />

Log ν (Hz)<br />

zZ<br />

5000 AU<br />

500 AU<br />

Figure 1.1: The figure shows the formation <strong>of</strong> a single star and the evolution <strong>of</strong> the circum-<br />

stellar material. The left panels show the spectral energy distribution <strong>of</strong> the system at different<br />

50 AU<br />

evolutionary stages; the right panels show the corresponding system geometry.<br />

5 AU


1.1 Star formation 7<br />

the mid-infrared. Class II stars are characterized by a SED composed by the protostellar<br />

emission and the flux arising from the circumstellar disk composed by gas and dust;<br />

in this phase most <strong>of</strong> the circumstellar material is in the disk. The SED peaks in the<br />

near infrared and decreases roughly as a power law from the mid infrared to millimeter<br />

wavelengths. Finally, for the Class III objects the emission <strong>of</strong> the circumstellar material<br />

becomes negligible and the SED has a pure photospheric shape.<br />

The protostellar phase ends when the core has either accreted onto the star+disk or<br />

dissipated by winds and jets. At this point the mass <strong>of</strong> the central star is practically<br />

fixed and the mass accretion rate ˙M decreases to very low values (typically ˙M< ∼ 10 −7 M⊙<br />

yr −1 ). The star evolves by slow contraction on the Kelvin-Helmholtz timescale (tKH∼<br />

GM 2 ⋆ /R⋆L⋆) toward the ZAMS, where the stellar luminosity is due to the hydrogen<br />

burning. During this <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> phase, the stellar luminosity is dominated by<br />

the gravitational energy released in the contraction while the accretion luminosity is<br />

negligible.<br />

Pre-<strong>main</strong> <strong>sequence</strong> stars are located on the Herzsprung-Russell diagram in a region<br />

to the right <strong>of</strong> the ZAMS and evolutionary tracks have been computed by several au-<br />

thors (Stahler and Palla, 2005). By comparing the position <strong>of</strong> any individual star to the<br />

evolutionary tracks, it is possible to determine its age and mass (a more direct way to<br />

determine the mass <strong>of</strong> the central star will be discussed in Sec. 1.2). In this way it has<br />

been found that the age <strong>of</strong> <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars range from 10 5 to 10 7 yr. It is also<br />

generally accepted to call T Tauri stars (TTS) the low-mass (M⋆< 1.5M⊙) <strong>pre</strong>-<strong>main</strong> se-<br />

quence stars, while Herbig Ae/Be stars (HAe/Be, since the work <strong>of</strong> George H. Herbig in<br />

1960) are the intermediate mass <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars (1.5M⊙< M⋆< 8M⊙). Higher-<br />

mass stars reach the ZAMS while still accreting and do not have a <strong>pre</strong>-<strong>main</strong>-<strong>sequence</strong><br />

phase (Beuther et al., 2006).<br />

Both TTS and HAe/Be stars show a strong stellar activity in the form <strong>of</strong> continuum<br />

excess emission in the UV and IR, emission lines, strong photometric and spectroscopic<br />

variability, ejection <strong>of</strong> matter as wind and jets (see the review <strong>of</strong> Bally et al., 2006). In<br />

many cases the activity is strictly related to the <strong>pre</strong>sence <strong>of</strong> circumstellar <strong>disks</strong>, which


8 Scientific overview<br />

are the subject <strong>of</strong> this thesis.<br />

1.2 Disk structure in the <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> phase<br />

Circumstellar <strong>disks</strong> are an essential component <strong>of</strong> the process <strong>of</strong> stellar formation.<br />

Nearly all stars are thought to be born with circumstellar <strong>disks</strong> (Beckwith and Sar-<br />

gent, 1996; Hillenbrand et al., 1998) and direct evidence <strong>of</strong> <strong>disks</strong> exist in a increasing<br />

number <strong>of</strong> <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars <strong>of</strong> all masses.<br />

Even if spectroscopically and spatially resolved <strong>observations</strong> <strong>of</strong> TTS and HAe stars<br />

have given a wealth <strong>of</strong> information on circumstellar disk structures, no theory or model<br />

can simultaneously account for all the physical processes occurring in circumstellar<br />

<strong>disks</strong> – viscous accretion onto the central star, mass loss due to outflow, irradiation by<br />

the central star, interaction with the stellar wind and magnetic field, turbulent mixing <strong>of</strong><br />

material, dust grain growth, gradual settling <strong>of</strong> the dust towards the disk mid plane and<br />

planetary formation – and many questions about the structure and evolution <strong>of</strong> circum-<br />

stellar <strong>disks</strong> are still open.<br />

The simple case <strong>of</strong> an accretion disk where the mass accretion rate ˙M is constant<br />

and the viscous accretion is the only source <strong>of</strong> heating has been discussed by Shakura<br />

and Sunyaev (1973), while Lynden-Bell and Pringle (1974) also discussed the time-<br />

dependent case. In these models, viscous stresses within the disk transport material<br />

towards the star and angular momentum to the disk outer regions in a viscous timescale<br />

tν= r 2 /ν, depending on the viscosityν.<br />

Observation <strong>of</strong> a large sample <strong>of</strong> <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars has shown that after the<br />

infall phase is finished (in 10 5 years), the total disk mass decreases with time and most <strong>of</strong><br />

disk material is dispersed after about 10 7 years (Meyer et al., 2006). After Shu (1992),<br />

it has been clear that molecular viscosity alone was too small to account for the disk<br />

lifetime and an additional viscosity should be introduced to explain the <strong>observations</strong>.<br />

To avoid having to solve the problem <strong>of</strong> viscosity in detail, Shakura and Sunyaev<br />

introduced theα<strong>pre</strong>scription in which the viscosity is ex<strong>pre</strong>ssed by the relationν=αhcs


1.2 Disk structure in the <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> phase 9<br />

where h is the <strong>pre</strong>ssure scale height <strong>of</strong> the disk, cs is the sound speed and the parameter<br />

α contains all the uncertainties on the real nature <strong>of</strong> the viscosity. Assuming that the<br />

material orbits with a keplerian rotation around the star and that the disk is in hydrostatic<br />

equilibrium in the vertical direction, the viscous time scale, which is the time to move<br />

matter from r to the star, is<br />

tν∼ 3×10 6<br />

<br />

0.01 r<br />

5/4 yr (1.1)<br />

α 100AU<br />

which, forα = 0.01, is roughly consistent with the fact that <strong>disks</strong> have in general<br />

disappeared by the time stars reach the <strong>main</strong>-<strong>sequence</strong>.<br />

A useful quantity to describe how the mass is distributed in a disk is the surface<br />

densityΣ(r) defined as the integral <strong>of</strong> the mass densityρ(r, z) in the vertical direction<br />

z. In the case <strong>of</strong> a steady disk, i.e. characterized by a constant mass accretion rate,<br />

the equation <strong>of</strong> mass conservation into the disk implies thatΣ∝r −3/4 α −1 . As it will<br />

be discussed in Chapter 7, high angular resolution millimetric <strong>observations</strong> <strong>of</strong> a large<br />

number <strong>of</strong> HAe and TTS <strong>disks</strong> indicate that the surface density generally decreases with<br />

the distance from the central star as the power low r −0.8 − r −1 , indicating thus an almost<br />

constant value <strong>of</strong>αthrough all the disk.<br />

Among the many disk instabilities suggested to explain the observed value <strong>of</strong>α(Pa-<br />

paloizou and Lin, 1995; Balbus, 2003), the magnetorotational instability (MRI), <strong>pre</strong>sent<br />

in a weakly magnetized disk, is at the <strong>pre</strong>sent the most accepted mechanism to drive<br />

turbulent viscosity in <strong>disks</strong> and transport angular momentum outwards (Velikhov, 1959;<br />

Chandrasekhar, 1960; Balbus and Hawley, 1991; Stone and Pringle, 2001). However,<br />

many authors claim that neither thermal ionization, cosmic ray ionization, nor X-rays<br />

are able to provide a sufficient number <strong>of</strong> free electron to have MRI operating in the<br />

disk-mid plane where most <strong>of</strong> the disk mass and angular momentum are confined (Gam-<br />

mie, 1996; Glassgold et al., 1997; Ilgner and Nelson, 2006). Indeed, in the case <strong>of</strong> TTS<br />

<strong>disks</strong>, Gammie (1996) has shown that the ionization fraction is sufficient to couple the<br />

magnetic field to the gas only very close to the central star (r< ∼ 0.1 AU) or in a superficial<br />

layer <strong>of</strong> the disk. Between these active layers, there is a dead layer corresponding to the


10 Scientific overview<br />

disk regions close to the mid plane where the magnetic field and the gas are not coupled<br />

and the MRI can not operate.<br />

A second efficient process <strong>of</strong> angular momentum transfer that may operate in <strong>pre</strong>-<br />

<strong>main</strong> <strong>sequence</strong> disk is gravitational instability (Tomley et al., 1991; Laughlin and Bo-<br />

denheimer, 1994; Bertin, 1997; Bertin and Lodato, 1999; Lodato and Bertin, 2001 and<br />

2003; see also the recent review <strong>of</strong> Durisen et al., 2006). Gravitational instability may<br />

dominate the angular momentum transfer in the earliest stages <strong>of</strong> the stellar evolution,<br />

when the mass <strong>of</strong> the circumstellar disk may be comparable with that <strong>of</strong> the central<br />

star. In the more evolved evolutionary phase typical <strong>of</strong> TTS and HAe stars, most <strong>of</strong> the<br />

circumstellar mass is accreted in the central star with the result that the measured disk<br />

masses are usually below the value required for gravitational instability. The existing<br />

estimates <strong>of</strong> the mass <strong>of</strong> the <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> <strong>disks</strong> are in most cases obtained from<br />

the measure <strong>of</strong> the optically thin dust emission at millimeter wavelength. This method<br />

is however based on strong assumptions on the dust opacity and it can be affected by<br />

large uncertainties (see also the discussion in Chapter 7). Thirty years after the work <strong>of</strong><br />

Shakura and Sunyaev, the nature <strong>of</strong> the disk viscosity is still matter <strong>of</strong> debate.<br />

For anα-disk (a disk based on theα<strong>pre</strong>scription), the temperature at each radius<br />

r is Td ∝ r −3/4 , obtained by equating the heating rate, due to viscous dissipation, to<br />

the cooling rate, computed assuming that the disk radiates as a black body at the local<br />

temperature Td. However, for most disk around TTS and HAe stars the heating by<br />

stellar radiation is dominant over viscous heating and the disk temperature depends on<br />

the angle at which the stellar radiation strikes the disk surface. Since the circumstellar<br />

material is composed by a large amount <strong>of</strong> dust and gas, all the stellar flux intercepted<br />

by the disk is absorbed and re-emitted at longer wavelength, between the near infrared<br />

and the millimeter. Disk where the stellar radiation is the dominant source <strong>of</strong> heating<br />

are refereed to as reprocessing <strong>disks</strong>. Kenyon and Hartmann (1987) recognized that<br />

reprocessing <strong>disks</strong> in hydrostatic equilibrium are flared, i.e., the disk opening angle<br />

increases with r. The flaring geometry allows the disk to capture a significant portion<br />

<strong>of</strong> the stellar radiation at large radii where the disk is cool, increasing the mid to far


1.2 Disk structure in the <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> phase 11<br />

infrared emission and explaining the strong infrared excess observed around TTS and<br />

HAe stars. The resulting temperature pr<strong>of</strong>ile is flatter than in aα-disk and typically<br />

Td∝ r −1/2 . At any given radius, the vertical dependence <strong>of</strong> the density on z is obtained<br />

by integrating the equation <strong>of</strong> hydrostatic equilibrium, which gives<br />

ρ(r, z)=ρ(r, 0) exp(−z 2 /2h 2 ), (1.2)<br />

where the disk <strong>pre</strong>ssure scale is h∝r 9/7 . The density decreases rapidly with z, so that<br />

the matter is concentrated on the disk mid plane.<br />

The SED <strong>of</strong> a disk can be computed in first approximation assuming that each disk<br />

annulus emits as a black body at the effective temperature Td, so that the observed flux<br />

at any given frequencyνis given by the relation<br />

Fν= cosθ<br />

D 2<br />

Rd<br />

Rin<br />

Bν(Td)(1−e −τν )2πrdr, (1.3)<br />

where Rin and Rd are the inner and outer disk radius, D is the distance from the observer,<br />

θ is the disk inclination with respect to the line <strong>of</strong> sight (θ=0means face-on <strong>disks</strong>), Bν<br />

is the black body emission. The optical depthτν is<br />

τν= 1<br />

cosθ kνΣ(r) (1.4)<br />

where kν is the total (gas+dust) opacity at the frequencyν. For conditions typical <strong>of</strong><br />

HAe and TTS, the <strong>main</strong> portion <strong>of</strong> the energy is emitted in the wavelength range from<br />

∼1µm to 100µm, where the SED usually shows a slow dependence on the wavelength<br />

(see Fig. 1.2). At longer wavelengths the SED is characterized by a power law pr<strong>of</strong>ile<br />

with a slope depending on the grain properties and optical depth as discussed in Sec. 1.3.<br />

As direct con<strong>sequence</strong> <strong>of</strong> the radial monotonic decrease <strong>of</strong> the disk temperature,<br />

<strong>observations</strong> at different wavelengths map disk emission at distances from the central<br />

star which increase with the observational wavelength. Between 1µm and 7µm, the<br />

emission is dominated by the disk regions very close to the central star (r< ∼ 0.5 AU)<br />

and characterized by temperatures higher than about 500–1000 K. The study <strong>of</strong> this<br />

innermost region <strong>of</strong> the disk is particularly interesting because it is the locus where<br />

the gas is supposed to accrete on the star, where the interaction between the stellar


12 Scientific overview<br />

Figure 1.2: From Dullemond et al. (2006). Build-up <strong>of</strong> the SED <strong>of</strong> a flaring circumstellar disk<br />

and the origin <strong>of</strong> various components: the near infrared bump is supposed to originate in the<br />

puffed-up inner rim, the infrared dust features (as the silicate ones between 10µm and 20µm)<br />

from the warm surface layer, and the underlying continuum from the deeper and cooler disk<br />

regions. Typically the near and mid-infrared emission comes from small radii, while the far-<br />

infrared and the millimeter emission come from the outer disk regions.<br />

radiation and the disk is stronger and where the magnetic field is supposed to drive the<br />

<strong>of</strong>ten observed outflows and jets that propagate perpendicularly to the disk plane. In the<br />

innermost region <strong>of</strong> the disk, where the temperature is higher than∼1500 K, most <strong>of</strong> the<br />

dust sublimates (Pollack et al., 1994) and the opacity undergoes a strong discontinuity<br />

(Natta et al., 2001; Dullemond et al., 2001). If the gas inside the dust sublimation radius


1.2 Disk structure in the <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> phase 13<br />

absorbs a negligible amount <strong>of</strong> stellar radiation, the dust inner rim is perpendicularly<br />

illuminated by the central star and is expected to be hotter and higher, i.e. with an<br />

higher <strong>pre</strong>ssure scale, than a standard reprocessing disk. The resulting puffed-up inner<br />

rim dominates the near-infrared emission and naturally explains the strong near-infrared<br />

excess observed in most <strong>of</strong> the SEDs <strong>of</strong> HAe and TTS stars. The <strong>pre</strong>sence <strong>of</strong> a puffed-up<br />

inner rim can influence the structure <strong>of</strong> the outer regions <strong>of</strong> the disk shadowing the dust<br />

from the direct stellar radiation. Meeus et al. (2001, 2003) divided the SEDs <strong>of</strong> HAe<br />

stars into two groups: those with strong far-infrared flux (group I) and those with weak<br />

infrared flux (group II), suggesting that the latter are characterized by a disk that lies<br />

in the shadow <strong>of</strong> the puffed-up inner rim. Although the real existence <strong>of</strong> the puffed-up<br />

inner rim is still under debate (Vinkovic et al., 2006), model <strong>pre</strong>dictions are in good<br />

agreement with the spatially resolved near infrared interferometric <strong>observations</strong> <strong>of</strong> a<br />

number <strong>of</strong> HAe and TTS stars (see the review <strong>of</strong> Millan-Gabet et al. 2006). The second<br />

part <strong>of</strong> the thesis (Chapters 3, 4 and 5) is dedicated to the discussion <strong>of</strong> the structure <strong>of</strong><br />

the puffed-up inner rim and to the analysis <strong>of</strong> interferometric <strong>observations</strong> <strong>of</strong> the inner<br />

disk.<br />

The structure <strong>of</strong> the gaseous disk inside the dust sublimation radius is still poorly<br />

known (see the review <strong>of</strong> Najita et al., 2006). Existing high resolution spectroscopic<br />

<strong>observations</strong> <strong>of</strong> CO and water near infrared emission lines indicate that the gas extends<br />

very close to the central star, probably until the distance at which the gaseous disk<br />

is truncated by the stellar magnetic field lines (Shu et al., 1994). Inside this point,<br />

typically located at few stellar radii, the gas moves along the magnetic field lines and<br />

accretes on the central star emitting the UV radiation usually observed in HAe and TTS<br />

spectra as UV veiling (Hartman et al., 1998; Bouvier et al., 2006). The accreting gas<br />

is also supposed to be responsible for the hydrogen recombination lines (Hα, Brγ, Paβ,<br />

etc) characteristic <strong>of</strong> most HAe and TTS. Both the amount <strong>of</strong> veiling and the hydrogen<br />

line intensities depend on the accreting gas density and are used to measure the mass<br />

accretion rate (Muzerolle et al., 2001; Natta et al., 2006). Chapter 6 is dedicated to<br />

the spectrally resolved near infrared interferometric <strong>observations</strong> <strong>of</strong> the gas in the dust-


14 Scientific overview<br />

depleted inner disk.<br />

The mid-infrared part <strong>of</strong> the SED comes from disk regions at distances between<br />

∼5 AU and∼50 AU, where the disk opacity and the thermal emission are dominated<br />

by the dust. The dust density is generally high enough that the disk is optically thick<br />

to its own emitted radiation from near to far infrared and the resulting SED is in first<br />

approximation the sum <strong>of</strong> many black bodies with different effective temperatures. A<br />

closer look at the physics <strong>of</strong> a reprocessing disk reveals however that the disk surface<br />

temperature is generally higher than the effective temperature Td (Calvet et al., 1991;<br />

Chiang and Goldreich, 1997). Dust grains in the surface layers are directly exposed<br />

to the stellar radiation and are hotter than dust grains in the disk interior which are<br />

heated only by the diffused infrared emission <strong>of</strong> the disk itself. As a con<strong>sequence</strong>, the<br />

thermal emission from this surface layer produces emission features in good agreement<br />

with nearly all the TTS and HAe stars spectra (Meeus et al., 2001; Kessler-Silacci et<br />

al., 2006). The analysis <strong>of</strong> spectroscopic features, <strong>main</strong>ly in mid-infrared, give a lot <strong>of</strong><br />

information <strong>of</strong> the dust composition and size and will be discussed in more detail in the<br />

next section.<br />

Finally, particularly interesting for the study <strong>of</strong> circumstellar disk is the range <strong>of</strong> long<br />

wavelengths, from sub-millimeter to radio, where the dust opacity can be approximated<br />

by a power law, kν∝λ −β (Beckwith and Sargent, 1991), and the disk is optically thin to<br />

its own thermal radiation (τν < ∼ 1). In this case, the observed flux Fν is directly related<br />

to the dust surface densityΣ(see Eq. 1.3 and Eq. 1.4) and both the disk mass and the<br />

parameterβcan be derived from <strong>observations</strong> at different wavelengths (Mannings and<br />

Sargent, 1997; Testi et al., 2001 and 2003; Natta et al., 2004). Millimeter interferome-<br />

ters allow to resolve spatially the continuum <strong>disks</strong> <strong>observations</strong> and both spatially and<br />

spectrally the molecular lines. Using CO vibrational transitions, it has been confirmed<br />

that in most <strong>of</strong> the observed TTS and HAe stars the gas is orbiting the central star with<br />

velocity patterns in good agreement with keplerian rotation. In this case, the gas veloc-<br />

ity gives a direct measure <strong>of</strong> the mass <strong>of</strong> the central object, which is independent <strong>of</strong> the<br />

position <strong>of</strong> the star on the HR diagram and <strong>of</strong> the evolutionary tracks. Taking advantage


1.3 Grain growth and planetary formation 15<br />

<strong>of</strong> varying opacity <strong>of</strong> the different transitions <strong>of</strong> CO and its isotopes, is possible to probe<br />

disk properties at different depth and calculate the relative CO abundance. For some<br />

nearby stars, the continuum <strong>observations</strong> show asymmetry in the radial dust distribu-<br />

tion which can be caused by dynamical perturbations <strong>of</strong> unknown companions or giant<br />

planets (Pietú et al., 2005; Corder et al., 2005). All these points will be discussed in<br />

more detail in Chapter 7 which describes new millimeter <strong>observations</strong> <strong>of</strong> the HAe star<br />

HD 163296.<br />

1.3 Grain growth and planetary formation<br />

Circumstellar <strong>disks</strong> around TTS and HAe are composed for about 99% <strong>of</strong> gas and for<br />

only 1% <strong>of</strong> solid dust grains. Nevertheless, as seen in the <strong>pre</strong>vious section, the dust<br />

component is responsible for the observed disk thermal emission, which re<strong>main</strong>s a key<br />

diagnostic tool to detect <strong>disks</strong> and characterize their structure. Dust grains dominate the<br />

disk opacity, control the disk temperature and geometry, regulate the chemical reactions<br />

<strong>of</strong> molecular species as CO and H2O in the colder parts <strong>of</strong> <strong>disks</strong>, and are the building<br />

block for the formation <strong>of</strong> planetesimals and planets within <strong>disks</strong>.<br />

From the diffuse interstellar medium to Class I objects, dust grains are, and re<strong>main</strong>,<br />

a mixture <strong>of</strong> amorphous silicates and carbons, with a size distribution from∼ 0.01µm<br />

to∼ 0.2µm (Draine et al., 2003; Bianchi et al., 2003; Kessler-Salacci et al., 2005). The<br />

situation changes drastically in circumstellar <strong>disks</strong> in which they grow many order <strong>of</strong><br />

magnitude in size forming kilometers-size body (planetesimals) and even planets.<br />

While the growth from sub-micron to micron sizes is probably driven by collisional<br />

aggregation in Brownian velocity fields (Blum, 2004), it is still under debate if the for-<br />

mation <strong>of</strong> meter size objects occurs through collisional aggregation (see the review <strong>of</strong><br />

Dominik et al., 2006) or gravitational instabilities in the over-dense dust regions <strong>of</strong> the<br />

disk (Youdin and Shu, 2002; Rice et al., 2004 and 2006). Modelling the grain growth<br />

requires a complex analysis <strong>of</strong> the dust and gas coupling, dust sticking properties and<br />

velocity fields within the disk. While growing in mass, particles start to decouple from


16 Scientific overview<br />

the gas and sediment on the disk mid plane, increasing the collisional efficiency and<br />

probably reaching a dynamical equilibrium in which large particles form smaller ones<br />

through destructive collisions and vice versa. Given the extreme complexity <strong>of</strong> the<br />

problem, different models gives <strong>of</strong>ten contradictory results and the formation <strong>of</strong> plan-<br />

etesimals re<strong>main</strong> at the moment an open question.<br />

As con<strong>sequence</strong> <strong>of</strong> the grain growth and settling, circumstellar <strong>disks</strong> are <strong>pre</strong>dicted<br />

to be vertically stratified with small grains in the uppermost layers <strong>of</strong> the disk and large<br />

grains in the disk mid plane. Furthermore, grain growth should be much slower in<br />

the outer regions <strong>of</strong> the disk than in the regions closer to the star, which should be<br />

completely depleted <strong>of</strong> small grains. The situation becomes even more complicated if<br />

we consider that inward radial circulations, due for example to the disk viscosity and/or<br />

to the dust drag due to the gas, bring material close to the star where dust sublimates.<br />

At the same time, opposite radial circulations, due, for example, to the stellar wind<br />

or to the magnetorotational instability, move the inner material outward permitting the<br />

ricondensation <strong>of</strong> grains and the formation <strong>of</strong> crystalline structures (Gail, 2004).<br />

Despite all the theoretical difficulties in understanding the dust evolution, grain prop-<br />

erties in <strong>disks</strong> can be explored with a large number <strong>of</strong> techniques and over large range<br />

<strong>of</strong> wavelengths. It is worth to stress that since the <strong>disks</strong> around <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars<br />

have high optical depth, <strong>observations</strong> at optical and infrared wavelengths allow to ob-<br />

tain information only on grains located in the very superficial layer <strong>of</strong> the disk where<br />

τ≃1. Such <strong>observations</strong> are thus sensitive only to a very small fraction <strong>of</strong> the dust mass<br />

which is concentrated on the disk mid plane (see Eq. 1.2). An important exception to<br />

this general rule is re<strong>pre</strong>sented by near-infrared interferometric <strong>observations</strong> which are<br />

able to spatially resolve the regions close to the star where the dusty disk is truncated<br />

by the dust evaporation. We will show in Chapter 4 that in this case, the measure <strong>of</strong> the<br />

location <strong>of</strong> the dust evaporation radius gives information on dust grains <strong>pre</strong>sent in the<br />

disk mid plane.<br />

Particularly important to the study <strong>of</strong> the chemical composition, size and structure<br />

<strong>of</strong> the dust grains is the mid-infrared spectral region (between∼ 3µm and∼ 100µm),


1.3 Grain growth and planetary formation 17<br />

rich in vibrational resonances <strong>of</strong> many dust species. In <strong>disks</strong> around HAe and TTS,<br />

this dust features are generally seen in emission and are thought to originate in the<br />

optically thin disk surface layer, directly illuminated by the stellar radiation (Calvet et<br />

al., 1991; Chiang and Goldreich, 1997). The more common features observed around<br />

TTS and HAe stars are those <strong>of</strong> amorphous and crystalline silicates, located around<br />

10µm and 20µm. Recent Spitzer data (e.g. Furlan et al., 2006) show a large variation<br />

in the strength and shape <strong>of</strong> this features, confirming the correlation firstly identified by<br />

van Boekel et al. (2005): weaker is the feature flatter is its shape. This is inter<strong>pre</strong>ted in<br />

terms <strong>of</strong> the growth <strong>of</strong> silicate grains from sub-micron to micron sizes; if grains grow<br />

further, the silicate emission disappears.<br />

The <strong>pre</strong>sence <strong>of</strong> crystalline grains around HAe and TTS stars implies that the pris-<br />

tine material underwent a drastic change in its structure due to some, not yet known,<br />

energetic process. Using spectrally resolved mid-infrared interferometry, van Boekel et<br />

al. (2004) showed for the first time that the disk surface inside 1-2 AU around three HAe<br />

stars is much more crystalline than the outer disk surface, supporting the idea that crys-<br />

tallization occurs in the inner disk (Gail, 2004). However crystalline silicates are also<br />

<strong>pre</strong>sent at larger distances (∼ 20 AU), and in the comets in our Solar System, suggesting<br />

that the radial drift can have an effective role in the dust evolution within circumstellar<br />

<strong>disks</strong>.<br />

At millimetric and centimetric wavelengths <strong>disks</strong> become optically thin to their own<br />

thermal radiation and the <strong>observations</strong> can probe the bulk <strong>of</strong> the dust mass concentrated<br />

on the disk mid plane. The SED <strong>of</strong> the continuum emission at these wavelengths can be<br />

directly related to the grain emissivity which, in turn, depends on the grain properties.<br />

The millimetric dust opacity can be approximated by a power law, k∼λ −β , where the<br />

exponentβdepends strongly on the grain size distribution and on its upper limit. In<br />

a size distribution composed only by grains larger than the observing wavelength, i.e.<br />

centimeter size grains, the dust opacity is grey andβ=0, whileβis close or larger than<br />

unity only for sub-millimeter grains (see Fig. 1.3). If the disk emission is optically thin<br />

(τν≪1) and the Reyleigh-Jeans approximation is verified (Bν(T)∝λ −2 ), the emitted


18 Scientific overview<br />

Figure 1.3: From Natta et al. (2006). Opacity indexβfor grains with a size distribution n(a)∝<br />

a −q between amin and amax. The indexβis computed between 1 and 7 mm for compact segregated<br />

spheres <strong>of</strong> olivine, organic materials, water ice (Pollack et al., 1994), and plotted as function <strong>of</strong><br />

amax. Different curves correspond to different values <strong>of</strong> q, as labelled. In all cases, amin≪ 1 mm.<br />

The dotted curve showsβfor grains with single size amax. The histogram on the right shows the<br />

distribution <strong>of</strong> the values <strong>of</strong>βderived from millimeter <strong>observations</strong>. About 60% <strong>of</strong> the objects<br />

haveβ≤1, and amax > ∼ 1 cm.<br />

flux is ex<strong>pre</strong>ssed by the power law Fν ∝ λ −α . Using Eq. 1.3 and Eq. 1.4 it can be<br />

demonstrated using that in this caseα=β+2. This simple relation allow to translate<br />

the observed value <strong>of</strong>αinto the opacity power low indexβ. To verify the fundamental<br />

assumption that the millimetric disk emission is optically thin, the emission itself must<br />

be spatially resolved and the disk size measured (Testi et al., 2001 and 2003).<br />

There are at <strong>pre</strong>sent several <strong>disks</strong> whose millimeter emission has been resolved with<br />

interferometers (Natta et al., 2004; Rodmann et al., 2006) and in Chapter 7 we will<br />

<strong>pre</strong>sent new multi-wavelengths <strong>observations</strong> <strong>of</strong> the HAe star HD 163296. The obtained


1.4 Outline and aims <strong>of</strong> the thesis 19<br />

values <strong>of</strong>βvaries from 1.6 (very similar to the interstellar medium and Class 0/I objects)<br />

to 0.5, with most <strong>of</strong> the objects withβ< ∼ 1, corresponding to the <strong>pre</strong>sence <strong>of</strong> centimeter<br />

size grains (Natta et al., 2004; Draine, 2006).<br />

In order to constrain dust growth models, the next step is to check for variations <strong>of</strong><br />

the dust properties as a function <strong>of</strong> the radius. Our observation <strong>of</strong> HD 163296 are the<br />

first attempt to obtain the multi wavelengths high angular resolution and high sensitivity<br />

<strong>observations</strong> required for such analysis and we refer to the Chapter 7 for the results.<br />

The major actual limitation to understand planetary formation is that none <strong>of</strong> the<br />

discussed techniques is sensitive to bodies larger than few centimeters. In principle,<br />

planetesimals <strong>of</strong> few kilometers in size could be detected through the dynamical pertur-<br />

bations exerted on the surrounding material. Particularly interesting is the discovery <strong>of</strong><br />

clumps in the dust density distribution in the disk around the close-by HAe star AB Aur<br />

(Corder et al., 2005; Pietú et al., 2005). On the other hand, the detection <strong>of</strong> meter-size<br />

bodies is at the moment practically impossible.<br />

1.4 Outline and aims <strong>of</strong> the thesis<br />

The aim <strong>of</strong> this thesis is to study the structure <strong>of</strong> circumstellar <strong>disks</strong> around <strong>pre</strong>-<strong>main</strong><br />

<strong>sequence</strong> stars. Taking advantage <strong>of</strong> high spatial resolution interferometric <strong>observations</strong><br />

we have focused our attention on the gas and dust properties both in the inner disk, at<br />

fractions <strong>of</strong> AU from the central star, and in the outer disk regions, at R>50 AU.<br />

The thesis is divided into five parts: a general introduction and a general review <strong>of</strong><br />

astronomical interferometry (this Chapter and Chapter 2), a study <strong>of</strong> the dust in the inner<br />

disk (Chapter 3, 4, 5), a study <strong>of</strong> the gas in the inner disk (Chapter 6), a study <strong>of</strong> gas and<br />

dust in the outer disk (Chapter 7) and the final remarks (Chapter 8 and 9).<br />

Chapter 2 describes the basis <strong>of</strong> interferometry, similarities and differences between<br />

the near-infrared and millimeter interferometry and a <strong>pre</strong>sentation <strong>of</strong> the modelling tech-<br />

niques used to analyse interferometric <strong>observations</strong>.<br />

In Chapter 3 we describe the structure <strong>of</strong> the innermost part <strong>of</strong> circumstellar <strong>disks</strong>.


20 Scientific overview<br />

We propose a model in which the dusty component <strong>of</strong> the disk is truncated by the dust<br />

evaporation forming a puffed-up inner rim. Introducing the dependence <strong>of</strong> the dust<br />

evaporation temperature on the gas density, we <strong>pre</strong>dict that the bright side <strong>of</strong> the rim<br />

is curved, rather than vertical, as expected when a constant evaporation temperature<br />

is assumed. We demonstrate that both the radial position <strong>of</strong> the rim and the emitted<br />

flux depend on the dust grain size. We finally compute synthetic images <strong>of</strong> the curved<br />

rim seen under different inclinations and show how images become more and more<br />

asymmetric for increasing inclinations.<br />

Chapter 4 describes the analysis <strong>of</strong> interferometric <strong>observations</strong> <strong>of</strong> six HAe stars in<br />

the framework <strong>of</strong> our puffed-up inner rim model. Assuming that astronomical silicates<br />

are the most abundant dust component, we found that in four cases dust grains larger<br />

than few microns are required by, or consistent with the <strong>observations</strong>. We constrain the<br />

inclination <strong>of</strong> the inner disk and, where possible, compare our results with the values <strong>of</strong><br />

the outer disk existing in the literature.<br />

In Chapter 5 we discuss new developments <strong>of</strong> the puffed-up inner rim model that<br />

followed the publication <strong>of</strong> the results <strong>pre</strong>sented in Chapter 3.<br />

In Chapter 6 we discuss the first results <strong>of</strong> spectrally resolved interferometric obser-<br />

vation (VLTI/AMBER) <strong>of</strong> the Brγ emission <strong>of</strong> the HAe star HD 104237. We point out<br />

that the line emission and the continuum infrared emission, arising from the puffed-up<br />

inner rim, have similar size scales. We conclude that the line emission is not compatible<br />

with standard magnetospheric accretion models and most likely originate in a compact<br />

wind, starting from a region at 0.2-0.5 AU from the central star.<br />

Chapter 7 describes recent multi wavelengths interferometric <strong>observations</strong> <strong>of</strong> the<br />

HAe star HD 163296, performed with the Sub-Millimeter Array, the Plateau de Bure<br />

Interferometer and the Vary Large Array, between 0.87 mm and 7 mm. Molecular CO<br />

emissions show a very large circumstellar disk with a velocity pattern in good agree-<br />

ment with Keplerian rotation. The mass <strong>of</strong> the central star and the disk parameters are<br />

determined. Comparing the CO and continuum emission, we suggest that the dust prop-<br />

erties undergo a sharp discontinuity at about 200 AU from the star and discuss possible


1.4 Outline and aims <strong>of</strong> the thesis 21<br />

explanations.<br />

Chapter 8 summarizes the <strong>main</strong> results discussed in the thesis while in the Chapter 9<br />

we discuss possible future new <strong>observations</strong> and developments.


CHAPTER 2<br />

2.1 A bit <strong>of</strong> history<br />

Introduction to interferometry<br />

Interferometry is based on wave properties <strong>of</strong> light as first observed by Thomas Young<br />

in 1803 through his famous two-slit experiment: if monochromatic light from a distant<br />

point source propagates through two-slits, see Fig. 2.1, the resulting illumination pattern<br />

is composed by bright and dark fringes, due respectively to the constructive and destruc-<br />

tive interference between the secondary waves produced along the slits. The condition<br />

for constructive interference implies that the fringe spacing∆Θ <strong>of</strong> the intensity distri-<br />

bution is inversely proportional to the projected separation between the two slits, called<br />

baseline B, in units <strong>of</strong> the radiation wavelengthλ:<br />

∆Θ= λ<br />

. (2.1)<br />

B<br />

The corresponding fringe spatial frequency (fringes per unit angle) is thus<br />

u= B<br />

. (2.2)<br />

λ<br />

Imagine now another point-source <strong>of</strong> light (<strong>of</strong> equal brightness, but incoherent with the<br />

first) located at an angle <strong>of</strong>θ=λ/2B from the first source (see right panel <strong>of</strong> Fig. 2.1).


24 Introduction to interferometry<br />

baseline=b<br />

Point source<br />

at infinity<br />

Incoming plane waves<br />

∆θ =<br />

2 slits<br />

Fringe spacing<br />

λ /b radians<br />

Interference pattern<br />

(Visibility = 1)<br />

λ<br />

= Wavelength<br />

Point sources<br />

at infinity<br />

separated by<br />

1/2 the fringe<br />

spacing<br />

2 sine waves<br />

destructively<br />

interfere<br />

(Visibility = 0)<br />

Figure 2.1: From Monnier (2003). Young’s two-slit interference experiment (monochromatic<br />

light) is <strong>pre</strong>sented to illustrate the basic principle behind stellar interferometry. On the left is the<br />

case for a single point-source, on the right a double source with an angular separation half <strong>of</strong> the<br />

fringes spacing. The interference patterns shown re<strong>pre</strong>sent the intensity distribution.<br />

In this case, the two illumination patters are out <strong>of</strong> phase <strong>of</strong> 180 ◦ , cancelling each other<br />

out and <strong>pre</strong>senting an uniformly illuminated screen. This simple experiment shows a<br />

fundamental property <strong>of</strong> interference, namely that the contrast <strong>of</strong> fringes is a function<br />

<strong>of</strong> the geometry <strong>of</strong> the source.<br />

The application <strong>of</strong> interferometry in astronomy has been firstly proposed by Fizeau<br />

in 1868, who suggested to measure the fringes pattern produced by the stellar light to


2.1 A bit <strong>of</strong> history 25<br />

measure stellar diameters. The practical implementation <strong>of</strong> this technique arrived many<br />

years later with Michelson who, in 1891, measured the angular diameter <strong>of</strong> Jupiter’s<br />

satellites. For the following thirty years astronomical interferometry stalled; only in<br />

1921, Michelson and Pease successfully measured the diameter <strong>of</strong> the super-giant star<br />

Betelgeuse using an interferometer at the Mount Wilson Observatory. While techni-<br />

cal difficulties delayed the growth <strong>of</strong> optical interferometry until the mid-1970’s, radio<br />

interferometry had a fast development after the second world war. In 1946, Ryle and<br />

Vonderg constructed a radio analogue <strong>of</strong> the Michelson interferometer and soon discov-<br />

ered a number <strong>of</strong> new cosmic radio sources. During 1950s and 60s, Ryle and Hewish<br />

developed the aperture synthesis method at radio wavelengths which allows the source<br />

image reconstruction combining the signals from many different pairs <strong>of</strong> antennas. In<br />

the 70s, they applied this method to reproduce a one-mile effective aperture using an ar-<br />

ray <strong>of</strong> three antennas with a diameter <strong>of</strong> 60 feet each. A common technique in aperture<br />

synthesis imaging is to use the rotation <strong>of</strong> the Earth to increase the number <strong>of</strong> different<br />

antenna pairs included in an observation. In the last forty years, aperture synthesis has<br />

been widely applied in radio interferometry with the result that astronomical images <strong>of</strong><br />

good quality can be now obtained (see Chapter 7)<br />

Despite the improvement in the radio do<strong>main</strong>, optical interferometry has only been<br />

developed from the mid-1970’s with the advent <strong>of</strong> new technology: detectors, actuators,<br />

servo-control, etc. Labeyrie (1975) was the first to directly combine the light from two<br />

separated small telescopes at the Nice observatory. In the 1980s, the aperture synthe-<br />

sis technique was extended to visible and infrared astronomy by the Cavendish Astro-<br />

physics Group, providing the first very high resolution images <strong>of</strong> nearby stars. These<br />

development allowed the planning and construction in the mid-90s <strong>of</strong> several optical in-<br />

terferometers. In particular, with the ESO Very Large Telescope Interferometer (VLTI),<br />

the Keck Interferometer (KI) and the CHARA array, optical interferometry is becoming<br />

a familiar technique for an increasing number <strong>of</strong> astronomers.


