astronomy c eso 2004 astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · we briefly...

17
A&A 415, 643–659 (2004) DOI: 10.1051/0004-6361:20034061 c ESO 2004 Astronomy & Astrophysics Transport processes and chemical evolution in steady accretion disk flows M. Ilgner 1 , Th. Henning 2 , A. J. Markwick 3 , and T. J. Millar 3 1 Computational Physics, Auf der Morgenstelle 10, 72076 T¨ ubingen, Germany 2 Max Planck Institute for Astronomy, K¨ onigstuhl 17, 69117 Heidelberg, Germany 3 Department of Physics, UMIST, PO Box 88, Manchester M60 1QD, UK Received 9 July 2003 / Accepted 12 November 2003 Abstract. We study the influence of mass transport processes on the chemical evolution in a protoplanetary accretion disk. Local transport processes by advection as well as global transport processes by diusion are taken into account. Concerning the multi-component system only diusion in the vertical direction was taken into account. Depending on the transport properties, dierent schemes are explored to couple/decouple the physical and chemical evolution. Our model is based on a simplified description of hydrodynamics in terms of a steady 1+1-D-α-disk model and includes the kinetics of an extended chemical network of about 250 species. We restrict our calculations to the inner planet formation zone within a distance to the central star of 10 AU. Vertical mixing does change the global chemical evolution as it is demonstrated in detail through a discussion of the chemistry of sulphur-bearing molecules. In addition, the influence of the local disk structure on the chemical evolution is analysed. Our results demonstrate that the chemical evolution is influenced globally by mass transport processes. However, in addition to mass transport processes, information about the local conditions, which determine the kinetics, is still needed in order to analyze the chemical evolution. Key words. solar system: formation – stars: circumstellar matter – ISM: molecules – ISM: abundances – ISM: evolution – protoplanetary disks 1. Introduction It is dicult to give a quantitative comparision between the pre- vious models describing chemical evolution in protoplanetary accretion disks; the chemical models are full of dierent inten- tions. For instance, the models of the Heidelberg group (Bauer et al. 1997; Finocchi & Gail 1997; Finocchi et al. 1997; Gail 1998, 2001) include treatments of ice mantle evaporation, dust evaporation, three-body reactions, ionisations by cosmic rays, radionuclides and UV photons. Later, in addition to annealing and carbon combustion the model of Gail (2002) accounts for carbon dust oxidation. The approach used by Willacy et al. (1998) and Aikawa et al. (1999) is similar to Bauer et al. but does not account for dust destruction processes. Later, Aikawa & Herbst (1999, 2001) and Aikawa et al. (2002) added deuterium chemistry to their chemical model. The influence of the ionisation processes such as X-rays on the chemistry were introduced by Aikawa & Herbst (1999), and were studied in more detail by Markwick et al. (2002). Furthermore, Willacy & Langer (2000) discussed the importance of photoprocessing. Send oprint requests to: M. Ilgner, e-mail: [email protected] However, summarizing the improvements of recent re- search the models mainly focused attention on upgrading and extending the chemical networks as well as the implementation of additional kinetic processes. A thorough treatment of physical processes within proto- planetary disks in parallel to chemical simulations was not the focus of many previous studies. This kind of imbalance was finally broken by Willacy et al. (1998) who introduced a 1+1- D-disk model in order to evaluate the chemical evolution based on that model. In a series of simulations, concerning the chem- ical evolution in accretion disks, the advective mass transport was taken into account, cf. Duschl et al. (1996), Bauer et al. (1997), Finocchi et al. (1997), Finocchi & Gail (1997), Willacy et al. (1998), Aikawa et al. (1999), Aikawa & Herbst (1999), and Markwick et al. (2002). However, vertical mass transport processes caused by diusion were not considered. This is the first of a series of papers aimed at dealing with mass transport processes in accretion disks in the framework of chemical evolution. The fluid in the model is continuously injected at the outer edge of the disk and the flow (and chem- istry) is followed until the lower radial bound of our calcu- lation, 1 AU. At smaller radii, as the material accretes on to the central protostar, temperatures are very large and we would

Upload: others

Post on 10-Nov-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

A&A 415, 643–659 (2004)DOI: 10.1051/0004-6361:20034061c© ESO 2004

Astronomy&

Astrophysics

Transport processes and chemical evolutionin steady accretion disk flows

M. Ilgner1, Th. Henning2, A. J. Markwick3, and T. J. Millar3

1 Computational Physics, Auf der Morgenstelle 10, 72076 Tubingen, Germany2 Max Planck Institute for Astronomy, Konigstuhl 17, 69117 Heidelberg, Germany3 Department of Physics, UMIST, PO Box 88, Manchester M60 1QD, UK

Received 9 July 2003 / Accepted 12 November 2003

Abstract. We study the influence of mass transport processes on the chemical evolution in a protoplanetary accretion disk.Local transport processes by advection as well as global transport processes by diffusion are taken into account. Concerning themulti-component system only diffusion in the vertical direction was taken into account. Depending on the transport properties,different schemes are explored to couple/decouple the physical and chemical evolution. Our model is based on a simplifieddescription of hydrodynamics in terms of a steady 1+1-D-α-disk model and includes the kinetics of an extended chemicalnetwork of about 250 species. We restrict our calculations to the inner planet formation zone within a distance to the centralstar of 10 AU.Vertical mixing does change the global chemical evolution as it is demonstrated in detail through a discussion of the chemistryof sulphur-bearing molecules. In addition, the influence of the local disk structure on the chemical evolution is analysed. Ourresults demonstrate that the chemical evolution is influenced globally by mass transport processes. However, in addition to masstransport processes, information about the local conditions, which determine the kinetics, is still needed in order to analyze thechemical evolution.

Key words. solar system: formation – stars: circumstellar matter – ISM: molecules – ISM: abundances – ISM: evolution –protoplanetary disks

1. Introduction

It is difficult to give a quantitative comparision between the pre-vious models describing chemical evolution in protoplanetaryaccretion disks; the chemical models are full of different inten-tions. For instance, the models of the Heidelberg group (Baueret al. 1997; Finocchi & Gail 1997; Finocchi et al. 1997; Gail1998, 2001) include treatments of ice mantle evaporation, dustevaporation, three-body reactions, ionisations by cosmic rays,radionuclides and UV photons. Later, in addition to annealingand carbon combustion the model of Gail (2002) accounts forcarbon dust oxidation.

The approach used by Willacy et al. (1998) and Aikawaet al. (1999) is similar to Bauer et al. but does not accountfor dust destruction processes. Later, Aikawa & Herbst (1999,2001) and Aikawa et al. (2002) added deuterium chemistry totheir chemical model. The influence of the ionisation processessuch as X-rays on the chemistry were introduced by Aikawa &Herbst (1999), and were studied in more detail by Markwicket al. (2002). Furthermore, Willacy & Langer (2000) discussedthe importance of photoprocessing.

Send offprint requests to: M. Ilgner,e-mail: [email protected]

However, summarizing the improvements of recent re-search the models mainly focused attention on upgrading andextending the chemical networks as well as the implementationof additional kinetic processes.

A thorough treatment of physical processes within proto-planetary disks in parallel to chemical simulations was not thefocus of many previous studies. This kind of imbalance wasfinally broken by Willacy et al. (1998) who introduced a 1+1-D-disk model in order to evaluate the chemical evolution basedon that model. In a series of simulations, concerning the chem-ical evolution in accretion disks, the advective mass transportwas taken into account, cf. Duschl et al. (1996), Bauer et al.(1997), Finocchi et al. (1997), Finocchi & Gail (1997), Willacyet al. (1998), Aikawa et al. (1999), Aikawa & Herbst (1999),and Markwick et al. (2002). However, vertical mass transportprocesses caused by diffusion were not considered.

This is the first of a series of papers aimed at dealing withmass transport processes in accretion disks in the frameworkof chemical evolution. The fluid in the model is continuouslyinjected at the outer edge of the disk and the flow (and chem-istry) is followed until the lower radial bound of our calcu-lation, 1 AU. At smaller radii, as the material accretes on tothe central protostar, temperatures are very large and we would

Page 2: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

644 M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows

need to take into account the physical and chemical destructionof the dust grains. It is this steady injection of material whichallows us to adopt a Langrangian description so that our derivedchemical distributions represent the steady state distributions ofmolecules throughout the disk. While the present paper is ded-icated to steady accretion disk flows, the case of a non-steadyaccretion disk flow will be considered in a companion paper(Ilgner & Henning, in preparation).

We briefly summarise the basic concepts of accretion diskphysics and astrochemistry as far as these concepts are neededfor understanding the model developed here. The outline of thepaper is as follows. In Sects. 2 and 3 we explain the physicsand kinetics governing the chemical evolution of accretiondisks. In Sect. 4 different techniques are introduced to modelthe advection transport and the transport by diffusion in areacting flow system. Finally, in Sect. 5, we discuss our resultsand summarise our conclusions in Sect. 6.

2. Disk physics

2.1. Transport in disks

The dynamics of accretion disks are closely related to thetransport of energy and angular momentum. Although thereis still a fundamental problem on the origin of this transportprocess (hydrodynamical vs. magneto-hydrodynamical origin),one has to have a very efficient transport mechanism such asturbulent motion. The idea that disks might be turbulent toproduce the anomalous viscosity seems to be feasible sinceBalbus & Hawley (1991) have given an explanation for the ori-gin of turbulence via the magneto-rotational instability (MRI).Nevertheless, MRI cannot work throughout the entire sincethere are “dead zones” where other processes must operate.

If one wants to study the chemical evolution within the ac-cretion flow one cannot incorporate the latest developments indisk modelling. Despite the improved computational resources,we are not able to perform 3D calculations taking into accounta complete network of chemical reactions. Thus, we are forcedto simplify the description of disk physics including the masstransport processes. For that reason, we have to consider thedevelopment of recent accretion disk theory in order to get anidea about how the physics might be simplified.

As soon as we move to a two-dimensional (2D) descriptionwe are not in a position to track comprehensively the causeswhich lead to turbulence because turbulence is a “rotational”(vorticities), and a three-dimensional phenomenon. Of course,one can infer a high value of the phenomenological turbulentkinematic viscosity νt due to the presence of a high Reynoldsnumber flow which is typically for turbulent flows. In orderto keep the transport properties of a turbulent flow in a 2Dmodel one might be attempted to replace the kinematic vis-cosity ν with the turbulent kinematic viscosity νt within theNavier Stokes equation. However, one has to remember thatturbulence is a dynamical phenomenon and not a feature of thefluid itself. The viscous stress tensor σ′rϕ (cylindrical coordi-nates r, ϕ, and z are assumed) is approximated by means of thestandard α-prescription given by Shakura & Sunyaev (1973),i.e., σ′rϕ =αP, where α and P denote a dimensionless quantity

and the gas pressure, respectively. Finally, a turbulence modeldescribing the turbulent kinematic viscosity νt is explicitly im-posed in the form νt ∼ αcsH (Pringle 1981), where cs and Hdenote the sound speed and a suitably defined disk height, re-spectively.

As stated above, two-dimensional models of accretion disksare qualitatively quite different from the three-dimensionalmodels in terms of the physical interpretation of turbulence.Under the assumption of an axisymmetric (i.e., ∂/∂ϕ = 0) hy-drodynamical 2D description of an accretion disk, we are espe-cially interested in the flow pattern. For instance the results ofKley et al. (1993), who studied the properties of convectivelyunstable disks, can be utilised to get general information aboutthe direction of the flow. As long as Schwarzschild’s criterionis satisfied, the matter inside a convective region sinks at lowtemperatures and rises at high temperatures, respectively. Thevalues of the vertically integrated local mass flow are charac-terised by strong variations. In addition, the vertical stratifica-tion is nonsymmetric with respect to the disk midplane. Thelimiting case that the disk is kept stable against convection re-sults in a laminar flow pattern. It has been demonstrated byKley & Lin (1992) that the variation of the radial inflow veloc-ity with height approaches a steady state. For such a system,the vertically integrated local mass flow reaches an equilibriumvalue and the radial velocities represent inflow.