26 Introduction to interferometry<br />

2.2 Basic principle <strong>of</strong> astronomical interferometry<br />

A fundamental property <strong>of</strong> interferometers is the angular resolution. Classical diffrac-<br />

tion theory has established the Reyleigh Criterion for defining the (diffraction-limited)<br />

resolution∆θ <strong>of</strong> a filled circular aperture <strong>of</strong> diameter D, at the observational wavelength<br />

λ:<br />

∆θ=1.22 λ<br />

D<br />

rad. (2.3)<br />

Considering a binary star system, this criterion corresponds to the angular separation on<br />

the sky where one stellar component is centered on the first null <strong>of</strong> the diffraction pattern<br />

<strong>of</strong> the other; the binary is then said resolved. A similar criterion can be defined for an<br />

interferometer: the same binary system is resolved by an interferometer if the fringe<br />

contrast goes to zero at the longest baseline. As shown in Fig. 2.1, this correspond to<br />

the angular separation∆θ=λ/2B.<br />

The fringe contrast is historically called visibility and, for the simple two-slit inter-<br />

ferometer considered here, can be written as<br />

|V|= Imax− Imin<br />

Imax+ Imin<br />

= Fringe amplitude<br />

, (2.4)<br />

Average Intensity<br />

where Imin and Imax denote the minimum and maximum intensity <strong>of</strong> the fringes. From<br />

this definition, it follows that the left and the right fringe patterns <strong>of</strong> Fig. 2.1 have vis-<br />

ibility <strong>of</strong> 1 and 0, corresponding respectively to non resolved and resolved sources.<br />

Moreover, if the source is shifted with respect to the optical axis <strong>of</strong> the system (see the<br />

right panel <strong>of</strong> Fig. 2.1), the corresponding fringe patter is also shifted by the same an-<br />

gleφ, called phase. Visibility and phase are usually ex<strong>pre</strong>ssed together as the complex<br />

visibility V=|V|e iφ , which completely defines a pattern <strong>of</strong> interference fringes.<br />

The importance <strong>of</strong> measuring complex visibilities derives from the van Cittert–<br />

Zernike theorem (see Thompson et al., 2001 for a complete discussion) which relates<br />

the fringe contrast to the Fourier components <strong>of</strong> the impinging brightness distribution.<br />

In particular, the theorem states that for spatially incoherent sources in the far field,<br />

the value <strong>of</strong> the visibility V, measured with a baseline B, is equal to the normalized


2.2 Basic principle <strong>of</strong> astronomical interferometry 27<br />

Figure 2.2: Schematic re<strong>pre</strong>sentation <strong>of</strong> an interferometer composed by the telescopes or anten-<br />

nas. If the source brightness distribution has a small angular extension, it can be approximated<br />

with a function I(l,m) in a plane tangential to the celestial sphere. The corresponding spatial<br />

frequency are measured in the uv plane as described in the text.<br />

Fourier transform <strong>of</strong> the sky brightness distribution corresponding to the fringe spatial<br />

frequency u=B/λ rad −1 ; the phase <strong>of</strong> the fringe pattern is also equal to the Fourier<br />

phase <strong>of</strong> the same spatial frequency component.<br />

In the real case <strong>of</strong> a bi-dimensional astronomical source whose radiation comes from<br />

a small portion <strong>of</strong> the sky, it is possible to define a system <strong>of</strong> Cartesian coordinates (l, m)<br />

on the tangent plane to the celestial sphere at the point where the object is located, and<br />

ex<strong>pre</strong>ss the surface brightness as a function I(l, m). In this case, the van Cittert–Zernike


28 Introduction to interferometry<br />

theorem can be mathematically ex<strong>pre</strong>ssed in the form<br />

<br />

−i2π(ul+vm) P(l, m)I(l, m)e dl dm<br />

V(u, v)= <br />

I(l, m) dl dm<br />

(2.5)<br />

where u=Bu/λ and v=Bv/λ are the spatial frequencies corresponding to the projec-<br />

tions <strong>of</strong> the baseline B along the orthogonal directions (u, v) (see Fig. 2.2), and P(l, m)<br />

is a function that describes the sensitivity <strong>of</strong> each element <strong>of</strong> the interferometer to the<br />

direction <strong>of</strong> arrival <strong>of</strong> the radiation, called primary beam.<br />

The <strong>pre</strong>vious equation can be inverted in the form<br />

<br />

P(l, m)I(l, m)∝ V(u, v)e +i2π(ul+vm) du dv (2.6)<br />

which stat that, known the primary beam P(l, m), it is possible to recover the sky bright-<br />

ness distribution I(l, m) <strong>of</strong> whatever source if a suitable number <strong>of</strong> measurements <strong>of</strong><br />

V(u, v) are obtained; the meaning <strong>of</strong> “suitable number” will be discussed in the follow-<br />

ing.<br />

Using the synthesis aperture technique, we can measure the visibility only in a finite<br />

number <strong>of</strong> places in the uv-plane. We can thus define a sampling function S (u, v), which<br />

is zero where no data have been taken. In this case, the image that we can recover is the<br />

the so-called dirty image<br />

I d <br />

(l, m)∝ S (u, v)×V(u, v)e +i2π(ul+vm) du dv∝ B d (l, m)∗(P(l, m)I(l, m)), (2.7)<br />

where B d (l, m) is the Fourier transform <strong>of</strong> the sampling function, commonly known as<br />

dirty beam. The dirty beam is the response on the interferometer to a point source and<br />

it corresponds to the Airy pattern for a single-dish circular telescope. The dirty image,<br />

obtained from the Fourier transform <strong>of</strong> the observed visibilities, is thus the convolu-<br />

tion <strong>of</strong> the brightness P(l, m)I(l, m) and the dirty beam B d (l, m). The sampling function<br />

is a linear filter that cuts all the not sampled spatial frequencies and the correspond-<br />

ing angular scales. Its choice should be therefore done with attention on the basis <strong>of</strong><br />

the <strong>pre</strong>sumed size and morphology <strong>of</strong> the source. In general, for a good image recon-<br />

struction, the sampling function should be as uniform as possible. On the contrary, if


2.3 Optical versus millimeter interferometry 29<br />

the sampling <strong>of</strong> the uv-plane is sparse and non uniform then the dirty beam is charac-<br />

terized by strong and numerous sidelobes; in these cases the brightness pr<strong>of</strong>ile <strong>of</strong> the<br />

source in dirty image results distort and hard to identify. Nevertheless, the dirty beam<br />

is determined by the known sampling <strong>of</strong> the uv-plane and the dirty image can be par-<br />

tially corrected through complex processes <strong>of</strong> deconvolution which assume “a priori”<br />

information about the source (see Cornwell et al., 1999, for a detailed discussion on the<br />

existing deconvolution techniques).<br />

Together with the sampling <strong>of</strong> the uv-plane, the sensitivity <strong>of</strong> an interferometer is<br />

the crucial quantity that define the range <strong>of</strong> source types that can be observed. For a<br />

single dish telescope, sensitivity and resolution are strongly coupled, depending both<br />

on the telescope diameter D, respectively through D 2 andλ/D. This is not true for<br />

an interferometer where the spatial resolution depends on the maximum baseline Bmax<br />

while the sensitivity depends <strong>main</strong>ly on the total collecting area, namely the number <strong>of</strong><br />

telescopes×the telescope’s diameters. The major limits to the sensitivity are due to the<br />

atmosphere, to the optical transmission and to the detector/background noise. All these<br />

problems depend strongly on the observational wavelength; they affects in a different<br />

way optical and radio <strong>observations</strong> and will be discussed in the following section.<br />

2.3 Optical versus millimeter interferometry<br />

Optical and millimeter interferometry are essentially the same technique used in two<br />

different wavelength do<strong>main</strong>s. Although they share the same fundamental properties<br />

their technical implementation exhibits some significant differences.<br />

Both optical and millimeter interferometers measure the complex visibility <strong>of</strong> the in-<br />

terference fringe pattern produced by separated apertures, called telescopes, siderostats<br />

or antennas. The principle <strong>of</strong> these two techniques are based on the van Citter-Zerniche<br />

theorem, i.e. the visibility is directly related to the Fourier transform <strong>of</strong> the spatial<br />

distribution <strong>of</strong> the brightness <strong>of</strong> the observed object (see Eq. 2.5). By using separated<br />

apertures, both techniques achieve high angular resolution <strong>observations</strong>, with a resolu-


30 Introduction to interferometry<br />

tion tens to hundreds <strong>of</strong> time larger than single aperture in the same wavelength do<strong>main</strong>.<br />

They both use arrays <strong>of</strong> telescopes and, as described in the <strong>pre</strong>vious section, have to find<br />

the most suitable telescope configuration to reach their objective. Moreover, both optical<br />

and radio interferometry require to calibrate the measured complex visibilities.<br />

Optical and millimeter interferometry differ, first <strong>of</strong> all, for the angular resolution<br />

that can be achieved, defined by the fringe spacingλ/2B. Fixed the maximum baseline<br />

B, the wavelength dependence implies that an interferometer observing in the near-<br />

infrared may achieve an angular resolution three orders <strong>of</strong> magnitude higher than at<br />

millimeter wavelengths. In practise, the existing millimeter interferometers are charac-<br />

terized by antennas configurations more extended that the optical array and the differ-<br />

ence is generally lower. For examples, while the VLTI has a maximum baseline <strong>of</strong>∼200<br />

m, corresponding to a resolution <strong>of</strong> few milli-arcsec in the near-infrared, the Plateau the<br />

Bure Interferometer can operate with a baseline <strong>of</strong>∼ 700 m and can achieve a resolution<br />

<strong>of</strong>∼0.3 ′′ at 1 mm; for the Very Large Array the resolution is∼0.2 ′′ at 3 cm (B=36000<br />

m). In future, with the Atacama Large Millimeter Array it will be possible to achieve a<br />

resolution <strong>of</strong>∼0.01 ′′ at 1 mm with a baseline <strong>of</strong> 15000 m.<br />

The atmosphere is the source <strong>of</strong> a second important difference. Due to the diver-<br />

sity <strong>of</strong> the temperature between the ground and the upper layers <strong>of</strong> the atmosphere,<br />

convection occurs and creates turbulent eddies, characterized by different refractive in-<br />

dices. When looking to objects through the atmosphere, the rays <strong>of</strong> light are therefore<br />

randomly deviated. The atmospheric turbulence is characterized by the spatial Fried’s<br />

parameter r0, which corresponds to the spatial scale <strong>of</strong> the atmospheric turbulence, and<br />

the coherence time t0, which define the temporal variation <strong>of</strong> the turbulence. At opti-<br />

cal wavelengths, typical values <strong>of</strong> r0 and t0 are in the range 0.1–1 m and 10-100 ms<br />

respectively; moreover, the physical analysis <strong>of</strong> the turbulence shows that both r0 and<br />

t0 depend on the wavelength asλ 6/5 . At optical wavelengths, the dominant effect <strong>of</strong><br />

atmospheric turbulence is the corrugation <strong>of</strong> the wavefront both at the level <strong>of</strong> each in-<br />

dividual aperture, seeing, and at the level <strong>of</strong> the interferometer, atmospheric piston. As<br />

in the case <strong>of</strong> single optical telescopes, interferometers use adaptive optics systems to


2.3 Optical versus millimeter interferometry 31<br />

correct for the seeing and stabilize the light on each aperture. Nevertheless, the optical<br />

path difference between two beams fluctuates in time and the fringes move back and<br />

forth on the plane <strong>of</strong> the detector, blending after the coherence time t0∼ 10−100 ms.<br />

This effect, called piston, can be corrected using a fringe tracker that measures the op-<br />

tical path difference at a frequency higher than t0 and sends the appropriate correction<br />

to a delay lines system the correct the incoming beams. The fringe tracker allows thus<br />

to stabilize the phase <strong>of</strong> the signal for integration times longer than t0 and to increase<br />

the signal-to-noise ratio on the detector; the absence <strong>of</strong> a fringe tracker on most <strong>of</strong> the<br />

available optical interferometers is one <strong>of</strong> the major causes <strong>of</strong> their low sensitivity.<br />

In the millimeter do<strong>main</strong> the situation is slightly easier. The spatial Fried’s param-<br />

eter, r0, is larger than the antenna sizes. Observations <strong>of</strong> each antenna are therefore<br />

diffraction limited and no seeing correction is required. However, millimeter observa-<br />

tions may be strongly affected by variable ionospheric and tropospheric conditions, e.g<br />

the time dependent water vapour column density which can introduce variations in the<br />

pathlength, and hence variations in the interferometric phase. This causes the loss <strong>of</strong><br />

visibility amplitude over the integration time and reduces the spatial resolution. Since<br />

at millimeter wavelengths the atmosphere coherence time t0 is <strong>of</strong> several minutes, it is<br />

possible to calibrate the visibility phase using the phase referencing technique which<br />

consists on short <strong>observations</strong> <strong>of</strong> a calibrator source with known phase within a few<br />

degrees <strong>of</strong> the target source every 5-10 minutes.<br />

The application <strong>of</strong> the phase referencing technique in the optical is very challenging,<br />

with the result that the absolute phase calibration can be hardly achieved even with the<br />

fringe tracker. Indirect and partial information on the absolute phase can however be<br />

obtained using the closure phase technique. The basic idea <strong>of</strong> this method, introduced<br />

by Jennison (1958), is founded on the principle that the sum <strong>of</strong> the observed phases<br />

around a close loop <strong>of</strong> at least three baselines, is equal to the sum <strong>of</strong> the intrinsic phases<br />

due only to the source.<br />

To understand this statement, we show in Fig. 2.3 a configuration <strong>of</strong> three telescopes


32 Introduction to interferometry<br />

Figure 2.3: Modified from Monnier (2003). The figure explain the principle behind the closure<br />

phase analysis. As described in the text, phase errors introduced at any telescopes causes equal<br />

but opposite phase shifts, cancelling out if the closure phase is calculated on a close loop <strong>of</strong> three<br />

or more telescopes.<br />

where, for each pair (i, j), the observed phase can be ex<strong>pre</strong>ssed by the relation<br />

φi j=φ0,i j+θi−θ j, (2.8)<br />

whereφ0,i j is the intrinsic source phase andθi,θ j are the phase shifts introduced by<br />

atmospheric variations and/or instrumental errors at each antenna. Defining the closure<br />

phase as the sum <strong>of</strong> the observed phases,β123=φ12+φ23+φ31, the Eq. 2.8 implies that<br />

the termsθi cancel each other out with the result that the closure phase is equal to the<br />

sum <strong>of</strong> the intrinsic phases:<br />

β123=φ12+φ23+φ31=φ0,12+φ0,23+φ0,31. (2.9)<br />

Closure phase measurements can be used to obtain information on the single Fourier<br />

phasesφ0,i j, which are required to reconstruct the source image, with complex methods


2.3 Optical versus millimeter interferometry 33<br />

whose description is out <strong>of</strong> the aim <strong>of</strong> this thesis (see Monnier, 2003). Nevertheless,<br />

closure phase has an important property that gives information on the morphology <strong>of</strong><br />

the source: if the source has a centro-symmetric brightness distribution, the closure<br />

phases is 0 ◦ for every choice <strong>of</strong> the telescopes configuration. It is easy to prove this by<br />

imaging the centro-symmetric brightness exactly at the center <strong>of</strong> the image: in this case<br />

every single phaseφ0,i j is indeed zero. Fixed the three telescopes array configuration,<br />

the closure phase increases with the asymmetry <strong>of</strong> the surface brightness <strong>of</strong> the source;<br />

a practical application <strong>of</strong> this method to the study <strong>of</strong> the structure <strong>of</strong> the inner regions <strong>of</strong><br />

circumstellar <strong>disks</strong> is shown in Chapters 5 and 9.<br />

A third difference between radio and optical interferometry comes from the type<br />

<strong>of</strong> detection and beam combination. In the radio do<strong>main</strong>, the signal detection occurs<br />

at the antenna level thanks to the heterodyne technique: the signal is coupled with a<br />

reference signal <strong>of</strong> high coherence and therefore one records the amplitude and the phase<br />

<strong>of</strong> the coming electric field. The signals from each antenna are then digitized and the<br />

combination takes place in the correlator. An electronic phase delay is applied to take<br />

into account the difference in path length between the two arms. In the optical do<strong>main</strong>,<br />

the heterodyne technique can not be applied and the interferometry is obtained by direct<br />

detection <strong>of</strong> interference fringes. The light beams are propagated to a central lab where<br />

the optical paths are equalized using delay lines, and are combined to form interference<br />

fringes which are recorded. Sensitivity <strong>of</strong> optical interferometers is strongly limited by<br />

the low throughput due to the large number <strong>of</strong> reflections between the telescopes and the<br />

final detector. Taking into account also the light loss in filters, dichroics and along the<br />

beam transport, the visible-light transmission results to be generally between 1% and<br />

10%.<br />

The direct con<strong>sequence</strong> <strong>of</strong> all these differences is that while images are routinely<br />

produced by radio interferometers, this is not the case in the optical do<strong>main</strong>. Although<br />

optical interferometers have made enormous progress in the last few years, in most <strong>of</strong><br />

cases we are still limited to the inter<strong>pre</strong>tation <strong>of</strong> visibility measurements.


34 Introduction to interferometry<br />

2.4 Introduction to modelling techniques<br />

Radio and optical interferometric <strong>observations</strong> can by analysed at two different levels.<br />

Where the quality <strong>of</strong> optical <strong>observations</strong> do not allow the image reconstruction, the data<br />

analysis has to be performed in the Fourier space, comparing the observed visibilities<br />

with synthetic values obtained starting from theoretical models. On the contrary, radio<br />

<strong>observations</strong> generally produce images and the analysis can thus be directly performed.<br />

Nevertheless, even if this latter method appears <strong>of</strong> easier application, it suffers strong<br />

limitation when the signal-to-noise ratio <strong>of</strong> the <strong>observations</strong> is low and/or the sampling<br />

<strong>of</strong> the uv-plane is sparse and not uniform. In this case, the dirty image is characterized<br />

by strong sidelobes that have to be removed with the deconvolution. However, decon-<br />

volution is a non-linear process which increases the noise level, specially in the case <strong>of</strong><br />

weak extended sources. This is <strong>of</strong>ten the case <strong>of</strong> <strong>observations</strong> <strong>of</strong> circumstellar <strong>disks</strong>,<br />

where the surface brightness decreases with the angular distance from the central star<br />

and the outermost disk regions are lost in the observational noise. The con<strong>sequence</strong> is<br />

that the comparison between disk models and cleaned images is <strong>of</strong>ten inaccurate.<br />

As for the optical interferometry, the best way to analyze the data is therefore to<br />

compare the observed visibilities with those <strong>pre</strong>dicted by our models. Weighting each<br />

visibility data point by its known thermal noise, it is possible to define how well the<br />

model fits the data adopting the usualχ 2 notation:<br />

χ 2 =Σi[(Remod,i− Reobs,i) 2 + (Immod,i−Imobs,i) 2 ]·Wi, (2.10)<br />

where Re∗,i and Im∗,i are the real and imaginary parts <strong>of</strong> the i-th observed (obs) or model-<br />

<strong>pre</strong>dicted (mod) visibility. The weight Wi, provided in the observed uv-table (the table<br />

containing the baselines Bu and Bv, the corresponding visibility and the weight), is de-<br />

rived from the system temperature Tsys, the spectral resolution∆ν, the integration time<br />

τ, the effective collecting area <strong>of</strong> one antenna Ae f f and the loss <strong>of</strong> efficiency introduced<br />

by the correlatorη:<br />

Wi= 1<br />

withσi=<br />

σ2 √ 2kTsys<br />

Ae f fη √ . (2.11)<br />

τ∆ν


2.4 Introduction to modelling techniques 35<br />

Models <strong>of</strong> possible circumstellar disk images can by schematically divided in two<br />

category: centro-symmetric models, for which the visibility can be ex<strong>pre</strong>ssed analyti-<br />

cally introducing the Bessel functions, and more complex asymmetric models, for which<br />

a numerical solution <strong>of</strong> the image Fourier transform is required.<br />

To the first class belong all the geometrical models such as point sources, uniform<br />

or Gaussian <strong>disks</strong> and rings, either circular or elliptical, and all their possible combina-<br />

tions, i.e, multiple systems composed by the sum <strong>of</strong> two or more <strong>of</strong> theme. When the<br />

model has a circular symmetry it is easier to switch to polar coordinates and ex<strong>pre</strong>ss the<br />

brightness <strong>of</strong> the image I(l, m) as a function I(ρ) <strong>of</strong> the radial coordinateρ= √ l 2 + m 2 .<br />

In this case, the corresponding visibility depends also only on the radial coordinate<br />

r= √ u 2 + v 2 , through the relation<br />

V(r)=2π<br />

∞<br />

0<br />

I(ρ)J0(2πρr)ρ dρ, (2.12)<br />

where J0 is the zeroth-order Bessel function <strong>of</strong> the first kind:<br />

J0(x)= 1<br />

2π<br />

2π<br />

0<br />

exp (−ix cosθ) dθ. (2.13)<br />

For a circular ring <strong>of</strong> radiusρ and infinitesimal thickness, the brightness can be ex-<br />

<strong>pre</strong>ssed through theδfunction as<br />

I(ρ)= 1<br />

δ(ρ−ρ0), (2.14)<br />

2πρ0<br />

and the corresponding normalized visibility is then:<br />

V(r)= J0(2πρ0r). (2.15)<br />

Any circularly symmetric function can be described as the integral <strong>of</strong> infinitesimally<br />

thickness rings with varying radius and intensity and, since the Fourier transform is<br />

addictive, its visibility is the integral <strong>of</strong> the corresponding visibility curves. In this<br />

way, the visibility <strong>of</strong> a uniform disk <strong>of</strong> angular diameterθand normalized integrated<br />

brightness is given by<br />

V(r)=2 J1(πθr)<br />

, (2.16)<br />

πθr


36 Introduction to interferometry<br />

where J1 is the first-order Bessel function <strong>of</strong> the first kind; while for a normalized Gaus-<br />

sian disk (remember that the Fourier transform <strong>of</strong> a Gaussian function is also a Gaussian<br />

function)<br />

V(r)=exp<br />

<br />

− (πθr)2<br />

<br />

, (2.17)<br />

4 ln 2<br />

whereθis the full width half maximum <strong>of</strong> the disk. Finally, the visibility <strong>of</strong> a uniform<br />

ring characterized by angular inner diametersθ and width f is<br />

V(r)=<br />

2<br />

πθr(2 f+f 2 ) {(1+ f )J1[(1+ f )πθr]− J1(πθr)}. (2.18)<br />

Circular symmetric models can be considered as a first approximation <strong>of</strong> face-on<br />

disk images and can be used to estimate the angular diameter <strong>of</strong> the emission. Ring<br />

models, in particular, have been widely applied in near-infrared interferometry to cal-<br />

culate the disk inner radius (Monnier and Millan-Gabet, 2002; Eisner et al., 2004; see<br />

the discussion in Chapter 3). In most cases, the disk inclination is however not known<br />

and the face-on assumption is not justified. Inclining a circular ring the resulting im-<br />

age is an ellipse. Through a transformation <strong>of</strong> coordinates, and taking advantage <strong>of</strong> the<br />

similarity properties <strong>of</strong> the Fourier transform, it is possible to re-conduce the elliptical<br />

case to the circular one and ex<strong>pre</strong>ss the visibility trough the Bessel functions. For an<br />

exhaustive description <strong>of</strong> analytical calculation <strong>of</strong> visibility we remand to Perrin and<br />

Malbet (2003).<br />

If the model image has a complex morphology, a numerical bi-dimensional compu-<br />

tation <strong>of</strong> the visibility is required. Thanks to the fast improvement <strong>of</strong> the technology,<br />

Fourier transform <strong>of</strong> images composed by millions <strong>of</strong> pixels can by today performed<br />

on a scale time <strong>of</strong> seconds or minutes, on a normal personal computer. This time is<br />

incredible fast if compared to the hours, or even days, required not more than 20 years<br />

ago. All the popular commercial s<strong>of</strong>tware for numerical analysis, i.e., IDL and Mat-<br />

Lab, contain dedicated routines for multi dimensional Fourier transform calculations.<br />

Another large number <strong>of</strong> codes is freely available on the web. In particular, all the data<br />

analyses discussed in this thesis have been performed using a free C subroutine library<br />

for computing discrete Fourier transform called FFTW (Fast Fourier Transform in the


2.4 Introduction to modelling techniques 37<br />

West); documentation, features and benchmarks can be found on the web site 1 . The<br />

input <strong>of</strong> this numerical recipe is a matrix <strong>of</strong> arbitrary size, containing the brightness<br />

values I(i, j) for each pixel <strong>of</strong> the model image, defined by discrete coordinates i and j<br />

(corresponding to the continuum coordinate l and m <strong>of</strong> Fig. 2.2). The output consists in<br />

a table containing the real and imaginary part <strong>of</strong> the visibility at the corresponding spa-<br />

tial frequencies. The comparison with the observed visibility can be finally <strong>pre</strong>formed<br />

interpolating the theoretical visibilities, calculated on a regular grid, in the points <strong>of</strong><br />

the uv-plane sampled during the observation and calculating theχ 2 as discussed at the<br />

beginning <strong>of</strong> this Section.<br />

If the <strong>observations</strong> are characterized by an uniform sampling <strong>of</strong> the uv-plane, with<br />

baselines in the interval Bu,v,min-Bu,v,max, we can expect to observe source features on an-<br />

gular scale sizes ranging from∼λ/Bu,v,max up to∼λ/Bu,v,min. To perform the comparison<br />

between models and <strong>observations</strong>, we must therefore create a model image able to re-<br />

produces such features. In particular, the grid spacing, i.e. the pixel sizes∆i and∆ j, and<br />

the number <strong>of</strong> pixels on each axis, Ni and N j, must allow the re<strong>pre</strong>sentation <strong>of</strong> all these<br />

scales. In terms <strong>of</strong> range <strong>of</strong> uv points sampled, the requirements are∆i≤λ/2Bu,max,<br />

∆ j≤λ/2Bv,max and Ni∆i≥λ/Bu,min, N j∆ j≥λ/Bv,min.<br />

1 http://www.fftw.org


Part II<br />

Dust in the inner disk


CHAPTER 3<br />

The shape <strong>of</strong> the inner rim in<br />

proto-planetary <strong>disks</strong><br />

This Chapter has been published in Astronomy & Astrophysics:<br />

A. Isella 1,2 and A. Natta 2 , 2005 A&A, 438, 899<br />

“The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong>”<br />

1 Dipartimento di Fisica, Univeristá di Milano, via Celoria 16, 20133 Milano, Italy<br />

2 INAF - Osservatorio Astr<strong>of</strong>isico di Arcetri, Largo E. Fermi 5, 50125, Firenze, Italy<br />

Abstract: This Chapter discusses the properties <strong>of</strong> the puffed-up inner rim that forms in<br />

circumstellar <strong>disks</strong> when dust evaporates. We argue that the rim shape is controlled by a funda-<br />

mental property <strong>of</strong> circumstellar <strong>disks</strong>, namely their very large vertical density gradient, through<br />

the dependence <strong>of</strong> grain evaporation temperature on gas density. As a result, the bright side <strong>of</strong><br />

the rim is curved, rather than vertical, as expected when a constant evaporation temperature<br />

is assumed. We have computed a number <strong>of</strong> rim models that take into account this effect in a<br />

self-consistent way. The results show that the curved rim (as the vertical rim) emits most <strong>of</strong> its


42 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

radiation in the near and mid-IR, and provides a simple explanation for the observed values<br />

<strong>of</strong> the near-IR excess (the “3µm bump” <strong>of</strong> Herbig Ae stars). Contrary to the vertical rim, for<br />

curved rims the near-IR excess does not depend much on the inclination, being maximum for<br />

face-on objects.We then computed synthetic images <strong>of</strong> the curved rim seen under different incli-<br />

nations; face-on rims are seen as bright, centrally symmetric rings on the sky; increasing the<br />

inclination, the rim takes an elliptical shape, with one side brighter than the other.<br />

3.1 Introduction<br />

The structure <strong>of</strong> the inner regions <strong>of</strong> circumstellar <strong>disks</strong> associated with <strong>pre</strong>-<strong>main</strong> se-<br />

quence stars is the subject <strong>of</strong> intense research. Interferometers working in the near<br />

infrared are providing the first direct information on the morphology <strong>of</strong> <strong>disks</strong> on scales<br />

<strong>of</strong> fractions <strong>of</strong> AU. They show that in the majority <strong>of</strong> cases the observed visibility curves<br />

are not well reproduced by flared disk models; rather, they are consistent with the emis-<br />

sion <strong>of</strong> a ring <strong>of</strong> uniform brightness, <strong>of</strong> radius similar to the dust evaporation distance<br />

from the star (Millan-Gabet et al., 2001; Tuthill et al., 2001).<br />

The interferometric results provide strong support for the idea that the inner disk<br />

structure deviates substantially from that <strong>of</strong> a flared disk because dust evaporation in-<br />

troduces a strong discontinuity in the opacity, which results in a puffed-up rim at the<br />

dust destruction radius, where dust is exposed directly to the heating stellar radiation.<br />

The idea <strong>of</strong> a puffed-up inner rim was proposed by Natta et al. (2001) and developed<br />

further by Dullemond et al. (2001, hereafter DDN01) for Herbig Ae stars, to account for<br />

the shape <strong>of</strong> the near-infrared excess <strong>of</strong> these stars (the “3-µm bump”). These authors<br />

pointed out that the rim also had the right properties to explain the early interferomet-<br />

ric results <strong>of</strong> Millan-Gabet et al. (2001). Recent theoretical work by Muzerolle et al.<br />

(2004) has shown that the condition required to produce a puffed-up inner rim are in-<br />

deed likely to exist in most Herbig Ae and T Tauri stars. The concept <strong>of</strong> such an inner<br />

rim has been widely used to inter<strong>pre</strong>t near-IR interferometric data for Herbig and T


3.1 Introduction 43<br />

Tauri stars (Eisner et al., 2004; Muzerolle et al., 2003; Colavita et al., 2003; Eisner et<br />

al., 2003; Monnier and Millan-Gabet, 2002; Millan-Gabet et al., 2001). Its effects on<br />

the disk structure and emission at larger radii have been discussed by Dullemond and<br />

Dominik (2004), who propose that the classification <strong>of</strong> Herbig Ae stars in two groups,<br />

based on the shape <strong>of</strong> the far-infrared excess (Meeus et al., 2001), can be inter<strong>pre</strong>ted as<br />

differences between objects where the outer disk emerges from the shadow <strong>of</strong> the inner<br />

rim and objects where this never happens.<br />

In spite <strong>of</strong> its success in accounting for a variety <strong>of</strong> <strong>observations</strong>, the actual structure<br />

<strong>of</strong> the rim has not been much discussed. DDN01 adopted for their models a very crude<br />

approximation, namely that the illuminated side <strong>of</strong> the rim is “vertical”, and that its<br />

photospheric height is controlled by radial heat diffusion behind the rim. Such a model,<br />

taken at face value, has the obvious disadvantage that the rim emission vanishes for<br />

objects seen face-on, for which the projection on the line <strong>of</strong> sight <strong>of</strong> the rim surface<br />

is null, and for objects seen edge-on, where the rim obscures its own emission. This<br />

is clearly inconsistent with <strong>observations</strong> <strong>of</strong> the SED, which show that all the Herbig<br />

Ae stars with <strong>disks</strong> have similar near-IR excess, regardless <strong>of</strong> their inferred inclination<br />

(Natta et al., 2001; Dominik et al., 2003).<br />

The vertical shape <strong>of</strong> the illuminated face <strong>of</strong> the rim is clearly not physical, as<br />

pointed out already by DDN01. Several effects are likely to “bend” the rim: among<br />

them, one can expect that radiation <strong>pre</strong>ssure on dust grains or dynamical instability,<br />

due to self-shadowing effects, could modify the illuminated face <strong>of</strong> the rim (Dulle-<br />

mond, 2000; DDN01). None <strong>of</strong> these suggestions, however, has been explored further.<br />

In this Chapter, we will discuss in detail a different process, not mentioned so far,<br />

which depends exclusively on the basic physics <strong>of</strong> dust evaporation, i.e., on the depen-<br />

dence <strong>of</strong> the evaporation temperature on gas density. Circumstellar <strong>disks</strong> are charac-<br />

terized by a very large variation <strong>of</strong> the density in the vertical direction, so that the dust<br />

evaporation temperature varies by several hundred degrees in a few scale heights; mov-<br />

ing vertically away from the disk mid plane along the rim, dust will evaporate at lower<br />

and lower temperatures, i.e., further away from the central star. This very simple effect


44 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

curves significantly the inner face <strong>of</strong> the rim, as we will describe in the following.<br />

The Chapter is organized as follows. In Sec. 3.2 we describe the model we use to<br />

compute the rim shape and its observational properties. The results are <strong>pre</strong>sented in<br />

Sec. 3.3, where we discuss also how the rim depends on dust properties. A discussion<br />

<strong>of</strong> the results, in view <strong>of</strong> the existing <strong>observations</strong>, follows in Sec. 3.4, and a summary<br />

is given in Sec. 3.5.<br />

3.2 Model description<br />

Our model <strong>of</strong> the inner rim <strong>of</strong> passive-irradiated flaring <strong>disks</strong> joins the two different<br />

analytical methods to solve the structure <strong>of</strong> a circumstellar disk proposed respectively<br />

by Calvet et al. (1991; 1992, hereafter C92) and Chiang and Goldreich (1997, hereafter<br />

CG97). The temperature in the rim atmosphere is determined using the analytical solu-<br />

tion <strong>of</strong> the problem <strong>of</strong> the radiation transfer as in C92, neglecting the heating term due to<br />

the mass accretion. The vertical structure <strong>of</strong> the rim is then computed in a way derived<br />

from CG97 and DDN01, adding a relation between the dust vaporization temperature<br />

and the gas density as proposed in Pollack et al. (1994). As a result we obtain a curved<br />

model for the inner rim whose features are described in the next paragraph.<br />

Although the ex<strong>pre</strong>ssions for the dust temperature derive from a first order solution<br />

<strong>of</strong> the radiation transfer equation, we found (see Appendix) good agreement with the<br />

correct numerical result in most cases. To zero order, we can compute the rim structure<br />

avoiding proper radiation transfer calculations.<br />

In the limit where the incident angleα<strong>of</strong> the stellar radiation onto the disk surface<br />

isα≪1, the equations <strong>of</strong> the temperature forτd = 0 andτd ≫ 1 (whereτd is the<br />

optical depth for the emitted radiation) are formally equal to the optically thin Ts and<br />

mid plane Ti temperatures introduced by CG97 (two-layer approximation). Therefore,<br />

while the two layer approximation is useful to study the structure and the emission<br />

features (e.g. silicate features at 10µm) <strong>of</strong> the flaring part <strong>of</strong> the disk (as in DDN01),<br />

it must be abandoned in modeling the inner rim, sinceα≃1. Nevertheless we can


3.2 Model description 45<br />

adapt the CG97 treatment <strong>of</strong> the vertical structure <strong>of</strong> the disk to the inner rim using the<br />

appropriate ex<strong>pre</strong>ssions for the dust temperature.<br />

In order to clarify this concept and to introduce the relation between the vaporization<br />

temperature <strong>of</strong> dust and the gas density, the basic equations are briefly summarized. We<br />

refer to the cited works for a physical discussion <strong>of</strong> the equations.<br />

We suppose that the disk is heated only by the stellar radiation and we callαthe<br />

incident angle between the radiation and the disk surface. The incident beam is absorbed<br />

exponentially as it penetrates the dusts and ifτd is the optical depth for the emitted<br />

radiation, the dust temperature T(τd) is given by the relation (C92)<br />

T 4 (τd)=T 4 2 <br />

R⋆<br />

1<br />

⋆ · µ(2+3µǫ)+ − 3ǫµ2 e<br />

2r ǫ (−τd/µǫ)<br />

<br />

(3.1)<br />

whereµ=sinα andǫ is an efficiency factor that characterizes the grain opacity, defined<br />

as the ratio <strong>of</strong> the Planck mean opacity <strong>of</strong> the grains at the local temperature T(τd) and<br />

at the stellar temperature T⋆, respectively. In writing the <strong>pre</strong>vious equation we have<br />

assumed that the scattering <strong>of</strong> the dust grain is negligible and that the Planck mean<br />

opacity is defined as:<br />

∞ 0<br />

KP(T)=<br />

Bν(T)kνdν<br />

∞ . (3.2)<br />

Bν(T)dν<br />

0<br />

Following Muzerolle et al. (2003), the continuum emission <strong>of</strong> the disk is assumed to<br />

originate from the surface characterized by the optical depthτd= 2/3.<br />

In the flaring part <strong>of</strong> the disk, for whichµ≪1, Eq. 3.1 can be rewritten in terms<br />

<strong>of</strong> the two layer approximation, proposed in CG97, in which the interior <strong>of</strong> the disk<br />

(withτd≫ 1) is heated to the temperature Ti by half <strong>of</strong> the stellar flux, while the other<br />

half <strong>of</strong> the stellar flux is re-emitted by the superficial layer heated at the optically thin<br />

temperature Ts:<br />

T 4 (τd≫ 1)≃T 4 i = µ<br />

2<br />

T 4 (τd= 0)≃T 4 s<br />

= 1<br />

ǫ<br />

2<br />

R⋆<br />

T<br />

r<br />

4 ⋆<br />

2<br />

R⋆<br />

T<br />

2r<br />

4 ⋆<br />

(3.3)<br />

(3.4)


46 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

Following DDN01 we assume that the dust component <strong>of</strong> the disk is truncated on the<br />

inside by dust evaporation, forming an inner hole <strong>of</strong> radius Rin. Inside this hole, only<br />

gas may exist but as long as the gas is optically thin to the stellar radiation (Muzerolle<br />

et al., 2004) we can neglect its absorbing effect. Near the inner radius Rin, since the rim<br />

surface is nearly perpendicularly exposed to the stellar radiation (µ≃1), Eq. 3.3 and<br />

3.4 must be replaced by the more general relations<br />

T 4 (τd≫ 1)≡T 4 ∞ =µ(2+3µǫ)<br />

T 4 (τd= 0)≡T 4<br />

0 =<br />

<br />

2µ+ 1<br />

ǫ<br />

derived as limiting cases from Eq. 3.1.<br />

2<br />

R⋆<br />

T<br />

2r<br />

4 ⋆ , (3.5)<br />

2<br />

R⋆<br />

T<br />

2r<br />

4 ⋆, (3.6)<br />

Since the dust in the rim is by definition close to the evaporation temperature, the<br />

value <strong>of</strong>ǫ is fixed and depends only on the grain absorbing cross section. Moreover, to<br />

first order, since the difference between T(τd= 0) and T(τd=∞) is never very large, at<br />

the inner radius (µ=1) the ratio between these two temperature depends only onǫand<br />

is given by:<br />

T0<br />

T∞<br />

=<br />

1/4 2ǫ+ 1<br />

. (3.7)<br />

ǫ(2+3ǫ)<br />

The rim is defined by the condition that the incoming stellar radiation is entirely<br />

absorbed by the intervening dust and in the following we will refer to the surface where<br />

τs= 1 to define its shape and location. Once the grain evaporation temperature Tevp is<br />

fixed, one can see from Eq. 3.1 that the rim has a sharp surface whenǫǫcr, the dust temperature increases with the optical depth and<br />

the transition to optically thick regimes is controlled by the geometrical dilution <strong>of</strong> the<br />

stellar radiation; in this case the transition fromτs= 0 to very high values occurs in a


3.2 Model description 47<br />

log(δR/R in )<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

-6<br />

0.2 0.4 0.6 0.8 1<br />

Figure 3.1: Width <strong>of</strong> the region over which 99% <strong>of</strong> the stellar radiation is absorbed as function<br />

<strong>of</strong>ǫ.<br />

relatively broader region (see Fig. 3.1); however, also in these cases the rim is located<br />

by theτs= 1 surface with an accuracy better than 10–20%.<br />

ε<br />

With the assumption that the disk is in hydrostatic equilibrium in the gravitational<br />

field <strong>of</strong> the central star and that is isothermal in the vertical direction z, the gas density<br />

distribution is ex<strong>pre</strong>ssed by the Gaussian relation<br />

where<br />

ρg(r, z)=ρg,0(r) exp(−z 2 /2h 2 ), (3.8)<br />

h<br />

r =<br />

T∞<br />

Tc<br />

r<br />

R⋆<br />

1/2<br />

(3.9)<br />

defines the relation between the <strong>pre</strong>ssure scale h and the interior temperature T∞ at a<br />

distance r from the central star. The temperature Tc is a measure <strong>of</strong> the gravitational<br />

field <strong>of</strong> the central star, ex<strong>pre</strong>ssed by<br />

Tc= GM⋆µg<br />

kR⋆<br />

whereµg is the mean molecular weight <strong>of</strong> the gas.<br />

(3.10)