Finally, one can describe the global transport processeswithin an accretion disk, taking into account only a few of theequations determining the 2D radiation-hydrodynamics. Thegeometry of the disk is considered in the “thin disk” approx-imation, assuming that the thickness of the disk is everywheresmall compared to its radius. Based on the equation of continu-ity, the z-component of the momentum equation, and the bal-ance equation of angular momentum one can derive a parabolicdifferential equation determining the dynamical evolution ofthe disk (Lynden-Bell & Pringle 1974):

∂Σ

∂t=

3r∂

∂r

[r

12∂

∂r

(νtΣr

12

)], (1)

where vr is given by

vr = −3

Σr12

∂r

(νtΣr

12

). (2)

The quantities Σ and vr denote the surface density and theradial drift velocity, respectively. The viscous shear stress isapproximated by the α-prescription of Shakura & Sunyaev(1973) and the motion in azimuthal direction is a Keplerianmotion.

2.2. Steady 1+1-D-disk model

Provided that νtΣ is a function of Σ itself, the steady-statecondition applied to Eq. (1) will only define the quantityνtΣ, not Σ. The steady-state behaviour of Eq. (1) is given byLynden-Bell & Pringle (1974):

νtΣ :=M3π

1 −(R

r

) 12

, (3)

Page 3: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows 645

where R refers to the radius of the star, assuming that the starrotates more slowly than Keplerian at the inner boundary of thedisk. In the steady-state case the evolution is regulated by oneparameter only, the mass accretion rate M.

The mass accretion rate M is constant for each ring of astationary, cylindrically symmetric Keplerian disk. Extractionof energy and angular momentum from test particles resultsin a spiraling inward motion and a constant mass transport to-wards the central object. The loss of angular momentum can beunderstood in terms of shearing stresses. The energy will dissi-pate, and the corresponding term in the dissipation function Φis given by:

Φ = νtΣr2 | ∇Ω |2 (4)

=3

GMM

r3

1 −(

Rr

) 12

,refering to dissipation of energy per unit time and per unit area.Ω denotes the Keplerian value of the angular velocity, G andM the gravitational constant and the mass of the central ob-ject, respectively. The dissipated energy will be radiated locallysomewhere above the disk midplane. Assuming that the radia-tion field is close to the blackbody value, i.e. πB(T ) = σRT 4

eff,one can calculate the radiation flux emergent from one side ofthe disk surface:

σRT 4eff =

38π

GMM

r3

1 −(

Rr

) 12

, (5)

where σR and B refer to the Stefan-Boltzmann constant andthe integrated Planck function. In addition, the energy flux atthe disk midplane (z = 0) vanishes because of reflection sym-metry.

The still unknown upper boundary z∗(r) can be deter-mined by considering the vertical disk structure which is de-termined by a two-point boundary problem. In particular, thez-components of mass density , pressure P, temperature T ,and the radiative flux of energy F of the steady solution of theequations of radiation hydrodynamics have to be calculated upto the disk surface:

dPdz= −gz (6)

dTdz=

− κRF

16σRλT3 for ∇rad ≤ ∇ad

gzTP∇conv for ∇rad > ∇ad

(7)

dFdz=

94νt

GMr3· (8)

The quantities gz, λ, and κR denote the z-component of the ac-celeration of gravity, the fluxlimiter, and the Rosseland meanopacity1. Further we used a compilation given by Semenovet al. (2003), in particular composite dust aggregates witha “normal” chemical composition refering to the iron abun-dance within the silicates, i.e., Fe/(Fe+Mg) = 0.3, were con-sidered. Energy transport by radiation as well as by convec-tion is assumed depending on the local ratio between the

1 See Appendix.

radiative and adiabatic temperature gradients ∇rad and ∇ad, re-spectively. Considering radiation transport, we used a flux lim-ited diffusion approximation of the radiation transport givenby Levermore & Pomraning (1981). Based on this theory themagnitude of the flux is regulated to be no greater than thedensity times the maximum transport speed. Finally, the sys-tem of equations can be closed with the equation of state of anideal gas.

In order to solve the coupled system of three first-order or-dinary differential Eqs. (6–8), three boundary conditions forP(z∗), T (z∗), and F(z∗) are needed. While the value of F canbe obtained from Eq. (5), the boundary values of P and T arelinked to the value of the optical depth τ∗ at z∗. In Appendixwe summarize the radiative transfer problem and the τ∗ = 2/3condition. Concerning the boundary condition for T , we getT (z∗)= Teff. The boundary problem can be reduced to find thephotospheric disk height z∗ by shooting methods. More than100 zones are needed to calculate the vertical disk structure.

The changes in radial structure of the disk are determinedby the viscous time scale. This time scale is much longer thanthe time scales for other relevant physical processes in the disk.Because of the difference between the viscous time scale andthe other time scales, one may decouple the radial and verticaldirection of a 2D model resulting in a 1+1-D-description.We assume implicitly that a static disk atmosphere will beestablished instantaneously during the viscous evolution. Ofcourse, the differences between a 2D and 1+1-D-descriptionstill remain. However, in the steady-state case, one can expectthe 1+1-D-model to represent the 2D-model.

2.3. Model calculations

To keep the underlying assumptions of the disk model valid,one is forced to restrict the inner as well as the outer bound-ary of the computational domain [rmin, rmax]. The disk has to bekept optically thick, i.e., τ(r, z) ≥ 2/3, and stable against self-gravitation. While the instability against self-gravitation can bemeasured by Toomre’s parameter Q (Toomre 1964), the opti-cal thickness at the upper boundary z∗ is fixed by the boundarycondition τ∗ = 2/3. The regions where the disk becomes opti-cally thin can be estimated by means of approximate relationsshown in Bell et al. (1997). The quantity rmax was chosen tobe 10 AU which does not necessarily mean an upper limit con-cerning τ and Q. Depending on the parameters of the system (α,M, opacity table) rmax can be extended by a factor of 2. We alsonote that the inner regions of the disks adjust to a quasi steadystate more quickly than the outer parts. Thus, the steady-stateassumption is easier fullfilled in the inner regions.

Purely for kinetic reasons the inner boundary rmin was fixedat 1 AU. Due to destruction of the dust particles by evapora-tion, the the opacity will change and the mathematical descrip-tion of the kinetics, based on a system of ordinary differentialequations (see Eq. (12)), is no longer suitable. It finally resultsin a system of differential algebraic equations. The only com-putations in this field have been carried out by the Heidelberggroup, e.g., Bauer et al. (1997).

Page 4: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

646 M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows

Fig. 1. The radial distribution of the photospheric temperatures (dottedline) in comparison to the model by D’Alessio et al. (1998) – (dashedline – for M= 10−8 M/yr, α= 0.01, and the opacity table derived byPollack et al. (1994). In addition, the temperature along the “super-heated” surface layer and the disk interior temperature after Chiang &Goldreich (1997) are plotted.

The focus of our disk model is not to improve existing diskmodels describing the vertical structure. In fact, we balancedthe recent developments of more sophisticated disk modelsagainst simpler models in order to include the main features.Our disk model is quite similiar to that of Bell et al. (1997);differences are additional heating sources. Ionization by ener-getic particles and the radioactive decay of 26Al were taken intoaccount (see e.g. D’Alessio et al. 1998).

Our model differs from the model of D’Alessio et al.concerning the influence of the stellar radiation on the diskstructure. In this study we did not consider stellar heatingsources. When we compare the physical parameters of ournon-irradiative disk model with that of the irradiative modelof D’Alessio for the regions between 1 and 10 AU, we findthat the differences in temperature, density and optical depthare not large (see Fig. 1) and are similiar to the differenceswhich are caused by, for example, adopting a different grainmodel and therefore different opacity model. That is to say,in the inner disk interior to 10 AU, viscous dissipation ratherthan radiation dominates the heating. The largest differencescan be found in the radial distribution of the photospheric tem-perature T (z∗), see Fig. 1. In addition, the photospheric heightz∗ is always below the height zs where the stellar radiation isdeposited. The temperature inversion caused by the stellar ra-diation field occurs only above the photospheric height z∗. Aninversion of temperature occurs in the simple disk model ofChiang & Goldreich (1997). Although this model does not ac-count for transport processes, it completes the disk models usedin astrochemical modelling of accretion disks, e.g., Aikawaet al. (2002) (D’Alessio disk model), Willacy & Langer (2000)(Chiang-Goldreich disk model). In the case of the radial tem-perature distribution shown in Fig. 1, we also plot the tempera-ture distribution of the “superheated” surface layer Tds and theinterior disk temperature Tint applying the description given byChiang & Goldreich. The stellar parameters are the same as thecorresponding parameters used in the dynamical disk models.

3. Thermodynamics

Chemical reactions as well as diffusion processes belong tothe class of irreversible processes. They are macroscopicallydescribed by non-equilibrium thermodynamics. Irreversibletransformations are distinguishable from reversible transforma-tions by the state function S , the entropy of the system. A localdescription of S requires a continuous function of space coordi-

nates, i.e., S =∫ VsdV , where and s denote the mass density

and the entropy per unit mass, respectively. In particular, therate of entropy change inside the system, i.e., diS/dt=

∫ VσdV ,

where σ denotes the entropy source strength or entropy pro-duction per unit volume and unit time, is related to irreversibleprocesses inside the volume element V . The entropy productionis always a non-negative quantity, i.e., σ ≥ 0, and the entropyproduction vanishes only for reversible processes.

A macroscopic expression for σ can be obtained byσ =

∑i JiXi, where Ji and Xi are the Cartesian components of

the independent thermodynamic fluxes and forces. The fluxesand forces are only up to 2nd tensorial order and, therefore, theentropy production σ can be split into four different productsof the fluxes and forces. The product of Ji and Xi can be deter-mined by scalar quantities as well as by (polar and axial) vec-tors, and by two symmetric tensors with zero trace. However,the forces and fluxes of different tensorial character do not cou-ple (Curie Principle). According to this classification, the en-tropy production σ caused by chemical reactions is related toscalar phenomena in contrast to vectorial phenomena like dif-fusion.

In the following two subsections, we introduce the phe-nomenological equations in order to get information about thethermodynamic forces driving chemical reactions and diffu-sion, respectively. In a first approximation, the phenomenolog-ical equations can be written as linear relations between thethermodynamical forces and fluxes

Ji =

n∑j=1

Li jX j, (9)

where the components of the tensor Li j are called phenomeno-logical coefficients. Under the assumption of the existenceof spatial symmetry properties one can demonstrate that theCartesian components of the fluxes depend only on a few ofthe components of the forces (Curie principle).

3.1. Chemical reactions

Within thermodynamics, chemical reactions are attributed to aclass of phenomenological equations which account for scalarfluxes and scalar forces. The scalar flux of a fixed chemicalreaction is given by

Ji = −1T

r∑j=1

Li jA j −1

3TLiv∇ · u (i = 1, . . . , r), (10)

where T is the temperature and A j refers to the chemical affinityof the reaction j introduced by

Ai =

n∑j=1

µ jν ji.

Page 5: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows 647

The quantities µ j denote the chemical potential of the compo-nent j; the quantity ν ji divided by the molecular mass Mi is pro-portional to the stoichiometric coefficient of the component i inreaction j. By means of the general form of the phenomeno-logical Eq. (9), one can specify the thermodynamical forcesXi =−Ai/T . The second term in Eq. (10) corresponds to cross-phenomena because the scalar flux may be caused by chemistryas well as by bulk viscosity. This phenomenon must be takeninto account in the case of an incompressible fluid or a fluidwith vanishing volume viscosity (Maxwell fluid).

The rate of change of the mass of component i is deter-mined by the scalar thermodynamical flux. In terms of the mo-lar concentration Ni =i/Mi (where i and Mi denote the massdensity and the molecular mass of component i, respectively)the law of mass action is given by

∂tNi =

r∑j=1

νi jJ j (i = 1, . . . , n), (11)

neglecting any macroscopic transport. Because of the prefereddescription, using molar concentrations, the quantity νi j is the

stoichometric coefficient and J j the rate of the chemical reac-tion j. With respect to chemical kinetics Eq. (11) can be writ-ten as

∂tNi =

r∑j=1

νi j

k+jm∏

l=1

N−νl j

l − k−j

n∏l=m+1

Nνl j

l

, (12)

where k+j and k−j refer to the rate constants for the forward andbackward reactions, respectively.