48 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

Note that the interior temperature <strong>of</strong> the disk T∞ depends on the incident angle<br />

α between the stellar radiation and the disk surface, through Eq. 3.5. The quantity<br />

µ=sinα accounts for the projection <strong>of</strong> a disk annulus on a plane perpendicular to the<br />

incident radiation and is given by the relation<br />

<br />

r<br />

dz(r)<br />

µ=sin arctan + arctan −<br />

z(r)<br />

dr<br />

π<br />

<br />

, (3.11)<br />

2<br />

where z(r) is the axisymmetric equation <strong>of</strong> the surface <strong>of</strong> the disk that we want to deter-<br />

mine. For the standard flaring model, the incident angleα is given by the relation<br />

sinα≡µ≃ 0.4R⋆<br />

r<br />

+ r d(H/r)<br />

dr<br />

, (3.12)<br />

where H is the photospheric height <strong>of</strong> the disk, defined as the height to which the optical<br />

depth <strong>of</strong> the disk to the stellar radiation isτs= 1, on a radially directed ray. For the inner<br />

rim we can retain this definition for H and place z(r)=H(r) in Eq. 3.11, where H is<br />

thus given by the relation<br />

KP(T⋆)<br />

µ<br />

∞<br />

H(r)<br />

ρd(z, r)dz=1. (3.13)<br />

As shown in DDN01, the ratioχ=H/h, between the photospheric and <strong>pre</strong>ssure<br />

height, is a dimensionless number <strong>of</strong> the order <strong>of</strong> 4-6, depending weakly on the angle<br />

α, the densityρd and the Planck mean opacity <strong>of</strong> dust at the stellar temperature T⋆.<br />

Finally, to obtain the dust density distribution from Eq. 3.8 we assume a constant ratio<br />

ρd/ρg= 0.01.<br />

We can follow CG97 and DDN01 to obtain a self-consistent solution <strong>of</strong> Eq. 3.2, 3.5,<br />

3.6, 3.8, 3.9, 3.11, 3.13 to determine the structure <strong>of</strong> the disk, as long as dust evaporation<br />

can be neglected. When dust evaporation is important, DDN01 have developed an ap-<br />

proximate solution <strong>of</strong> the rim/disk structure under the assumption <strong>of</strong> constant Tevp. The<br />

inner rim has a vertical surface toward the star located at the dust evaporation distance<br />

Rin. The vertical photospherical height H depends on the dust density behind the rim;<br />

since the rim is higher than the flaring disk, it casts a shadows over the disk. In this<br />

shadowed region, assuming that there is no external heating except the stellar radiation,


3.2 Model description 49<br />

the <strong>pre</strong>ssure height h depends only on the radial heating diffusion. The exact determi-<br />

nation <strong>of</strong> the structure <strong>of</strong> the rim in this diffusive region would thus require to solve<br />

the problem <strong>of</strong> radiation transport in two dimensions (see Dullemond, 2002). Since this<br />

goes well beyond our aims, we adopt the approximated relation used in DDN01<br />

d(rT 4 ∞ )<br />

dr<br />

≃− rT 4 ∞<br />

. (3.14)<br />

h<br />

Since h ∝ (T∞R 3 ) 1/2 (from Eq. 3.9), the temperature T∞ can be eliminated and for<br />

h≪R we obtain the relation for h behind the rim<br />

d(h/r)<br />

dr<br />

1<br />

=− . (3.15)<br />

8R<br />

The solution <strong>of</strong> Eq. 3.15 can therefore be used to determine the densityρd(r, z) behind<br />

the rim through Eq. 3.8.<br />

We now introduce the relation between the evaporation temperature <strong>of</strong> the dust and<br />

the gas density. The physical reasons for this effect can be easily understood thinking<br />

<strong>of</strong> evaporation as the process by which equilibrium between the gas <strong>pre</strong>ssure and the<br />

surface tension <strong>of</strong> the dust grains is reached: the higher the gas density, the higher will<br />

be the evaporation temperature. We adopt for the dust in the disk the model proposed by<br />

Pollack et al. (1994). In this model, the grains with the higher evaporation temperature<br />

are the silicates, which will therefore determine the location <strong>of</strong> the rim. Their evapora-<br />

tion temperature (see Table 3 in Pollack et al.) varies with the gas density roughly as a<br />

power law, <strong>of</strong> the kind:<br />

Tevp= Gρ γ g (r, zevp) (3.16)<br />

where G=2000 andγ=1.95·10 −2 . Sinceρg varies exponentially with z (see Eq. 3.8),<br />

for zevp= h the evaporation temperature <strong>of</strong> dust is only 1% smaller than that on the mid<br />

plane, but for zevp=H∼ 5h (for a reasonable value <strong>of</strong>χ) the difference is about 20%<br />

and dust evaporation takes place at a larger distance from the star, curving the rim.<br />

To obtain a self-consistent determination <strong>of</strong> the structure <strong>of</strong> the curved rim, we im-<br />

plemented a numerical method able to solve Eq. 3.16 together with the set <strong>of</strong> equa-<br />

tions 3.2, 3.5, 3.6, 3.8, 3.9, 3.11, 3.13. The distance <strong>of</strong> dust evaporation in the mid plane


50 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

Rin is computed for z=0 in the set <strong>of</strong> equations and is taken as the starting point <strong>of</strong><br />

the radial grid on which the rim structure is computed. As a result <strong>of</strong> the calculations,<br />

we obtain the location in the (R, z) plane <strong>of</strong> the rim surfaces characterized by a constant<br />

value <strong>of</strong> the optical depth. The surface forτs= 0 is thus the evaporation surface <strong>of</strong> dust<br />

grains, forτs=1 we obtain the surface relative to the photospheric height H, defined<br />

through the Eq. 3.13, while forτd=ǫ·τs= 2/3 we obtain the emitting surface <strong>of</strong> the<br />

rim.<br />

Moving on the rim surface away from the star, the incident angleαdecreases and,<br />

when it approaches zero, the determination <strong>of</strong> the rim shape becomes very difficult. This<br />

is <strong>main</strong>ly due to the fact that the described solution for the radiation transfer neglects<br />

the heat diffusion between contiguous annulus <strong>of</strong> the rim. Therefore both the mid plane<br />

temperature T∞ and the <strong>pre</strong>ssure height h <strong>of</strong> the rim goes unphysically to zero forα=0,<br />

according to Eq. 3.5 and 3.9. To avoid this unrealistic behaviour, we use the approxi-<br />

mated relation discussed <strong>pre</strong>viously (see Eq. 3.15) to determine the <strong>pre</strong>ssure height h in<br />

the region where the diffusion is the dominant heating source. The transition distance<br />

between the region <strong>of</strong> the rim heated by the star and those heated by the diffusion is de-<br />

termined imposing continuity <strong>of</strong> dh/dr. The photospherical height H is then determined<br />

as for the vertical rim.<br />

3.3 Results<br />

Using the model described in the <strong>pre</strong>vious section, we compute the structure <strong>of</strong> the inner<br />

rim for a disk heated by a star with temperature T⋆= 10000K, mass M⋆= 2.5M⊙ and<br />

luminosity L⋆= 47L⊙. We take a disk surface densityΣ(r)=2·10 3 (R/AU) −1.5 g cm −2 ,<br />

and a dust-to-gas mass ratio dust/gas=0.01. In our models, this value <strong>of</strong>Σcorresponds<br />

to a mid plane gas density <strong>of</strong> about 10 −8 g/cm 3 at 0.5AU from the star. Note that the<br />

results are not very sensitive to the exact value <strong>of</strong>Σ, as long as the inner disk re<strong>main</strong>s<br />

very optically thick.<br />

The dust properties are those <strong>of</strong> the astronomical silicates <strong>of</strong> Weingartner and Draine


3.3 Results 51<br />

Photospherical height, H (AU)<br />

0.12<br />

0.08<br />

0.04<br />

ε = 0.92<br />

ε = 0.58<br />

ε = 0.25<br />

ε = 0.14<br />

0<br />

0.5 0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4<br />

R (AU)<br />

ε = 0.08<br />

Figure 3.2: Photospherical height H (τs= 1) <strong>of</strong> the inner rim for different values <strong>of</strong>ǫ, as labelled.<br />

The curves have been computed for a star with T⋆= 10000K, M⋆= 2.5M⊙, L⋆= 47L⊙. Each<br />

curve ends where the rim becomes optically thin to the stellar radiation.<br />

(2001). In our models, we consider that all the grains have the same size and charac-<br />

terize their properties with the quantityǫ, the ratio <strong>of</strong> the mean Planck opacity at the<br />

evaporation temperature to that at the stellar temperature (see Eq. 3.2). The evaporation<br />

temperature <strong>of</strong> silicate grains vary from 1600K, for gas density <strong>of</strong>ρg= 10 −6 g/cm 3 , to<br />

1000K forρg= 10 −16 g/cm 3 (see Eq. 3.16). For the Weingartner silicates and a vapor-<br />

ization temperature Tevp∼ 1400 K,ǫ∼ 0.08 for grains <strong>of</strong> radius a=0.1µm, and grows<br />

to values <strong>of</strong> about unity for grain radii>5µm.<br />

3.3.1 The rim shape<br />

The shape <strong>of</strong> the rim is shown in Fig. 3.2, which plots the locus <strong>of</strong>τs = 1 (i.e., the<br />

photospheric height H) as function <strong>of</strong> R for different values <strong>of</strong>ǫ. The ending point <strong>of</strong><br />

the rim (Rout, Hout) is when the rim becomes optically thin at the stellar radiation.<br />

Forǫ ≥ǫcr ∼ 0.58, corresponding to silicate grains bigger than 1.3µm, the inner<br />

radius Rin, the outer radius Rout and the maximum photospheric height <strong>of</strong> the rim Hout,<br />

all vary very little, with values Rin≃0.50 AU, Rout≃ 0.65 AU and Hout≃ 0.09 AU. For<br />

smaller values <strong>of</strong>ǫ, the rim becomes steeper and the inner radius increases. Forǫ= 0.08,<br />

corresponding to grains with radius 0.1µm, the rim has Rin= 1.16 AU, Rout= 1.34 AU<br />

and Hout= 0.14 AU.


52 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

Effective Temperature, T eff (K)<br />

1400<br />

1350<br />

1300<br />

1250<br />

1200<br />

1150<br />

1100<br />

1050<br />

1000<br />

ε = 0.25<br />

ε = 0.58<br />

ε = 0.92<br />

950<br />

1 1.05 1.1 1.15 1.2 1.25<br />

R/Rin Figure 3.3: Effective temperature (i.e., T(τd= 2/3)) along the rim surface for different values<br />

<strong>of</strong>ǫ, as labelled. Each curve is plotted as function <strong>of</strong> R/Rin, where Rin is the distance <strong>of</strong> the rim<br />

from the star on the disk mid plane. Note that, in fact, not only R but also z changes along each<br />

curve, as shown in Fig. 3.2. Forǫ>ǫcr≃ 0.58 Te f f is almost independent <strong>of</strong> the grain opacity.<br />

The shape <strong>of</strong> the rim can be roughly characterized by the ratio <strong>of</strong> its maximum height<br />

Hout over width∆R=(Rout−Rin). This quantity, which is nominally infinity in a vertical<br />

rim, becomes in the curved rim models∼ 0.6 forǫ> ∼ ǫcr and is∼ 0.8 forǫ∼ 0.14−0.08.<br />

In other words, as grains grow, the inner rim approaches the star but the bending <strong>of</strong> the<br />

surface varies very little.<br />

For any given value <strong>of</strong>ǫ, the dust temperature along the rim is not constant (as in<br />

the vertical rim) but decreases from values <strong>of</strong> about 1400 K, typical <strong>of</strong> silicate evap-<br />

oration temperatures at density <strong>of</strong> 10 −8 g/cm 3 , to 1200 K for density <strong>of</strong> 10 −11 g/cm 3 .<br />

Fig. 3.3 plots the effective temperature <strong>of</strong> the rim (i.e., T(τd= 2/3)) along its surface<br />

for different values <strong>of</strong>ǫ. Forǫ> ∼ ǫcr the effective temperature is equal to the vaporiza-<br />

tion temperature; forǫ


3.3 Results 53<br />

ever, is sufficiently small that most <strong>of</strong> the rim emission occurs in the near-IR, as for the<br />

vertical rim.<br />

3.3.2 The rim SED<br />

The fraction <strong>of</strong> stellar luminosity intercepted by the rim, given by the ratio Hout/Rout,<br />

varies from 10%, for small values <strong>of</strong>ǫ, to 14%, forǫ≥ǫcr. Assuming that the rim is<br />

in thermal equilibrium, the intercepted radiation is equal to the total emitted flux FIR.<br />

For eachǫ, we have computed the spectral energy distribution <strong>of</strong> the rim emission for<br />

different inclination angles. As discussed in Sec. 3.2, the rim emission is computed<br />

as that <strong>of</strong> a blackbody at the local temperature along the rimτd = 2/3 surface. We<br />

will come back in the Appendix to this assumption, which, in any case, gives a very<br />

good approximation to the global properties <strong>of</strong> the rim emission (see also Muzerolle<br />

et al., 2003; C92). Most <strong>of</strong> the emission, as expected from the range <strong>of</strong> temperatures,<br />

occurs in the near-IR. We have computed the fraction <strong>of</strong> the stellar luminosity re-emitted<br />

by the rim in the wavelength range 1.25-7.0µm for different values <strong>of</strong>ǫ as function <strong>of</strong><br />

the inclination (Fig. 3.4). This near-IR excess peaks at zero inclination, where has values<br />

between∼ 10% (smaller grains) and∼ 20% (larger grains). As the inclination increases,<br />

the near-IR excess decreases slowly, reaching values between 5% and 8%, depending<br />

onǫ. For inclination higher then∼80 ◦ the rim emission is self-absorbed. Note that for<br />

the large grains, withǫ> ∼ ǫcr, the near-IR excess becomes almost independent <strong>of</strong>ǫ.<br />

The behaviour described in Fig. 3.4 is very different from what one obtains in the<br />

case <strong>of</strong> the vertical rim. As described in the <strong>pre</strong>vious section, neglecting the dependence<br />

<strong>of</strong> the evaporation temperature <strong>of</strong> grains from gas density results in the inner face <strong>of</strong> the<br />

rim to be vertical (as in DDN01). Withǫ=ǫcr and Tevp≃ 1400 K (equal to the evapo-<br />

ration temperature at the inner radius <strong>of</strong> the curved rim), the inner radius Rin is the same<br />

as for the curved rim; the photospheric height, evaluated using the approximation de-<br />

scribed by Eq. 3.15, is also the same (Hout= 0.09 AU); the fraction <strong>of</strong> stellar luminosity<br />

intercepted by the vertical rim is 17%. However, the value <strong>of</strong> FNIR observed for different<br />

inclination angles is very different, with a very strong (and opposite) dependence on i


54 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

L NIR /L *<br />

0.2<br />

0.16<br />

0.12<br />

0.08<br />

0.04<br />

0 0 10 20 30 40 50 60 70 80<br />

Inclination angle (deg)<br />

Figure 3.4: Near infrared emission <strong>of</strong> the curved rim (integrated between 1.25−7µm and<br />

normalized to L⋆) for different inclination angles and different values <strong>of</strong>ǫ. Starting from the<br />

bottom, the curves refer toǫ= 0.08 (long-dashed), 0.14 (short-dashed), 0.25 (dotted), 0.58=<br />

ǫcr (short-dash-dotted), 0.92 (long-dash-dotted). For the last two values <strong>of</strong>ǫ the emission is<br />

almost the same and the lines overlap. The continuum line plots the emission <strong>of</strong> the vertical rim,<br />

computed forǫ=ǫcr.<br />

(Fig. 3.4). In particular, FNIR vanishes for face-on <strong>disks</strong>, and is maximum (<strong>of</strong> the order<br />

<strong>of</strong> 20%) for very inclined systems, just before self-absorption sets in.<br />

There are also differences between the curved and vertical rim models concerning<br />

the shape <strong>of</strong> the <strong>pre</strong>dicted 3µm bump. While the vertical rim has a constant temperature<br />

<strong>of</strong> 1400K over all its surface, in the curved rim the temperature varies from 1400K on<br />

the mid plane to about 1200K at the outer edge (see Fig. 3.3) and the SED is broader<br />

than a single-temperature black body. In practice, however, this is only a minor effect.<br />

3.3.3 Rim images<br />

Fig. 3.5 shows synthetic images <strong>of</strong> the curved rim at 2.2µm for grains withǫ= 0.08<br />

(radius <strong>of</strong> about 0.1µm) andǫ=ǫcr= 0.58 (radius <strong>of</strong> about 1.3µm). As described in


3.3 Results 55<br />

Sec. 3.3.1 (see Fig. 3.2 and Fig. 3.3), the inner radius <strong>of</strong> the rim is larger for smaller<br />

grains and the surface brightness is lower, due to the lower effective temperature <strong>of</strong> the<br />

emitting surface.<br />

For the comparison with the vertical model <strong>of</strong> the rim, the left panels <strong>of</strong> Fig. 3.5<br />

show the images <strong>of</strong> the vertical rim, calculated forǫ= 0.58. The largest difference is<br />

at low inclinations: for i=0 ◦ (face-on <strong>disks</strong>) the vertical rim vanishes, as the projected<br />

emitting surface along the line <strong>of</strong> sight is zero, while the curved rim has a centrally<br />

symmetric ring shape. For a distance <strong>of</strong> 144pc, the inner radius <strong>of</strong> the ring, forǫ=ǫcr,<br />

is about 4 milliarcsecond (mas); its brightness peaks practically at Rin (Rpeak/Rin∼ 1.05)<br />

and decreases slowly (by about a factor <strong>of</strong> two) outward, until at Rout it drops to very<br />

low values. At this distance, the width <strong>of</strong> the rim (roughly its FWHM) is about 0.8 mas.<br />

For higher inclination the central symmetry is lost and the projected image is an<br />

ellipse with one edge brighter than the other. In general, however, the brightness distri-<br />

bution <strong>of</strong> the rounded rim is much more symmetric than that <strong>of</strong> the vertical rim.<br />

3.3.4 Grain size distribution<br />

The results shown so far have been computed assuming that all the grains have the same<br />

composition and size. This is in practice unrealistic, and one wonders how the results<br />

will change if grains with different properties are <strong>pre</strong>sent. Different grains absorb differ-<br />

ently the stellar radiation and reach very different equilibrium temperatures. The degree<br />

<strong>of</strong> complexity <strong>of</strong> the radiation transfer problem increases remarkably (see Wolfire and<br />

Cassinelli, 1986, 1987) and there is no approximate solution such as Eq. 3.1. In fact, to<br />

the best <strong>of</strong> our knowledge even numerical solutions are not available to describe accu-<br />

rately the transition region where some grains are cooler than evaporation and survive<br />

while others do not.<br />

We can however estimate the effect <strong>of</strong> a grain mixture in the following way. Let us<br />

consider for simplicity a classical MRN grain size distribution, characterized by a power<br />

law a −q with a the radius <strong>of</strong> the dust grains, q=3.5, and a varying between a minimum<br />

value amin= 0.01µm and a maximum value amax≫amin. From what we have discussed


56 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

Figure 3.5: Synthetic images <strong>of</strong> the curved rim for different values <strong>of</strong>ǫ and inclination angle<br />

(i = 0 ◦ is face-on); comparison with the vertical rim. The left and central panels show the<br />

images <strong>of</strong> the curved rim calculated forǫ= 0.08 (silicate grains with radius a=0.1µm) and<br />

ǫ=ǫcr= 0.58 (a=1.3µm). The right panels show the images for the vertical rim calculated<br />

forǫ= 0.58. The surface brightness <strong>of</strong> the rim, plotted in colors, is computed for a wavelength<br />

<strong>of</strong> 2.2µm. The stellar parameters are as in Sec. 3.3, the distance is d=144pc. The vertical rim<br />

for i=0 ◦ has zero brightness. Note that as discussed in Sec. 3.3 forǫ>ǫcr the images <strong>of</strong> the<br />

curved rim re<strong>main</strong> the same.<br />

in Sec. 3.2, if amax is smaller than 1.3µm (corresponding toǫ = ǫcr for the adopted<br />

silicate grains), all the grains have a maximum temperature atτd= 0 and the higher


3.4 Discussion 57<br />

the value <strong>of</strong>ǫ, the smaller the evaporation distance from the star. As soon as the largest<br />

grains can survive, the stellar radiation is rapidly absorbed and all the other grains will<br />

also survive. In this case, the shape <strong>of</strong> the inner rim is controlled by the largest grains in<br />

the distribution. The shapes in Fig. 3.2 and the emitted fluxes shown in Fig. 3.4 should<br />

not change significantly as long as one inter<strong>pre</strong>tsǫas relative to the largest grains.<br />

The situation is more complex if amax> 1.3µm, since the grains withǫ>ǫcr can<br />

survive near the star only in an optically thin regime. However, our single-grain models<br />

show that there is very little difference in the location and shape <strong>of</strong> the rim as soon as<br />

ǫ>ǫcr. This suggests that varying amax above the “transition” value (i.e., the value <strong>of</strong><br />

ǫ at which the dependence <strong>of</strong> T onτchanges from decreasing to increasing) will not<br />

change the rim properties any further, at least to zero order.<br />

It is also likely that grains have not just different size but also different chemical<br />

composition. In this case, the rim location and properties are determined by the dust<br />

species with the highest evaporation temperature, as long as its contribution to the disk<br />

opacity is sufficient to make the disk optically thick. As discussed in Sec. 3.2, in the<br />

Pollack et al. (1994) dust model silicates have the highest evaporation temperature. This<br />

is why we have performed all our calculations for silicate grains. However, one should<br />

keep in mind that, in some dust models, graphite contributes most <strong>of</strong> the opacity at<br />

short wavelengths; since graphite has a much higher evaporation temperature (Tevp∼<br />

2000 K) than silicates, its <strong>pre</strong>sence would move the rim much closer to the central star.<br />

Although the details <strong>of</strong> the rim shape may change, its curving, which is caused only by<br />

the dependence <strong>of</strong> Tevp on the density, will not disappear.<br />

3.4 Discussion<br />

3.4.1 The 3µm bump<br />

The curvature <strong>of</strong> the inner side <strong>of</strong> the rim has important con<strong>sequence</strong>s on the observable<br />

near-IR excess, as shown in Fig. 3.4, which shows that the <strong>pre</strong>dicted near-IR excess


58 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

Figure 3.6: Histogram <strong>of</strong> the observed values <strong>of</strong> LNIR/L⋆ for a sample <strong>of</strong> 16 Herbig Ae stars<br />

from Natta et al. (2001) and Dominik et al. (2003). Note that the Dominik et al. values<br />

are computed between 2 and 7µm, while those from Natta et al. between 1.25 and 7µm. The<br />

uncertainties are in most cases rather large, due to the subtraction <strong>of</strong> the photospheric flux, which<br />

contributes significantly to the observed ones at short wavelengths, and to variability (see Natta<br />

et al. for some examples).<br />

ranges between about 10 and 20%, and that there is no significant dependence on the<br />

inclination <strong>of</strong> the disk with respect to the observer. These results compare very well<br />

with the existing <strong>observations</strong>. Fig. 3.6 shows the observed values for a well-studied<br />

sample <strong>of</strong> Herbig Ae stars from Natta et al. (2001) and Dominik et al. (2003); <strong>of</strong> a total<br />

<strong>of</strong> 16 objects, only one (HD 142527) has LNIR/L⋆> 0.25, and one (HD 169142)< 0.09.<br />

The <strong>observations</strong> <strong>of</strong> the near-IR SEDs <strong>of</strong> Herbig Ae stars have also shown that there<br />

is no systematic variation <strong>of</strong> the near-IR excess with the inclination <strong>of</strong> the disk. The


3.4 Discussion 59<br />

sample <strong>of</strong> Fig. 3.6 includes some rather face-on objects such as AB Aur (LNIR/L⋆∼<br />

0.20, i∼20 ◦ − 40 ◦ ; Fukagawa et al., 2004) and HD 163296 (LNIR/L⋆∼ 0.21, i∼30 ◦ ;<br />

Grady et al., 2000) and some UXORS variable, such as UX Ori, WW Vul and CQ Tau,<br />

which are generally considered to be close to edge-on, and have values <strong>of</strong> LNIR/L⋆∼<br />

0.12−0.25. This, which has been a puzzle for the vertical rim, finds a natural physical<br />

explanation in our models, which <strong>pre</strong>dict that the near-IR excess depends little on the<br />

inclination and that, in particular, does not vanish for face-on objects. Actually, our<br />

models <strong>pre</strong>dict that the largest excesses should be seen in face-on objects, and it would<br />

be interesting to explore in detail if this is indeed the case. However, for such a study to<br />

be significant, one would need a large sample <strong>of</strong> objects with known inclinations, which<br />

is at <strong>pre</strong>sent not available.<br />

Another interesting aspect <strong>of</strong> our results is the dependence <strong>of</strong> LNIR/L⋆ on grain prop-<br />

erties. If taken at face-value, one should expect that only objects with relatively large<br />

grains (greater than about 1µm) can have large values <strong>of</strong> LNIR/L⋆, and that low val-<br />

ues <strong>of</strong> the near-IR excess can only occur in face-on objects with small grains. These<br />

properties <strong>of</strong> the rim emission are potentially important for a better understanding <strong>of</strong><br />

grain properties in the inner disk and should be pursued further in the future, combining<br />

<strong>observations</strong> <strong>of</strong> <strong>disks</strong> with well known inclinations with models that explore a broader<br />

range <strong>of</strong> grain properties than we considered in this Chapter.<br />

3.4.2 The rim radius<br />

One side product <strong>of</strong> our models is Rin, the distance from the star to the rim on the disk<br />

mid plane. For a face-on disk, Rin practically coincide with the position <strong>of</strong> the peak <strong>of</strong><br />

the near-IR brightness. For fixed values <strong>of</strong> the stellar parameters, Rin depends on the<br />

dust properties; its value is shown in Fig. 3.7 for different values <strong>of</strong>ǫ. Larger grains<br />

can survive closer to the star than smaller grains and, because <strong>of</strong> the inversion in the<br />

temperature gradient, Rin does not change much with the grain size once this exceeds<br />

the value for whichǫ=ǫcr, so that for L⋆= 47L⊙, Rin > ∼ 0.5AU. Rin scales roughly as<br />

L 0.5<br />

⋆ , so that one can expect Rin > ∼ 0.1 AU for L⋆∼ 2L⊙. These values, especially for


60 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

small grain sizes, are larger than the <strong>pre</strong>dictions <strong>of</strong> simple, optically thin calculations,<br />

because Eq. 3.1 correctly includes the effect <strong>of</strong> the diffuse radiation field. Note that if<br />

backward scattering cannot be neglected, the temperature atτ=0 will be even higher,<br />

increasing further the size <strong>of</strong> the inner hole (see Appendix).<br />

Values <strong>of</strong> Rin have been derived for a number <strong>of</strong> objects from near-IR interfero-<br />

metric <strong>observations</strong> (see, for example, Akeson et al., 2000 and 2002; Millan-Gabet et<br />

al., 2001; Monnier and Millan-Gabet, 2002; Colavita et al., 2003; Eisner et al., 2003,<br />

2004 and 2005). The results depend not only on the assumed disk model, but also on<br />

other quantities such as the stellar SED, the disk inclination etc., so that a direct com-<br />

parison with our values <strong>of</strong> Rin can be misleading. We note, however, that Rin ∼ 0.5<br />

AU for intermediate-mass objects is roughly consistent with the <strong>observations</strong>, suggest-<br />

ing that in many objects grains have grown to sizes a> ∼ 1.3µm (see also Monnier and<br />

Millan-Gabet, 2002).<br />

This conclusion, <strong>of</strong> course, needs to be taken with great caution. There are a number<br />

<strong>of</strong> effects that can change the model-<strong>pre</strong>dicted Rin, in addition to different grain sizes.<br />

For example, on one hand smaller values <strong>of</strong> Rin can be due to some low-density gas<br />

in the inner disk hole, able to absorb the UV continuum from the star (Monnier and<br />

Millan-Gabet, 2002; Akeson at al., 2005). Also, the <strong>pre</strong>sence <strong>of</strong> graphite, which has<br />

an evaporation temperature <strong>of</strong> about 2000 K, can decrease Rin. On the other hand, the<br />

<strong>pre</strong>sence <strong>of</strong> accretion at a significant rate can increase Rin by heating grains in the inner<br />

disk to temperatures higher than those produced by the photospheric radiation alone<br />

(Muzerolle et al., 2004).<br />

In practice, one needs to compare in detail the model <strong>pre</strong>dictions <strong>of</strong> the quantities<br />

measured with interferometers (visibility curves, their dependence on baseline and hour<br />

angle, phase measurements) for specific, well-known objects. We have performed such<br />

a study with the aim <strong>of</strong> assessing the impact <strong>of</strong> curved rim models on the inter<strong>pre</strong>tation<br />

<strong>of</strong> current (and future) interferometric <strong>observations</strong>. In particular, we have included a<br />

discussion <strong>of</strong> the effect <strong>of</strong> the asymmetries <strong>of</strong> the rim projected images when not face-<br />

on. The results are <strong>pre</strong>sented in the next Chapter.


3.5 Summary and conclusions 61<br />

R in (AU)<br />

1.3<br />

1.1<br />

0.9<br />

0.7<br />

0.5<br />

0.3<br />

0.2 0.4 0.6 0.8 1<br />

Figure 3.7: Behaviour <strong>of</strong> the inner radius <strong>of</strong> the rim for different values <strong>of</strong>ǫ, evaluated for the<br />

star and disk parameters described in Sec. 3.3. Note that forǫ> ∼ ǫcr∼ 0.58 (corresponding to<br />

silicate grains <strong>of</strong> 1.3µm) the radius <strong>of</strong> the inner rim is almost constant around 0.5 AU.<br />

3.4.3 LkHa101<br />

To the best <strong>of</strong> our knowledge, the only image <strong>of</strong> the inner region <strong>of</strong> a circumstellar disk<br />

is that obtained by Tuthill et al. (2001), for LkHa101 using Keck in the H and K bands.<br />

The results are reminiscent <strong>of</strong> our images, showing an elliptical ring with a side much<br />

brighter than the other. The similarity is very interesting but one should keep in mind<br />

that LkHa101 is an early B star with a luminosity between 500 and 50000 L⊙, depending<br />

on its distance. Our model may not apply to such an object (e.g., Monnier et al., 2005).<br />

3.5 Summary and conclusions<br />

In this Chapter, we discuss the properties <strong>of</strong> the inner puffed-up rim that forms in cir-<br />

cumstellar <strong>disks</strong> when dust evaporates. The existence <strong>of</strong> a rim has been claimed starting<br />

from the work <strong>of</strong> Natta et al. (2001), both on theoretical and observational grounds.<br />

Here, we investigate the shape <strong>of</strong> the illuminated face <strong>of</strong> the rim. We argue that this<br />

ε


62 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

shape is controlled by a fundamental property <strong>of</strong> circumstellar <strong>disks</strong>, namely their very<br />

large vertical density gradient, through the dependence <strong>of</strong> grain evaporation temperature<br />

on density. As a result, the bright side <strong>of</strong> the rim is naturally curved, rather than vertical,<br />

as expected when a constant evaporation temperature is assumed.<br />

We have computed a number <strong>of</strong> rim models that take into account this effect in a<br />

self-consistent way. A number <strong>of</strong> approximations have been necessary to perform the<br />

calculations, and we discuss their validity. The basic result (i.e., the curved shape <strong>of</strong> the<br />

rim illuminated face) appears to be quite robust.<br />

For a given star, the rim properties depend mostly on the properties <strong>of</strong> the grains, and<br />

very little on those <strong>of</strong> the disk itself, for example the exact value <strong>of</strong> the surface density.<br />

The distance <strong>of</strong> the rim from the star is determined by the evaporation temperature (at<br />

the density <strong>of</strong> the disk mid plane) <strong>of</strong> the dust species that has the highest evaporation<br />

temperature, as long as its opacity is sufficient to make the disk very optically thick;<br />

in the model <strong>of</strong> Pollack et al. (1994) <strong>of</strong> the dust in accretion <strong>disks</strong>, silicates have the<br />

highest evaporation temperature. Therefore, we have assumed in our models dust made<br />

<strong>of</strong> astronomical silicates, and varied their size over a large range <strong>of</strong> values. We find that<br />

the rim properties do not depend on size as soon as a> ∼ 1.3µm; the values <strong>of</strong> the rim radii<br />

observed with interferometers suggest that in many <strong>pre</strong>-<strong>main</strong>–<strong>sequence</strong> disk grains have<br />

grown to sizes <strong>of</strong> 1–fewµm at least.<br />

The curved rim (as the vertical rim) emits most <strong>of</strong> its radiation in the near and mid-<br />

IR, and provides a simple explanation for the observed values <strong>of</strong> the near-IR excess<br />

(the “3µm bump” <strong>of</strong> Herbig Ae stars). Unlike the vertical rim, for curved rims the<br />

near-IR excess does not depend much on the inclination, being maximum for face-on<br />

objects and only somewhat smaller for highly inclined ones. This is in agreement with<br />

the apparent similarity <strong>of</strong> the observed near-IR SED between objects seen face-on and<br />

close to edge-on.<br />

We have computed synthetic images <strong>of</strong> the curved rim seen under different inclina-<br />

tions. Face-on rims are seen as bright, centrally symmetric rings on the sky; increasing<br />

the inclination, the rim takes an elliptical shape, with one side brighter than the other.


3.6 Appendix 63<br />

However, the brightness distribution <strong>of</strong> curved rims re<strong>main</strong>s at any inclination much<br />

more centrally symmetric than that <strong>of</strong> vertical ones. In the next Chapter we will discuss<br />

the application <strong>of</strong> the curved rim models to the inter<strong>pre</strong>tation <strong>of</strong> near-IR interferometric<br />

<strong>observations</strong> <strong>of</strong> <strong>disks</strong>.<br />

3.6 Appendix<br />

In this appendix we discuss in detail some <strong>of</strong> the assumptions on radiation transfer made<br />

in building our rim models. We use as templates the results <strong>of</strong> a radiation transfer code<br />

developed by E. Krügel, which is described in Habart et al. (2004). The code considers<br />

a plane-parallel slab <strong>of</strong> dust, illuminated on one side by a star with properties as in Sec.<br />

3.3.<br />

Figure 3.8: Temperature as function <strong>of</strong> the optical depth to the stellar radiationτs for small<br />

grains (ǫ= 0.08) and very large grains (ǫ= 1). The dashed curves are the results <strong>of</strong> the radiation<br />

transfer code, the solid lines the temperature obtained from Eq. 3.1.


64 The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong><br />

3.6.1 Evaluation <strong>of</strong> Eq. 3.1 for single grains<br />

We show in Fig. 3.8 the comparison <strong>of</strong> the temperature derived from Eq. 3.1 with the<br />

results <strong>of</strong> the radiation transfer code for two cases, one withǫ= 0.08 (small grains) and<br />

one withǫ= 1 (very large grains). In both cases scattering is neglected. The temperature<br />

is plotted as function <strong>of</strong> the optical depth to the stellar radiationτs. One can see that Eq.<br />

3.1 gives the correct values <strong>of</strong> T forτs=0 and forτs=∞. For intermediate values <strong>of</strong><br />

τs, the results are in both cases accurate within 10%. This is more than adequate for the<br />

purposes <strong>of</strong> this Thesis.<br />

3.6.2 SED<br />

We compare now the SED <strong>of</strong> a black-body at temperature T(τd= 2/3) from Eq. 3.1, to<br />

the results <strong>of</strong> the radiation transfer code (dashed curves) for the same grains <strong>of</strong> Fig. 3.8.<br />

The approximation is very good for large grains. For small values <strong>of</strong>ǫ, the discrepancy<br />

is larger, and, in particular, one cannot reproduce any dust feature. However, also in the<br />

caseǫ= 0.08, the difference in LNIR/L⋆ is only <strong>of</strong> 4%.<br />

3.6.3 Scattering<br />

Scattering has the obvious effect <strong>of</strong> increasing the grain temperature at the slab surface<br />

and decreasing it at large optical depth. We compare in Fig. 3.9 the temperature pr<strong>of</strong>ile<br />

<strong>of</strong> grains withǫ= 0.08 when scattering is properly included in the radiation transfer<br />

(dashed line) rather than sup<strong>pre</strong>ssed (Qsc= 0.). The curves are labelled with the val-<br />

ues <strong>of</strong> ˜ω, defined as Qsc(1−g)/Qsc(1−g)+Qabs, where g measures the asymmetry <strong>of</strong><br />

the scattering phase function; ˜ω=0 if Qabs= 0 or if the scattering is forward peaked<br />

(g=1). All quantities are averaged over the stellar radiation field and the correspond-<br />

ing SEDs show that the value <strong>of</strong> LNIR/L⋆ decreases by about 25% when scattering is<br />

included.


3.6 Appendix 65<br />

Figure 3.9: Left panel: comparison <strong>of</strong> the SED computed with the radiation transfer code (dotted<br />

lines) and as a black-body at T(τd= 2/3) from Eq. 3.1 (solid line) forǫ= 0.08 andǫ= 1.0, as<br />

labelled. Right panel: Temperature pr<strong>of</strong>ile for grains withǫ= 0.08, when scattering is included<br />

Acknowledgements<br />

( ˜ω=0.7; dashed line) or sup<strong>pre</strong>ssed ( ˜ω=0.0; solid line).<br />

We would like to thank Endrik Krügel for having allowed us to make use <strong>of</strong> his radiation<br />

transfer code, Leonardo Testi, Kees Dullemond, Carsten Dominik, Giuseppe Lodato<br />

and Giuseppe Bertin for useful discussions. The authors acknowledge partial support<br />

by MIUR COFIN grant 2003/027003-001.