3.2. Diffusion

Diffusion is a vectorial phenomena. In general, the diffusionflow J

ai of the component i

Jai = Ni

(ui − ua

)(13)

is defined with respect to various reference velocitiesua =

∑aiui where ai refers to the normalized weights of the

component velocities ui.Only the most simple case will be considered here, neglect-

ing external forces and vectorial cross-phenomena. The ther-modynamical force is determined by Xi = (∇µi)

∣∣∣T,P/T where

µi denotes the chemical potential of component i. Because ofthe linear relations between the fluxes and thermodynamicalforces in Eq. (9), the diffusion flow of component i, which isgiven by

Jai = −

n−1∑j=1

Di j∇Nj, (14)

does not only depend on the gradient of the chemical potentialof component i. Therefore, we have to deal with a tensor Dikof (n − 1)2 diffusion coefficients. Only in the special case ofdiffusion in a binary system (n= 2), the tensor can be reducedto a scalar obtaining

Ja1 = −D ∇N1

D ≡ La11

N1N2

N1

a221T

∂µ1

∂N1

·

In addition, the relation between the diffusion coefficient Dand the phenomenological coefficient La

11 is shown in the lastformula.

In order to achieve a simple model of diffusion in a multi-component system, we are interested in decoupling the thermo-dynamical forces and fluxes. Decoupling can be forced if theentire system is approximated by the superposition of n self-diffusion models. Every self-diffusion model combines one car-rier material and only one tracer material assuming that bothcomponents can be characterized by the same dynamical prop-erties. Finally, the change of mass of component i caused bydiffusion results in

∂tNi = −∇ · J

ai (15)

= ∇ · (D ∇Ni

),

where D denotes the binary diffusion coefficient. However, thisform of the diffusion law, conventionally used in the kinetictheory of gases, is slightly different from, e.g., the descriptionwhich can be found in Xie et al. (1995). The different descrip-tions come from the chosen reference velocity ua and the nor-malized weights ai, respectively. The description given by Xieet al. refers to “relative diffusion flows” where the referencevelocity is measured with respect to the nth component.

Here we assume that the disk is in a turbulent state. Wenow have to specify the diffusion coefficient D. A viscous fluidwill support stresses which are not only directed normal to thesurface of the fluid element. Normal stresses as well as tan-gential stresses are represented by the viscous stress tensor σ′i j.The viscous stress tensor is part of the momentum flux den-sity tensor Πi j which accounts for the rate of momentum flowthrough the boundary. Remembering the momentum equation,one may interpret viscous friction by an exchange of momen-tum. Because of the structure of the momentum equation onecan draw an analogy between the exchange of momentum anda diffusion process where the kinematic viscosity ν acts as adiffusion coefficient.

The kinematic viscosity ν is now replaced by the turbu-lent kinematic viscosity νt. By means of the Schmidt num-ber Sc = νt/D, measuring quantitative difference betweenthe transport of material and momentum, one can determinethe diffusion coefficient of Eq. (15). For example, the Schmidtnumber is less than unity if the spreading rate of material isgreater than that of momentum.

However, the diffusion coefficient D is influenced by dustparticles mixed within the gas. Therefore, the coupling be-tween the gas and dust particles implies that one should de-fine an effective diffusion coefficient. However, recent stud-ies by Schrapler & Henning (in preparation) have shown thatthis effect on the diffusivity becomes important only for dustsizes above 0.01 cm, depending on the local position withinthe disk (r ≤ 10AU). Therefore, we can neglect the role ofdust and take the Schmidt number in our calculations to beSc= 1. Experimental data typically account for Schmidt num-bers Sc ≈ 0.7 (cf. Spalding 1971); our estimate of Sc = 1 willunderestimate the efficiency of turbulent material transport dueto mixing. Previous calculations of mixing in disks have alsoadopted Sc=1 (Gail 2001; Wehrstedt & Gail 2002).

Page 6: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

648 M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows

The eddy diffusion profile can be obtained by the diskmodel and the turbulence model, respectively. Assuming theα-turbulence model, the local value of the turbulent viscosityνt =αcs min[HP,H] is determined by the model itself where csdenotes the sound speed, HP the pressure scale height, and Hthe disk height. Therefore, a priori estimates of the turbulentviscosity are not needed, in contrast to the eddy diffusionprofiles in molecular clouds, cf. Rawlings & Hartquist (1997).

4. Modelling of a reacting flow system

In order to evaluate the importance of transport processes onthe chemical disk evolution, it may be useful to start with afew preliminary remarks. It is obvious that the modeling ofchemical disk evolution within a protoplanetary disk becomessimpler if one excludes mass transport processes. Remarkably,since the first studies concerning the chemical disk evolution,this simplified description is still quite popular, although mod-els including mass transport processes are already available,e.g., Duschl et al. (1996), Willacy et al. (1998), Aikawa et al.(1999). In addition, we have to point out that models with-out transport processes discussed by Aikawa & Herbst (1999,2001) and Willacy & Langer (2000) refer to a very specificapplication of disk models. While Willacy and Langer appliedthe Chiang-Goldreich model, Aikawa and Herbst utilized theso-called minimum-mass solar nebula model (Hayashi et al.1985). However, a protoplanetary disk has to include the accre-tion process onto the central object and, therefore, the accretionflow within the disk has be taken into account.

Transport processes can be distinguished into transport dueto advection and transport due to diffusion. This kind of dis-tinction can be introduced by applying Eq. (13) to the equationof continuity, which for component i is:

∂tNi + ∇ ·

(Niu

a) = −∇ · Jai + Ri , (16)

where Ri denotes the rate term for production and destructionof component i, respectively. The source term on the left handside, ∇ ·

(Niu

a), refers to advective transport while diffusive

transport is represented by the source term on the right handside, ∇ · Ja

i .

As far as the kinetics is determined by non-equilibriumthermodynamics, simulations of the chemical disk evolutionhave not yet been performed in the Eulerian description.Simulations, assuming chemical equilibrium, have beencarried out in order to study the influence of annealing andcarbon combustion on the disk evolution, see Gail (2001). Ina recent study, Gail (2002) extended his model, to include arestricted chemical network of 18 species in the gas phase.In this study only vertical mixing of chemical species isaccount for diffusion processes; diffusion in radial directionwould need a different formulation than we used, and wouldrequire a fully two-dimensional reaction-diffusion model.However, such work is in progress.