CHAPTER 4<br />

Large dust grains in the inner region <strong>of</strong><br />

circumstellar <strong>disks</strong><br />

This Chapter has been published in Astronomy & Astrophysics:<br />

A. Isella 1,2 , L. Testi 2 and A. Natta 2 , 2006 A&A, 451, 951<br />

“Large dust grain in the inner region <strong>of</strong> circumstellar <strong>disks</strong>”<br />

1 Dipartimento di Fisica, Univeristá di Milano, via Celoria 16, 20133 Milano, Italy<br />

2 INAF - Osservatorio Astr<strong>of</strong>isico di Arcetri, Largo E. Fermi 5, 50125, Firenze, Italy<br />

Abstract: Simple geometrical ring models account well for near-infrared interferometric<br />

<strong>observations</strong> <strong>of</strong> dusty <strong>disks</strong> surrounding <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars <strong>of</strong> intermediate mass. Such<br />

models demonstrate that the dust distribution in these <strong>disks</strong> has an inner hole and puffed-up<br />

inner edge consistent with theoretical expectations. In this Chapter, we reanalyze the available<br />

interferometric <strong>observations</strong> <strong>of</strong> six intermediate mass <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars (CQ Tau, VV Ser,<br />

MWC 480, MWC 758, V1295 Aql and AB Aur) in the framework <strong>of</strong> a more detailed physical<br />

model <strong>of</strong> the inner region <strong>of</strong> the dusty disk. Our aim is to verify whether the model will allow us<br />

to constrain the disk and dust properties. Observed visibilities from the literature are compared


68 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

with theoretical visibilities from our model. With the assumption that silicates are the most<br />

refractory dust species, our model computes self-consistently the shape and emission <strong>of</strong> the<br />

inner edge <strong>of</strong> the dusty disk (and hence its visibilities for given interferometer configurations).<br />

The only free parameters in our model are the inner disk orientation and the size <strong>of</strong> the dust<br />

grains. In all objects with the exception <strong>of</strong> AB Aur, our self-consistent models reproduce both<br />

the interferometric results and the near-infrared spectral energy distribution. In four cases,<br />

grains larger than∼1.2µm, and possibly much larger are either required by or consistent with<br />

the <strong>observations</strong>. The inclination <strong>of</strong> the inner disk is found to be always larger than∼30 ◦ , and<br />

in at least two objects much larger.<br />

4.1 Introduction<br />

Understanding the properties and evolution <strong>of</strong> the dust grains contained in proto-planetary<br />

<strong>disks</strong> around <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars is important because they are the seeds from which<br />

planets may form. We have now strong evidence that grains in <strong>disks</strong> are very different<br />

from the grains in the diffuse interstellar medium and in the molecular clouds from<br />

which <strong>disks</strong> form, as reviewed, e.g., by Natta et al. (2006). In many objects, observa-<br />

tions with millimeter interferometers have provided strong evidence that the grains in<br />

the outer and cooler regions <strong>of</strong> the disk (further than 50 AU from the star) have been<br />

hugely processed, and have grown from sub-micron sizes to millimeter and centimeter<br />

ones. Closer to the star, however, in the regions were planets are more likely to form,<br />

observational evidence has been confined to grains close to the disk surface. For these,<br />

which however account for a tiny fraction <strong>of</strong> the total dust mass, emission in the silicate<br />

features has shown a correlation between the shape <strong>of</strong> the feature and its strength that is<br />

inter<strong>pre</strong>ted as due to growth <strong>of</strong> the grains from size a∼0.1µm to a∼1µm (van Boekel<br />

et al., 2003 and 2004; Meeus et al., 2003). In this inner disk, the properties <strong>of</strong> the grains<br />

in the disk mid plane are still unknown.<br />

In the last few years, due to the new long baseline near-infrared interferometers,


4.2 Target stars and <strong>observations</strong> 69<br />

many important steps forward in the study <strong>of</strong> the internal regions <strong>of</strong> circumstellar <strong>disks</strong><br />

have occurred. The available near-infrared interferometric <strong>observations</strong> <strong>of</strong> T Tauri<br />

(TTS) and Herbig Ae (HAe) stars (Eisner et al., 2003 and 2004; Millan-Gabet et al., 2001;<br />

Tuthill et al., 2001; Monnier et al., 2005) confirm the idea that the inner disk properties<br />

are controlled by the dust evaporation process which produce a “puffed-up” inner rim at<br />

the dust destruction radius (Natta et al., 2001; Dullemond, Dominick and Natta, 2001,<br />

hereafter DDN01). In these models, the location and shape <strong>of</strong> the rim depends on the<br />

properties <strong>of</strong> grains located not on the disk surface but on its mid plane.<br />

As discussed in the <strong>pre</strong>vious Chapter, we have recently proposed models <strong>of</strong> the<br />

“puffed-up”inner rim which include a self-consistent description <strong>of</strong> the grain evapora-<br />

tion and its dependence on the gas density (Isella and Natta, 2005; hereafter IN05). IN05<br />

have explored a large range <strong>of</strong> grain properties, and discussed how the location <strong>of</strong> the<br />

rim depends on grain properties. In this Chapter, we will use the IN05 models to analyze<br />

the existing interferometric data <strong>of</strong> the best observed HAe stars to explore, in practice,<br />

the constraints on grain properties provided by this technique and their uncertainties.<br />

As a byproduct <strong>of</strong> the modeling process, one obtains also the orientation <strong>of</strong> the<br />

inner disk (i.e. its inclination with respect to the line <strong>of</strong> sight and its position angle);<br />

this can be compared with the orientation <strong>of</strong> the outer disk, obtained from millimeter<br />

<strong>observations</strong> <strong>of</strong> the molecular gas and dust emission and/or scattered light in the optical.<br />

The Chapter is organized as follows. In Sec. 4.2 we describe the available interfer-<br />

ometric <strong>observations</strong> <strong>of</strong> the target stars. The IN05 model for the inner rim is briefly<br />

summarized in Sec. 4.3 and used to fit the <strong>observations</strong> <strong>of</strong> the individual objects in<br />

Sec. 4.4. A comparison <strong>of</strong> the results with <strong>pre</strong>vious analysis <strong>of</strong> the same data is <strong>pre</strong>-<br />

sented in Sec. 4.5. Our results are discussed in Sec. 4.6. Conclusions follow in Sec. 4.7.<br />

4.2 Target stars and <strong>observations</strong><br />

Our sample is composed <strong>of</strong> six HAe stars (AB Aur, CQ Tau, VV Ser, MWC 480,<br />

MWC 758 and V1295 Aql), for which near-infrared interferometric <strong>observations</strong> ex-


70 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

Table 4.1: Stellar parameters. Stellar parameters are from Hillenbrand et al. (1992),<br />

van den Ancker et al. (1998), Chiang et al. (2001), Straizys et al. (1996), Mannings et<br />

al. (2000) and references therein.<br />

Source Spectral Type d T L M Av<br />

(pc) (K) (L⊙) (M⊙)<br />

AB Aur A0pe 144 9772 47 2.4 0.5<br />

MWC 480 A2/3ep+sh 140 8700 25 2.2 0.25<br />

MWC 758 A5IVe 230 8128 22 2.0 0.22<br />

CQ Tau A8 V/F2 IVea 100 8000 5 1.5 1.00<br />

VV Ser B9/A0 Vevp 260 10200 49 3.0 3.6<br />

V1295 Aql B9/A0 Vp+sh 290 8912 83 4.3 0.19<br />

ist in the literature. Tab. 4.1 summarizes the physical properties <strong>of</strong> the target stars. All<br />

the stars are classified as young stellar objects with masses ranging from 1.5 to 4.3 solar<br />

mass and a spectral type between A0/B9 and A8/F2. CQ Tau and VV Ser belong to the<br />

family <strong>of</strong> UXORs and are characterized by large and irregular variability.<br />

We use visibility measurements <strong>of</strong> the target stars from the literature, obtained with<br />

interferometric <strong>observations</strong> carried out with PTI (Palomar Testbed Interferometer) in<br />

K band (λ0 = 2.2µm,∆λ=0.4µm) described in Eisner et al. (2004). For AB Aur<br />

and V1295Aql, IOTA <strong>observations</strong> are also available (Millan-Gabet et al., 2001) for the<br />

K’(λ0= 2.16µm,∆λ=0.32µm) and H (λ0= 1.65µm,∆λ=0.30µm) bands.<br />

4.3 Model description<br />

We use a model based on the assumption that the near-infrared emission <strong>of</strong> HAe stars<br />

originates in the “puffed up” inner rim which forms in the circumstellar disk at the dust<br />

evaporation radius (Natta et al., 2001; DDN01).<br />

In IN05 we revised the concept <strong>of</strong> the “puffed up”inner rim, introducing the depen-


4.3 Model description 71<br />

Figure 4.1: Sketch <strong>of</strong> the structure <strong>of</strong> the inner part <strong>of</strong> a proto-planetary disk. Panel (a) shows<br />

a disk with an inner puffed up rim that casts a shadow over the outer part <strong>of</strong> the disk. Farther<br />

than 5-10AU from the central star, the flaring disk emerges from the rim shadow. Panel (b)<br />

<strong>pre</strong>sents the image <strong>of</strong> the inner rim, computed with the IN05 models for a star with an effective<br />

temperature <strong>of</strong> 10000K and disk inclination <strong>of</strong> 30 ◦ . Panel (c) shows a vertical (edge on) section<br />

<strong>of</strong> the inner rim. The curvature <strong>of</strong> the surface <strong>of</strong> the rim is caused by the variation <strong>of</strong> the dust<br />

evaporation temperature with the height above the disk mid plane (see text).<br />

dence <strong>of</strong> the dust evaporation temperature on the local gas density, following the dust<br />

model <strong>of</strong> Pollack et al. (1994). The <strong>main</strong> result is that the surface <strong>of</strong> the rim <strong>pre</strong>sents a<br />

curved shape (see Fig. 4.1), whose features are summarized in the following.


72 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

4.3.1 The dust evaporation and the “puffed-up” inner rim<br />

Following the suggestion <strong>of</strong> Natta et al. (2001), we assume that the dust component<br />

<strong>of</strong> a circumstellar accreting disk is internally truncated by the dust evaporation process,<br />

forming an inner hole <strong>of</strong> radius Revp inside <strong>of</strong> which only gas can survive. If the radiation<br />

absorption due to this inner gas is negligible (as is <strong>of</strong>ten the case; see, e.g., Muzerolle et<br />

al., 2004), dust evaporation occurs where the equilibrium temperature Td <strong>of</strong> grains em-<br />

bedded in the unattenuated stellar radiation field, equals their evaporation temperature<br />

Tevp. In IN05 we used an analytical solution <strong>of</strong> the radiation transfer problem (Calvet<br />

et al., 1991 and 1992) to calculate the grain temperature inside the disk and we showed<br />

that evaporation occurs at a distance from the star that can be ex<strong>pre</strong>ssed as:<br />

<br />

<br />

2<br />

1500<br />

Revp[AU]=0.034·<br />

Tevp<br />

L⋆<br />

L⊙<br />

<br />

2+ 1<br />

<br />

, (4.1)<br />

ǫ<br />

where L⋆ is the stellar luminosity andǫ is the ratio <strong>of</strong> the Planck mean opacity at Tevp<br />

to that at the stellar effective temperature T⋆,ǫ = κP(Tevp)/κP(T⋆). The quantityǫ<br />

measures the cooling efficiency <strong>of</strong> the grains; it depends on the wavelength dependence<br />

<strong>of</strong> the absorption efficiency <strong>of</strong> the grains and varies with grain composition and size.<br />

If the dust in the proto-planetary disk is composed <strong>of</strong> different types <strong>of</strong> grains, the<br />

location and structure <strong>of</strong> the inner rim depends on the properties <strong>of</strong> the grains with the<br />

highest evaporation temperature. In the dust model proposed by Pollack et al. (1994),<br />

the most refractory grains are silicates for which Tevp is given by the relation<br />

Tevp(r, z)=2000·[ρg(r, z)] 0.02 , (4.2)<br />

valid for the gas densityρg in the range between 10 −18 g/cm 3 and 10 −5 g/cm 3 . In the<br />

following analysis, we will therefore assume that the inner disk dust is made <strong>of</strong> silicates,<br />

with optical properties given by Weingartner and Draine (2001); thus,ǫ is uniquely<br />

defined by the grain radius a, and we will use a, rather thanǫ, as a model parameter.<br />

Assuming that the proto-planetary disk is in hydrostatic equilibrium in the gravita-<br />

tional field <strong>of</strong> the central star and that it is isothermal in the vertical direction z, the gas


4.3 Model description 73<br />

densityρg(r, z) has its maximum value on the mid plane and decreases with z as<br />

ρg(r, z)=ρg,0(r) exp(−z 2 /2h(r) 2 ), (4.3)<br />

where h is the <strong>pre</strong>ssure scale height <strong>of</strong> the disk. The mid plane density can be ex<strong>pre</strong>ssed<br />

as a power-law <strong>of</strong> rρg,0(r)=ρg,0(r0)(r0/r) γ , withγ<strong>of</strong> the order <strong>of</strong> 2–3 (see, e.g., Chiang<br />

and Goldreich, 1997).<br />

The decrease <strong>of</strong>ρg with z, combined with Eq. 4.2, implies that the silicate evapora-<br />

tion temperature varies from, i.e.,∼1500 K on the mid plane (assuming a typical gas<br />

density <strong>of</strong>∼ 10 −7 g/cm 3 ) to∼ 1000 K at z/h=6.4 and∼ 800 K at z/h=8. Since Tevp<br />

decreases with z, it is immediately clear from Eq. 4.1 that the distance from the star at<br />

which dust evaporates increases with z, describing a curved surface as shown in Fig. 4.1.<br />

The dependence <strong>of</strong> Tevp on the gas density is an important factor when computing<br />

the shape <strong>of</strong> the rim in the vertical direction, where the gas density varies by many<br />

orders <strong>of</strong> magnitude while the distance from the star is practically unchanged. In the<br />

radial direction, we expect relatively small variations <strong>of</strong>ρg,0, for any reasonable value <strong>of</strong><br />

the disk mass, so that the distance <strong>of</strong> the rim from the star, measured in the mid plane,<br />

is practically independent <strong>of</strong> the gas density.<br />

The emission <strong>of</strong> the rim is computed assuming that it originates from the surface<br />

characterized by an effective temperature Te f f = T(τd = 2/3), whereτd is the opti-<br />

cal depth for the emitted radiation. The Te f f surface, therefore, defines the observed<br />

location and shape <strong>of</strong> the rim. In IN05, we discussed how the Tevp and the Te f f sur-<br />

faces behave for small and large silicate grains, and showed that the Te f f surface moves<br />

closer to the star for increasing grain size until a critical value, which for silicates is<br />

about 1.2µm. Larger grains produce rims with Te f f surfaces practically independent <strong>of</strong><br />

a. Therefore, for a fixed stellar luminosity, silicates with a∼1.2µm give the minimum<br />

value <strong>of</strong> the distance <strong>of</strong> the rim from the star. Conversely, if the measured rim distance<br />

is equal to this minimum value, one can derive from it only a lower limit (∼ 1.2µm) to<br />

the grain size.<br />

The rim emission peaks at near-infrared wavelengths. Atλ< ∼ 5−7µm, one can assume<br />

that the observed flux is the sum <strong>of</strong> the stellar+rim emission, with only negligible


74 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

contribution from the outer disk (see, e.g., DDN01). We model the stellar photospheric<br />

flux using standard Kurucz model atmospheres.<br />

4.3.2 Visibility model<br />

Due to the limited coverage <strong>of</strong> the uv-plane <strong>of</strong> the existing near-infrared interferometers,<br />

it is not possible at <strong>pre</strong>sent to recover full images from the available data, and one has<br />

to resort to the analysis <strong>of</strong> the visibilities on given interferometric baselines.<br />

Starting from the synthetic images <strong>of</strong> the inner rim (see Fig. 3.5 at pg. 56), we<br />

compute the <strong>pre</strong>dicted visibility values using a Fast Fourier Transform recipe. For face-<br />

on inclination, due to the circular symmetry <strong>of</strong> the rim image, the visibility depends only<br />

on the length <strong>of</strong> the baseline B. For inclination greater than zero, the image <strong>of</strong> the rim<br />

has an “elliptical” shape: the minor axis decreases with increasing inclination and the<br />

upper half <strong>of</strong> the rim becomes brighter than the lower part. For baselines oriented along<br />

the direction <strong>of</strong> the minor axis <strong>of</strong> the rim image, the visibility decreases more slowly<br />

than for those oriented along the major axis. For all other orientations <strong>of</strong> the baseline,<br />

the visibility will have values intermediate between these two (see Fig. 4.3). Moreover,<br />

due to the Earth rotation during the observation, the baseline corresponding to a fixed<br />

telescope pair moves in the uv plane describing an ellipse. Along this ellipse, each point<br />

is related to the hour angle HA <strong>of</strong> the target object in the sky. In the next section we use<br />

the V 2 -B plot and V 2 -HA plot, to show how the models fit the <strong>observations</strong>.<br />

The visibility model takes into account the emission <strong>of</strong> the central star, modeled as<br />

a uniform disk <strong>of</strong> radius R⋆. If F⋆ and Fd are respectively the stellar and the inner rim<br />

flux at the wavelength <strong>of</strong> the observation, the total visibility is given by the relation:<br />

V 2 2 F⋆V⋆+ FdVd<br />

= , (4.4)<br />

F⋆+ Fd<br />

where V⋆ and Vd are the visibility values <strong>of</strong> the star and <strong>of</strong> the disk. Note that for the<br />

PTI configuration (baselines between 84m and 100m) V⋆ is in all cases very closed to 1.


4.4 Comparison with the <strong>observations</strong> 75<br />

4.4 Comparison with the <strong>observations</strong><br />

4.4.1 Model parameters<br />

Once the stellar and dust properties are fixed, the model-<strong>pre</strong>dicted visibilities depend on<br />

the dust grain radius a, which completely defines the rim structure, and two parameters<br />

(observational parameters in the following) that describe the orientation <strong>of</strong> the disk,<br />

namely the inclinationιand position angle PA.<br />

For each star, we firstly compute the <strong>pre</strong>dicted rim structure varying the size <strong>of</strong> the<br />

grains from very small to very large values. As discussed in Sec. 4.3, we assume that sil-<br />

icates are the most refractory component; we take the optical properties <strong>of</strong> astronomical<br />

silicates defined by Weingartner and Draine (2001). Other disk parameters (i.e., mass<br />

and density radial pr<strong>of</strong>ile) can be neglected in this analysis. We fixρg,0(Rrim)∼10 −7<br />

g/cm 3 which gives a total disk mass <strong>of</strong> about 0.1M⊙, for a nominal value <strong>of</strong>γ=2.5 and<br />

an outer disk radius <strong>of</strong> the disk <strong>of</strong> 200 AU.<br />

Once the structure <strong>of</strong> the “puffed up” inner rim is calculated, the <strong>pre</strong>dicted visibility<br />

depends on the orientation <strong>of</strong> the disk in the sky, defined by the inclinationι<strong>of</strong> the mid<br />

plane <strong>of</strong> the disk with respect to the line <strong>of</strong> sight and its position angle PA, measured<br />

from north to east and relative to the major axis <strong>of</strong> the projected image <strong>of</strong> the disk on<br />

the sky. The inclination is defined so thatι=0 ◦ identifies a face-on disk whileι=90 ◦<br />

corresponds to an edge-on disk. For inclinations higher than 80 ◦ the rim emission is<br />

likely absorbed by the outer regions <strong>of</strong> the disk and the IN05 model can not be applied.<br />

In practice, we compute visibility models for each object varying a,ιand PA in-<br />

dependently. We then select the best models calculating the reducedχ 2 between the<br />

visibility data and the theoretical values calculated at the same points in the uv plane.<br />

The observed near-infrared fluxes, and the IOTA data when available, are then “a poste-<br />

riori” used to check the quality <strong>of</strong> the fit and, when possible, to reduce the degeneracy<br />

due to the small number <strong>of</strong> visibility points. For some stars, the existing data do not<br />

constrain the parameters, but still define a range, outside <strong>of</strong> which the fit to the data is<br />

very poor.


76 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

Tab. 4.2 shows in column 2 the best values <strong>of</strong> the astronomical silicate radius a and,<br />

in column 3, the corresponding values <strong>of</strong> the radius <strong>of</strong> the inner rim Rrim. The two<br />

observational parameters i and PA are given in columns 4 and 5, respectively. Note that<br />

the free parameters are in boldface; Rrim is a derived quantity.<br />

4.4.2 MWC 758<br />

PTI visibilities are fitted by a family <strong>of</strong> models, with parameters varying between the<br />

two extreme cases shown in Fig. 4.4. In one case, the disk has small grains <strong>of</strong> radius<br />

a=0.17µm,ι=48 ◦ and PA=134 ◦ ; in the other, big grains with a≥1.2µm,ι=40 ◦<br />

and PA=145 ◦ . Models with a values within this range will fit the observed visibilities<br />

equally well, provided that we varyιand PA in an appropriately way.<br />

However, if we consider also the constraints set by the SED at near-infrared wave-<br />

lengths, we find that only models with big grains fit it reasonably well (see the right<br />

panel <strong>of</strong> Fig. 4.4). The best-fitting model (χ 2 r<br />

= 2.0) has then a≥1.2µm, inner rim<br />

radius is Rrim= 0.32AU, rim effective temperature (at z=0) is 1460K. The near-infrared<br />

flux, LNIR, integrated between 2µm and 7µm is 25% <strong>of</strong> the total stellar luminosity, sim-<br />

ilar to the observed value. Once we fix a, the formal uncertainties onι, estimated from<br />

the surface where the reducedχ 2 equalsχ 2 min + 1, are quite small,±3◦ . More realistic<br />

uncertainties are <strong>of</strong> the order <strong>of</strong> 10 ◦ for bothιand PA.<br />

Note that in MWC 758 the PTI visibilities define quite well the orientation <strong>of</strong> the<br />

disk, even when the SED is not used to constrain the grain size. In particular, the<br />

inclination cannot be lower than about 30 ◦ .<br />

4.4.3 VV Ser<br />

The results for the star VV Ser are shown in Fig. 4.5. As for MWC 758, the inter-<br />

ferometric <strong>observations</strong> allow different sets <strong>of</strong> model parameters. Namely, we obtain<br />

similar values <strong>of</strong> the reducedχ 2 (∼ 1.2) for all grain sizes a> ∼ 0.4µm. Over this range<br />

<strong>of</strong> a, inclination and position angle vary in the intervals 45 ◦ − 80 ◦ and 60 ◦ − 120 ◦ , re-


4.4 Comparison with the <strong>observations</strong> 77<br />

spectively, with lower inclinations for larger grains. The correlation betweenιand PA<br />

is very strong, and the uncertainties in these two parameters re<strong>main</strong> very large even for<br />

fixed a.<br />

Although the fit is never very good, the VV Ser SED is better accounted for by large<br />

grains (see the right panel <strong>of</strong> Fig. 4.5) and in Table 2 we show the best values <strong>of</strong> the<br />

parameters for a≥1.2µm. The rim effective temperature is 1400 K, the near-infrared<br />

excess is 21% <strong>of</strong> L⋆.<br />

4.4.4 CQ Tau<br />

The limited number <strong>of</strong> visibility points does not allow us to constrain all the parameters<br />

<strong>of</strong> the disk. Fig. 4.6 shows two models, with the same level <strong>of</strong> confidence (χ 2 r<br />

∼ 1); the<br />

two models have similar orientations (inclinations <strong>of</strong> 52 ◦ and 46 ◦ with position angles<br />

<strong>of</strong> 168 ◦ and 164 ◦ respectively) but very different grain sizes (a=0.3µm and a≥1.2µm)<br />

and radii <strong>of</strong> the inner rim (0.25AU and 0.16AU, respectively). For a ≥ 1.2µm, the<br />

effective temperature <strong>of</strong> the rim (at z=0) is Te f f = 1480K, while Te f f = 1050 for<br />

a=0.3µm. The SEDs <strong>of</strong> the two models are compatible with the observed fluxes, with<br />

LNIR= 13%−18% L⋆. All the models with a within this range, and similarιand PA,<br />

have similarχ 2 values. Outside this range, models give a much poorer fit to the data.<br />

As for MWC 758, the visibility data constrain well the orientation <strong>of</strong> the disk on the<br />

sky. The inclination, in particular, has to be quite large, 40 ◦< ∼ ι< ∼ 55 ◦ .<br />

4.4.5 V1295 Aql<br />

The PTI <strong>observations</strong> <strong>of</strong> V1295 Aql are characterized by a very small number <strong>of</strong> visi-<br />

bility points and the disk parameters are hardly constrained. Even adding the IOTA data<br />

does not help due to the big errors that affect these <strong>observations</strong>. Note also that PTI and<br />

IOTA <strong>observations</strong> are performed at different wavelengths, K and H respectively.<br />

As for CQ Tau, we show the models for the two extreme sets <strong>of</strong> parameters that give<br />

an equally good fit (Fig. 4.7) withχ 2 r<br />

∼ 1. All the intermediate combinations <strong>of</strong> a,ι


78 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

and PA can explain the <strong>observations</strong> as well. In order to fit the values <strong>of</strong> visibility, an<br />

inclination ranging between 40 ◦ and 65 ◦ is required. More face-on systems can not in<br />

general reproduce the visibility s<strong>pre</strong>ad in the IOTA data and the V 2 − HA behaviour <strong>of</strong><br />

the PTI points. The grain radius varies from a≥1.2µm (Rrim= 0.7 AU, Te f f= 1400 K)<br />

to a=0.3µm (Rrim= 1.2 AU, Te f f= 970 K) while the position angle can not be defined<br />

at all.<br />

The right panel <strong>of</strong> Fig. 4.7 shows the comparison between the observed and the<br />

<strong>pre</strong>dicted SED. More inclined <strong>disks</strong> can in general reproduce better the photometric<br />

measurements around 1.5µm. The near-infrared emission <strong>of</strong> disk with an inclination <strong>of</strong><br />

less than 50 ◦ is peaked at about 4µm and can not reproduce the infrared excess between<br />

1µm and 2µm; in both models, the integrated near-infrared flux is LNIR∼ 20% L⋆.<br />

4.4.6 MWC 480<br />

Also in this case, due to the narrow range <strong>of</strong> available baselines, the PTI visibilities <strong>of</strong><br />

MWC 480 are consistent with different sets <strong>of</strong> parameters at the same level <strong>of</strong> confidence<br />

(χ 2 r<br />

∼ 2). Fig. 4.8 shows the two extreme disk configurations characterized by quite<br />

similar parameters for the dust grain size (a=0.2µm−0.3µm; Rrim=0.63–0.53 AU and<br />

Te f f≃ 1250 K), but very different values <strong>of</strong> the inclination (ι=35 ◦ andι=60 ◦ ) and the<br />

position angle (ψ=60 ◦ andψ=168 ◦ ). Several intermediate configurations reproduce<br />

the observed data as well. The degeneracy can not be removed even using the SED<br />

(Fig. 4.8) which is similar in the two models (LNIR/L⋆= 0.18%−0.14%) and it is only<br />

roughly consistent with the photometric values.<br />

4.4.7 AB Aur<br />

AB Aur is the only HAe star that cannot be fitted with the IN05 models. The PTI<br />

visibilities require a face-on rim, consistent with the inclination derived from large-scale<br />

images in scattered light (Grady et al., 1999; Fukagawa et al., 2004) and at millimeter<br />

wavelengths (Corder et al., 2005; Piétu et al., 2005). For these inclinations, the V 2 data


4.5 Comparison with <strong>pre</strong>vious analysis 79<br />

imply a very small inner radius, about two times smaller that the smallest Rrim obtained<br />

using the IN05 model (see Fig. 4.2). If, to put the discrepancy in a more physical<br />

context, we take Tevp as a free parameter, we find good agreement with the PTI data for<br />

Tevp∼ 2800 K (dashed line), a value by far too high not only for silicates but also for<br />

any other type <strong>of</strong> grains (e.g., Pollack et al. 1994).<br />

The situation becomes even less clear if we consider also the IOTA <strong>observations</strong><br />

(squared points), since they seem to indicate the <strong>pre</strong>sence <strong>of</strong> a more inclined disk with an<br />

inner radius between 0.26 AU and 0.51 AU. No additional information can be obtained<br />

from the analysis <strong>of</strong> the spectral energy distribution, since all the inner rim models<br />

with an effective temperature between 1500 K and 2500 K are compatible with the<br />

photometric data. We will come back to AB Aur in Sec. 4.6.<br />

4.5 Comparison with <strong>pre</strong>vious analysis<br />

Fits to the same interferometric data analyzed in Sec. 4.4 have been obtained by Eisner<br />

et al. (2004, hereafter E04) assuming a toroidal shape for the puffed-up inner rim, based<br />

on the simplified DDN01 model. In these fits, the free parameters are the location <strong>of</strong> the<br />

rim Rrim and the two observational parameters,ι and PA. The E04 results are shown in<br />

Tab. 4.2.<br />

We note that for three objects (MWC 758, VV Ser and CQ Tau) the E04 inclinations<br />

are in agreement within the errors with the values obtained with the IN05 model, while<br />

for the other two (V1295 Aql and MWC 480) the E04ιestimates are consistent with<br />

the lowest value <strong>of</strong> the range derived in this Chapter.<br />

The largest differences are in the derived values <strong>of</strong> Rrim: the IN05 inner radii are<br />

always larger than E04 results, with a maximum difference <strong>of</strong> a factor∼ 3 if we consider<br />

our maximum Rrim in the MWC 480 system. While for CQ Tau the two values are almost<br />

the same, for all the other stars the difference is a factor 1.5 and 2. This discrepancy is<br />

<strong>main</strong>ly due to the difference between IN05 and E04 models. In particular, in IN05,<br />

the curved shape <strong>of</strong> the emitting surface is self-consistently calculated, allowing a more


80 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 50 100 150 200<br />

Baseline (m)<br />

PTI<br />

IOTA<br />

IN05 - R in =0.51<br />

E04 - R in =0.26<br />

Figure 4.2: V 2 data for AB Aur plotted in function <strong>of</strong> the baseline. The square and circle points<br />

refer respectively to the IOTA and PTI <strong>observations</strong> in K band. The solid line is relative to the<br />

best <strong>pre</strong>diction <strong>of</strong> the self-consistent model <strong>of</strong> inner rim (a≥1.2µm, Rrim=0.51 AU,ι=20 ◦ ).<br />

The dashed line is obtained with the IN05 model using a≥1.2µm with an ad hoc Tevp= 2800<br />

K (Rrim= 0.26 AU).<br />

correct determination <strong>of</strong> the dependence <strong>of</strong> the rim emission on the inclination <strong>of</strong> the<br />

disk.<br />

Moreover, the IN05 model takes into account the effect <strong>of</strong> the radiation transport<br />

within the disk, even if in an approximate way (see Appendix <strong>of</strong> the <strong>pre</strong>vious Chapter).<br />

This supplementary heating is neglected by E04, who calculate the dust temperature<br />

taking into account only the direct stellar radiation. The ratio between the two values <strong>of</strong><br />

Rrim is given by the relation<br />

Rrim<br />

ˆRrim<br />

= ˆǫ (2+1/ǫ), (4.5)<br />

where the ˆǫ, ˆRrim and ˆT 2 evp are the values used by E04 and the inner radius ˆRrim is given


4.6 Discussion 81<br />

by the relation<br />

ˆRrim= 1<br />

ˆT 2 <br />

L⋆<br />

4πσ<br />

evp<br />

1<br />

. (4.6)<br />

ˆǫ<br />

Assuming the same value <strong>of</strong> the dust emissivity (ǫ=ˆǫ), the ratio Rrim/ ˆRin is∼ 1 for<br />

ǫ


82 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

lar medium (a=0.01−0.1µm, Weingartner and Draine, 2001), confirming that grain<br />

growth has taken place in the innermost disk regions (van Boekel et al., 2004).<br />

Even if the <strong>pre</strong>dicted near-infrared excess agrees well with the photometric observa-<br />

tions, some interesting differences exist between the theoretical and the observed spec-<br />

tral energy distributions. With the exception <strong>of</strong> CQ Tau, the <strong>pre</strong>dicted SEDs always<br />

peak at a wavelength slightly longer than found in the <strong>observations</strong>: the flux at short<br />

wavelengths (between 1.5µm and 2.2µm) is thus generally underestimated while the<br />

flux between 2.2µm and 7µm is overestimated. This may be due to the fact that in our<br />

models the SED is computed assuming that each point on the surface <strong>of</strong> the rim emits<br />

as a black body at the local effective temperature. This approximation is energetically<br />

correct but may not reproduce the exact wavelength dependence <strong>of</strong> the emitted radiation<br />

(see Appendix in Chapter 3).<br />

The rim models fail only in the case <strong>of</strong> AB Aur, where silicates, <strong>of</strong> whatever size,<br />

produce rims that are too distant from the star to be consistent with the <strong>observations</strong>. As<br />

shown in Sec. 4.4.7, PTI and IOTA data give somewhat contradictory results, and more<br />

interferometric data are clearly required. However, unless further <strong>observations</strong> drasti-<br />

cally change the <strong>pre</strong>sent picture, the discrepancy between the rim model <strong>pre</strong>dictions and<br />

the data is highly significant, and some <strong>of</strong> the basic underlying assumptions need to be<br />

changed. It is possible that in AB Aur grains more refractory than silicates dominate the<br />

dust population in the inner disk; however, the Tevp required (∼ 2800 K) is too high for<br />

any dust species likely <strong>pre</strong>sent in <strong>disks</strong> (e.g., Pollack et al., 1994).<br />

It is more likely that gas in the dust-depleted inner region absorbs a significant frac-<br />

tion <strong>of</strong> the stellar radiation, shielding the dust grains which are therefore cooler than in<br />

our models. This requires a high gas density in the inner disk, as expected if the accre-<br />

tion rate is high; in general, the accretion rates <strong>of</strong> HAe stars (including AB Aur) are low<br />

enough to ensure that the gaseous disk re<strong>main</strong>s optically thin (Muzerolle et al. 2004).<br />

However, our knowledge <strong>of</strong> the accretion properties and gas <strong>disks</strong> <strong>of</strong> HAe stars is very<br />

poor, and should be investigated further.<br />

AB Aur may be more than just an oddity. The <strong>pre</strong>sence <strong>of</strong> optically thick gas inside


4.6 Discussion 83<br />

the inner rim has been proposed to explain the near-infrared interferometric <strong>observations</strong><br />

<strong>of</strong> some very bright Herbig Be stars, for which the visibility data suggest inner disk<br />

radii many times too small to be consistent with the puffed-up rim models (Malbet et<br />

al., 2006; Monnier et al., 2005). If this is the case, AB Aur could be the low-luminosity<br />

tail <strong>of</strong> the same phenomenon, which, given its small distance and large brightness, could<br />

be used to understand a whole class <strong>of</strong> objects.<br />

4.6.2 Inclination and position angle<br />

Inclination and PA are well constrained by the existing data only in one case (MWC 758).<br />

We want to stress, however, that values <strong>of</strong> the parameters outside the ranges given in<br />

Tab. 4.2 do not fit the data at all. In particular, there are no objects consistent with<br />

face-on <strong>disks</strong>, or in general with a centro-symmetric brightness distribution. This rules<br />

out models, such as those <strong>of</strong> Vinkovic et al. (2006), where most <strong>of</strong> the near-infrared<br />

flux is contributed by a spherically symmetric shell around the star, rather than by a<br />

circumstellar disk.<br />

Two stars (CQ Tau and VV Ser) belong to the group <strong>of</strong> UXOR variables, which<br />

are inter<strong>pre</strong>ted as objects with <strong>disks</strong> seen close to edge-on (Grinin et al., 2001; Natta<br />

and Whitney, 2000; Dullemond et al., 2003). We derive for them large inclinations, in<br />

agreement with this inter<strong>pre</strong>tation.<br />

For some <strong>of</strong> our targets, there are in the literature estimates <strong>of</strong> the orientation <strong>of</strong><br />

the outer disk on the plane <strong>of</strong> the sky obtained with millimeter interferometers (Testi el<br />

al., 2001 and 2003; Manning and Sargent, 1997; Piétu et al., 2005; Corder et al., 2005).<br />

These determinations refer to the outer disk, i.e., to spatial scales <strong>of</strong> 50–100 AU at<br />

least. The comparison with the values derived in the near-infrared for the inner disk (on<br />

scales <strong>of</strong> less than 1 AU) can provide information on possible distortions in the disk, e.g.<br />

variations <strong>of</strong> the inclination with radius. For the three <strong>disks</strong> with millimeter data (MWC<br />

758, CQ Tau and MWC 480), there is agreement (within the uncertainties) between the<br />

inclination obtained by infrared and millimeter <strong>observations</strong>. However, it is certainly<br />

<strong>pre</strong>mature to exclude the existence <strong>of</strong> disk distortions, given the large uncertainties that


84 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

affect both the millimeter and the near-infrared estimates. More accurate interferometric<br />

<strong>observations</strong> in the two wavelength ranges and self-consistent models <strong>of</strong> the disk at all<br />

physical scales are required.<br />

An interesting case is that <strong>of</strong> VV Ser, whose disk has recently been imaged as a<br />

shadow seen against the background emission in the 11.3 PAH feature (Pontoppidan et<br />

al., 2006); These authors derive an inclination (<strong>of</strong> the outer disk) <strong>of</strong> about 70 ◦ and a<br />

position angle <strong>of</strong> 13 ◦ ± 5 ◦ . While the inclination is consistent with the upper limit <strong>of</strong><br />

the range we obtain for the inner disk, the position angle is <strong>of</strong>f by almost 90 ◦ . This<br />

discrepancy is intriguing, and deserves further investigation.<br />

4.6.3 Improving the model constrains<br />

Near-infrared interferometric <strong>observations</strong> <strong>of</strong> <strong>disks</strong> around <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars are<br />

still few and sparse. It is clear from our analysis that even in the most favorable cases<br />

more visibility data at different baselines are necessary to narrow the range <strong>of</strong> possible<br />

disk inclinations and grain properties.<br />

Given the huge demands <strong>of</strong> telescope time that interferometric <strong>observations</strong> require,<br />

it is useful to make use <strong>of</strong> model <strong>pre</strong>dictions in <strong>pre</strong>paring the <strong>observations</strong> and in choos-<br />

ing the baseline configurations that can constrain the disk structure.<br />

CQ Tau re<strong>pre</strong>sents an example <strong>of</strong> how the IN05 model can be used in this context. To<br />

better constrain the inner rim structure, one will need <strong>observations</strong> with baselines longer<br />

than 130m (available with the VLT interferometer), for which the <strong>pre</strong>dicted values <strong>of</strong> the<br />

squared visibility parameters are very different for different disk models (see Fig. 4.6).<br />

On the other hand, <strong>observations</strong> with baseline shorter than 60m could better constrain<br />

the inner radius <strong>of</strong> MWC 480, since a degeneracy in the models is <strong>pre</strong>sent at longer<br />

baselines.<br />

V1295 Aql re<strong>pre</strong>sents a still different case, in which the degeneracy in the values <strong>of</strong><br />

the <strong>pre</strong>dicted squared visibility can be removed observing at distant hour angles for the<br />

same baseline configurations, in order to determine the visibility variations at the same<br />

baseline, due to the inclination <strong>of</strong> the disk.


4.7 Summary and conclusions 85<br />

4.7 Summary and conclusions<br />

In this Chapter we have analyzed the near-infrared interferometric <strong>observations</strong> <strong>of</strong> the<br />

six best observed HAe stars using the rim models developed by Isella and Natta (2005),<br />

discussed in Chapter 3. Our aim was to explore the potential <strong>of</strong> near-infrared interfer-<br />

ometry to constrain the properties <strong>of</strong> the grains in the inner <strong>disks</strong> <strong>of</strong> these stars.<br />

The basic assumptions <strong>of</strong> the IN05 rim models are that the inner disk structure is<br />

controlled by the evaporation <strong>of</strong> dust in the unattenuated stellar radiation field, as ex-<br />

pected if the gaseous <strong>disks</strong> have low optical depth, and that the most refractory grains<br />

are silicates. The IN05 self-consistent models for the puffed-up inner rim reproduce<br />

both the interferometric <strong>observations</strong> and the near-infrared spectral energy distribution<br />

<strong>of</strong> all the objects we have studied, with the exception <strong>of</strong> AB Aur, which we have briefly<br />

discussed.<br />

For the five stars where we are able to obtain a good fit to the data, we can estimate<br />

the grain sizes in the rim, i.e., in the mid plane <strong>of</strong> the inner disk. We find that in four<br />

cases grains larger than∼1.2µm are either required by or consistent with the data. Only<br />

in one case do we find that the existing data require a∼0.2−0.3µm. Note that this<br />

value <strong>of</strong> a=1.2µm is a lower limit to the grain size: grains can be much larger, since<br />

the rim location and shape do not change significantly if the grains grow further.<br />

As a result <strong>of</strong> the model-fitting, one derives also the inclination and position angle <strong>of</strong><br />

the disk on the plane <strong>of</strong> the sky. We find that, in general, these parameters are not well<br />

constrained by the existing data. However, in all cases we can fit, inclinations lower<br />

than 30 ◦ are not consistent with the <strong>observations</strong> and the surface brightness distribution<br />

can not be circularly symmetric. This rules out a spherical envelope as the dominant<br />

source <strong>of</strong> the near-infrared emission.<br />

For some objects, estimates <strong>of</strong> the inclination <strong>of</strong> the outer disk have been obtained<br />

from millimeter interferometric <strong>observations</strong>; within the uncertainties, they agree with<br />

the values obtained for the inner disk.<br />

Our analysis shows that near-infrared interferometry is a very powerful tool for un-


86 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong><br />

derstanding the properties <strong>of</strong> the inner <strong>disks</strong>, in particular when combined with physical<br />

models <strong>of</strong> these regions. However, at <strong>pre</strong>sent the existing data are for many objects still<br />

too sparse in their coverage <strong>of</strong> the uv plane to allow an accurate determination <strong>of</strong> the<br />

disk parameters. We expect that this will be improved in the future. In this context,<br />

since near-infrared interferometry is and will re<strong>main</strong> a very time demanding technique,<br />

we stress the importance <strong>of</strong> using physical models <strong>of</strong> the inner region <strong>of</strong> the disk in<br />

planning future <strong>observations</strong>.<br />

Acknowledgements<br />

We are indebted to Josh Eisner and Rafael Millan-Gabet for providing us with the PTI<br />

and IOTA data. The authors acknowledge partial support for this project by MIUR PRIN<br />

grant 2003/027003-001.


IN05 model Eisner et al. (2004) outer disk<br />

Source a Rrim ι PA Rrim ι PA ι PA<br />

(µm) (AU) (deg) (deg) (AU) (deg) (deg) (deg) (deg)<br />

MWC 758 ≥ 1.2 0.32 40 145 0.21 36 +3<br />

−2 127 +4<br />

−3 46 116<br />

VV Ser ≥ 1.2 0.54 50−70 60−120 0.47 42 +6<br />

−2 166 +17<br />

−6 72±5 13±5 (b)<br />

CQ Tau 0.3−≥1.2 0.16−0.25 40−55 145−190 0.23 48 +3<br />

−4 106 +4<br />

−5 63 +10<br />

−15 2±13 (c)<br />

V1295 Aql 0.3−≥1.2 0.7−1.2 40−65 0.55 23 +15<br />

−23<br />

MWC 480 0.2−0.3 0.53−0.63 30−65 0.23 28 +2<br />

−1 145 +9<br />

−6 20−40 147−180 (a,d)<br />

AB Aur impossible to fit 0.25 8 +7<br />

−8 15−35 (e, f,g,h)<br />

50−110<br />

Table 4.2: Best fitting model parameters. From column 2 to 5 are reported the best fit parameters for the “puffed-up” inner rim, obtained<br />

with the IN05 model: the grain radius a, the radius <strong>of</strong> the inner rim Rrim, the inclinationι and the position angle PA. The free parameters<br />

<strong>of</strong> the model are <strong>pre</strong>sented in bold face. Columns 6,7 and 8 show the values <strong>of</strong> the radius <strong>of</strong> the inner rim, the inclination and the position<br />

angle, obtained by Eisner et al. (2004). Finally, the last two columns show the available estimates <strong>of</strong> inclination and position angle for<br />

the external region <strong>of</strong> the disk: (a) Mannings and Sargent (1997); (b) Pontoppidan et al. (2006); (c) Testi et al. (2001 and 2003); (d)<br />

Simon et al. (2000); (e) Fukagawa et al. (2004); ( f ) Grady et al. (1999); (g) Corder et al. (2005); (h) Piétu et al. (2005).<br />

+6 (a)<br />

−5<br />

4.7 Summary and conclusions 87


Face-on disk image<br />

Baseline (m)<br />

V 2<br />

1<br />

0.5<br />

0<br />

Inclined disk image<br />

Baseline (m)<br />

Figure 4.3: Predicted visibilities for two different inclinations <strong>of</strong> the inner rim. The two panels on the left show the <strong>pre</strong>dicted image for<br />

a face-on rim (ι=0 ◦ ) and the relative visibility squared V 2 : since the image is circularly symmetric, the visibility squared values are the<br />

same for every baseline orientation. The two panels on the right show the <strong>pre</strong>dicted image for an inclined rim (ι=60 ◦ ). In this case, the<br />

values <strong>of</strong> V 2 depend on the baseline orientations: V 2 decreases rapidly for baselines oriented along the major axis <strong>of</strong> the image (solid<br />

line) while V 2 decreases slowly for a baseline oriented along the minor axis (large and small dashes). For the intermediate orientation<br />

(dashed line), V 2 is in between the two extreme values.<br />

V 2<br />

1<br />

0.5<br />

0<br />

88 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong>


V 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 50 100 150 200<br />

Baseline (m)<br />

MWC758<br />

V 2<br />

0.40<br />

0.20<br />

0.40<br />

0.20<br />

0.40<br />

0.20<br />

NS<br />

NW<br />

SW<br />

-3 -2 -1 0 1<br />

hour angle<br />

Figure 4.4: The left and central panels show K-band V 2 data for MWC 758 respectively plotted as function <strong>of</strong> the baseline and <strong>of</strong> the<br />

hour angle, for the three different PTI baseline orientations (NS, baseline length <strong>of</strong> 110 m in direction North-South; NW, 86 m direction<br />

North-West; SW, 87 m direction South-West). PTI measurements (Eisner et al., 2004) are shown by dots. The right panel shows the<br />

spectral energy distribution <strong>of</strong> MWC 758. The de-reddened photometric fluxes are from Eisner et al. (2004, filled circles), Malfait et<br />

al. (1998, empty circles), van den Ancker et al. (1998, filled squares) and Cutri et al. (2003, empty squares); the dotted line shows<br />

the photospheric stellar flux. The solid lines show the <strong>pre</strong>dictions <strong>of</strong> the best-fitting model with small grains (a=0.17µm,ι=48 ◦ ,<br />

PA=134 ◦ ); the dashed lines the best-fitting model with big grains (a≥1.2µm,ι=40 ◦ , PA=145 ◦ ). As described in Sec. 4.4.2, all<br />

the intermediate disk configuration can reproduce the <strong>observations</strong> at almost the same level <strong>of</strong> confidence. Each curve on the left panel<br />

corresponds to a different orientation <strong>of</strong> the baseline on the plane <strong>of</strong> the sky (see Fig. 4.3). The stellar parameters are given in Tab. 4.1.<br />

λ F λ (erg cm 2 s -1 )<br />

1e-07<br />

1e-08<br />

1e-09<br />

0.5<br />

1<br />

λ (μm)<br />

5<br />

4.7 Summary and conclusions 89


V 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 50 100 150 200<br />

Baseline (m)<br />

VVSer<br />

V 2<br />

0.40<br />

0.20<br />

0<br />

0.60<br />

0.40<br />

0.20<br />

0.40<br />

0.20<br />

NS<br />

NW<br />

SW<br />

0<br />

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5<br />

hour angle<br />

Figure 4.5: Same as Fig. 4.4 for VV Ser. The solid curves show the model <strong>pre</strong>dicted values <strong>of</strong> the K-band V 2 for a rim with a=0.52µm,<br />

ι=65 ◦ and PA=80 ◦ ; the dashed curves for a rim model with a≥1.2µm,ι=55 ◦ and PA=115 ◦ . The photometric data are from<br />

Eisner et al. (2004, filled circles), Cutri et al. (2003, empty squares), Hillenbrand et al. (1992, empty squares), Rostopchina et al. (2001,<br />

filled squares). Stellar parameters in Tab. 4.1.<br />

λ F λ (erg cm 2 s -1 )<br />

1e-07<br />

1e-08<br />

1e-09<br />

0.5<br />

1<br />

λ (μm)<br />

5<br />

90 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong>


V 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 50 100 150 200<br />

Baseline (m)<br />

CQTau<br />

V 2<br />

0.40<br />

0.20<br />

0.40<br />

0.20<br />

0.40<br />

0.20<br />

NS<br />

NW<br />

SW<br />

-2.5 -2 -1.5 -1 -0.5 0 0.5 1 1.5<br />

hour angle<br />

Figure 4.6: Same as Fig. 4.4 for CQ Tau. The solid curves show the <strong>pre</strong>dictions <strong>of</strong> the model with a=0.3µm (Rrim= 0.25 AU),ι=52 ◦<br />

and PA=168 ◦ ; the dashed curves the model with a≥1.2µm (Rrim=0.16 AU),ι=46 ◦ and PA=164 ◦ . The photometric data are<br />

from Eisner et al. (2004, empty squared), Glass and Penston (1974, filled squares), ISO catalogue (empty circles) and Natta et al. (2001,<br />

filled triangles) and references therein.<br />

λ F λ (erg cm 2 s -1 )<br />

1e-08<br />

1e-09<br />

0.5<br />

1<br />

λ (μm)<br />

5<br />

4.7 Summary and conclusions 91


V 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 50 100 150 200<br />

Baseline (m)<br />

V1295Aql<br />

V 2<br />

0.30<br />

0.20<br />

0.10<br />

0.30<br />

0.20<br />

NW<br />

SW<br />

0.10<br />

-0.5 0 0.5 1 1.5<br />

hour angle<br />

Figure 4.7: Same as Fig. 4.4 for V1295 Aql. The solid lines plots the results <strong>of</strong> a model with a≥1.2µm (Rrim= 0.7 AU),ι=63 ◦ and<br />

PA=162 ◦ . The dashed lines a model with a=0.3µm (Rrim= 1.2 AU),ι=40 ◦ and PA=80 ◦ . The photometric measures are from<br />

Eisner et al. (2004, filled circles), Malfait et al. (1998, open circles), Kilkenny et al. (1985, filled squared), Glass and Penston (1974,<br />

open squared). The short baseline points in the left panel (empty squares) are from IOTA, while the central panel shows only the PTI<br />

data.<br />

λ F λ (erg cm 2 s -1 )<br />

1e-07<br />

1e-08<br />

1e-09<br />

0.5<br />

1<br />

λ (μm)<br />

5<br />

92 Large dust grains in the inner region <strong>of</strong> circumstellar <strong>disks</strong>


V 2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0 50 100 150 200<br />

Baseline (m)<br />

MWC480<br />

V 2<br />

0.40<br />

0.20<br />

0.40<br />

0.20<br />

NW<br />

SW<br />

-3 -2 -1 0 1<br />

hour angle<br />

Figure 4.8: Same as Fig. 4.4 for MWC 480. The solid lines are for the disk model characterized by a=0.2µm (Rrim= 0.63 AU),ι=35 ◦<br />

and PA=30 ◦ . The dashed lines are relative to a disk model characterized by a=0.3µm (Rrim=0.53 AU),ι=60 ◦ and PA=102 ◦ . The<br />

photometric data are from Eisner et al. (2004, filled circles), Malfait et al. (1998, open circles), Cutri et al. (2003, empty squares) and<br />

van den Ancker et al. (1998, filled squares).<br />

λ F λ (erg cm 2 s -1 )<br />

1e-07<br />

1e-08<br />

1e-09<br />

0.5<br />

1<br />

λ (μm)<br />

5<br />

4.7 Summary and conclusions 93


CHAPTER 5<br />

More on the “puffed–up” inner rim<br />

After that Natta et al. (2001) and Tuthill et al. (2001) independently suggested that<br />

the dusty component <strong>of</strong> <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> circumstellar <strong>disks</strong> have to be internally<br />

truncated at the dust evaporation radius, many efforts have been made to understand the<br />

possible implications <strong>of</strong> this effect on the circumstellar disk structure.<br />

Dullemond, Dominik and Natta (2001, hereafter DDN) were the firsts to include the<br />

dust evaporation in a self-consistent numerical solution <strong>of</strong> the disk structure, obtaining<br />

the puffed-up inner rim model summarized in Fig. 5.1. Even if the model was based on<br />

some very crude approximations (i.e, negligible viscous and self–diffused disk heating,<br />

negligible gas absorption inside the dust evaporation radius, black-body dust grains,<br />

constant dust evaporation temperature), its <strong>pre</strong>dictions were in good agreement not only<br />

with the observed spectral energy distributions <strong>of</strong> many HAe and TTS stars, but also<br />

with the first interferometric measurements <strong>of</strong> the disk inner radii <strong>of</strong> Millan-Gabet et<br />

al. (2001). Moreover, the <strong>pre</strong>sence <strong>of</strong> disk shadowed regions, shielded from the stellar<br />

radiation by the rim, has been suggested to explain the different mid-infrared spectral<br />

slopes observed in a large sample <strong>of</strong> HAe stars (Meeus et al. 2001).<br />

The major problem <strong>of</strong> the DDN model was the <strong>pre</strong>dicted vertical rim surface and<br />

the consequent strong dependence <strong>of</strong> the rim emission on the disk inclination angle (see


96 More on the “puffed–up” inner rim<br />

rim radiation<br />

star<br />

inner hole shadowed region flaring disk<br />

inner rim<br />

CG97<br />

flaring disk<br />

R in<br />

H rim<br />

diffusion<br />

projection <strong>of</strong><br />

the wall<br />

diffusion<br />

R fl<br />

H cg<br />

disk surface<br />

disk atmosphere<br />

Figure 5.1: Sketch <strong>of</strong> the “vertical puffed-up inner rim” model from Dullemond, Dominick<br />

and Natta (2001). Tacking into account the dust evaporation that occurs close to the central<br />

star, the dusty component <strong>of</strong> the disk is truncated at radius Rin. At this distance, the dust is<br />

perpendicularly heated by the stellar radiation and is hotter than in the classical flaring disk<br />

model labeled with CG97 (from Chiang and Goldreich, 1997). The <strong>pre</strong>ssure scale height <strong>of</strong> the<br />

disk at Rin is thus higher than in CG97, giving rise to a “puffed-up” inner rim that appears as a<br />

“vertical wall”. The disk regions behind the rim lay in the rim shadow with a vertical structure<br />

driven only by the heating radial diffusion into the disk. At larger distances and out <strong>of</strong> the rim<br />

shadow, the disk has the same properties <strong>of</strong> the flaring disk model <strong>of</strong> CG97.<br />

Fig.3.4). It required that all the observed HAe and TTS stars had an inclination <strong>of</strong>∼60 o ,<br />

which was both physically improbable and in contrast with other observational evidence<br />

which indicated a face-on disk around AB Aur.<br />

The inner rim model published in Isella and Natta (2005, IN05) removed some <strong>of</strong><br />

the approximation <strong>of</strong> the DDN model. In particular, we (i) introduced a realistic dust<br />

R


5.1 Fuzzy and sharp rims 97<br />

opacity to calculate the evaporation distance, (ii) considered the additional heating term<br />

due to the radiation diffused by the disk itself, (iii) introduced the dependence <strong>of</strong> the dust<br />

evaporation temperature on the surrounding gas density to compute a more realistic rim<br />

shape.<br />

As a con<strong>sequence</strong> <strong>of</strong> the first point, we showed that it is possible to calculate the size<br />

<strong>of</strong> the dust <strong>pre</strong>sent in the inner disk from the observed inner disk radius; the results have<br />

been discussed in Chapter 3 and 4. The latter two points have been the subject <strong>of</strong> follow-<br />

up works recently appeared in the literature. The <strong>main</strong> results are briefly summarized in<br />

the following.<br />

5.1 Fuzzy and sharp rims<br />

Introducing the additional heating term due to the radiation field diffused by the disk–<br />

itself, we found that the rim atmosphere, defined as the region in which 99% <strong>of</strong> the<br />

stellar radiation is absorbed, can be either very sharp or fuzzy, depending on the size<br />

<strong>of</strong> the dust grains <strong>pre</strong>sent into the disk. Assuming a nominal star <strong>of</strong> L⋆= 50L⊙ and<br />

Teff=10000 K, our calculations show that silicate grains smaller than about 1µm are<br />

<strong>main</strong>ly heated by the stellar radiation field and that their temperature decreases along the<br />

rim atmosphere, starting from the evaporation temperature atτ=0. In this case, given<br />

the high density <strong>of</strong> the dust, all the stellar radiation is absorbed in a very narrow region if<br />

compared with the radius <strong>of</strong> the inner rim (see Fig. 3.1 at pg. 47). On the contrary, larger<br />

grains are more efficiently heated by the diffused infrared radiation originating from the<br />

disk interior and their temperature tends to increase in the rim atmosphere, starting from<br />

the value atτ=0at which the grains reach the evaporation temperature (see Fig. 3.8 at<br />

pg. 63). Since the grain temperature cannot be higher than the evaporation temperature,<br />

large grain can thus survive only in an optically thin atmosphere (τ∼0). In this case,<br />

the rim atmosphere can be very extended and comparable with the inner rim radius (Fig.<br />

3.1).<br />

Using a different formalism, the dependence <strong>of</strong> the rim atmosphere width on the dust


98 More on the “puffed–up” inner rim<br />

Figure 5.2: From Vinkovic (2006). Dust dynamics in protoplanetary <strong>disks</strong> brings dust closer to<br />

the star than the inner edge <strong>of</strong> optically thick dusty disk. Such a disk structure is possible if this<br />

zone is populated only by big grains and it is vertically optically thin. These conditions enable<br />

disk thermal cooling and dust survival. The resulting sublimation zone spans from the optically<br />

thin to the optically thick inner disk radius. Its physical size is∼0.2 AU in Herbig Ae stars and<br />

∼0.03 AU in T Tau stars.


5.2 New shapes for the rim 99<br />

grain size has been recently confirmed by Vinkovic (2006) who extended the calculation<br />

to the case <strong>of</strong> a mixture <strong>of</strong> grain sizes. He showed that dust grains <strong>of</strong> different size have<br />

different temperature at the surface <strong>of</strong> the rim (whereτ=0), while their temperature is<br />

the same in the disk interior. The exact grain size composition in the rim atmosphere<br />

depends on the amount <strong>of</strong> the external flux because different grain sizes sublimate at<br />

different distance from the star. Moreover, he confirmed that a large optically thin sub-<br />

limation zone (∼ 0.2 AU for Teff=10000 K) may exist in <strong>pre</strong>sence <strong>of</strong> micron size grains<br />

(see Fig. 5.2). Note that this value is almost the same obtained with the single-size<br />

approximation adopted in IN05 (see Fig. 3.1 at pg. 47).<br />

The gas <strong>pre</strong>sent inside the rim atmosphere may be enriched by metals coming from<br />

the sublimated dust, which can favour a further grain growth by recondensation on the<br />

surface <strong>of</strong> existing grains. The author also suggest that the sublimation zone may play an<br />

important role in dusty wind processes, since its small optical depth and small distance<br />

from the central star should result in gas properties that are more susceptible to non-<br />

gravitational forces capable <strong>of</strong> launching a dusty wind (such as magnetic fields) than<br />

the rest <strong>of</strong> the dusty disk. This conclusion is particularly interesting if combined with<br />

the indication that a disk wind launched just in front <strong>of</strong> the inner rim may be responsible<br />

for the Brγ emission detected in HD 104237 (see Chapter 6).<br />

5.2 New shapes for the rim<br />

The shape <strong>of</strong> the puffed-up inner rim in the IN05 model is driven by the dependence<br />

<strong>of</strong> the dust evaporation temperature on the surrounding disk density. Close to the mid-<br />

plane, where the density are high, the dust evaporation temperature is also high; moving<br />

away from the mid plane, the gas density rapidly decreases and the corresponding evap-<br />

oration temperature also drops. For fixed stellar parameters and dust properties, the<br />

resulting dust evaporation surface is therefore curved rather than vertical, as <strong>pre</strong>dicted<br />

by the DDN model.<br />

Recently, Tannirkulam et al. (2007; hereafter T07) have shown that dust growth


100 More on the “puffed–up” inner rim<br />

Rim height[A.U]<br />

0.15<br />

0.10<br />

0.05<br />

IN05 values for density dependent dust sublimation model<br />

TORUS values for density dependent dust sublimation model<br />

TORUS values for dust-segregation model (this work)<br />

_ _ _ _ Analytic estimate for dust-segregation model<br />

large<br />

grains<br />

dust<br />

segregation<br />

small<br />

grains<br />

0.00<br />

0.0 0.5 Radius[A.U] 1.0 1.5<br />

Figure 5.3: Courtesy <strong>of</strong> A. Tannirkulam, to be published in Tannirkulam at al. (2007). The<br />

figure shows the height <strong>of</strong> the inner rim above the disk mid plane. The rim is defined as the<br />

τ=1surface (forλ=5500Å) computed along radial lines from the central star. The solid lines<br />

labelled with “small grains” and “large grains” indicate the IN05 model calculated for single<br />

grain sizes <strong>of</strong> 0.1µm and 1.2µm respectively. The diamond indicate the rim calculated with the<br />

TORUS code for the same grain sizes and without the dust settling (only the density-dependent<br />

sublimation temperature is <strong>pre</strong>sent). The stars indicate the rim surface calculated with TORUS<br />

introducing the dust settling (the dust segregation model). Finally, the dashed line is an analytical<br />

estimate <strong>of</strong> the dust evaporation front for the dust segregation model.<br />

and settling can curve the rim surface on larger scales than the ones produced by the<br />

density-dependent sublimation temperature. A number <strong>of</strong> observational (i.e. Retting et<br />

al., 2006) and theoretical (Dullemond and Dominik, 2004; Tanaka et al., 2005) results<br />

have shown that during the disk evolution, small dust grains tend to coagulate in larger<br />

grains and settle towards the disk mid plane. In absence <strong>of</strong> strong vertical turbulence the<br />

disk should therefore be characterized by a vertical grain size stratification dominated<br />

by large grains in the mid plane and small ones in the disk upper layers. Since the dust<br />

sublimation distance depends on the grain size (i.e., larger grains can survive closer to<br />

the central star), dust settling may thus lead to a significant rim curvature. Fig. 5.3 shows<br />

the results obtained by T07 using a Monte Carlo approach to calculate the structure <strong>of</strong><br />

an inner rim, where the vertical dust size stratification has been parametrized following<br />

the <strong>pre</strong>dictions <strong>of</strong> theoretical grain evolution models (Dullemond and Dominick, 2004).<br />

In absence <strong>of</strong> dust settling, the rim shape is in good agreement with the IN05 models


5.3 New observational results 101<br />

(see the solid lines and the empty diamonds in the figure). On the contrary, when the<br />

settling operates, the rim shape is strongly curved (see the stars in the figure). The<br />

rim dominates the near-infrared disk emission and it appears as a thick ring, larger<br />

than the IN05 model, when seen at face-on inclinations. At higher inclination the T07<br />

model <strong>pre</strong>dicts more symmetric images than those showed in Fig. 3.5, corresponding<br />

to smaller closure phase values when observed with near infrared interferometers. One<br />

should note that the dust settling parametrization used by T07 is valid in absence <strong>of</strong><br />

vertical turbulent motions which mix the circumstellar material and destroy the grain<br />

size stratification. If a significant degree <strong>of</strong> vertical turbulence exists (see i.e. Ilgner and<br />

Nelson, 2006), the IN05 rim shapes would be more correct. In general, the inner rim<br />

should have a shape in between the to extreme cases <strong>pre</strong>dicted by IN05 and T07. In any<br />

case, the dusty disk inner rim will have a fuzzy and significantly curved surface, very<br />

different from the “vertical wall” <strong>pre</strong>dicted by the first models <strong>of</strong> Dullemond, Dominick<br />

and Natta (2001).<br />

5.3 New observational results<br />

Given the spatial resolution required to observed the inner regions <strong>of</strong> nearby <strong>pre</strong>-<strong>main</strong><br />

<strong>sequence</strong> disk, only near-infrared interferometry allow to clarify the real structure <strong>of</strong><br />

the dusty disk inner rim. As we have discussed in Chapters 2 and 4, all the current<br />

results are based on measuring the module <strong>of</strong> fringe visibility, while the visibility phase<br />

is required to unambiguously detect deviation from simple symmetries. While atmo-<br />

spheric turbulence corrupts the direct measurement <strong>of</strong> the fringe phase, by combining<br />

three or more telescopes it is possible to measure the closure-phase, a phase quantity<br />

that depends only on the source intrinsic phase. Today, only three facilities (the Palo-<br />

mar Testbed Interferometer, PTI; the Infrared Optical Telescope Array, IOTA; and the<br />

VLT Interferometer ,VLTI) allow closure phase measurements and IOTA is the only one<br />

that has produced useful <strong>observations</strong> <strong>of</strong> young stellar objects so far. In particular, Mon-<br />

nier et al. (2006) have recently performed a closure-phase survey <strong>of</strong> 8 nearby HAe stars


102 More on the “puffed–up” inner rim<br />

Milliarcseconds<br />

Milliarcseconds<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

HD 45677<br />

Elliptical Ring Model<br />

15 10 5 0 -5 -10 -15<br />

Milliarcseconds<br />

HD 45677<br />

Skewed Ring Model<br />

15 10 5 0 -5 -10 -15<br />

Milliarcseconds<br />

Data Vis2<br />

Data Vis2<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.0 0.2 0.4<br />

Model Vis2<br />

0.6 0.8<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.0 0.2 0.4<br />

Model Vis2<br />

0.6 0.8<br />

Data Closure Phases (degs)<br />

Data Closure Phases (degs)<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-40 -30 -20 -10 0 10<br />

Model Closure Phases (degs)<br />

10<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-40 -30 -20 -10 0 10<br />

Model Closure Phases (degs)<br />

Figure 5.4: From Monnier at al. (2006). Top left panel shows the best-fit elliptical ring model to<br />

the IOTA data for the star HD 45677 (north is up and east is left). Top middle panels compares<br />

the observed visibility data to the model data. Top right panel compares the observed closure<br />

phases with the model closure phases. This kind <strong>of</strong> models re<strong>pre</strong>sents spherical or elliptical<br />

envelopes as <strong>pre</strong>dicted by Vinkovic et al. (2006). Since this model is constrained to be cen-<br />

trosymmetric, all closure phases are identically zero and cannot fit the <strong>observations</strong>. The bottom<br />

panels <strong>pre</strong>sent the best-fit “skewed ring model”. The closure phases are fairly well explained by<br />

a strong northwest skew. Some data points are not well explained by these models, suggesting<br />

a patchier dust distribution. This model is similar to what <strong>pre</strong>dicted by Isella and Natta (2005)<br />

and Tannirkulam et al. (2007).


5.3 New observational results 103<br />

with an angular resolution <strong>of</strong> about 5×12 milli-arcsec at 1.65µm. The data show the<br />

<strong>pre</strong>sence <strong>of</strong> a resolved continuum emission in agreement with dust depleted inner holes<br />

with radius between∼0.2–0.5 AU. In the case <strong>of</strong> AB Aur, RY Tau and HD 163296, the<br />

measured visibilities show that a small fraction <strong>of</strong> the observed emission (


Part III<br />

Gas in the inner disk


CHAPTER 6<br />

Constraining the wind launching region<br />

in Herbig Ae stars: AMBER/VLTI<br />

spectroscopy <strong>of</strong> HD 104237<br />

This chapter will appear in Astronomy & Astrophysics<br />

and can be found at www.arxiv.org/abs/astro-ph/0606684.<br />

E. Tatulli 2 , A. Isella 1,2 , A. Natta 2 , L. Testi 2 and the AMBER consortium 3 , A&A in <strong>pre</strong>ss<br />

“Constraining the wind launching region in Herbig Ae stars:<br />

AMBER/VLTI spectroscopy <strong>of</strong> HD 104237”<br />

1 Dipartimento di Fisica, Univeristá di Milano, via Celoria 16, 20133 Milano, Italy<br />

2 INAF - Osservatorio Astr<strong>of</strong>isico di Arcetri, Largo E. Fermi 5, 50125, Firenze, Italy<br />

3 http://amber.obs.ujf-grenoble.fr<br />

Abstract: We investigate the origin <strong>of</strong> the Brγ emission <strong>of</strong> the Herbig Ae star HD 104237<br />

on Astronomical Unit (AU) scales. Using AMBER/VLTI at a spectral resolutionR=1500 we<br />

spatially resolve the emission in both the Brγ line and the adjacent continuum. The visibility


108 AMBER/VLTI spectroscopy <strong>of</strong> HD 104237<br />

does not vary between the continuum and the Brγ line, even though the line is strongly detected<br />

in the spectrum, with a peak intensity 35% above the continuum. This demonstrates that the line<br />

and continuum emission have similar size scales. We assume that the K-band continuum excess<br />

originates in a “puffed-up” inner rim <strong>of</strong> the circumstellar disk, and discuss the likely origin<br />

<strong>of</strong> Brγ. We conclude that this emission most likely arises from a compact disk wind, launched<br />

from a region 0.2-0.5 AU from the star, with a spatial extent similar to that <strong>of</strong> the near infrared<br />

continuum emission region, i.e, very close to the inner rim location.<br />

6.1 Introduction<br />

The spectra <strong>of</strong> <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars <strong>of</strong> all masses show prominent strong and broad<br />

emission lines <strong>of</strong> both hydrogen and metals. These lines trace the complex circumstellar<br />

environment that characterizes this evolutionary phase, and are very likely powered<br />

by the associated accretion <strong>disks</strong>. The emission lines are used to infer the physical<br />

properties <strong>of</strong> the gas, and to constrain its geometry and dynamics. Their exact origin,<br />

however, is not known. The hydrogen lines, in particular, may originate either in the<br />

gas that accretes onto the star from the disk, as in magnetospheric accretion models<br />

(Hartmann et al., 1994), or in winds and jets, driven by the interaction <strong>of</strong> the accreting<br />

disk with a stellar (Shu et al., 1994) or disk (Casse et al., 2000) magnetic field. For<br />

Herbig Ae stars, it is additionally possible that they form in the inner disk (Tambovtseva<br />

et al., 1999).<br />

For all models, emission in the hydrogen lines is <strong>pre</strong>dicted to occur over very small<br />

spatial scales, a few AUs at most. To understand the physical processes that occur at<br />

these scales, one needs to combine very high spatial resolution with enough spectral res-<br />

olution to resolve the line pr<strong>of</strong>ile. AMBER, the three-beam near-IR recombiner <strong>of</strong> the<br />

VLTI (Petrov et al., 2003), simultaneously <strong>of</strong>fers high spatial and high spectral resolu-<br />

tion, with the sensitivity required to observe <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars. These capabilities<br />

were recently demonstrated by Malbet et al. (2006), who successfully resolved the lu-


6.1 Introduction 109<br />

λ Fλ (erg cm 2 s -1 λ Fλ (erg cm<br />

)<br />

2 s -1 )<br />

1e-07<br />

1e-07<br />

1e-08<br />

1e-08<br />

1e-09<br />

1e-09<br />

0.4 0.4 0.6 0.60.80.81<br />

1 1.5 1.52<br />

2 3 3 4 45<br />

56<br />

67<br />

7<br />

λ λ (μm)<br />

HD104237 ISAAC/AMBER spectrum<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

?64<br />

2.164<br />

?66<br />

2.166<br />

x0= 2.1??E+3<br />

λ (μm)<br />

?68<br />

2.168<br />

Visibility<br />

1.0<br />

0.5<br />

0.0<br />

2.140 2140 2.160 2160 2.180 2180<br />

Figure 6.1: Left: SED <strong>pre</strong>dicted by the “puffed-up” rim model used to normalise the K-band<br />

continuum visibility. The contributions <strong>of</strong> the A star (long-dashed line), the K3 star (dotted line)<br />

and the rim (short dashed line) are shown (see the text for the stellar parameters). The obser-<br />

vations are from 2MASS (filled circles), Hipparcos (filled squares), and Malfait et al. (1998,<br />

empty circles). An extinction <strong>of</strong> Av= 0.31 (van den Ancker et al., 1998) has been used to cor-<br />

rect the data for interstellar absorption. Center: Comparison <strong>of</strong> Brγ observed with AMBER in<br />

the photometric channels (solid line) and ISAAC (dotted line); the dashed line shows the ISAAC<br />

spectrum smoothed to the spectral resolution <strong>of</strong> AMBER. Right: Visibility <strong>of</strong> HD 104237 as a<br />

function <strong>of</strong> wavelength. The continuum has been normalised using the star+rim model, as<br />

described in the text. The vertical line shows the Brγ wavelength.<br />

minous Herbig Be system MWC 297 in both the continuum and the Brγ line. They<br />

showed that the line emission originates from an extended wind, while the continuum<br />

infrared excess traces a dusty accretion disk.<br />

In this Chapter we concentrate on a lower mass, less active system, the Herbig Ae<br />

system HD 104237. The central emission line star, <strong>of</strong> spectral type between A4V and<br />

A8, is surrounded by a circumstellar disk, which causes the infrared excess emission<br />

(Meeus et al., 2001) and drives a jet seen in Ly-α images (Grady et al., 2004). The<br />

optical spectrum shows a rather narrow Hα emission with a P-Cygni pr<strong>of</strong>ile (Feigelson<br />

et al., 2003). The disk is seen almost pole-on (i=18 ◦ +14<br />

−11 ; Grady et al., 2004), consistent<br />

with the low value <strong>of</strong> v sin i (12 km s −1 ; Donati et al., 1997). Donati et al. (1997)<br />

detected a stellar magnetic field <strong>of</strong> 50 G. Böhm et al. (2004) have revealed the <strong>pre</strong>sence<br />

<strong>of</strong> a very close companion <strong>of</strong> spectral type K3, orbiting with a period <strong>of</strong>∼ 20 days, and<br />

λ (μm)


110 AMBER/VLTI spectroscopy <strong>of</strong> HD 104237<br />

whose bolometric luminosity is 10 times fainter than the one <strong>of</strong> the central star. In the<br />

near infrared do<strong>main</strong>, spatially unresolved ISAAC <strong>observations</strong> (Garcia Lopez et al.,<br />

2006) show a strong Brγ emission line, with a peak flux 35% above the continuum.<br />

6.2 Observations and data reduction<br />

AMBER observed HD 104237 on 26 February 2004 on the UT2-UT3 baseline <strong>of</strong> the<br />

VLTI, which corresponds to a projected length <strong>of</strong> B=35 m. The instrument was set<br />

up to cover the [2.121, 2, 211]µm spectral range with a spectral resolution <strong>of</strong>R=1500,<br />

which resolved the pr<strong>of</strong>ile <strong>of</strong> the Brγ emission line at 2.165µm. The data consist <strong>of</strong><br />

10 exposures <strong>of</strong> 500 frames, with an integration time <strong>of</strong> 100ms for each frame. The<br />

integration time is a trade-<strong>of</strong>f between gathering enough flux in the photometric channel<br />

and <strong>pre</strong>venting excessive contrast loss due to fringe motion during the integration. At the<br />

time <strong>of</strong> these early <strong>observations</strong> the star magnitude <strong>of</strong> K=4.6 was close to the sensitivity<br />

limit <strong>of</strong> the medium spectral resolution mode, and as a result fringes are only visible in<br />

≈ 25% <strong>of</strong> the frames. In the photometric channels the Brγ line is under the noise level <strong>of</strong><br />

the individual frames, but it is well detected when averaging them and it contributes 35%<br />

<strong>of</strong> the total flux (Fig. 6.1, central panel). For comparison we overlay a higher spectral<br />

resolution spectrum (R=8900, taken with ISAAC approximately one year before the<br />

VLTI <strong>observations</strong>), as well as its smoothing to the AMBER spectral resolution. The<br />

two spectra are consistent with each other, within the combined typical line variability<br />

<strong>of</strong> Herbig Ae stars and calibration uncertainties.<br />

Data reduction followed standard AMBER procedures (Tatulli et al. 2006). To<br />

optimize the signal to noise ratio (SNR) <strong>of</strong> the visibility derived from these relatively<br />

poor-quality data, we <strong>pre</strong>-selected frames with individual SNR greater than 1.5, ensuring<br />

that fringes are <strong>pre</strong>sent in all short exposures that enter the visibility.<br />

For these <strong>observations</strong> the visibility amplitude could not be calibrated to an absolute<br />

scale using an astronomical calibrator (unresolved star), due to non-stationary vibrations<br />

in the optical train <strong>of</strong> the UT telescopes. With a 100ms integration time these vibrations


6.2 Observations and data reduction 111<br />

randomly degrade the fringe contrast <strong>of</strong> the individual frames by a large factor, and the<br />

statistics <strong>of</strong> that attenuation depends on uncontrolled factors (exact telescope and delay<br />

line pointing, recent environmental history, etc) that do not similarly affect the target<br />

star and its calibrator. Fortunately the contrast loss from vibrations is achromatic across<br />

our small relative bandpass, and as a con<strong>sequence</strong> the differential visibility, that is, the<br />

relative visibility between the spectral channels, is unaffected. Our dataset therefore<br />

allows us to investigate the visibility across the Brγ emission line compared to that in<br />

the adjacent continuum. The differential visibilities are accurate to approximately 5%.<br />

The excess near infrared continuum emission in Herbig Ae stars most likely origi-<br />

nates in the innermost regions <strong>of</strong> their dusty circumstellar <strong>disks</strong>, at the dust sublimation<br />

radius (Eisner et al., 2005; Isella et al., 2006). To approximately normalise the visi-<br />

bilities, we thus scaled the observed continuum value to the <strong>pre</strong>dictions <strong>of</strong> appropriate<br />

theoretical models <strong>of</strong> the disk inner rim emission (Isella and Natta, 2005). We computed<br />

the structure <strong>of</strong> the “puffed-up” inner rim as described in Isella et al. (2006). We used<br />

for the two stars the following parameters: Te f f= 8000K, L⋆=30L⊙ for the primary and<br />

Te f f = 4700K, L⋆=3L⊙ for the secondary star. We adopted a distance <strong>of</strong> D=115 pc,<br />

and a disk mass surface density <strong>of</strong>Σ(r)=2·10 3 · r −3/2 g/cm 2 (with r in AU).<br />

Fig. 6.1 (left) shows that the model using micron size astronomical silicates pro-<br />

duces a very good fit <strong>of</strong> the SED. The star fluxes contribute approximately 30% (respec-<br />

tively 20% and 10% for the primary and secondary star, which gives a binary flux ratio<br />

<strong>of</strong>∼ 0.5 in the K band) <strong>of</strong> the total 2.15µm flux, and the inner rim appears as a bright<br />

ring with radius Rrim= 0.45AU (3.8 mas at D=115 pc). The resulting model visibility<br />

on the 35m baseline is V = 0.38 (Fig 6.1, right). Note that the orbital period <strong>of</strong> the<br />

spectroscopic binary is 20 days. This leads to an average separation <strong>of</strong>δs=0.15 AU<br />

which corresponds to an angular separation <strong>of</strong> 1.2 mas. The central system (A star+K3<br />

companion) is therefore completely unresolved on the current baseline, and both stars<br />

are located inside the dust evaporation radius.


112 AMBER/VLTI spectroscopy <strong>of</strong> HD 104237<br />

6.3 Results and discussions<br />

Within the fairly small error bars, the visibility does not change across the Brγ emission<br />

line. This result is robust and puts strong constraints on the relative spatial extent <strong>of</strong> the<br />

line and continuum emission regions, demonstrating that they have very similar apparent<br />

sizes. We use this constraint to probe the processes responsible for the the Brγ emission<br />

in this star, and consider in turn the three <strong>main</strong> mechanisms usually invoked to inter<strong>pre</strong>t<br />

the hydrogen line emission in <strong>pre</strong>-<strong>main</strong>-<strong>sequence</strong> stars. We translate each mechanism<br />

to simple geometrical models <strong>of</strong> specific spatial extension, with the line strength fixed<br />

at the observational value, and evaluate the resulting visibility across the line 1 .<br />

Magnetospheric accretion: in such a model matter falls on the star along magnetic<br />

field lines, and the base <strong>of</strong> this accreting flow is (approximately) inside the corota-<br />

tion radius (Muzerolle et al., 2004). For HD 104237 we find Rcorot= 0.07AU, using<br />

v sin i = 12km/s (Donati et al., 1997) and an inclination <strong>of</strong> i = 18 ◦<br />

(Grady et al.,<br />

2004). Note that this value is, at our spatial resolution level, similar to the separation<br />

<strong>of</strong> the binary. Adopting Rstar and Rcorot as the inner and outer limits <strong>of</strong> the magne-<br />

tospheric accreting region, Fig. 6.2 (upper panel) demonstrates that Brγ emission is<br />

then confined much closer to the star than the dusty rim. The <strong>pre</strong>dicted visibility there-<br />

fore increases significantly in the line, contrary to the observational result. Explaining<br />

the observed visibility with magnetospheric accretion requires the corotation radius to<br />

approach the inner rim, which would need an unrealistically lower stellar rotational ve-<br />

1 Note that this analysis is valid as long as the continuum emission is resolved. In our case the contin-<br />

uum is calibrated by a model and not by an unresolved star, and therefore there might be a chance that the<br />

visibility <strong>of</strong> the continuum is close to 1. However, this peculiar scenario appears very unlikely since the<br />

puffed-up rim model has been shown to be well re<strong>pre</strong>sentative <strong>of</strong> the very close environment <strong>of</strong> Herbig Ae<br />

stars over a large luminosity range (Monnier et al., 2005). Furthermore, the hypothesis <strong>of</strong> a resolved con-<br />

tinuum emission for HD 104237 is strongly supported by its measured accretion rate <strong>of</strong> 10 −8 M⊙ (Grady<br />

et al., 2004). Indeed, Muzerolle et al. (2004) showed that for weak accretion rates ( ˙M


6.3 Results and discussions 113<br />

V<br />

V<br />

V<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

•<br />

•<br />

•<br />

2.14 2.15 2.16 2.17<br />

wavelength (μm)<br />

Figure 6.2: Comparison between the observed visibilities (empty square with error bars) and the<br />

<strong>pre</strong>dictions (solid curves) <strong>of</strong> the simple geometrical models for the Brγ emission (sketched in the<br />

same panels). The observed visibilities are scaled to match the continuum value <strong>pre</strong>dicted by the<br />

“puffed-up” inner rim model (Isella and Natta, 2005) as described in Sec. 6.2. The continuum<br />

emission arises both from the stellar photosphere (≈20%) and from the dusty disk inner rim,<br />

located at the dust evaporation distance Rrim=0.45 AU and which appears as the bright gray<br />

scale ring. The Brγ emission regions are shown as grid surfaces. The three panels illustrate the<br />

different models discussed in Sec. 6.3: the upper panel re<strong>pre</strong>sents the magnetospheric accretion<br />

model in which the Brγ emission originates very close to the star, inside the corotational radius<br />

Rcorot=0.07 AU; in the middle panel the Brγ emission originates between Rcorot and the rim<br />

radius Rrim=0.45 AU, re<strong>pre</strong>senting the gas within the disk model; the bottom panel shows the<br />

outflowing wind model, in which the emission is confined close to the inner rim, between∼ 0.2<br />

AU and∼0.5 AU.