4.1. Advection

In this subsection we focus only on the advective transport ofchemical species radially. As long as the turbulent mixing ofspecies is excluded, the computational domain can be splittedinto a set of horizontal layers. Once vertical mixing is takeninto account, the vertical grid is needed.

Instead of the Eulerian description, the Lagrangian de-scription will be used here to study the time variation of theLagrangian fluid element. The Eulerian description can betransformed into the Lagrangian description by means of thesubstantial derivative (d/dt) which is simply related to the timederivative (∂/∂t) computed at fixed gridpoints in space via

ddt=∂

∂t+ (u∇) .

Thus, the advective term on the left hand side of Eq. (16) canbe replaced by Ni∇ · ua.

As yet, the coupling between a particular fluid parcel fol-lowing the motion of that parcel and the chemical evolution isonly apparent from semi-analytical 1D disk models. Because ofthe algebraic nature of equations for the disk structure, the ki-netics can be attached quite easily to the parcel of gas within theaccretion flow as has been done by Duschl et al. (1996), Baueret al. (1997), Finocchi et al. (1997), Finocchi & Gail (1997),and Aikawa et al. (1999). However, with respect to the numer-ical scheme, one should exercise caution when integrating thesystem of stiff differential equations which is mainly done byimplicite multi-step techniques such as the BDF method (cf.Strehmel & Weiner 1995). In addition to the stability of the nu-merical scheme, implicite multi-step techniques take advantageof increasing step sizes in time, which are in general many or-ders of magnitude longer than the initial step size. Consideringonly one single integration step, the shift between the previouspoint tn and the new point tn+1 can become quite large. Thismeans that important physical processes, taking place mean-while, may not be resolved because of the numerically deter-mined step size control and order selection.

Because of the non-algebraic nature of equations, deter-mining the 1+1-D-disk structure, the chemical and the physicaldisk evolution along the streamline cannot be solved simulta-neously as it was done in the case of semianalytical 1D-diskmodels. Therefore, some kind of reasonable decoupling of ki-netics and advective transport is needed.

Returning to the extended discussion concerning the im-plementation of a Lagrangian fluid element in a 1D-chemicalmodel, we are able to specify a condition for decoupling quiteeasily. Adopting a cylindrical coordinate system, the spiral mo-tion of the Lagrangian fluid element towards the central objectcan be analysed in terms of the vectorial components of theaccretion flow. Concerning the azimuthal component, the mo-tion is assumed to be on Keplerian orbits. It can be shown thatthe azimuthal component of the velocity is highly supersonicwhile the value of the radial component is much less than thesound speed. Therefore, it seems reasonable to decompose thetrajectory into a Keplerian orbit which is possessed of a smallradial drift. In the case of steady 1+1-D-α-disk models the ra-dial component of the velocity is always directed inwards.

Page 7: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows 649

Fig. 2. Typical decomposition of a trajectory within a steady 1+1-D-α-disk model from 10 to 1 AU; mod1ELIS. More than 366 gridpointsare needed to construct the grid. The change in the physical quanti-ties moving from one gridpoint to the neighboring gridpoint is 9% onaverage.

Following the trajectory within two coaxial cylinders withradii r1, r2 we impose the condition that the variation of the ra-dial component of the accretion flow vr vanishes vertically. Thisassumption is in agreement with the thin-disk approximationdetermining vr by Eq. (2). Because of the assumed Keplerianmotion, we have to take into account only the radial compo-nent of the accretion flow. As long as the physical conditionswithin the two radii are quite similiar there is no need to sustainthe coupling of kinetics and advective transport. In particular,those physical quantities important in influencing the chemicalevolution are the temperature T (r, z), the mass density (r, z),and the optical depth τ(r, z). The decoupling mechanism wasapplied on a distance ∆r=r2− r1 where the local change withinthe T , , and τ profiles is less than 10%. For example, the de-composed trajectory of the simulation mod1ELIS is shown inFig. 2. After calculating the time t needed to follow the trans-port from r1 to r2, we first solved the kinetic equations withinthe vertical domain at r1 for ∆t = t. Second, the fluid elementwas transported by advection to r2 taking the equation of con-tinuity into account.

We calculated the chemical evolution along three differenttrajectories from 10 AU to 1 AU with the parameters given inTable 1. In the last plot of Fig. 3, the time needed for advec-tive transport is plotted vs. the radial position. As we shall see,the results presented in the next section can be classified ac-cording to local disk structure and trajectory. With respect tothe chemical evolution comparisons can be made of i) similiarlocal disk structure and different trajectories and ii) different lo-cal disk structure and similiar trajectories. The latter case refersto simulations mod2ELIS and mod3ELIS while mod1ELIS andmod3ELIS represent the former.

4.2. Diffusion

The previous discussion might lead one to separate the ki-netics and vertical transport processes. However, there is asubtle difference concerning the transport properties. In a

Table 1. Parameter of the presented simulations determining the vis-cous evolution of the disk.

simulation M [M/yr] α

mod1ELIS 1×10−7 1×10−2

mod2ELIS 5×10−8 1×10−2

mod3ELIS 1×10−7 5×10−3

multi-component system, the various species will move due todiffusion at different diffusion velocities, cf. Eqs. (13) and (14).In particular, the bulk velocity of the mixture is determined bythe local composition of the mixture itself. Splitting techniquessuch as operator splitting were applied, e.g., by Rawlings &Hartquist (1997), in order to solve the reaction-diffusion model.However, the problems related to splitting strategies are appar-ent from the lack of an uniform time scale, independent of thekinetics.

Instead of decoupling of kinetics and diffusion one canintegrate Eq. (16) simultaneously. The discretised equationsare given by semi-discretisation of the vertical components ofEq. (16) adapted for a Lagrangian description. Finally, the sys-tem of n partial differential equations was transformed into a setof n × nz ordinary differential equations where n and nz denotethe number of species and the number of vertical gridpoints,respectively. Although the dimension of the system may eas-ily become quite large, the system itself is characterized by asparse matrix. Depending on the used discretisation scheme forapproximating the spatial derivatives, the system is determinedby a banding structure at least.

Because of the global transport processes within areaction-diffusion model, boundary conditions have to befixed. Apparently the condition refering to the disk’s midplane(z = 0) is determined by the geometry of the disk itself.Furthermore, we assumed that the gradients in the species’abundances vanish beyond the photospheric height z∗.

4.3. Chemical network

The disk chemical model is based on that of Markwick et al.(2002). The species set consists of a total of n = 242 speciesincluding 69 mantle components2 made up of the elements H,He, C, N, O, S, Si, Fe and Mg. About 2500 reactions connectthese species.

The model contains various types of reactions in the gasphase: ion-neutral and neutral-neutral reactions as well as pho-toprocesses. The interaction between the species in the gasphase and the chemically reactive mantle components ontograin surfaces is described by adsorption and desorption mech-anisms. Adsorption is caused by freezing of species ontograins. The reverse processes, thermal desorption and cosmicray heating, desorb species from grain surfaces.

Expressions for the rate coefficients k±, refering to re-actions in the gas phase as well as to adsorption processes

2 Here, the so-called “two-phase” approach was used, taking onlythe two reactive phases into account (gas phase and surface phase).

Page 8: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

650 M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows

Fig. 3. Temperature distribution of a steady 1+1-D-α-disk model refering to 3 different simulations mod1ELIS, mod2ELIS, and mod3ELIS(clockwise from top left). The temperature contours for 100 K and 500 K are shown. The figure on the right bottom refers to the trajectories ofeach simulation within the computational domain.

onto grains, can be obtained by macroscopic considerations.If σc and u denote the cross-section and the velocity of themolecules, respectively, the mean collision rate coefficient isgiven by<σcu>. Once averaged over a Maxwellian velocity dis-tribution, one gets the Arrhenius type of rate coefficients whichis typically used in astrochemical databases. For the adsorp-tion process, one has additionally to take the sticking probabil-ity into account which depends on the charge of the adsorbingparticle, cf. Umebayashi & Nakano (1980).

The mobility of the adsorbed species on the surface de-pends on the surface structure. Therefore, in addition to bi-molecular processes one has to consider other microscopic ef-fects such as the characteristic (lattice)-vibration frequency inorder to describe the rate coefficients related to desorption pro-cesses on grains. With respect to thermal desorption as well asto cosmic ray heating the kinetic equations are a variation ofthe Polanyi-Wigner equation, cf. Hasegawa et al. (1992) andHasegawa & Herbst (1993).

Sources of ionisation in the disk include energetic particlesformed by the radioactive decay of 26Al, interstellar UVphotons, and ionisation due to cosmic rays. The rate of thelatter process depends on the local value of the surface densityΣ(r, z), cf. Umebayashi & Nakano (1981).

5. Results

The chemical evolution was evaluated along the trajectory from10 AU to 1 AU. The decomposition of the trajectory needed todecouple kinetics and advective transport requires more than366 gridpoints radially (mod1ELIS, cf. Fig. 2). The number ofgridpoints in the other simulations mod2ELIS and mod3ELIShave the same order of magnitude. The computational domainwas divided vertically into an equidistant grid with nz = 40.As soon as vertical mixing was included, the Crank-Nicholsenscheme was applied to discretise the spatial derivatives. Finally,the vode-integrator developed by Brown et al. (1988) was usedto solve the system of nonlinear ordinary differential equations.The initial abundances of the species at 10 AU were taken fromWillacy et al. (1998).

The presentation of our results will be made on three differ-ent levels in order to highlight different aspects of the chemicalevolution in disks. On the first level, we will analyse the dataobtained by a simulation without vertical mixing. To some ex-tent we will use this to identify the main formation and destruc-tion processes of particular species. Once the chemical evolu-tion is determined by a reaction-diffusion model, it is easierto draw conclusions concerning the influence of mixing pro-cesses on the chemical evolution (second level). Finally, wewill clarify the influence of the disk dynamics on the chemical

Page 9: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows 651

Fig. 4. Chemical evolution of CS and SO2 along the trajectory from 10 AU to 1 AU; mod1ELIS without vertical mixing. The contour lines referto the abundance values relative to the total amount of hydrogen: 10−10,−13,−16 and additionally 10−7 for SO2.

evolution. On this third level, we analyse the chemical evolu-tion with respect to different disk parameters determining thedynamics of mod1ELIS-mod3ELIS, while the disk parametershave been kept constant for the previous discussion.

At each level, we will consider the chemical evolution ofthe CS molecule. Because of its large dipole moment, CS is ex-pected to trace high-density regions. Based on measurements ofhigher rotational CS transitions one might get better constraintson the chemical and physical structure of disks. In addition,since CS is detected in the atmospheres of comets – cf., e.g.,Snyder et al. (2001) – one may be able to compare the earlyevolutionary stages of protostellar disks and the early stagesof the Solar Nebula. The ongoing use of CS as a diagnosticof different star-forming regions should be seen against thisbackground – cf., e.g., Blake et al. (1992), Dutrey et al. (1997),Launhardt et al. (1998), Qi (2000), and Doty et al. (2002).

The analysis of CS cannot be decoupled from the chemistryof other sulphur-bearing molecules. Therefore, the evolutionof other species closely connected to the evolution of CS suchas SO, SO2, HCS+, HCO+, and NH3 has to be discussed aswell.

5.1. Chemical evolution based on a pure kinetic model

We start the discussion with the question of where most of thesulphur is locked. Sulphur-bearing molecules may stick on thedust surfaces due to adsorption processes. In particular, sulphuris mainly channSte eled into mantle components in low tem-peratures regions. Note that in general more than 50% of thereservoir of sulphur is assumed to be bound within the graincore itself, especially in Troilite (FeS), cf. Pollack et al. (1994)and Finocchi et al. (1997). For increasing temperatures, one ob-tains a rapid rise of sulphur-bearing molecules in the gas phase.Now most of the sulphur is locked in CS and SO2 molecules,cf. Fig. 4.

The conclusions drawn by Doty et al. (2002), Willacy et al.(1998), and Langer et al. (2000) are confirmed by our calcula-tions. Doty et al. emphasise that sulphur is mainly channeledinto SO2 in regions with T ≈ 100–400 K while CS serves asa good measure of sulphur in the cool exterior regions. The

simulations of Willacy et al., which take only the evolutionalong the disk midplane into account, have shown that SO2 isthe dominant sulphur-bearing species. In addition, because ofthe Lagrangian description, we may characterise the high abun-dance of CS by a disk chronology analogy (Langer et al. 2000).They argued that an increased abundance of CS is associatedwith early states of chemical evolution.