114 AMBER/VLTI spectroscopy <strong>of</strong> HD 104237<br />

locity (v


6.4 Conclusions 115<br />

note however that the launching region <strong>of</strong> an X-wind, driven by the stellar magnetic field<br />

(Shu et al., 1994), is close to the corotation radius and too small to be consistent with the<br />

<strong>pre</strong>sent data. Making the X-wind acceptable needs one to relax the assumption that most<br />

<strong>of</strong> the line emission originates near the launching point, and instead have the brightest<br />

Brγ region a factor 5−8 further away from the star. The disk-wind scenario in contrast<br />

has its launching point a few tenths <strong>of</strong> an AU from the star and needs no adjusting.<br />

This <strong>pre</strong>ference for a disk-wind is consistent with the size <strong>of</strong> the jet launching regions<br />

inferred from the rotation <strong>of</strong> TTauri jets (C<strong>of</strong>fey et al., 2004).<br />

Our analysis assumes that the continuum originates in a dusty “puffed-up” ring. That<br />

model has been shown to match Herbig Ae and T Tauri stars over a large luminosity<br />

range (Monnier et al., 2005), but it obviously needs to be verified for the specific case <strong>of</strong><br />

HD 104237, by obtaining calibrated interferometric <strong>observations</strong> over a few baselines.<br />

6.4 Conclusions<br />

We have <strong>pre</strong>sented interferometric <strong>observations</strong> <strong>of</strong> the Herbig Ae star HD 104237, ob-<br />

tained with the AMBER/VLTI instrument withR=1500 high spectral resolution. The<br />

visibility is identical in the Brγ line and in the continuum, even though the line re<strong>pre</strong>-<br />

sents 35% <strong>of</strong> the total 2.165µm flux. This implies that the line and continuum emission<br />

regions have the same apparent size.<br />

Scaling the continuum visibility with a “puffed-up” inner rim model, and using sim-<br />

ple models to describe the Brγ emission, we have shown that the line emission is un-<br />

likely to originate in either magnetospheric accreting columns <strong>of</strong> gas or in the gaseous<br />

disk. It is much more likely to come from a compact outflowing disk wind launched<br />

in the vicinity <strong>of</strong> the rim, about 0.5 AU from the star. This does not <strong>pre</strong>clude accretion<br />

from occurring along the stellar magnetic field detected by Donati et al. (1997), and<br />

accreting matter might even dominate the optical hydrogen line emission, but our ob-<br />

servations show that the bulk <strong>of</strong> the Brγ emission in HD 104237 is unlikely to originate<br />

in magnetospheric accreting matter.


116 AMBER/VLTI spectroscopy <strong>of</strong> HD 104237<br />

Our results show that AMBER/VLTI is a powerful diagnostic <strong>of</strong> the origin <strong>of</strong> the<br />

line emission in young stellar objects. Observations <strong>of</strong> a consistent sample <strong>of</strong> objects<br />

will strongly constrain the wind launching mechanism.<br />

Acknowledgements<br />

This work has been partly supported by the MIUR COFIN grant 2003/027003-001 and<br />

025227/2004 to the INAF-Osservatorio Astr<strong>of</strong>isico di Arcetri. This project has benefited<br />

from funding from the French Centre National de la Recherche Scientifique (CNRS)<br />

through the Institut National des Sciences de l’Univers (INSU) and its Programmes Na-<br />

tionaux (ASHRA, PNPS). The authors from the French laboratories would like to thank<br />

the successive directors <strong>of</strong> the INSU/CNRS. C. Gil work was supported in part by the<br />

Fundação para a Ciência e a Tecnologia through project POCTI/CTE-AST/55691/2004<br />

from POCTI,with funds from the European program FEDER.<br />

This work is based on <strong>observations</strong> made with the European Southern Observatory<br />

telescopes. We made use <strong>of</strong> the ASPRO observation <strong>pre</strong>paration tool from the Jean-<br />

Marie Mariotti Center in France, the SIMBAD database at CDS, Strasbourg (France)<br />

and the Smithsonian/NASA Astrophysics Data System (ADS). The data reduction s<strong>of</strong>t-<br />

wareamdlib is freely available on the AMBER site http://amber.obs.ujf-grenoble.fr.<br />

It has been linked to the free s<strong>of</strong>tware Yorick 2 to provide the user friendly interface<br />

ammyorick.<br />

2 ftp://ftp-icf.llnl.gov/pub/Yorick


Part IV<br />

The structure <strong>of</strong> the outer disk


CHAPTER 7<br />

The keplerian disk <strong>of</strong> HD 163296<br />

This chapter describes the analysis <strong>of</strong> recent millimeter interferometric <strong>observations</strong> <strong>of</strong><br />

the HAe star HD 163296. While the data reduction and the implementation <strong>of</strong> disk<br />

models required to analyse the data are complete, part <strong>of</strong> the analisys <strong>of</strong> the results is<br />

still in progress. The results described in the following will appear in a forthcoming<br />

paper (Isella A., Testi L., Natta A., Neri R., Wilner D, Qi C., 2007).<br />

7.1 Introduction<br />

Millimeter and submillimeter interferometers are providing an increasingly detailed de-<br />

scription <strong>of</strong> <strong>disks</strong> around <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars <strong>of</strong> solar (T Tauri stars; TTS) and in-<br />

termediate mass (Herbig Ae; HAe). Both the dust continuum emission and the emission<br />

in molecular lines are observed and spatially resolved in a number <strong>of</strong> <strong>disks</strong>, yielding<br />

information on the disk density and temperature, the dust properties and the gas chem-<br />

istry and dynamics (e.g., Natta et al., 2006; Dutrey et al., 2006 and references therein).<br />

The number <strong>of</strong> well-studied <strong>disks</strong> is still very small, practically restricted to the most<br />

massive and luminous ones; still, it is clear that <strong>disks</strong> differ from one another. Recently,


120 The keplerian disk <strong>of</strong> HD 163296<br />

it has been reported evidence <strong>of</strong> spiral structures in AB Aur, a 2-3 Myr old intermedi-<br />

ate mass star, and <strong>of</strong> deviations from keplerian rotation (Piétu et al., 2005; Corder et<br />

al., 2005); the classical TTS LkCa15 has a large inner hole <strong>of</strong> size∼ 50 AU depleted<br />

<strong>of</strong> dust, while the HAe star MWC 480 has a smooth disk with an optically thick (at<br />

millimeter wavelengths) inner region <strong>of</strong> radius∼35 AU (Piétu et al., 2006). Both spiral<br />

structures and large gaps are evidence <strong>of</strong> dynamical perturbations, possibly due to the<br />

<strong>pre</strong>sence <strong>of</strong> large planets or/and to the effects <strong>of</strong> the disk self-gravity (see the case <strong>of</strong><br />

IRAS 16293-2422B, Rodriguez et al., 2005) which however may favour the planetary<br />

formation (see the review <strong>of</strong> Durisen et al., 2006).<br />

The existence <strong>of</strong> unperturbed and distorted <strong>disks</strong> among <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars<br />

suggests that the planet formation is actively occurring during this evolutionary stage,<br />

leaving detectable marks on their parent <strong>disks</strong>. It is therefore important to study in detail<br />

as many <strong>disks</strong> as possible, in order to characterize their basic properties and to detect<br />

deviations from the simple patterns <strong>of</strong> homogeneous <strong>disks</strong> in keplerian rotation.<br />

We report in this Chapter a detailed study <strong>of</strong> the disk associated to the HAe star<br />

HD 163296, using <strong>observations</strong> in the continuum and CO lines obtained with three<br />

different interferometers, namely the VLA at 7 mm, the Plateau the Bure Interferometer<br />

at 1.3 and 2.6 mm and SMA at 0.85 mm. HD 163296 is a star <strong>of</strong> spectral type A1,<br />

mass <strong>of</strong> roughly 2.3 M⊙, distance 122 pc. Early OVRO <strong>observations</strong> (Mannings and<br />

Sargent, 1997) have shown the <strong>pre</strong>sence <strong>of</strong> a disk with a minimum mass∼ 0.03 M⊙ and<br />

evidence <strong>of</strong> rotation from the CO lines. The disk is seen in scattered light by Grady<br />

et al. (2000), with radius <strong>of</strong>∼ 500 AU; it has an associated jet seen in Ly-α with HST,<br />

extending on both sides <strong>of</strong> the disk orthogonally to it. Natta et al. (2004) found evidence<br />

<strong>of</strong> evolved dust in the outer disk <strong>of</strong> HD 163296 by comparing the VLA 7 mm flux to the<br />

OVRO <strong>observations</strong>. The results we <strong>pre</strong>sent here have much higher spatial resolution<br />

and wavelength coverage than what <strong>pre</strong>viously known. They will allow us to measure<br />

accurately the dynamics <strong>of</strong> the disk as well as the disk and dust properties and to test<br />

the capability <strong>of</strong> disk models to account for the <strong>observations</strong>. As we will show, they<br />

suggest that in HD 163296 it is not the inner disk which is dust-depleted but rather the


7.2 Observations and data reduction 121<br />

outer disk, outside about 200 AU.<br />

The structure <strong>of</strong> the Chapter is as follows. Sec. 7.2 will describe the <strong>observations</strong>.<br />

The results will be <strong>pre</strong>sented in Sec. 7.3, where we will derive some <strong>of</strong> the disk param-<br />

eters. A more detailed analysis, using self-consistent disk models <strong>of</strong> the dust and CO<br />

line emission will be <strong>pre</strong>sented in Sec. 7.4; Sec. 7.5 contains the results, which will be<br />

further discussed in Sec. 7.6. Summary and conclusions follow in Sec. 7.7.<br />

7.2 Observations and data reduction<br />

HD 163296 was observed with the Plateau de Bure Interferometer (PBI), the NRAO<br />

Vary Large Array (VLA) and the Sub-Millimeter Array (SMA).<br />

7.2.1 PBI <strong>observations</strong><br />

The IRAM/PBI <strong>observations</strong> were carried over the 2003/2004 winter season. The six 15<br />

m dishes were used in the most extended configuration providing a baseline coverage<br />

between 25 and 400 m. The corresponding angular resolutions are reported in Tab. 7.1.<br />

The receivers were tuned to observe the 12 CO J=2–1 line and the nearby continuum at<br />

1.3 mm, while at 2.7 mm the 13 CO J=1–0, and C 18 O J=1–0 lines were observed along<br />

with the continuum. Bandpass, complex gain and amplitude calibrations were ensured<br />

by <strong>observations</strong> <strong>of</strong> standard IRAM calibrators. The phase stability was excellent dur-<br />

ing our <strong>observations</strong> and only a minimal amount <strong>of</strong> editing <strong>of</strong> the data was necessary.<br />

All calibrations were performed using the standard CLIC suite <strong>of</strong> programmes within<br />

the GILDAS s<strong>of</strong>tware package. The calibrated uv data were then exported for the sub-<br />

sequent analysis. The accuracy <strong>of</strong> the flux density scale calibration is expected to be<br />

within 20% at these wavelengths.


122 The keplerian disk <strong>of</strong> HD 163296<br />

7.2.2 SMA <strong>observations</strong><br />

The SMA 1 <strong>observations</strong> <strong>of</strong> HD 163296 was made on August 23rd, 2005 using the<br />

Compact Configuration <strong>of</strong> seven <strong>of</strong> the 6 meter diameter antennas, which provided 21<br />

independent baselines ranging in length from 8 to 80 meters. The SMA digital correlator<br />

was configured with a narrow band <strong>of</strong> 512 channels over 104 MHz, which provided 0.2<br />

MHz frequency resolution, or 0.18 km s −1 velocity resolution at 0.87 mm (345 GHz),<br />

and the full correlator bandwith was 2 GHz. The weather was good withτ(1.3 mm)<br />

around 0.06 and the double-sideband (DSB) system temperature were between 200 and<br />

500 K. The source HD 163296 was observed from HA -3 to 4.5. Calibration <strong>of</strong> the<br />

visibility phases and amplitudes was achieved with <strong>observations</strong> <strong>of</strong> the quasar 1921-<br />

293, typically at intervals <strong>of</strong> 25 minutes. Observations <strong>of</strong> Uranus provided the absolute<br />

scale for the flux density calibration and the uncertainties in the flux scale are estimated<br />

to be 20%. The data were calibrated using the MIR s<strong>of</strong>tware package 2 .<br />

7.2.3 VLA <strong>observations</strong><br />

HD 163296 was observed at the VLA as part <strong>of</strong> a larger survey for 7 mm disk emission<br />

around Herbig Ae stars (see Natta et al., 2004). Data were obtained with the array in<br />

the C and D configurations, in several occasions from Dec 2001 through May 2003.<br />

Accurate pointing was ensured by hourly pointing sessions at 3.6cm on a bright extra-<br />

galactic object. The array <strong>of</strong>fered baselines from the shadowing limit through∼ 3.4 km,<br />

although some data was obtained in a hybrid DnA configuration, all the data from the<br />

longer baselines had to be rejected due to large phase fluctuations that could not be<br />

corrected. The resulting uv-plane coverage <strong>of</strong>fered an angular resolution <strong>of</strong> 0.5 ′′ and a<br />

maximum recoverable size <strong>of</strong> the order <strong>of</strong>∼40 ′′ . All the data was edited and calibrated<br />

using standard recipes in AIPS. The short term complex gain variations were corrected<br />

1 The Submillimeter Array is a joint project between the Smithsonian Astrophysical Observatory and<br />

the Academia Sinica Institute <strong>of</strong> Astronomy and Astrophysics and is funded by the Smithsonian Institu-<br />

tion and the Academia Sinica<br />

2 http://cfa-www.harvard.edu/cqi/mircook.html


7.3 Observational results 123<br />

λ (mm) beam FWHM PA flux (mJy)<br />

0.87 3 ′′ .14×2 ′′ .21 161 o 1910±20<br />

1.3 1 ′′ .95×0 ′′ .42 7 o 705±12<br />

2.8 3 ′′ .3×0 ′′ .94 8 o 77.0±2.2<br />

7 1 ′′ .71×0 ′′ .81 172 o 4.5±0.5<br />

Table 7.1: Integrated flux and beam size <strong>of</strong> the <strong>observations</strong> <strong>of</strong> HD 163296 performed with the<br />

SMA (at 0.87 mm), PBI (at 1.3mm and 2.7 mm) and VLA (at 7 mm).<br />

using frequent (few minutes cycle) <strong>observations</strong> <strong>of</strong> the quasar 1820−254, while the flux<br />

scale was set observing the VLA calibrator 1331+305. This procedure is expected to be<br />

accurate within∼ 15% at 7 mm.<br />

7.3 Observational results<br />

The continuum maps obtained at 0.87 mm, 1.3 mm, 2.8 mm and 7 mm are shown in<br />

Fig. 7.1, the emission <strong>of</strong> the 12 CO J=2–1 transition is shown in Fig. 7.2, the emission<br />

<strong>of</strong> the 12 CO J=3–2 transition is shown in Fig. 7.3 and the emission <strong>of</strong> the 13 CO J=2–1<br />

transition in Fig. 7.4. At the wavelength corresponding to the C 18 O J=1–0 transition we<br />

did not detected any emission; the continuum integrated fluxes are reported in Tab. 7.1.<br />

With simple physical assumptions, these <strong>observations</strong> allow to determine some fun-<br />

damental parameters <strong>of</strong> the star+disk system as the stellar and the disk masses, the<br />

contribution to the observed fluxes <strong>of</strong> free-free gas emission, the wavelength depen-<br />

dence <strong>of</strong> the observed flux at millimeter wavelengths and the related dust grain opacity.<br />

7.3.1 Disk morphology and apparent size<br />

The 12 ′′ ×12 ′′ continuum maps <strong>of</strong> HD 163296 are shown in Fig. 7.1. At all wavelengths,<br />

the peak is coincident with the position <strong>of</strong> the optical star as measured from Hipparcos<br />

and, given the respective synthesized beams FWHM (see Tab. 7.1), the emission is


124 The keplerian disk <strong>of</strong> HD 163296<br />

Figure 7.1: Continuum maps <strong>of</strong> HD 163296 at 0.87 mm (SMA <strong>observations</strong>), 1.3 mm and 2.8<br />

mm (PBI) and 7 mm (VLA), starting form the left. In the 0.87 mm map, in order to highlight the<br />

extended morphology <strong>of</strong> the disk, the first contour level corresponds to 30 mJy (3σ), the second<br />

to 6σ and the inner contour levels are spaced by 10σ. At longer wavelength the contour levels<br />

are all spaced by 3σ, corresponding to 16 mJy at 1.3 mm, 3.3 mJy at 2.8 mm and 0.75 mJy at<br />

7 mm. The small boxes show the relative synthesized beams. The integrated fluxes, the beam<br />

dimensions and orientations are summarized in Tab.7.1.<br />

resolved and elongated approximately in the east-west direction. Approximating the<br />

source with a circularly symmetric geometrically thin disk and taking into account the<br />

beam shape, the observed aspect ratio <strong>of</strong> the level contours implies an inclination <strong>of</strong> the<br />

disk plane from the line <strong>of</strong> sight <strong>of</strong> about 45 o ± 20 o and a position angle <strong>of</strong> 120 o ± 30 o ,<br />

roughly in agreement with the values obtained by Mannings and Sargent (1997) using<br />

marginally resolved OVRO <strong>observations</strong>.<br />

In the continuum map, the maximum elongations <strong>of</strong> the 3σ contours along the more<br />

resolved direction varies between 4 ′′ at 0.87 mm to 0.9 ′′ at 7 mm. The fainter contour<br />

levels are not centrally symmetric, showing a flux excess in the east half <strong>of</strong> the image,<br />

better visible in the 0.87 and 1.3 mm maps. Both the disk size and the morphology will<br />

be discussed in more detail in Sec. 7.6.1.<br />

7.3.2 Disk kinematic<br />

The emission maps <strong>of</strong> the 12 CO J=2-1 (at 1.3 mm) and J=3-2 (at 0.87 mm), and the<br />

13 CO J=1-0 (at 2.8 mm) transitions are shown in Fig. 7.2, 7.3 and 7.4 respectively; the


7.3 Observational results 125<br />

Figure 7.2: Velocity channel maps <strong>of</strong> the 12 CO J=2–1 line emission around HD 163296. The<br />

LSR velocity (km s −1 ) is indicated in the upper left corner <strong>of</strong> each panel. The angular resolution<br />

(synthesized beam), indicated in the small boxes, is 1.92 ′′ × 0.42 ′′ at PA 7 o ; the contour spacing<br />

is 0.23 Jy/beam corresponding to 3σ. The last two panels show the integrated intensity (contour<br />

levels spaced by 0.4 Jy/beam km s −1 ) and the velocity field (contour levels from 3 km s −1 to 9<br />

km s −1 spaced by 0.5 km s −1 ) respectively.


126 The keplerian disk <strong>of</strong> HD 163296<br />

Figure 7.3: Velocity channel maps <strong>of</strong> the 12 CO J=3-2 line emission around HD 163296. The<br />

LSR velocity (km s −1 ) is indicated in the upper left corner <strong>of</strong> each panel. The angular resolution<br />

(synthesized beam), indicated in the small boxes, is 3.1 ′′ ×2.03 ′′ at PA 161 o ; the contour spacing<br />

is 1.5 Jy/beam corresponding to 3σ. The last two panels show the integrated intensity (contour<br />

levels spaced by 5 Jy/beam km s −1 ) and the velocity field (contour levels from 3 km s −1 to 9 km<br />

s −1 spaced by 0.5 km s −1 ) respectively.


7.3 Observational results 127<br />

Figure 7.4: Velocity channel maps <strong>of</strong> the 13 CO J=1–0 line emission around HD 163296. The<br />

LSR velocity (km s −1 ) is indicated in the upper left corner <strong>of</strong> each panel. The angular resolution<br />

(synthesized beam), indicated in the small boxes, is 3.28 ′′ × 0.94 ′′ at PA 7 o ; the contour spacing<br />

is 0.09 Jy/beam corresponding to 3σ. The last two panels show the integrated intensity (contour<br />

levels spaced by 0.12 Jy/beam km s −1 ) and the velocity field (contour levels from 3 km s −1 to 9<br />

km s −1 spaced by 0.5 km s −1 ) respectively.


128 The keplerian disk <strong>of</strong> HD 163296<br />

velocity-position diagrams along the line <strong>of</strong> nodes are shown in Fig. 7.5. For all the<br />

molecular transitions, the emission is resolved, showing a velocity pattern typical <strong>of</strong> an<br />

inclined rotating disk characterized by a position angle <strong>of</strong> about 130 o (a more <strong>pre</strong>cise<br />

estimate <strong>of</strong> the position angle will be <strong>pre</strong>sented in Sec. 7.5). The velocity-position<br />

diagram calculated along this direction (see Fig. 7.5) shows a well defined “butterfly<br />

shape” typical <strong>of</strong> keplerian rotation. A first estimate <strong>of</strong> the mass <strong>of</strong> the central object<br />

and <strong>of</strong> the dimension <strong>of</strong> the disk can be obtained by comparing the observed velocities<br />

with the keplerian law:<br />

vǫ= C·ǫ −1/2 , (7.1)<br />

whereǫ is the angular distance from the central star and vǫ is the component <strong>of</strong> the disk<br />

rotational velocity along the line <strong>of</strong> sight. If the stellar mass M⋆ is in solar unit, the<br />

stellar distance d in parsec andθis the disk inclination (θ=0 means pole-on disk),<br />

the constant C≃30 √ M⋆/d sinθ is the component along the line <strong>of</strong> sight <strong>of</strong> the disk<br />

rotational velocity (in km/sec) atǫ= 1 ′′ . As shown in Fig. 7.5 the envelopes <strong>of</strong> both<br />

the 12 CO and 13 CO emissions are in agreement with C≃ 2.7±0.4 km/sec. This value<br />

corresponds to a stellar mass <strong>of</strong> 2.0±0.5 M⊙ for an inclination <strong>of</strong> 45 o (obtained in the<br />

<strong>pre</strong>vious section) and d=122 pc; the systemic velocity is 5.8 Km/sec. Note finally that<br />

the disk rotation is clearly observed at least up to a distance <strong>of</strong> about 4 ′′ , corresponding<br />

to a minimum disk outer radius <strong>of</strong> about 500 AU. We will discuss in more detail the<br />

determination <strong>of</strong> the stellar mass and the disk outer radius in Sec. 7.5, using a detailed<br />

model for the continuum and CO molecular emission.<br />

7.3.3 Free-free contribution and spectral index<br />

The continuum spatially integrated fluxes <strong>of</strong> HD 163296 and the corresponding spectral<br />

energy distribution are given in Tab. 7.1 and shown in Fig. 7.6, respectively. In addition<br />

to our measurements (full squares), <strong>observations</strong> at 7, 13, 36 and 61 mm are from Natta<br />

et al. (2004 and references therein, empty circles) while <strong>observations</strong> at 0.75, 0.8, 0.85,<br />

1.1 and 1.3 mm are from Mannings (1994, empty squares).


7.3 Observational results 129<br />

Figure 7.5: Velocity-position plots along the plane <strong>of</strong> the disk for the C 12 O J=2-1 (upper panel),<br />

the C 12 O J=3-2 (middle panel) and the C 13 O J=1-0 (lower panel) transitions. The angular <strong>of</strong>fset<br />

is measured with respect to the phase center <strong>of</strong> the <strong>observations</strong> corresponding to the position <strong>of</strong><br />

the central star. The contour levels are spaced by 2σ corresponding to 0.14 Jy/beam, 1 Jy/beam<br />

and 0.06 Jy/beam respectively. The cross in the lower left <strong>of</strong> each panel gives the angular and<br />

spectral resolution <strong>of</strong> the corresponding map. The thick solid lines marks the border where<br />

emission is expected for a Keplerian disk inclined by 45 o and rotating around a 2.0 M⊙ point<br />

source; the external and internal dashed lines correspond to stellar masses <strong>of</strong> 2.5 M⊙ and 1.5 M⊙<br />

respectively. The horizontal and vertical straight dashed lines mark the systemic velocity (5.8<br />

km/sec) and the position <strong>of</strong> the continuum peak.


130 The keplerian disk <strong>of</strong> HD 163296<br />

F λ (mJy)<br />

10000<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

Free-free emission<br />

1 10 100<br />

λ (mm)<br />

Figure 7.6: Spectral energy distribution at mm-cm wavelengths <strong>of</strong> HD 163296. Our new ob-<br />

servations are shown with full squares, data from Mannings (1994) with empty squares and data<br />

from Natta et al. (2004 and references therein) with empty circles. Observed fluxes at 36 mm<br />

and 61 mm have been used to calculate the free-free contribution (dashed line) to the observed<br />

flux. At 7 mm, the free-free subtracted flux are shown with crosses, while at shorter wave-<br />

length the the free-free contribution is negligible. The resulting millimetric spectral indexαmm<br />

(Fλ∝λ −αmm , see the solid line) is 3.0±0.1.<br />

Assuming that the observed flux at 36 and 61 mm is dominated by free-free emission<br />

from a wind, the free-free contribution at 7 mm corresponds to 1.2 mJy, while it is<br />

negligible at shorter wavelengths.<br />

After the subtraction <strong>of</strong> the free-free component, the millimeter spectral indexαmm<br />

(Fλ∝λ −αmm ) calculated between 1 mm and 7 mm is 3.0±0.1, slightly higher that the<br />

valueαmm= 2.6±0.2 obtained by Natta et al. (2004) using VLA and OVRO fluxes only.<br />

7.3.4 Disk mass<br />

Assuming that the dust emission is optically thin at millimeter wavelength, the measured<br />

flux can be used to estimate the product <strong>of</strong> the disk total mass M times the dust opacity


7.4 Disk Models 131<br />

kν, through the relation:<br />

M· kν=<br />

2 Fνd<br />

, (7.2)<br />

Bν(T)<br />

where d is the source distance and T is the typical temperature <strong>of</strong> the emitting dust. At<br />

millimeter wavelength, the dust opacity can by parametrized by a power low<br />

kν= 0.01·(ν/229GHz) β cm 2 g −1 , (7.3)<br />

where the normalization at 229 GHz (corresponding toλ=1.3 mm) assumes a dust/gas<br />

mass ratio <strong>of</strong> 0.01 (Beckwith et al., 1990, 1991). Taking a characteristic gas temperature<br />

for the outer disk <strong>of</strong> an early A star <strong>of</strong> 30K (Natta et al. 2000), the circumstellar material<br />

has mass <strong>of</strong> 0.12 M⊙. This rough mass determination, that does not take into account the<br />

<strong>pre</strong>sence <strong>of</strong> optically thick emission from the inner part <strong>of</strong> the disk, nor <strong>of</strong> temperature<br />

gradients, will be discussed in more detail in Sec. 7.5.2.<br />

With the assumption <strong>of</strong> optically thin emission, the slope <strong>of</strong> the dust opacity isβ=<br />

α−2∼1, whereαis the spectral index obtained in the <strong>pre</strong>vious section.<br />

7.4 Disk Models<br />

The disk parameters derived in Sec. 7.3 under a number <strong>of</strong> very crude assumptions<br />

can only provide order-<strong>of</strong>-magnitude estimates. A more quantitative analysis requires<br />

to compare the <strong>observations</strong> to model <strong>pre</strong>dictions. We chose to perform the compari-<br />

son using the observed visibilities (rather than reconstructed images), as discussed in<br />

Sec. 2.4<br />

This methodology requires to decide “a priori” which family <strong>of</strong> models is likely to<br />

describe the observed object. In view <strong>of</strong> the results described in Sec. 7.3, we model<br />

the millimeter emission <strong>of</strong> HD 163296 (continuum and CO lines) as coming from a cir-<br />

cumstellar disk. We assume that the disk is heated by the stellar radiation only, and that<br />

any viscous contribution can be neglected. This is very likely a good approximation,<br />

given the relatively low accretion rate measured in HD 163296 (∼ 10 −7 M⊙/yr; Garcia


132 The keplerian disk <strong>of</strong> HD 163296<br />

Lopez et al., 2006) and that the millimeter emission is dominated by the emission <strong>of</strong><br />

the outer disk, where the stellar heating is always dominant. The disk is in hydrostatic<br />

equilibrium and flared; to start with, we consider that dust and gas are fully mixed, and<br />

that no decoupling occurs. The mid plane temperature and disk scale height are com-<br />

puted using the 2-layer approximation <strong>of</strong> Chiang and Goldreich (1997), as developed<br />

by Dullemond et al. (2001). Similar models have been used in Testi et al. (2003) and<br />

Natta et al. (2004), to analyze the (sub)millimeter emission <strong>of</strong> a number <strong>of</strong> Herbig Ae<br />

stars. We refer to these papers for a more detailed description.<br />

Once the stellar properties are known, the disk structure is completely characterized<br />

by the following parameters: the disk mass (Md), the disk inner and outer radii (Rin and<br />

Rd), the dependence <strong>of</strong> the surface density on radius (Σ∝r −p ) and the properties <strong>of</strong><br />

dust on the disk surface and mid plane. Note that some <strong>of</strong> these parameters are not well<br />

constrained by the emission at millimeter wavelengths, which depends very little, e.g.,<br />

on Rin and the surface dust properties. In addition, the observed emission depends on<br />

the orientation <strong>of</strong> the disk with respect to the observer, which is characterized by the<br />

inclinationι <strong>of</strong> the disk with respect to the line <strong>of</strong> sight (ι=0 for face-on <strong>disks</strong>) and the<br />

position angle PA.<br />

7.4.1 Continuum emission<br />

The continuum emission at millimeter and submillimeter wavelengths is computed by<br />

ray integration as in Dullemond et al. (2001). We describe the midplane dust opacity<br />

at long wavelengths as a power-law <strong>of</strong> indexβ, withβafree parameter, as in Eq. 7.3.<br />

The inner disk radius is the dust sublimation radius, as in the rim models <strong>of</strong> Isella et al.<br />

(2006) for large (∼ 1µm) grains, and it is equal to 0.45 AU; the dust on the disk surface<br />

is as in Natta et al. (2004). Neither <strong>of</strong> these two quantities is relevant for the following<br />

analysis.


7.4 Disk Models 133<br />

7.4.2 CO emission<br />

The observed CO emission originates in the outer surface layers <strong>of</strong> the disk, at heights<br />

that depend on the optical depth <strong>of</strong> the specific transition. Once the disk structure is<br />

specified, as described above, one needs to compute, at each radius, the gas temperature<br />

pr<strong>of</strong>ile in the vertical direction. This is a complex problem, whose results depend on<br />

a number <strong>of</strong> not well known properties, among them the X-ray field and the role <strong>of</strong><br />

very small grains in heating the gas (e.g., Dullemond et al. 2006). Therefore, we use a<br />

parametric description assuming that for each CO transition the excitation temperature<br />

is the same at all z and can be described as a power-law <strong>of</strong> r in the form:<br />

Tline= Tline(r0)(r/r0) −q . (7.4)<br />

The assumption <strong>of</strong> a constant excitation temperature along the vertical direction is cor-<br />

rect for very optically thick lines, and/or if the velocity gradient along the line <strong>of</strong> sight<br />

is very large. In both cases, the contribution at each wavelength from the line <strong>of</strong> sight<br />

under consideration comes from a small region only, where the optical depth in the line<br />

is <strong>of</strong> order unity. For any CO line, the height to which Tline(r) refers is therefore differ-<br />

ent, depending on the density and velocity structure, as well as on the disk inclination<br />

angle. The values <strong>of</strong> Tline(r0) and q for each CO transition are free parameters. A similar<br />

procedure has been used by Dutrey et al. (1994) and Dartois et al. (2003) in their study<br />

<strong>of</strong> the CO emission <strong>of</strong> T Tauri <strong>disks</strong>.<br />

In the analysis <strong>of</strong> CO <strong>observations</strong> we have assumed 12 CO/H=7.0e-5 and 13 CO/H<br />

= 1.0e-6, which correspond to an isotopomer ratio 12 CO/ 13 CO=70 (Beckwith and<br />

Sargent, 1990; Dutrey et al., 1996; and reference therein)<br />

As described in detail in the Appendix, we have developed a code which computes,<br />

for each CO transition, the line intensity and pr<strong>of</strong>ile as function <strong>of</strong> the disk parameters,<br />

namely the inclination and PA, the surface density pr<strong>of</strong>ile, the disk outer radius. In<br />

addition, we assume that the disk is in keplerian rotation around the central star and<br />

vary the stellar mass independently for each line.


134 The keplerian disk <strong>of</strong> HD 163296<br />

Parameters Continuum 12 CO J=2-1 13 CO J=1-0<br />

M⋆ (M⊙) 2.6 a 2.6 +0.3<br />

−0.5 2.6±0.6<br />

PA 120 o ± 20 o 128 o ± 5 o 130 o ± 8 o<br />

Incl 40 o ± 12 o 45 o ± 5 o 50 o ± 8 o<br />

Rout (AU) 200±15 550±50 500±80<br />

p 0.81±0.01 0.6 +0.3<br />

−0.1 1.0±0.5<br />

Σ0 (10AU) (g/cm 2 ) 46±4 90±70 4 +12<br />

−3<br />

β 1.0±0.1 1.0 a 1.0 a<br />

χ 2 r 2.6 1.17 1.15<br />

Table 7.2: Parameter <strong>of</strong> the disk structure relative to the best fit models for the continuum and the<br />

CO molecular emissions as described in Sec. 7.5. For each parameters uncertainties are given at<br />

7.5 Results<br />

7.5.1 Method <strong>of</strong> analysis<br />

1σ level. a Fixed parameter.<br />

The <strong>observations</strong> <strong>of</strong> HD 163296 have been analyzed by comparing the observed uv-<br />

tables with synthetic uv-tables obtained from the models for the continuum and CO<br />

emission described in Sec. 7.4. At each point in the synthetic uv-plane we have associ-<br />

ated the corresponding weight in the real uv-plane. The analysis is thus independent <strong>of</strong><br />

image reconstruction and cleaning procedures, and takes into account the true sensitivity<br />

<strong>of</strong> the three different interferometers.<br />

We have in total seven sets <strong>of</strong> data corresponding to four continuum wavelengths<br />

and three CO lines. For each set aχ 2 minimization has been performed exploring a<br />

wide region <strong>of</strong> the space <strong>of</strong> model parameters. In all cases except the 12 CO J=3-2 line,<br />

for which the analysis is still in progress, we find a single minimum <strong>of</strong>χ 2 . For each


7.5 Results 135<br />

Figure 7.7: Maps <strong>of</strong> the residuals relative to the best fit model for the continuum emission (see<br />

Fig. 7.1 for the <strong>observations</strong>). The contour level are spaced by 3σ corresponding to 30 mJy at<br />

0.87mm, 16 mJy at 1.3 mm, 3.3 mJy at 2.8 mm and 0.75mJy at 7 mm. The small boxes show<br />

the relative synthesized beams.<br />

parameter, the 1σ uncertainties are estimated asχ 2 1σ =χ2 m+ √ 2n, where n is the number<br />

<strong>of</strong> degrees <strong>of</strong> freedom andχ 2 m is theχ2 value relative to the best fit model.<br />

7.5.2 Continuum emission<br />

The continuum data at the four observed wavelengths (0.85, 1.3, 2.7, 7 mm) have been<br />

analyzed together, to estimate the best disk parameters and their uncertainty. We have<br />

varied the two parameters which define the position <strong>of</strong> the disk on the plane <strong>of</strong> the sky,<br />

namely the inclination and position angle, as well as the three physical parameters which<br />

affect the continuum dust emission, namely the disk outer radius Rout, the slope p <strong>of</strong> the<br />

surface density pr<strong>of</strong>ile and the its valueΣ0 at 10 AU. Since the continuum emission<br />

has a very weak dependence on the mass <strong>of</strong> the central star, we fixed M⋆=2.6 M⊙ (see<br />

Sec. 7.5.3).<br />

Tab. 7.2 shows the parameter values <strong>of</strong> the best fitting model; the residuals corre-<br />

sponding to the four different wavelength are shown in Fig. 7.7. They have been recon-<br />

structed from the residuals in the uv-plane with the same procedure used to obtain the<br />

images in Fig. 7.1. Residual contours are generally lower than 3σ with the exception <strong>of</strong><br />

the 0.87 and 1.3 mm maps, where a flux asymmetry in the east half <strong>of</strong> the map (see also


136 The keplerian disk <strong>of</strong> HD 163296<br />

Sec. 7.3.1) is clearly visible. This structure, not detected at longer wavelengths, requires<br />

more resolved <strong>observations</strong> to be investigated in more details.<br />

The position angle (120 o ±20 o ) and inclination (40 o ±12 o ) are in good agreement with<br />

the rough estimate <strong>of</strong> Sec. 7.3.1. The position angle is not very well constrained by the<br />

continuum data alone, due to the spatial resolution generally larger than 1 ′′ and to the<br />

very elongated beam at 1.3 mm; within the uncertainty, it is in agreement with <strong>pre</strong>vious<br />

estimates in the literature. On the contrary, the inclination is significantly lower than the<br />

∼60 o obtained by Mannings and Sargent (1997) using less resolved OVRO <strong>observations</strong>,<br />

and by Grady et al. (2000), analysing HST images in scattered light. In both cases the<br />

uncertainties may be very large and it is not clear how significant is the discrepancy.<br />

The surface density radial pr<strong>of</strong>ile has a slope p=0.81±0.01 and the disk outer<br />

radius is Rout= 200±15 AU. The very good constraints on both parameters is due to<br />

the favourable orientation <strong>of</strong> the beam at 1.3 mm with the maximum resolution (0.42 ′′ )<br />

in almost the same direction <strong>of</strong> the major axis <strong>of</strong> the disk. Note that larger values <strong>of</strong> Rout<br />

are not consistent with the data, irregardless <strong>of</strong> the value <strong>of</strong> p.<br />

The slope <strong>of</strong> the dust opacity derived from the model fit isβ=1.0±0.1. The derived<br />

value <strong>of</strong>βconfirms the <strong>pre</strong>sence <strong>of</strong> very large grains in the HD 163296 circumstellar<br />

disk (Natta et al. 2004).<br />

The disk total mass (for the adopted opacity given in Eq. 7.3) is 0.05±0.01 M⊙,<br />

equal to the value obtained by Natta et al. (2004) and Mannings & Sargent (1997), but<br />

more than twice lower than the crude estimate <strong>of</strong> Sec. 7.3.4. Note that the disk mass is<br />

computed assuming that the surface density pr<strong>of</strong>ile derived from the millimeter data at<br />

large radii extends smoothly inward to the dust sublimation radius. Given the low value<br />

<strong>of</strong> p, this is not a strong assumption (most <strong>of</strong> the mass is at large radii), but we cannot<br />

rule out that more mass than we estimate is contained in the optically thick inner disk<br />

regions at r< ∼ 10 AU.


7.5 Results 137<br />

Figure 7.8: Comparison between the observed and the model <strong>pre</strong>dicted 12 CO J=2-1 emis-<br />

sion. The upper panel shows the position-velocity diagram for the 12 CO J=2-1 transition (as in<br />

Fig. 7.5). The lower panel shows the residuals relative to the best fit model parameters reported<br />

7.5.3 CO emission<br />

in Tab. 7.2. The contour levels are at 2σ as in Fig. 7.5.<br />

The analysis <strong>of</strong> the CO emission has been carried out separately for the different CO<br />

transitions and the corresponding best fit parameters are given in Tab. 7.2. At the mo-<br />

ment, the <strong>observations</strong> <strong>of</strong> the 12 CO J=3-2 emission are still under analysis. However, we<br />

think that these data will not contribute to a better constraint <strong>of</strong> the disk structure, given<br />

the low spatial resolution <strong>of</strong> the SMA <strong>observations</strong> As for the continuum, the solutions<br />

<strong>of</strong> the model fitting are unique with the reducedχ 2 close to 1.<br />

The observed 12 CO J=2-1 emission is well fitted by a keplerian disk orbiting a cen-<br />

tral star with mass M⋆ = 2.6 +0.3<br />

−0.5 M⊙. Fig. 7.8 shows the position-velocity residuals<br />

obtained subtracting the best fit model from the observed uv–table: no evidence <strong>of</strong> non-


138 The keplerian disk <strong>of</strong> HD 163296<br />

keplerian rotation or stellar outflow is detected, within the actual instrumental sensitiv-<br />

ity. Both the position angle (128 o ±5 o ) and the inclination (45 o ± 5 o ) are in agreement<br />

with the values obtained from the continuum. Since the line is optically thick, the con-<br />

straint on the gas surface density is poor with p=0.6 +0.3<br />

0.1 andΣ0= 90±70 g/cm 2 . The<br />

inferred outer radius <strong>of</strong> the disk is 550± 50 AU, more than three times larger than the<br />

value obtained by the continuum and similar to the result obtained by Thi et al. (2004).<br />

The disk outer radius will be discussed in detail in Sec. 7.6.1.<br />

The model fit to the 13 CO J=1-0 line gives results consistent with those obtained<br />

from the 12 CO J=2-1 line with exception <strong>of</strong> the value <strong>of</strong>Σ0 which is significantly<br />

smaller. This discrepancy may be due to a depletion <strong>of</strong> the 13 CO and will be discussed<br />

in Sec. 7.6.2. In general, the parameters constraint obtained from the 13 CO is worse, as<br />

expected given the lower resolution <strong>of</strong> the <strong>observations</strong>.<br />

The radial temperature pr<strong>of</strong>ile <strong>of</strong> the 12CO has a slope q=0.5 +0.2<br />

−0.1 and a value <strong>of</strong><br />

40 +2<br />

−5 K at 100AU. The radial temperature pr<strong>of</strong>ile <strong>of</strong> the 13CO has a slope q=0.8±0.4<br />

and a value <strong>of</strong> 30±10 K at 100AU. We will comment on the gas physical conditions in<br />

Sec. 7.6.<br />

7.6 Discussion<br />

The results <strong>pre</strong>sented in Sec. 7.5 show that all the <strong>observations</strong> are consistent with the<br />

emission <strong>of</strong> a circumstellar disk in keplerian rotation around a star <strong>of</strong> 2.6 +0.2<br />

−0.4 M⊙, assuming<br />

the Hipparcos distance <strong>of</strong> 122 pc. Within the error, the stellar mass is in agreement<br />

with the value <strong>of</strong> 2.3 M⊙ (Natta et al., 2004) obtained from the location <strong>of</strong> the star on<br />

the HR diagram, using Palla and Stahler (1993) evolutionary tracks; the corresponding<br />

stellar age is <strong>of</strong> about 5 Myr. The good keplerian rotation pattern and the low disk mass<br />

seem also to indicate that the disk self-gravity may have small effects on the HD 163296<br />

disk structure and evolution, while deviations from keplerian rotation can be <strong>pre</strong>sent in<br />

more massive <strong>disks</strong> (Lodato and Bertin, 2003; Cesaroni et al., 2005)<br />

The disk has a moderate inclination with respect to the line <strong>of</strong> sight (45 o ±4 o ) with a


7.6 Discussion 139<br />

Rout<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

CO<br />

Continuum<br />

0.5 0.6 0.7 0.8 0.9 1<br />

Figure 7.9:χ 2 contours in the plane p–Rout relative to the best fit models for the continuum and<br />

the 12 CO J=2–1 <strong>observations</strong>, as labelled in figure. The contours correspond to uncertainties <strong>of</strong><br />

1, 3 and 10σ, whereχ 2 1σ =χ2 m +√ 2n andχ 2 m<br />

p<br />

is the minimun value corresponding to the disk<br />

parameters reported in Tab. 7.2.<br />

position angle <strong>of</strong> 128 o ±4 o . Both the dust continuum emission and the CO lines indicate<br />

a rather shallow surface density pr<strong>of</strong>ile,Σ∝r −0.81±0.01 . The discrepancies observed in<br />

the disk outer radius andΣ0 will be discussed in the following.<br />

7.6.1 The disk outer radius<br />

The model fitting (Tab. 7.2) shows that the value <strong>of</strong> disk outer radius inferred from<br />

the continuum dust emission is three times smaller than the value obtained from the


140 The keplerian disk <strong>of</strong> HD 163296<br />

CO analysis. Since our method takes into account the sensitivity limits <strong>of</strong> the different<br />

interferometric <strong>observations</strong>, this discrepancy can not be explained by the fact that the<br />

disk outer regions have a continuum surface brightness below the sensitivity limit. In<br />

other words, if we extend the disk model that fit the continuum to the outer radius <strong>of</strong> the<br />

CO, we <strong>pre</strong>dict a continuum emission between 200 AU and 550 AU which should be<br />

easily detected in our <strong>observations</strong>. This point is illustrated in Fig. 7.6.1, which shows<br />

theχ 2 contours relative to 1, 3 and 10σ in the p-Rout plane for the 1.3 mm continuum<br />

and the 12 CO J=2-1 line. The two regions are clearly separated even at the 10σ level<br />

and no variation <strong>of</strong> the other model parameters can reduce the discrepancy.<br />

To reconcile CO and dust <strong>observations</strong>, we find that it is necessary to introduce a<br />

drop in the the continuum flux <strong>of</strong> a factor>30 at a distance <strong>of</strong> about 200 AU. With this<br />

drop, the millimeter fluxes at larger r will be below the actual instrumental sensitivity<br />

and will be lost in the observational noise. What can be the origin <strong>of</strong> such a drop?<br />

In the optically thin regime, the continuum flux emitted at distance r from the star<br />

depends on the mass surface densityΣ(r), the dust/gas ratioΠ(r), the dust opacity Kν(r)<br />

and the midplane dust temperature T(r) through the relation<br />

Fν(r)∝Σ(r)·Π(r)· Kν(r)·T(r). (7.5)<br />

It is possible that a different disk geometry (e.g., a lower flaring angle) gives a mid-<br />

plane dust temperature lower than the values <strong>pre</strong>dicted by our disk model. A lower<br />

temperature limit, <strong>of</strong>∼10 K, is however imposed by the equilibrium with the interstellar<br />

radiation field. Given that our models <strong>pre</strong>dict temperatures <strong>of</strong> 20-30 K in the millimeter<br />

emitting regions, one can reduce Fν by a factor 3 at most. Moreover, it is difficult to<br />

see how the temperature can have a large discontinuity at 200 AU. In any case, we are<br />

exploring the effects <strong>of</strong> varying the degree <strong>of</strong> flaring on the model <strong>pre</strong>dictions.<br />

A second possibility to explain the observed flux depletion is that the dust opacity<br />

Kν(r) at distance larger than 200 AU is much lower because most <strong>of</strong> the grains have<br />

grown into bodies large than about 1 m. This, however, contradicts theoretical calcu-<br />

lations which <strong>pre</strong>dict that meter size objects, wherever they form, should drift inwards<br />

with a velocity <strong>of</strong> about 1 AU/century (see the review <strong>of</strong> Dominik et al., 2006)


7.6 Discussion 141<br />

Finally, a more likely possibility is that the dust is depleted in the outer disk due to<br />

dynamical perturbation between the disk and an unknown companion orbiting the star<br />

at large distance. Such a body may also account for the asymmetric shape detected in<br />

the continuum emission (Fig. 7.7). The <strong>pre</strong>sence <strong>of</strong> a giant planet, or a brown dwarf,<br />

orbiting in the outer disk <strong>of</strong> HD 163296 has also be suggested by Grady et al. (2000),<br />

in order to explain the dark line observed in the scattered light HST images between<br />

300 AU and 350 AU from the central star. While planets are invoked to explain the<br />

large inner gaps observed in the dust distribution in the so called “transitional <strong>disks</strong>”<br />

(i.e., Calvet et al. 2005), HD 163296 will be, if confirmed by future <strong>observations</strong>, the<br />

first case in which a sub-stellar mass companion is found in the outer disk <strong>of</strong> a <strong>pre</strong>-<strong>main</strong><br />

<strong>sequence</strong> star.<br />

7.6.2 The gaseous disk<br />

The values <strong>of</strong>Σ0 reported in Tab. 7.2 indicate that the 13 CO J=1–0 emission is consistent<br />

with a ratio 13 CO/H∼ 10 −7 , about a factor 10 lower than what found in molecular<br />

clouds. We note that a similar depletion in the 13 CO abundance seems to be common in<br />

a number <strong>of</strong> observed TTS (see Dutrey et al., 1994, 1996). Since the 13 CO J=1–0 line<br />

is more optically thin than the 12 CO J=2–1, it probes the gas conditions deeper into the<br />

disk. Therefore, a possible explanation may be that part <strong>of</strong> the 13 CO is condensated onto<br />

dust grains in the colder disk regions close to the disk midplane. A second possibility,<br />

recently discussed by Jonkheid et al. (2006), is that the ratio 12 CO/ 13 CO may depend on<br />

on the dust settling. More quantitative tests about these hypotesys are in progress.<br />

Our models show that both the 12 CO and 13 CO lines are optically thick in the<br />

HD 163296 disk. The surface density is therefore poorly constrained and the mass<br />

<strong>of</strong> the gaseous disk is uncertain by more than one order <strong>of</strong> magnitude, in the interval<br />

0.01-0.8 M⊙. Extrapolating the gas surface density obtained from the continuum dust<br />

emission to the CO radius, we obtain a disk mass <strong>of</strong> 0.2 M⊙, about 8% <strong>of</strong> the stellar<br />

mass and four times higher that the disk mass obtained from the dust emission. If this<br />

is correct, HD 163296 has a rather large and massive disk and self-gravity may produce


142 The keplerian disk <strong>of</strong> HD 163296<br />

significant effects in the outher and colder regions <strong>of</strong> the disk. Note that this estimate<br />

relies on the assumed 1.3 mm opacity value, which sets the normalization <strong>of</strong> the surface<br />

density pr<strong>of</strong>ile at r∼10 AU, i.e., where the 1.3 millimeter dust continuum emission<br />

becomes optically thin. Also in this case, a more detailed analisys <strong>of</strong> the results is still<br />

in progress.<br />

7.7 Summary and conclusions<br />

This Chapter <strong>pre</strong>sents new <strong>observations</strong> <strong>of</strong> the disk <strong>of</strong> HD 163296 in the dust continuum<br />

and CO lines obtained with the VLA (7 mm continuum), PBI (1.3 mm and 2.8 mm<br />

continuum, 12 CO J=2-1 and 13 CO J=1-0 lines) and SMA (0.87 mm continuum and<br />

12 CO J=3-2 line). The disk is well resolved in both lines and continuum.<br />

We have compared the <strong>observations</strong> to the <strong>pre</strong>dictions <strong>of</strong> self-consistent disk models.<br />

We find that the disk, as seen in CO lines, is very large (R = 540±40 AU), with<br />

a keplerian rotation pattern consistent with a central mass <strong>of</strong> 2.6 +0.2<br />

−0.4 M⊙. Within the<br />

observational errors, there is no evidence <strong>of</strong> non-keplerian motions and/or significant<br />

turbulent broadening. We obtain a disk inclination <strong>of</strong> 45 o ±4 o , lower than the value<br />

<strong>of</strong>∼60 o found in literature, while the position angle <strong>of</strong> 128 o ±4 o is in agreement with<br />

<strong>pre</strong>vious results.<br />

The dust opacity has a power law dependence on wavelengthκ ∝λ −β withβ=<br />

1.0±0.1 in the interval 0.87-7 mm. This value is similar to what has been measured in<br />

a number <strong>of</strong> spatially resolved <strong>disks</strong> <strong>of</strong> HAe and TTS (e.g., Natta et al. 2004, Rodmann<br />

et al. 2006), and is very likely an indication that the bulk <strong>of</strong> the solid material in these<br />

<strong>disks</strong> has coagulated into very large bodies, <strong>of</strong> millimeter and centimeter size (Natta et<br />

al. 2006). Within the accuracy <strong>of</strong> our data, we find no significant variation <strong>of</strong>βwith r.<br />

The continuum <strong>observations</strong> constrain the surface density pr<strong>of</strong>ile (Σ∝r 0.81±0.01 ) for<br />

r


7.7 Summary and conclusions 143<br />

A comparison <strong>of</strong> the disk properties derived from the dust continuum and the CO<br />

lines at r


144 The keplerian disk <strong>of</strong> HD 163296<br />

Appendix<br />

CO emission<br />

In order to analyze the CO emission we developed a numerical code that solve the<br />

general formulation <strong>of</strong> the radiation transfer equation along each direction between the<br />

observer and the emitting source. If s is the linear coordinate along the line <strong>of</strong> sight,<br />

increasing from the observer (s≡0) towards the source, the observed emission in each<br />

direction is given by the relation<br />

Iν=<br />

∞<br />

0<br />

Sν(s)e −τν (s)Kν(s)ds, (7.6)<br />

where the optical depth is defined in each point s through<br />

τν(s)=<br />

s<br />

0<br />

Kν(s ′ )ds ′<br />

(7.7)<br />

and Kν is the absorbing coefficient <strong>of</strong> the interstellar medium. Given the high gas densi-<br />

ties in the <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> circumstellar <strong>disks</strong>, we can assume that all the CO levels<br />

corresponding to the rotational transitions under investigation are thermalized. In this<br />

case the source function can be approximated by the Planck function<br />

Sν(s)= Bν(TCO(s))= 2hν3<br />

c 2<br />

1<br />

, (7.8)<br />

exp(hν/kTCO(s))−1<br />

depending only on the local temperature <strong>of</strong> the gas TCO (c, h and k are respectively the<br />

light speed, the Plank and Boltzman constant).<br />

The absorbing coefficient <strong>of</strong> the circumstellar medium is due both to gaseous CO<br />

and dust: Kν(s)=K CO<br />

ν (s)+K d ν (s). For the dust Kd ν (s)=ρ(s)·kν, whereρ(s) is the local<br />

density <strong>of</strong> the circumstellar material (gas+dust) and kν is the dust absorbing coefficient<br />

for total mass unit given in Eq. 7.3. The CO absorbing coefficient is given by the relation<br />

K CO<br />

ν (s)=nl(s)·σν(s) (7.9)<br />

where nl(s) is the total number <strong>of</strong> CO molecules at the lower level l <strong>of</strong> the transition and<br />

σν(s) is the CO absorbing cross section.


7.7 Summary and conclusions 145<br />

Figure 7.10: Schematic re<strong>pre</strong>sentation <strong>of</strong> the frame <strong>of</strong> reference adopted to calculate the CO<br />

emission arising from a rotating disk. The disk mid plane and the observer lie respectively on<br />

the (x, y) and (y, z) planes;θ is the disk inclination; d is the distance between the observer and<br />

the central star; s is the linear coordinate along the line <strong>of</strong> sight increasing from the observer<br />

(s≡0) towards the emitting source. Assuming that the material within the disk is subject to the<br />

keplerian rotation around the central star, we call vφ the velocity <strong>of</strong> a mass element at distance r<br />

and vk the projection <strong>of</strong> vφ along the line <strong>of</strong> sight.<br />

Calling m0 andχCO the mean molecular weight <strong>of</strong> the gas and the fraction <strong>of</strong> CO<br />

<strong>pre</strong>sent in the gas respectively, the number <strong>of</strong> molecules nl(s) is given by the Boltzman<br />

equation<br />

ρ(s)<br />

nl(s)=χCO<br />

m0<br />

· gl e−El/kTCO(s) , (7.10)<br />

Z(TCO(s))<br />

where gl= 2l+1 is the statistical weight <strong>of</strong> the lower level l <strong>of</strong> the transition, El=<br />

(1/2)l(l+1)kT1 is the level energy, T1 is the temperature equivalent to the transition<br />

energy and Z(TCO(s)) is the partition function at the gas temperature TCO(s). Following<br />

Beckwith and Sargent (1993, and references therein), the absorbing cross sectionσν(s)<br />

can be ex<strong>pre</strong>ssed in term <strong>of</strong> the integrated cross section <strong>of</strong> the transitionσ0 through the<br />

relation<br />

σν(s)=σ0·φν(s)·(1−e −hν/kTCO(s) ), (7.11)


146 The keplerian disk <strong>of</strong> HD 163296<br />

whereφν(s) is the intrinsic line pr<strong>of</strong>ile<br />

φν(s)= c<br />

1/2 <br />

mCO<br />

mCO<br />

· exp<br />

2πkTCO(s) 2kTCO(s) ·∆2 <br />

v,<br />

(7.12)<br />

and<br />

ν0<br />

σ0= 8π3 kT1<br />

h 2 c<br />

(l+1) 2<br />

2l+1 µ2 . (7.13)<br />

In the <strong>pre</strong>vious equations, mCO is the CO molecular weight,ν0 is the rest frequency <strong>of</strong><br />

the molecular transition,∆v is the difference between the velocity vobs= (c/ν0)(ν−ν0),<br />

corresponding to the frequencyν, and the component <strong>of</strong> the gas velocity along the line<br />

<strong>of</strong> sight, vk(s),µis the dipole moment <strong>of</strong> the CO molecule. Note that writing Eq. 7.12,<br />

we assumed that the intrinsic line width depends only on the thermal velocity dispersion<br />

in the gas and that turbulent motion are negligible.<br />

As shown in Fig. 7.5 and described in Sec. 7.3.2, the observed velocity patterns in<br />

the CO transitions are in very good agreement with the keplerian rotation <strong>of</strong> the disk.<br />

We can thus assume that gas moves on circular orbits around the star characterized by a<br />

tangential velocity<br />

<br />

GM⋆<br />

vφ(r)= , (7.14)<br />

r<br />

where r is the radius <strong>of</strong> the orbit and M⋆ is the stellar mass. To calculate the velocity<br />

vk(s), component <strong>of</strong> the tangential velocity vφ(r) along the line <strong>of</strong> sight, it is useful to<br />

define a coordinates system centered on the star as shown in Fig. 7.10: the (x, y) plane<br />

corresponds to the disk mid plane; the observer lies in the (y, z) plane and its position<br />

it is defined by the inclinationθand the distance d from the star; each point <strong>of</strong> the<br />

circumstellar space can be defined through the cylindrical coordinates r= x 2 + y 2 ,<br />

φ=arctan (y/x) and z. Since d≫r, we can write<br />

vk(s)vφ(r)·cosφ·sinθ. (7.15)<br />

Finally, the velocity vk(s) can be calculated for each direction, knowing the geometrical<br />

transformations between the coordinate s along the line <strong>of</strong> sight and the cylindrical<br />

coordinates r andφ.


7.7 Summary and conclusions 147<br />

In order to solve the described set <strong>of</strong> equations, we thus need an ex<strong>pre</strong>ssion for the<br />

circumstellar mass densityρ(s) and the temperature <strong>of</strong> the emitting gas TCO(s). In both<br />

cases we can assume the cylindrical symmetry and writeρ(s)≡ρ(r, z) and TCO(s)≡<br />

TCO(r).<br />

For the emitting gas temperature, we choose the parametrization<br />

TCO(r)=TCO(r0)(r/r0) −q , (7.16)<br />

while the mass density is calculated assuming that the disk is in hydrostatic equilibrium<br />

and vertically isothermal in the inner region at the mid plane temperature Tm(r), through<br />

the relation<br />

ρ(r, z)=ρ0(r)·e −z2 /2h 2 (r) , (7.17)<br />

valid between the disk inner and outer radii Rin and Rout. The density on the disk mid<br />

planeρ0(r) can by ex<strong>pre</strong>ssed as function <strong>of</strong> the surface mass densityΣ(r)=Σ0(r/r0) −p<br />

through the relation<br />

ρ0(r)=<br />

Σ(r)<br />

√ 2πh(r) . (7.18)<br />

Finally, the <strong>pre</strong>ssure scale h(r) is given by the relation<br />

<br />

2r<br />

h(r)=<br />

3kTm(r) , (7.19)<br />

GM⋆m0<br />

where Tm(r) is obtained solving the structure <strong>of</strong> a stellar-irradiated passive disk as de-<br />

scribed in the <strong>pre</strong>vious section.<br />

Note that the emitting gas temperature TCO has been parametrized independently <strong>of</strong><br />

the disk mid-plane temperature Tm, which governs the density structure <strong>of</strong> the disk. As<br />

pointed out by Dartois et al. (2003), the 12 CO J=2-1 transition is in fact a good tracer<br />

<strong>of</strong> the “CO disk surface” where the gas temperature is very different from the disk mid<br />

plane.<br />

The resulting model, produces brightness maps for each frequency (velocity) that<br />

can be compared with the <strong>observations</strong> <strong>pre</strong>sented in Fig. 7.2 and 7.4.


Part V<br />

Conclusions and future prospects


CHAPTER 8<br />

Summary and Conclusions<br />

This thesis discusses a study <strong>of</strong> circumstellar <strong>disks</strong> based on high spatial resolution in-<br />

terferometric <strong>observations</strong> <strong>of</strong> intermediate (Herbig Ae; HAe) and low mass (T Tauri;<br />

TTS) <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars. The aim was to improve our knowledge <strong>of</strong> the structure<br />

and evolution <strong>of</strong> the circumstellar material by determining the composition, the size,<br />

the density distribution, the temperature and the kinematic <strong>of</strong> gas and dust. The thesis<br />

has been developed following two complementary lines: the characterization <strong>of</strong> disk re-<br />

gions at fractions <strong>of</strong> AU from the central star through near-infrared interferometry (Part<br />

II and III) and the study <strong>of</strong> the outer disk regions through millimeter interferometry (Part<br />

IV). In order to analyse the <strong>observations</strong>, I developed physical models <strong>of</strong> the disk struc-<br />

ture and emission, realized their numerical implementations and compared the model<br />

<strong>pre</strong>dictions with the data. The <strong>main</strong> results are summarized in the following.<br />

8.1 The inner disk<br />

The physical definition and the numerical implementation <strong>of</strong> models for the near-infrared<br />

emission arising from the circumstellar dust is described in Chapter 3. Following the<br />

idea <strong>of</strong> Natta et al. (2001), we confirm that the spectral energy distribution between


152 Summary and Conclusions<br />

∼1 and∼5µm <strong>of</strong> most <strong>of</strong> the observed HAe and TTS is dominated by the puffed-up<br />

inner rim, which forms at the dust evaporation radius. We argue that the rim shape<br />

is controlled by the large vertical density gradient through the dependence <strong>of</strong> the dust<br />

evaporation temperature on the surrounding gas density. The dust density (and the cor-<br />

responding evaporation temperature) is maximum on the disk mid-plane and decreases<br />

in the vertical direction. As a result, the bright side <strong>of</strong> the rim is curved, rather than<br />

vertical, as expected when a constant evaporation temperature is assumed. Contrary to<br />

the vertical rim model proposed by Dullemond et al. (2001), the near-infrared excess <strong>of</strong><br />

our model does not depend much on the disk inclination, being maximum for face-on<br />

objects. For a given star, the location <strong>of</strong> the inner rim depends mostly on the grain size<br />

<strong>of</strong> the dust species with the highest evaporation temperature. For astronomical silicates,<br />

the rim radius is maximum for grains <strong>of</strong>∼0.01µm and it reaches a minimum value for<br />

grains <strong>of</strong>∼ 1µm (assuming a typical HAe star with L⋆= 50L⊙ and Teff=10000 K). In<br />

this range <strong>of</strong> grain sizes the dust temperature is maximum on the evaporation surface<br />

and decreases rapidly moving inside the disk, i.e., with increasing optical depth. As a<br />

con<strong>sequence</strong>, the rim has a very narrow optically thin atmosphere, which, for a typi-<br />

cal Herbig Ae star, has width∼ 10 −6 AU. On the contrary, grains larger than∼ 1µm<br />

are <strong>main</strong>ly heated by the radiation field diffused by the disk its-self; their temperature<br />

increases with increasing optical depth and the dust evaporation occurs in a relatively<br />

wide optically thin atmosphere (∆r∼10 −2 AU). The con<strong>sequence</strong>s <strong>of</strong> this effect have<br />

been discussed further in Chapter 5.<br />

For a large number <strong>of</strong> models, we have computed synthetic images <strong>of</strong> the curved rim<br />

seen under different inclinations. Face-on rims are always seen as bright centrally sym-<br />

metric rings, while the images become more skewed as the inclination increases. The<br />

comparison with the interferometric <strong>observations</strong> <strong>of</strong> six <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars (CQ<br />

Tau, VV Ser, MWC 480, MWC 758, V1295 Aql and AB Aur) is discussed in Chap-<br />

ter 4. For each star, we performed aχ 2 test between the measured and the theoretical<br />

visibilities varying independently the model parameters, namely the dust size, the disk<br />

inclination and position angle. In all objects with the exception <strong>of</strong> AB Aur, our self-


8.1 The inner disk 153<br />

consistent model reproduce both the interferometric results and the near-infrared spec-<br />

tral energy distribution. In four cases we found that silicate grains larger than∼ 1µm<br />

are either required by or consistent with the data. Only in the case <strong>of</strong> MWC 480 grains<br />

<strong>of</strong> about 0.2–0.3µm are required to explain the large disk inner radius. Our analysis<br />

shows that the dust <strong>pre</strong>sent in the disk mid plane within 1 AU from the star has grown<br />

from the ISM size (∼0.01µm) to micron, or larger, radii.<br />

In the case <strong>of</strong> AB Aur, our model can not reproduce the available <strong>observations</strong>,<br />

which require an inner disk radius too small to be due to dust evaporation in the unatten-<br />

uated stellar radiation field. As a possible explanation, we suggest that the gas <strong>pre</strong>sent in<br />

the dust-depleted inner region absorbs a significant fraction <strong>of</strong> the stellar radiation, <strong>pre</strong>-<br />

venting the dust grain from the evaporation. One should also keep in mind that AB Aur<br />

is a very complex object. Recently, it has been the target <strong>of</strong> a large number <strong>of</strong> interfero-<br />

metric <strong>observations</strong> with the Infrared Optical Telescope Array (IOTA) by Millan-Gabet<br />

et al. (2006). The measured visibilities and closure phases indicate the <strong>pre</strong>sence <strong>of</strong> an<br />

asymmetry in the circumstellar environment, probably due to a compact thermal emis-<br />

sion at spatial scales <strong>of</strong> 1-4 AU. Moreover, both near-infrared coronographic images<br />

(Fukagawa et al. 2004) and millimeter interferometric <strong>observations</strong> (Piétu et al., 2005)<br />

have shown the existence <strong>of</strong> spiral density waves on scales <strong>of</strong> few tens <strong>of</strong> AU from the<br />

central star. All these results indicate a complex morphology <strong>of</strong> the material close to<br />

AB Aur which requires to be investigated further.<br />

The third part <strong>of</strong> the thesis (Chapter 6) is dedicated to the analysis <strong>of</strong> spectrally<br />

resolved AMBER/VLTI interferometric <strong>observations</strong> <strong>of</strong> the Brγ emission line <strong>of</strong> the<br />

HAe star HD 104237. The line is quite strong in this object, with a peak intensity <strong>of</strong> 35%<br />

<strong>of</strong> the continuum. The data show that the visibility does not vary between the continuum<br />

and the line. This implies that the line and the continuum emission arise at roughly the<br />

same distance from the central star. We assume that the near-infrared continuum excess<br />

originates from the puffed-up inner rim <strong>of</strong> the circumstellar disk and discuss the origin<br />

<strong>of</strong> the Brγ line. We conclude that this emission most likely comes from the base <strong>of</strong> a<br />

compact disk wind localized between 0.2–0.5 AU from the star, just in front <strong>of</strong> the dusty


154 Summary and Conclusions<br />

disk inner rim. Our conclusions are in contrast with the magnetospherical accretion<br />

model (Hartmann et al. 1998), which <strong>pre</strong>dicts that the hydrogen lines originate in the<br />

gas accretion flow at few stellar radii from the central star. Our results, if confirmed<br />

by future <strong>observations</strong>, will thus bring a new insight into the processes that govern the<br />

accretion <strong>of</strong> material on the central star during the HAe <strong>disks</strong> evolution.<br />

8.2 The outer disk<br />

The last part <strong>of</strong> the thesis (Chapter 7) is dedicated to the analysis <strong>of</strong> the <strong>observations</strong><br />

<strong>of</strong> the HAe star HD 163296, obtained with different interferometers (Plateau de Bure,<br />

SMA and VLA) in the wavelength interval between 0.87 and 7 mm. The <strong>observations</strong><br />

show the <strong>pre</strong>sence <strong>of</strong> a spatially resolved disk both in the continuum (0.87, 1.3 and<br />

2.8 mm) and in the CO lines ( 12 CO J=3–2 and J=2–1, 13 CO J=1-0). Comparing the<br />

measured visibilities with the <strong>pre</strong>dictions <strong>of</strong> self-consistent disk models we obtain a<br />

disk inclination <strong>of</strong> 45 o ±4 o and a position angle <strong>of</strong> 128 o ±4 o . The exponentβ<strong>of</strong> the dust<br />

opacity (κ∝λ −β ) between 0.87 and 7 mm isβ=1.0±0.1 and indicates the <strong>pre</strong>sence <strong>of</strong><br />

very large dust grains, <strong>of</strong> millimeter and centimeter size (Natta et al. 2006).<br />

Seen in the CO lines, the disk has an outer radius <strong>of</strong> 540±40 AU and is characterized<br />

by a keplerian rotation pattern consistent with a central star <strong>of</strong> 2.6 +0.2<br />

−0.4 M⊙, in agreement<br />

with the value <strong>of</strong> 2.3 M⊙ obtained by the stellar position on the HR diagram. Within the<br />

observational errors, there is no evidence <strong>of</strong> non-keplerian motions and/or significant<br />

turbulent broadening.<br />

The continuum <strong>observations</strong> constrain the surface density pr<strong>of</strong>ile (Σ∝r 0.81±0.01 ) for<br />

r


8.3 Conclusions 155<br />

to explain scattered light HST <strong>observations</strong>. HD 163296 is the first disk with evidence<br />

<strong>of</strong> a big discontinuity in the millimetric dust emission at large radii, and it is clearly<br />

worth it to investigate it further (see Sec. 9.2).<br />

8.3 Conclusions<br />

In this thesis I have shown that astronomical interferometry is a powerful technique<br />

to investigate the properties <strong>of</strong> circumstellar <strong>disks</strong> around nearby <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong><br />

stars. Near-infrared interferometric <strong>observations</strong>, albeit still challenging and telescope–<br />

time consuming, probe the gas and dust distribution at fractions <strong>of</strong> AU from the central<br />

star and allow the investigation <strong>of</strong> important physical processes responsible for the disk<br />

evolution and the planet formation, namely the grain growth, the gas accretion on the<br />

central star and the launching mechanism <strong>of</strong> the observed outflows. On the other hand,<br />

long baseline millimeter interferometry re<strong>main</strong>s a fundamental technique to study the<br />

gas kinematic and the dust properties in the outer disk regions, on scales <strong>of</strong> tens <strong>of</strong> AU.<br />

Due <strong>main</strong>ly to the limitations <strong>of</strong> the available near-infrared interferometers, direct<br />

<strong>observations</strong> <strong>of</strong> the inner disk have been limited so far to the few brightest objects and<br />

it is not known if the inferred properties are common to most <strong>of</strong> the <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong><br />

stars. In Chapter 9 I will discuss some ideas for future investigations.<br />

Finally, I would like to note that the self-consistent disk models, developed and dis-<br />

cussed in my thesis, have explored the importance <strong>of</strong> physical processes in circumstellar<br />

disk not considered before. Some follow-up works, which have appeared in the litera-<br />

ture after the publication <strong>of</strong> the Isella and Natta (2005) paper, have been briefly reviewed<br />

in Chapter 5.


CHAPTER 9<br />

Future developments<br />

As my thesis project developed, there were more open questions that attracted my atten-<br />

tion. In some cases, they become proposals for observing time, which, however, have<br />

not been implemented yet. I describe briefly these future developments in the following.<br />

9.1 Gas in the inner disk <strong>of</strong> Herbig Ae stars<br />

Even if the dust evaporation is responsible for the dusty disk truncation at fraction <strong>of</strong><br />

AU from the central star, about 99% <strong>of</strong> the circumstellar material is settled in a gaseous<br />

disk that may extend inward until the stellar radius. If the central star has a magnetic<br />

field, the gaseous disk may be truncated at few stellar radii and the gas captured by the<br />

magnetic field lines, giving rise to magnetospheric accretion columns. The accreting<br />

gas eventually hits the star, producing accretion shocks on the stellar surface. This<br />

magnetospherical accretion model has been successfully applied to classical TTS, where<br />

strong magnetic fields have been measured (see the review <strong>of</strong> Bouvier et al., 2006 at<br />

PPV).<br />

The situation is more confused in the case <strong>of</strong> HAe stars since the strength <strong>of</strong> the stel-<br />

lar magnetic field is very uncertain (Hubrig et al., 2006) and stellar winds can strongly


158 Future developments<br />

interact with the accreting gas. In the relatively large dust depleted inner region that<br />

characterizes HAe stars (R∼0.5 AU), the gas density and temperature are correlated<br />

with the mass accretion rate ˙M: the higher is ˙M, the warmer and denser is the gas. For<br />

˙M∼ 10 −7 M⊙/yr−10 −8 M⊙/yr, the gas temperature in the dust depleted inner disk ranges<br />

between∼1500 K, at the dust evaporation distance, and∼4000 K very close to the star<br />

(Muzerolle et al., 2004). In this range <strong>of</strong> temperatures, CO and water should exist. Wa-<br />

ter vapor in particular should be abundant since it is the second molecular reservoir <strong>of</strong><br />

oxygen after CO in the gas phase at temperatures below the dissociation temperature<br />

(∼ 2500 K) and above the water-ice sublimation temperature (∼ 150 K). Between 1000<br />

K and 2500 K, H2O is a strong molecular coolant, as well as the dominant source <strong>of</strong><br />

infrared opacity in the case in which dust grains are not <strong>pre</strong>sent (Najita et al., 2000).<br />

The <strong>pre</strong>sence <strong>of</strong> warm molecular gas in the circumstellar environment has been es-<br />

tablished both for low and high mass young stellar objects. The fundamental CO emis-<br />

sion at 4.6µm has been frequently detected in TTS (Najita et al., 2003), where it has<br />

been modelled as coming from the warm gas (∼1000 K) on the dusty disk surface. The<br />

CO overtone emission (bandhead at 2.3µm) requires on the other hand the <strong>pre</strong>sence<br />

<strong>of</strong> gas at 2000 K–4000 K, it is weaker than the fundamental one and it appears more<br />

commonly among higher luminosity objects. In TTS, it has been detected only in few<br />

percent (∼ 6%) <strong>of</strong> the observed stars (Najita et al., 2003). High spectral resolution ob-<br />

servations <strong>of</strong> the CO overtone emission show in most cases symmetric double-peaked<br />

pr<strong>of</strong>iles, strongly supporting the idea that the lines originate in the inner part (∼0.3AU)<br />

<strong>of</strong> a rotating disk (Najita et al., 2006 and references therein). The absence <strong>of</strong> red or<br />

blue-shifted absorption similar to what seen in the Balmer line <strong>of</strong> TTS, seems to rule<br />

out the possibility that the lines are emitted in winds or funnel flows. High resolution<br />

spectra provide a strong evidence <strong>of</strong> the rotation <strong>of</strong> the inner disk and are crucial to un-<br />

derstand how the gas accretes on the central star. Like the CO overtone transitions, the<br />

ro-vibrational transitions <strong>of</strong> water are also expected to probe the high density conditions<br />

in <strong>disks</strong>. Emission from hot water (see Fig. 9.1) has been detected in the near-infrared in<br />

a few stars (both low and high mass) that also show CO overtone emission (Carr et al.,


9.1 Gas in the inner disk <strong>of</strong> Herbig Ae stars 159<br />

Figure 9.1: Adapted from Thi & Bik (2005). ISAAC medium resolution (R=8900) spectrum <strong>of</strong><br />

08576nr292. The lower panel shows the observed spectrum. The box marks the region <strong>of</strong> the<br />

spectrum where the steam features are found. The top panel shows the blow-up <strong>of</strong> this region.<br />

The upper histogram is the continuum subtracted observed flux, normalized to the maximum<br />

strength at∼ 2.280µm. The middle one shows the best fit discussed in the paper; the lower curve<br />

is the synthetic spectrum before degradation to the resolution <strong>of</strong> the observed spectrum. The<br />

strong detected lines are labeled 1 to 10. Solid bars in the lower part <strong>of</strong> each panel show the<br />

spectral intervals that can be covered with CRIRES <strong>observations</strong>.


160 Future developments<br />

2004; Najita et al., 2000; Thi and Bik, 2005). Emission from steam (2.280µm-2.293µm)<br />

has been reported in <strong>disks</strong> around the low-mass young star SVS 13 (Carr et al., 2004),<br />

and the strongly accreting TTS DG Tau (Najita et al., 2000). In velocity–resolved spec-<br />

tra the water lines are much narrower than the CO lines: this is consistent with both<br />

water and CO emission originating in a keplerian rotating disk with a decreasing radial<br />

temperature pr<strong>of</strong>ile. Given the lower dissociation temperature <strong>of</strong> water (2500 K) com-<br />

pared to CO (4000 K), CO is expected to extend inward to smaller radii than water, i.e.,<br />

to higher velocities and temperatures. Spectral synthesis modeling <strong>of</strong> the detected CO<br />

and water steam features shows, however, that the water abundances relative to the CO<br />

is a factor 2-10 below the <strong>pre</strong>dictions <strong>of</strong> chemical equilibrium (Carr et al., 2004; Thi<br />

and Bik, 2005). This has at the moment no clear explanation but suggests that the UV<br />

field and/or X rays may play an important role in the gas chemistry.<br />

Since these results have been obtained on a very small sample <strong>of</strong> objects, we have<br />

recently proposed a study <strong>of</strong> a sample <strong>of</strong> 6 HAe stars using the high resolution spectro-<br />

graph CRIRES (R=30000) at the VLT in order to detect both the ro-vibrational steam<br />

emission and the overtone CO emission.<br />

For a typical HAe star (L⋆= 50L⊙) the dust depleted inner disk has a radius about<br />

10 times larger than for a TTS (L⋆=0.5⊙). The corresponding gas emitting area is thus a<br />

factor 100 larger than for TTS. The total mass <strong>of</strong> the emitting gas is also correspondingly<br />

higher. Both effects should make the observation <strong>of</strong> CO and steam emission easier in<br />

HAe than in TTS, making HAe stars the best choice to study the gas in the inner disk.<br />

Our 6 target stars are part <strong>of</strong> a large sample <strong>of</strong> 35 HAe stars for which the mass<br />

accretion rate has been derived from the Brγ luminosity (Garcia Lopez et al., 2006).<br />

They are characterized by similar mass accretion rate, i.e. similar density <strong>of</strong> the gaseous<br />

inner disk, but different value <strong>of</strong> the UV radiation field (see Tab. 9.1). For all the stars<br />

the UV flux inside the dust-depleted inner disk is due only to the stellar radiation (for<br />

˙M∼10 −7 the accretion UV flux is negligible) and increases <strong>of</strong> more than three orders<br />

<strong>of</strong> magnitude from the later to the earlier stellar spectral type; the X ray radiation is also<br />

likely negligible compared to the UV one. In this way we hope to study the chemistry <strong>of</strong>


9.2 Probing planet formation through interferometric <strong>observations</strong> 161<br />

Name SpT L⋆/L⊙ lg ˙M (a) Revp(AU) (b) G(Revp) (c) G(0.5AU) (c)<br />

HD 142527 F6 70 -7.2 0.5 2.2e+7 2.2e+7<br />

HD 144432 F0 23 -7.1 0.4 7.1e+8 4.5e+8<br />

HD 169142 A5 15 -7.4 0.2 2.7e+9 4.3e+8<br />

HD 179218 B9 500 -6.7 1.2 9.6e+9 5.5e+10<br />

HD 163296 A1 36 -7.1 0.3 1.5e+10 5.e+9<br />

VV Ser A2 63 -6.5 0.4 3.1e+10 2.0e+10<br />

Table 9.1: (a) The mass accretion rates are calculated from the Brγ line intensity (Garcia Lopez<br />

et al., 2006); (b) dust evaporation distance calculated as described in Isella et al. (2006); (c)<br />

integrated UV-flux (912Å


162 Future developments<br />

tions <strong>of</strong> the dust emission have shown that the dust undergoes drastic changes during<br />

the disk evolution. In particular, starting from a dimension <strong>of</strong>∼ 0.01µm, typical <strong>of</strong> the<br />

interstellar medium, dust grows many orders <strong>of</strong> magnitude to form the centimeter grains<br />

observed around low (T Tauri; TTS) and intermediate (Herbig Ae; HAe) <strong>pre</strong>-<strong>main</strong> se-<br />

quence stars (see the review <strong>of</strong> Natta et al. (2006) and the discussion in Chapter 7).<br />

While growing in mass and size, particles start to decouple from the gas and sediment<br />

on the disk mid-plane. For bodies larger than about 1 km, self-gravity dominates and<br />

the orbital dynamic is no more affected by the gas drag. Called planetesimals, these<br />

particles move on keplerian orbits around the central star, do not participate to the in-<br />

ward drift <strong>of</strong> the smaller material and can attract each other through their mutual gravity,<br />

aiding further growth into planet-size objects. While this core-accretion scenario (see<br />

the review <strong>of</strong> Lissauer and Stevenson, 2006) is a widely accepted theory <strong>of</strong> planet for-<br />

mation, no clear observational confirmation exists so far.<br />

Information about planet formation can be obtained through the analysis <strong>of</strong> the<br />

known exoplanets. Starting from 1995, more than 170 planetary systems have been<br />

discovered around close-by <strong>main</strong> <strong>sequence</strong> stars 1 ; the corresponding fraction <strong>of</strong> stars<br />

with planets is about 6% (Udry et al., 2006). Most <strong>of</strong> them, detected through radial<br />

velocity measurements, host giant planets (M> 0.02M jup) orbiting inside∼ 4 AU from<br />

stars <strong>of</strong> late F, G and early K spectral types. Only in four cases (Chauvin et al., 2005;<br />

Guenther et al., 2005; Biller et al., 2006), all discovered through direct images, plan-<br />

ets have been found to orbit at distances larger than 100 AU. Due to the strong biases<br />

that affect the detection techniques, these number are poorly re<strong>pre</strong>sentative <strong>of</strong> the initial<br />

phase <strong>of</strong> planetary formation. Nevertheless, they raise many questions: is the small frac-<br />

tion <strong>of</strong> discovered planetary systems real? If yes, how can it be related with the common<br />

<strong>pre</strong>sence <strong>of</strong> circumstellar <strong>disks</strong> around <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> stars? Are the circumstellar<br />

<strong>disks</strong> all proto-planetary <strong>disks</strong>? And again, where do planets form in <strong>disks</strong>? How do<br />

they interact with the surrounding material? Is the planetary formation responsible for<br />

the disk disappearance? Albeit challenging, investigation <strong>of</strong> these questions through<br />

1 from “the Extrasolar Planets Encyclopedia” at http://exoplanet.eu/


9.2 Probing planet formation through interferometric <strong>observations</strong> 163<br />

<strong>observations</strong> <strong>of</strong> <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> circumstellar <strong>disks</strong> is extremely important.<br />

At infrared and millimeter wavelengths, planetesimals and proto-planets behave like<br />

black bodies characterized by a temperature between tens to hundreds degrees, depend-<br />

ing on the distance from the central star. Their total emitting area is so small that the<br />

corresponding thermal emission is well below the actual instrumental sensitivity and<br />

overwhelmed by the surrounding gas and dust. As a result, the direct detection <strong>of</strong> plan-<br />

ets around <strong>pre</strong>-<strong>main</strong> stars is actually impossible. However, we can infer their <strong>pre</strong>sence<br />

through the observation <strong>of</strong> the dynamical perturbations that massive bodies induce both<br />

on the material surrounding them and on the global disk structure.<br />

Examples <strong>of</strong> global perturbations are constituted by the so called transitional <strong>disks</strong> in<br />

which a planet may be responsible for the clearing out <strong>of</strong> the inner disk. Planets may also<br />

be responsible for circular gaps in the gas distribution observed in few HAe and TTS<br />

(Calvet et al., 2005). Moreover, the <strong>pre</strong>sence <strong>of</strong> a planetary companion is suggested in<br />

the case <strong>of</strong> HD 163296 by our millimetric <strong>observations</strong> (Chapter 7) and by HST images<br />

(Grady et al., 2000). Local perturbations induced by a forming planet can be observed<br />

in the form <strong>of</strong> small scale enhancements <strong>of</strong> the millimetric flux, corresponding to a local<br />

increase <strong>of</strong> the dust density. Such structures have been observed in the HAe star AB Aur<br />

(Corder et al., 2005; Piétu et al. 2005). For HD 163296 we also detected asymmetry in<br />

the millimeter flux compatible with dynamical perturbations by an unknown companion<br />

(Chapter 7).<br />

Due to the spatial resolution <strong>of</strong> existing millimeter interferometers and adaptive op-<br />

tics systems, all these <strong>observations</strong> refer to disk regions at distances larger than about<br />

30 AU from the central star. In apparent contradiction with the discussed <strong>observations</strong>,<br />

planetary formation theories suggests that planets should form inside about∼20 AU,<br />

where the gas density is larger, and then eventually migrate outward (Levison et al.,<br />

2006). At the distance <strong>of</strong> the nearby star forming regions (∼150 pc), this correspond<br />

to an angular size smaller than 0.02 arcsec and can be achieved only through near- and<br />

mid-infrared interferometry.<br />

In the next future it will be therefore fundamental to increase the few available near-


164 Future developments<br />

Figure 9.2: Analysis <strong>of</strong> the existing <strong>observations</strong>. The left and right panels show K-band V 2<br />

data for MWC 758 respectively plotted as function <strong>of</strong> the baseline and <strong>of</strong> the hour angle, for the<br />

three different PTI baseline orientations (NS, baseline length <strong>of</strong> 110m in direction North-South;<br />

NW, 86m direction North-West; SW, 87m direction South-West). PTI measurements (Eisner et<br />

al. 2004) are shown by dots. The solid lines show the <strong>pre</strong>dictions <strong>of</strong> the best-fitting model with<br />

small grains (a=0.17µm); the dashed lines the best-fitting model with big grains (a≥1.2µm);<br />

grain sizes in this range can reproduce the <strong>observations</strong> at almost the same level <strong>of</strong> confidence.<br />

Each curve on the left panel corresponds to a different orientation <strong>of</strong> the baseline on the plane <strong>of</strong><br />

the sky (see Chapter 4 for more details).<br />

infrared <strong>observations</strong> <strong>of</strong> <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> <strong>disks</strong> and combine these data with interfer-<br />

ometric <strong>observations</strong> at longer wavelengths. In this way, infrared and millimeter ob-<br />

servations will allow to characterize the disk structure at all distances from the central<br />

star, from the dust evaporation radius, at fraction <strong>of</strong> AU, until the disk outer radius, at<br />

hundreds <strong>of</strong> AUs.<br />

As part <strong>of</strong> an AMBER/VLTI Guarantee Time Observation program led by the Astro-<br />

physical Observatory <strong>of</strong> Arcetri (PI: Antonella Natta), a sample <strong>of</strong> more than 30 HAe/Be<br />

stars will be observed in the following two years; the <strong>observations</strong> <strong>of</strong> a first subset <strong>of</strong> 8<br />

objects have been scheduled for the ongoing ESO observing period. Using a triplet <strong>of</strong><br />

VLT Auxiliary Telescopes, we will obtain three visibility values, spanning on a baseline<br />

range between∼30 and∼80 m, and one closure phase measurement. Even if these<br />

<strong>observations</strong> alone will be not sufficient to constrain the structure <strong>of</strong> the inner disk, they


9.2 Probing planet formation through interferometric <strong>observations</strong> 165<br />

Figure 9.3: Simulated images <strong>of</strong> the circumstellar disk puffed-up inner rims and relative values<br />

<strong>of</strong> the closure phase for the VLTI configuration UT1-UT3-UT4. The left panel shows the <strong>pre</strong>-<br />

dicted image <strong>of</strong> the puffed-up inner rim, obtained assuming the <strong>pre</strong>sence <strong>of</strong> micron size grains<br />

for an inclination <strong>of</strong> 40 ◦ and a position angle <strong>of</strong> 145 ◦ ; the relative visibility is shown with dashed<br />

line in Fig. 9.2 and 9.4. The central panel shows the <strong>pre</strong>dicted image for the disk model with<br />

small grains (a=0.17µm) and the same inclination and position angle; the relative visibility is<br />

shown with solid lines in Fig. 9.2 and 9.4. The right panel shows the closure phase for the two<br />

models as a function <strong>of</strong> the hour angle. Due to the different size <strong>of</strong> the rim, the closure phase<br />

<strong>of</strong> the model with small grains (solid line) is much larger than the value for the big grain model<br />

(dashed line).<br />

will constitute a solid data-base that we plan to extend with future <strong>observations</strong>.<br />

The first object we have selected for a detailed follow-up is the HAe MWC 758, al-<br />

ready discussed in Chapter 4. This well studied <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> star (spectral index<br />

A5IVe, luminosity 22L⊙, mass 2.0M⊙, distance 230pc) has a strong IR excess (Malfait<br />

et al., 1998), modelled as due to a flared disk with an inner rim supposed to be respon-<br />

sible for the observed [OI] 6300Å emission (Acke et al., 2005). The circumstellar disk<br />

has also been resolved using millimetric interferometric <strong>observations</strong> by Manning et al.<br />

(1997), who inferred a disk inclination <strong>of</strong> 46 ◦ and a position angle <strong>of</strong> the major axis <strong>of</strong><br />

the disk <strong>of</strong> about 116 ◦ . As discussed in Sec. 4.4.2, the available PTI visibilities can be<br />

fitted by a family <strong>of</strong> inner rim models, with parameters varying between the two extreme


166 Future developments<br />

cases shown in Fig. 9.2. While the inferred inclination and position angle <strong>of</strong> the disk<br />

are very similar to the values obtained from the millimetric <strong>observations</strong>, models with<br />

grain sizes varying by a factor 10 (between 0.1µm and 1µm) fit the <strong>observations</strong> equally<br />

well. Our model shows that it is however possible to solve the existing ambiguity and<br />

constrain the grain size using new AMBER-VLTI <strong>observations</strong>. In particular, we <strong>pre</strong>dict<br />

that the closure phase measurements will be a powerful diagnostic to constrain the grain<br />

size, being strongly dependent on the disk geometrical structure, as shown in Fig. 9.3.<br />

The <strong>observations</strong> <strong>of</strong> MWC 758 have been proposed to ESO/VLTI and are scheduled for<br />

the ongoing ESO observing period.<br />

Given the huge amount <strong>of</strong> time required to obtain useful near-infrared interferomet-<br />

ric <strong>observations</strong>, it will be fundamental to combine data obtained with different interfer-<br />

ometers (e.g. the Keck Interferometer and the Palomar Testbed Interferometer) in order<br />

to improve the uv-plane coverage and obtain a good constrain <strong>of</strong> different theoretical<br />

models.<br />

Important information on the disk structure can also be obtained from mid-infrared<br />

interferometric <strong>observations</strong> around 10µm. At this wavelength, the disk emission is<br />

dominated by the dust inside 5-10 AU from the central star, where most planets are<br />

supposed to form. The measured visibility can be compared with classical disk models<br />

(e.g. passive irradiated disk models) to determine the disk structure and orientation.<br />

Puffed-up inner rim models <strong>pre</strong>dict that disk regions between 1 and 10 AU lay in the<br />

rim shadow (see Fig. 7.10); disk shadowed regions, being cooler and denser than the<br />

ones exposed to stellar radiation, are supposed to favour planet formation. Shadowed<br />

<strong>disks</strong> differ from full-flared <strong>disks</strong> by a lower mid-infrared continuum flux and have<br />

been invoked to explain different shapes <strong>of</strong> the spectral energy distribution <strong>of</strong> HAe stars<br />

(Meeus et al. 2001). Using VLTI or/and KI, it will be possible to confirm or rule out<br />

this theory (Leinert et al., 2004; van Boekel et al., 2004 and 2005).<br />

Finally, new millimeter interferometric <strong>observations</strong> will be fundamental to perform<br />

a multi-scale study <strong>of</strong> <strong>pre</strong>-<strong>main</strong> <strong>sequence</strong> <strong>disks</strong>. Thanks to the new extended configura-


9.2 Probing planet formation through interferometric <strong>observations</strong> 167<br />

tion <strong>of</strong> Plateau de Bure and the new CARMA array 2 , angular resolutions <strong>of</strong> about 0.3<br />

and 0.1 arcsec (at 1mm) can be respectively achieved. This will allows us to resolve<br />

the nearby circumstellar <strong>disks</strong> on scale sizes <strong>of</strong>∼20-40 AU (at distances <strong>of</strong> 150 pc) and<br />

eventually detect planet forming regions.<br />

2 https://www.mmarray.org/


Figure 9.4: Simulated <strong>observations</strong> <strong>of</strong> two different inner disk models, using different telescope configurations. The left panel shows<br />

the <strong>pre</strong>dicted squared visibilities as a function <strong>of</strong> the baseline length: solid and dashed lines refer respectively to the model with small<br />

(a=0.17µm) and large (a≥1.2µm) dust grains. Solid and empty circles indicate the visibilities for the VLT UT1-UT3-UT4 telescopes<br />

configuration at four different hour angles. The central panel shows the variation <strong>of</strong> the visibility with the hour angle for each baseline.<br />

Finally, the panel on the right shows the uv-plane coverage (dots) obtained with the chosen configuration. The <strong>pre</strong>dicted closure phases<br />

are plotted in Fig. 9.3<br />

168 Future developments


Published papers<br />

• Isella, A.; Testi, L.; Natta, A., “Large dust grains in the inner region <strong>of</strong> circum-<br />

stellar <strong>disks</strong>”, 2006, A&A, 451, 951<br />

• Tatulli, E.; Isella, A.; Natta, A.; Testi, L.; Marconi, A.; and the AMBER consor-<br />

tium, “Constraining the wind launching region in Herbig Ae stars: AMBER/VLTI<br />

spectroscopy <strong>of</strong> HD104237”, to appear in A&A (astro-ph/0606684)<br />

• Isella, A.; Natta, A., “The shape <strong>of</strong> the inner rim in proto-planetary <strong>disks</strong>”, 2005,<br />

A&A, 438, 899<br />

• Malbet, F.; Benisty, M.; De Wit, W. J.; Kraus, S.; Meilland, A.; Millour, F.;<br />

Tatulli, E.; Berger, J. -P.; Chesneau, O.; and the AMBER consortium, “Disk and<br />

wind interaction in the young stellar object MWC 297 spatially resolved with<br />

VLTI/AMBER”, to appear in A&A (astro-ph/0510350)<br />

In <strong>pre</strong>paration<br />

• Isella, A; Testi, L; Natta, A.; Neri, R.; Wilner, D.; and Qi C., “The keplerian disk<br />

<strong>of</strong> HD 163296”


BIBLIOGRAPHY<br />

[2005] Acke B., van den Ancker M. E., Dullemond C. P., 2005, A&A, 436, 209<br />

[2000] Akeson R.L., Ciardi D.R., van Belle G.T., Creech-Eakman M.J., Lada E.A. 2000, ApJ<br />

543, 313<br />

[2002] Akeson R.L., Ciardi D.R., van Belle G.T., Creech-Eakman M.J., 2000, ApJ 566, 1124<br />

[2005] Akeson R.L, Walker C.H., Wood K., Eisner J.A., et al., 2005, ApJ, 622, 440<br />

[1993] André P, Ward-Thompson D.; Barsony M., 1993, ApJ, 406, 122<br />

[1986] Arquilla R, Goldsmith P.F., 1986, ApJ, 303, 356<br />

[1991] Balbus S.A.; Hawley J.F., 1991, ApJ, 376, 214<br />

[2003] Balbus S.A., 2003, ARA&A, 41, 555<br />

[2006] Bally J., Reipurth B., Davis C.J., 2006, in Protostars and Planets V, B. Reipurth, D.<br />

Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, in <strong>pre</strong>ss.<br />

[1990] Beckwith S.V.W., Sargent A.I., Chini R.S., Guesten R., 1990, AJ, 99, 924<br />

[1991] Beckwith S.V.W., Sargent A.I., 1991, ApJ, 381, 250<br />

[1993] Beckwith S.V.W., Sargent A.I., 1993, ApJ, 402, 280<br />

[1996] Beckwith S.V.W., Sargent A.I., 1996, Nature, 383, 139<br />

[1997] Bertin G., 1997, ApJ, 478, 71<br />

[1999] Bertin G., Lodato G., 1999, A&A, 350, 694<br />

[2003] Bianchi S., Goncalves J., Albrecht M., Caselli P., Chini R., Galli D., Walmsley M.,<br />

2003, A&A, 399, 43<br />

[2006] Biller B.A., Close L.M., Li A., Marengo M., Bieging J.H., Hinz P.M., et al., 2006,<br />

ApJ, 647, 464<br />

[2004] Blum J., Schräpler R., 2004, Phys. Rev. Lett., vol. 93


172 BIBLIOGRAPHY<br />

[2000] Bodenheimer P., Burkert A., Klein R.I., Boss A.P., 2000, in Protostars and Planets IV,<br />

Mannings V., Boss A., and Russel S.S Eds., University <strong>of</strong> Arizona Press, Tucson, p.<br />

675<br />

[2004] Böhm T., Catala C., Balona L., Carter, B., 2004, A&A, 427, 907<br />

[1995] Boss A.P, Myhill E.A., 1995, ApJ, 451, 218<br />

[2006] Bouvier J., Alencar S. H. P., Harries T. J., Johns-Krull C.M., Romanova M.M., 2006,<br />

in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong><br />

Arizona Press, Tucson, in <strong>pre</strong>ss.<br />

[1991] Calvet N., Patiño A., Magris G.C., D’Alessio P. 1991, ApJ 380, 617C<br />

[1992] Calvet N., Magris G.C., Patiño A., D’Alessio P. 1992, Rev. Mexicana Astron. Astr<strong>of</strong>.,<br />

24, 27<br />

[2005] Calvet N., D’Alessio P., Watson D.M., Franco-Hernández R., Furlan E., Green J.,<br />

2005, ApJ, 630, 185<br />

[2004] Carr J.S., Tokunaga A.T., Najita J., 2004, ApJ, 603, 213;<br />

[2002] Caselli P., Walmsley C.M., Zucconi A., Tafalla M., Dore L., Myers P.C., 2002, ApJ,<br />

572, 238<br />

[2000] Casse F., Ferreira J., 2000, A&A, 353, 1115<br />

[2005] Cesaroni R., Galli D., Lodato G., Walmsley C.M., Zhang Q., 2006, in Protostars and<br />

Planets V, B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson,<br />

in <strong>pre</strong>ss.<br />

[1960] Chandrasekhar S., 1960, Plasma Physics (Chicago: Univ. Chicago Press)<br />

[2005] Chauvin G., Lagrange A.-M., Dumas C., Zuckerman B., Mouillet D., et al., 2005,<br />

A&A, 438, 25<br />

[1997] Chiang E.I., Goldreich P. 1997, ApJ 490, 368<br />

[2001] Chiang E.I., Joung M.K., Creech-Eakman M.J., Qi C., Kessler J.E., Blake G.A., van<br />

Dishoeck E.F. 2001, ApJ, 547, 1077<br />

[2004] C<strong>of</strong>fey D., Bacciotti F., Woitas J., Ray T.P., Eislöffel J., 2004, ApJ, 604, 758<br />

[2003] Colavita M. et al, 2003, ApJ, 592, 83<br />

[2005] Corder S., Eisner J., Sargent A.I. 2005, ApJ, 622L, 133C<br />

[1999] Cornwell T., Braun R., Briggs D., 1999, in Sysnthesis Imaging in Radio Astronomy<br />

II, Taylor G.B., Carilli C.L., Perley R.A eds, ASP Conference Series, Vol. 180.<br />

[2006] Crutcher R.M., Troland T.H., 2006, IAU Symp. ,237, 25<br />

[2003] Cutri R.M. et al. 2003, yCat, 2246, 0<br />

[2003] Dartois E., Dutrey A., Guilloteau S. 2003, A&A, 399, 773<br />

[2003] Dominik C., Dullemond C.P., Waters L.B.F.M., Walch S. 2003, A&A, 398, 607


BIBLIOGRAPHY 173<br />

[2006] Dominik C., Blum J., Cuzzi J.N., Wurm G., 2006, in Protostars and Planets V, B.<br />

Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, in <strong>pre</strong>ss.<br />

[1997] Donati J.-F., Semel M., Carter B. D., Rees D. E., Collier Cameron A., 1997, MNRAS,<br />

291, 658<br />

[1985] Draine B.T. 1985, ApJS 57, 587D<br />

[2003] Draine B.T., 2003, ARA&A, 41, 241<br />

[2006] Draine B.T., 2006, ApJ, 636, 1114<br />

[2000] Dullemond C.P. 2000, A&A, 361, L17<br />

[2001] Dullemond C.P., Dominik C., Natta A. 2001, ApJ 560, 957<br />

[2002] Dullemond C.P. 2002, A&A, 395, 853<br />

[2003] Dullemond C.P., van den Ancker M.E., Acke B., van Boekel R. 2003, ApJ, 594, 47<br />

[2004] Dullemond C.P., Dominik C. 2004, A&A, 417, 159<br />

[2006] Dullemond C.P., Hollenbach D., Kamp I., D’Alessio P. 2006, in Protostars and Planets<br />

V, B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, 2006,<br />

in <strong>pre</strong>ss.<br />

[2006] Durisen R.H., Boss A.P., Mayer L., Nelson A.F., Quinn T., Rice W.K.M., 2006, in Protostars<br />

and Planets V, B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona<br />

Press, Tucson, 2006, in <strong>pre</strong>ss.<br />

[1993] Dutrey A., Duvert G., Castets A., Langer W.D., Bally J., Wilson R.W., 1993, A&A,<br />

270, 468<br />

[1994] Dutrey A., Guilloteau S., Simon M., 1994, A&A, 286, 149<br />

[1996] Dutrey A., Guilloteau S., Duvert G., Prato L., Simon M., Schuster K., Ménard F.,<br />

1996, A&A, 309, 493<br />

[2006] Dutrey A., Guilloteau S., Ho P. 2006, in Protostars and Planets V, B. Reipurth, D.<br />

Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, 2006, in <strong>pre</strong>ss.<br />

[2003] Eisner J.A., Lane B.F., Akeson R.L., Hillebrand L.A., Sargent A.I. 2003, ApJ 588,<br />

360<br />

[2004] Eisner J.A., Lane B.F., Hillebrand L.A., Akeson R.L., Sargent A.I. 2004, ApJ 613,<br />

1049<br />

[2005] Eisner J.A., Hillebrand L.A., White R.J., Akeson R.L., Sargent A.I. 2005, astroph/0501308<br />

[2003] Feigelson E.D., Lawson W.A., Garmire G.P., 2003, ApJ, 599, 1207<br />

[2004] Fukagawa et al., ApJ, 605, 53<br />

[2006] Furlan E., Hartmann L., Calvet N., D’Alessio P., Franco-Hernández R., Forrest W.J.,<br />

et al., ApJS, 165, 568


174 BIBLIOGRAPHY<br />

[2004] Gail H.-P., 2004, A&A, 413, 571<br />

[1996] Gammie C.F., 1996, ApJ, 457, 355<br />

[2006] Garcia Lopez R., Natta A., Testi L., Habart E. 2006, A&A, in <strong>pre</strong>ss (astro-ph 0609032)<br />

[1974] Glass I.S., Penston M.V. 1974, MNRAS, 167, 237G<br />

[1997] Glassgold A.E., Najita J., Igea J., 1997, ApJ, 480, 344<br />

[2006] Goodwin S., Kroupa P., Goodman A., Burkert A, 2006, in Protostars and Planets V,<br />

B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, 2006,<br />

in <strong>pre</strong>ss<br />

[1999] Grady C.A., Woodgate B., Bruhweiler F.C., Boggess A., Plait P., Lindler D.L.,<br />

Clampin M., Kalas P. 1999, ApJ, 523, L151<br />

[2000] Grady C.A., Devine D., Woodgate B., Kimble R., Bruhweiler F. C., et al., 2000, ApJ,<br />

544, 859<br />

[2004] Grady C.A., Woodgate B., Torres C.A.O.; Henning Th., Apai D.; Rodmann J., et al.,<br />

2004, ApJ, 608, 809<br />

[2001] Grinin V.P., Kozlova O.V., Natta A., Ilyin I., Tuominen I., Rostopchina A.N.,<br />

Shakhovskoy D.N. 2001, A&A, 379, 482<br />

[2005] Guenther E.W., Neuhäuser R., Wuchterl G., Mugrauer M., Bedalov A., Hauschildt P.,<br />

2005, AN, 326, 958<br />

[2004] Habart E., Natta A., Krügel E. 2004, A&A, 427, 179<br />

[1968] Habing H.J., 1968, Bulletin <strong>of</strong> the Astronomical Institutes <strong>of</strong> the Netherlands, 19, 421<br />

[1994] Hartmann L., Hewett R., Calvet N., 1994, ApJ, 426, 669<br />

[1998] Hartmann L., Calvet N., Gullbring E., D’Alessio P., 1998, ApJ, 495, 385<br />

[2006] Hartmann L., D’Alession P., Calvet N., Muzerolle J., 2006, ApJ, 648, 484<br />

[1992] Hillenbrand L.A., Strom S.E., Vrba F.J., Keene J. 1992, ApJ, 397,613<br />

[1998] Hillenbrand L.A., Strom S.E., Calvet N., Merrill K.M., Gatley I., 1998, Aj, 116, 1816<br />

[2006] Hubrig S., Yudin R.V., Schöller M., Pogodin M. A., A&A, 446, 1089<br />

[1987] Kenyon S.J., Hartmann L., 1987, ApJ, 323, 714<br />

[2005] Kessler-Silacci J.E., Hillenbrand L.A., Blake G.A., Meyer M.R., 2005, ApJ, 662, 404<br />

[2006] Kessler-Silacci J.E., Augereau J.-C., Dullemond C.P., Geers V., Lahuis F., et al., 2006,<br />

ApJ, 639, 275<br />

[1985] Kilkenny D., Whittet D.C.B, Davies J.K., Evans A., Bode M.F., Robson E.I., Banfield<br />

R.M. 1985, SAAOC, 9, 55K<br />

[2006] Jonkheid B., Dullemond C.P., Hogerheijde M.R., van Dishoeck E.F., 2006, to appear<br />

in A&A, astro-ph/0611223


BIBLIOGRAPHY 175<br />

[2006] Ilgner M., Nelson R.P.,2006, ApJ, 445, 223<br />

[2005] Isella A., Natta A. 2005, A&A, 438, 899I<br />

[2006] Isella A., Testi L., Natta A., 2005, A&A, 451, 951<br />

[2007] Isella A., Testi L., Natta A., Neri R., Wilner D, Qi C., 2007, in <strong>pre</strong>paration<br />

[1984] Lada C.J., Wilking B.A., 1984, ApJ, 287, 610<br />

[1994] Laughlin G., Bodenheimer P., 1994, ApJ, 436, 335<br />

[2004] Leinert Ch., van Boekel R., Waters L. B. F. M., Chesneau, O., Malbet, F., et al., 2004,<br />

A&A, 423, 537<br />

[2006] Levison H.E., Morbidelli A., Gomes R., Backman D., 2006, in Protostars and Planets<br />

V, B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, in<br />

<strong>pre</strong>ss.<br />

[2004] Li Z., Nakamura F., 2004, ApJ, 609, 83<br />

[2006] Lissauer J.J., Stevenson D.J., 2006, in Protostars and Planets V, B. Reipurth, D. Jewitt,<br />

and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, in <strong>pre</strong>ss.<br />

[2001] Lodato G., Bertin G., 2001, A&A, 375, 455<br />

[2003] Lodato G., Bertin G., 2003, A&A, 408, 1015<br />

[1974] Lynden-Bell D., Pringle J.E., 1974, MNRAS, 168, 603<br />

[2006] Malbet F., Benisty M., De Wit W.J., Kraus S., Meilland A., Millour F. et al. 2006,<br />

astro-ph/0510350<br />

[2006] Mayer M.R., Backman D.E., Weinberger A., Wyatt M.C., 2006, in Protostars and<br />

Planets V, B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson,<br />

in <strong>pre</strong>ss.<br />

[1998] Malfait K., Boegaert E., Waelkens C. 1998, A&A, 331, 211<br />

[1994] Mannings, V. 1994, MNRAS, 271, 587<br />

[1997] Mannings V., Sargent A.I., 1997, ApJ, 490, 792<br />

[2000] Mannings V., Sargent A.I., 2000, ApJ, 529, 391<br />

[2003] Matsumoto T., Hanawa T., 2003, ApJ, 595, 913<br />

[2001] Meeus G., Waters L.B.F.M., Bouwman J., van den Ancker M.E., Waelkens C., Malfait<br />

K. 2001, A&A, 365, 476<br />

[2003] Meeus G., Sterzik M., Bouwman J., Natta A. 2003, A&A, 409, 25<br />

[2001] Millan-Gabet R., Schloerb P. F., Traub W. A. 2001, ApJ, 546, 358<br />

[2006] Millan-Gabet R., Monnier J.D., Berger J.-P., Traub W.A., et al., 2006, ApJ, 645, 77<br />

[2006] Millan-Gabet R., Malbet F., Akeson R., Leinert C., Monnier J.D., Waters R., 2006,<br />

in Protostars and Planets V, B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong>


176 BIBLIOGRAPHY<br />

Arizona Press, Tucson, in <strong>pre</strong>ss.<br />

[2002] Monnier J.D., Millan-Gabet R. 2002, ApJ 597, 694<br />

[2003] Monnier J.D., 2003, Reports on Progress in Physics, 66, 789<br />

[2005] Monnier J.D., Millan-Gabet R., Billmeier R., Akeson R.L., Wallace D. et al. 2005,<br />

ApJ, 624, 832M<br />

[2006] Monnier J.D., Berger J.-P., Millan-Gabet R., Traub W.A., et al., 2006, ApJ, 647, 444<br />

[2001] Muzerolle J., Calvet N., Hatrmann L., 2001, ApJ, 550, 994<br />

[2003] Muzerolle J., Calvet N., Hartmann L., D’Alessio P. 2003, ApJ 597, 149<br />

[2004] Muzerolle J., D’Alessio P., Calvet N., Hartmann L. 2004, ApJ, 617, 406<br />

[1998] Nakano T., 1998, ApJ, 494, 587<br />

[2000] Najita J., Edwards S., Basri G., Carr J., 2000, in Protostars and Planets IV, Mannings<br />

V., Boss A., and Russel S.S Eds., University <strong>of</strong> Arizona Press, Tucson, p. 457.<br />

[2003] Najita J., Carr J.S., Mathieu R.D, 2003, ApJ, 589, 931;<br />

[2006] Najita J., Carr J.S., Glassgold A.E., Valenti J.A., 2006, in Protostars and Planets V, B.<br />

Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, in <strong>pre</strong>ss.<br />

[1988] Natta A., Giovanardi C., Palla F., 1988, ApJ, 332, 921<br />

[2000] Natta A., Grinin V.P., & Mannings V. 2000, in Protostars and Planets IV, V. Mannings,<br />

A.P. Boss & S.S. Russell Eds., (Tucson: Univ. <strong>of</strong> Arizona Press), p. 559<br />

[2000] Natta A., Whitney B.A. 2000, A&A, 364, 633<br />

[2001] Natta A., Prusti T., Neri R., Grinin V. P., Mannings V. 2001, A&A, 371, 186<br />

[2004] Natta A., Testi L., Neri R., Shepherd D. S., Wilner D. J. 2004, A&A, 416, 179<br />

[2006] Natta A., Testi L., Calvet N., Henning T., Waters R., Wilner D. 2006, in Protostars<br />

and Planets V, B. Reipurth, D. Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press,<br />

Tucson, 2006, in <strong>pre</strong>ss.<br />

[1999] Ostriker E.C., Gammie C.F., Stone J.M., 1999, ApJ, 513, 259<br />

[1993] Palla F., Sthaler S.W., 1993, ApJ, 418, 414<br />

[1995] Papaloizou J.C.B., Lin D.N.C., 1995, ARA&A, 33, 505<br />

[2003] Perrin G., Malbet F., 2003, Observing with the VLT interferometer, G. Perrin and F.<br />

Malbet. eds., Vol. 6<br />

[2003] Petrov R.G., Malbet F.; Weigelt G.; Lisi F.; Puget P.; Antonelli P., et al., 2003, SPIE,<br />

4838, 924<br />

[2005] Piétu V., Guilloteau S., Dutrey A. 2005, A&A, 443, 945<br />

[2006] Piétu V., Guilloteau S., Dutrey A., Chapillon E., Pety J. 2006, A&A, in <strong>pre</strong>ss (astro-ph<br />

0610200)


BIBLIOGRAPHY 177<br />

[1994] Pollack J.B., Hollenbach D., Beckwith S., Simonelli D.P., Roush T., Fong W. 1994,<br />

ApJ 421, 615<br />

[2006] Pontoppidan K.M., Dullemond C.P., Blake G.A., Adwin Boogert A.C., van Dishoeck<br />

E.F., 2006, to appear in ApJ (astro-ph/0610384)<br />

[2002] Press W. H., Teukolsky S. A., Vetterling W. T., Flannery B. P. 2002, Numerical<br />

Recipes in C++, Cambrige University Press<br />

[2006] Retting T.,Brittain S., Simon T., Gibb E., Balsara D.S., Tilley D.A., Kulesa C., 2006,<br />

ApJ, 646, 342<br />

[2004] Rice W.K.M., Lodato G., Pringle J.E., Armitage P.J., Bonnell I., 2004, MNRAS, 355,<br />

543<br />

[2006] Rice W.K.M., Lodato G., Pringle J.E., Armitage P.J., Bonnell I., 2006, MNRAS, 372,<br />

L9<br />

[2006] Rodmann J., Henning Th., Chandler C.J., Mundy L.G., Wilner D.J., 2006, A&A, 446,<br />

211<br />

[2005] Rodríguez L.F.; Loinard L., D’Alessio P., Wilner D.J., Ho P.T.P., 2005, ApJ, 621, 133<br />

[2001] Rostopchina A.N., Grinin V.P., Shakhovskoi D.N. 2001,Arep, 45, 51<br />

[1973] Shakura N.I., Sunyaev R.A., 1973, A&A, 24 337<br />

[1994] Shu F., Najita J., Ostriker E., Wilkin F., Ruden S., Lizano, S., 1994, ApJ, 429, 781<br />

[2000] Simon M., Dutrey A., Guilloteau S. 2000, ApJ, 545, 1034<br />

[1978] Spitzer L., 1978, Journal <strong>of</strong> the Royal Astronomical Society <strong>of</strong> Canada, Vol. 72, p.349<br />

[2005] Stahler S., Palla F., 2005, in The Formation <strong>of</strong> Stars, book review, Wiley ed.<br />

[2001] Stone J.M., Pringle J.E., 2001, MNRAS, 322, 461<br />

[1996] Straizys V., Cernis K, Bartasiute S. 1996, Baltic Astronomy, 5, 125<br />

[1999] Tambovtseva L.V., Grinin V.P., Kozlova O.V., 1999, Astrophysics, 42<br />

[2005] Tanaka H., Himeno Y., Ida S., 2005, ApJ, 625, 414<br />

[2007] Tannirkulam A., Harries T.J., Monnier J.D., 2007, submitted to ApJ<br />

[2006] Tatulli E., et al., 2006, to appear in A&A, (astro-ph/0602346)<br />

[1984] Terebey S, Shu F.H., Cassen P., 1984, 286, 529<br />

[2001] Testi L., Natta A., Shepherd D.S., Wilner D.J. 2001, A&A, 554, 1087T<br />

[2003] Testi L., Natta A., Shepherd D.S., Wilner D.J. 2003, A&A, 403, 323T<br />

[2004] Thi W.-F., van Zedelh<strong>of</strong>f G.-J., van Dishoeck E.F, 2004, A&A, 425, 955<br />

[2005] Thi W.-F., Bik A., 2005, A&A, 438, 557<br />

[2005] Thi W.-F., van Dalen B., Bik A., Waters L.B.F.M., 2005, A&A, 430, 61<br />

[2003] Thiébaut E., Garcia P.J.V., Foy, R., 2003, Ap&SS, 286, 171


178 BIBLIOGRAPHY<br />

[2001] Thompson A.R., Moran J.M., Swenson G.W., 2001, Interferometry and Synthesis in<br />

Radio Astronomy, Wiley ed.<br />

[1991] Tomley L., Cassen P., Steiman-Cameron T., 1991, ApJ, 382, 530<br />

[2001] Tuthill P.G., Monnier J.D., Danchi W.C., 2001, Nature, 409<br />

[2006] Udry S., Fischer D., Queloz D., 2006, in Protostars and Planets V, B. Reipurth, D.<br />

Jewitt, and K. Keil Eds., University <strong>of</strong> Arizona Press, Tucson, in <strong>pre</strong>ss.<br />

[2003] van Boekel R., Waters L.B.F.M., Dominik C., Bouwman J., de Koter A., Dullemond<br />

C.P., Paresce F. 2003, A&A, 400, L21<br />

[2004] van Boekel R., Min M., Leinert Ch., Waters L.B.F.M., Richichi A. et al. 2004, Nature,<br />

432, 479<br />

[2005] van Boekel R., Dullemond C. P., Dominik, C., 2005, A&A, 441, 563<br />

[1998] van den Ancker M.E., de Winter D., Tjin A Djie H.R.E 1998, A&A, 330, 145V<br />

[1959] Velikhov E.P., 1959: Sov. Phys. JETP 36, 1398<br />

[2006] Vinkovic D., Ivezic Z., Tomislav J., Moshe E. 2006, ApJ, 636, 348<br />

[2006] Vinkovic D., 2006, ApJ, 651, 906<br />

[2001] Weingartner J.C., Draine B.T. 2001, ApJ, 548, 296<br />

[1986] Wolfire M.G., Cassinelli J.P., 1986, ApJ, 310, 207<br />

[1987] Wolfire M.G., Cassinelli J.P., 1987, ApJ, 319, 850<br />

[2002] Youdin A.N., F.H. Shu, 2002, ApJ, 580, 494<br />

[1974] Zuckerman B., Evans N.J., 1974, ApJ, 192, 149


Ringraziamenti<br />

Al termine di questi tre anni di Dottorato i miei piú sinceri ringraziamenti vanno ad<br />

Antonella Natta, Leonardo Testi e Giuseppe Bertin per aver reso materialmente possible<br />

questo lavoro, per le sem<strong>pre</strong> stimolanti discussioni e l’inestimabile aiuto.<br />

Un grazie speciale a Malcolm Walmsley per la cordialitá, la simpatia e gli utili con-<br />

sigli; ad Eric Tatulli per la <strong>pre</strong>ziosa collaborazione, sperando che possa continuare a<br />

lungo; a Riccardo Cesaroni per la pazienza e la disponibilitá nel rispondere alle mie nu-<br />

merose richieste di aiuto; a tutti i compagni di pranzo per i piacevoli momenti trascorsi<br />

insieme.<br />

Ringrazio e saluto con piacere tutto il personale dell’Osservatorio di Arcetri, a par-<br />

tire dai nostri direttori Francesco Palla e Marco Salvati, tutto il personale ricercatore,<br />

tecnico ed amministrativo per la cordialitá e la disponibilitá sem<strong>pre</strong> dimostrata.<br />

Infine, saluto con affetto tutti gli amici distribuiti tra Milano e Firenze, la mia famiglia<br />

e sopratutto mia moglie Rossana per non avermi (ancora) buttato fuori di casa!

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!