We will now follow the trajectory from 10 to 1 AU, whichcrosses regions with different physical conditions. At first therelative abundance of CS decreases while following the ac-cretion flow toward inner regions. In particular, in the coldertemperature regions, close to the photospheric disk height, theamount of available sulphur in the gas phase drops due toadsorption onto the grain surfaces. For the vertically adjacentregions, the temperature will increase. Thus, the efficiency ofreactions in the gas phase rises. Compared to the initial abun-dance at 10 AU carbon monosulfide becomes more abundantin regions represented by the contour lines of 10−16 and 10−13,cf. Fig. 4. The major routes to CS in the disk are related to theevaporation of H2CS and H2S from grain mantles. Reaction ofthese molecules with C+ form CS and HCS+, respectively. Inaddition, proton transfer reactions with H2CS followed by thedissociative recombination of H3CS+ also produces CS. In theinner region of the disk, the most important reactions involveHCS+

HCS+ + e− −→ CS + H. (R 1)

Because of the inverse temperature dependence of its rate co-efficient (k ∝ T−3/4), reaction (R 1) gives rise to higher abun-dances of CS especially in colder regions. Of course, an en-hanced abundance of HCS+ needed to produce CS can be foundin these regions, cf. Fig. 5. The destruction of CS via protona-tion of HCO+

HCO+ + CS −→ HCS+ + CO (R 2)

cannot compensate the production of CS via reaction (R 1).Once one passes into regions with temperatures above 100

K, desorption processes become more efficient and an enrichedgas phase chemistry can be expected. An increased abundanceof CS within the contour line of 10−10 (cf. Fig. 4) is attributed

Page 10: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

652 M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows

to the evaporation of mantle molecules back to the gas phase.Of course, the adsorbed counterpart of CS, CSads., becomes lessabundant, cf. Fig. 5.

In addition to desorption, the change in the CS abundancecan be explained by differences within the gas phase chem-istry. The destruction of CS is influenced by oxygen chem-istry. Because of the enhanced abundance of atomic oxygen(cf. Fig. 5, region between 3 and 5 AU) CS is efficiently re-moved by

CS + O −→ CO + S (R 3)

returning sulphur to atomic form. Similar dependencies werereported by Charnley (1997) and Doty et al. (2002), althoughdifferent environments were considered. However, most of theavailable sulphur in the gas phase is now locked in SO2. Theincreased abundance of CS for r ≤ 2 AU is related to the en-hanced abundances of ammonia, NH3, which has a very largeproton affinity. The proton transfer reaction between ammoniaand the thioformyl cation, HCS+,

NH3 + HCS+ −→ CS + NH+4 (R 4)

reforms CS. Note that the corresponding rate coefficient isindependent of the temperature. It is remarkable that ammoniacan contribute significantly to the increased abundance ofCS although the coupling between nitrogen and sulphurchemistries, through NS, is weak. The implications of thismay be of particular interest to the unexpected distribution ofCS and NH3 seen in the observations of a few star-formingregions, cf. Taylor et al. (1996).

5.2. Chemical evolution based on a reaction-diffusionmodel

The previous model demonstrated that vertical abundance gra-dients of the gas phase species develop. Therefore, the gradi-ents of the chemical potentials cause a diffusion flow.The simulations in this subsection differ from the previoussimulations only in respect to the inclusion of the reaction-diffusion process. Neither the chemical network including theinitial abundances nor the disk parameters were changed.

At the beginning, the chemical evolution along the trajec-tory is quite similiar to that with vertical mixing neglected.In the particular case of CS, differences in the vertical dis-tribution become significant only for r ≤ 8.5 AU, cf. Fig. 6.For r > 8.5 AU diffusion processes cannot really influence thechemical evolution significantly because of the low abundancesof molecules in the gas phase. Nevertheless, the already estab-lished gradients are removed by diffusion. Apart from the gra-dients, the diffusion velocity also depends on the local valueof the diffusion coefficient. Concerning CS, one expects to ob-serve the most effective mixing near the disk midplane, thustransporting CS-molecules to regions above the disk midplane.Here, because thermal adsorption processes are favoured at lowtemperatures, the molecules will freeze out onto grain surfaces.Therefore, the contour lines of the adsorbed counterpart CSads.are shifted to higher values of z, cf. Fig. 6. There is no en-richment of CS between 8 and 9 AU as was found in the pure

kinetic model, cf. Fig. 5. In comparision to the results of theprevious level, less CS molecules were adsorbed onto grainsin these regions, and thus less mantle material is available fordesorption.

Between 4 and 6 AU the destruction of CS in the gas phasevia reaction (R 3) becomes more effective because of the in-creased abundance of atomic oxygen, cf. Fig. 7. The oxygenitself is formed mainly by the reaction

N + NO −→ N2 + O. (R 5)

Unlike the previous model, the destruction of CS takes place inthe inner disk regions although there is ammonia available atr≤1.5 AU to recycle CS via reaction (R 4). However, becauseof the low HCS+ abundance, cf. Fig. 7, the formation rates arenot sufficient to overcome the loss rates with atomic oxygen.With respect to a chronological classification one may argue,due to the earlier appearance of the peak in CS, that CS andNH3 are associated with earlier and later stages of the chemi-cal evolution, respectively. Langer et al. (2000) have drawn thesame conclusions.

Finally, relating to the previous model, we discuss whichsulphur-bearing species in the gas phase contributes most tothe entire reservoir of sulphur. Compared to the results of theprevious model, cf. Fig. 4, SO2 seems to be already a goodmeasure of sulphur abundance in the gas phase at larger radii.Additionally, the abundance of SO2 molecules drops in the veryinner regions in sharp contrast to the previous model becauseof the low abundance of the OH radical which is more easilyforced into H2O than into SO2 through the reaction

SO + OH −→ SO2 + H. (R 6)

On the other hand, the abundance of atomic sulphur does notdecrease due to the contributions of the reactions (R 3) and

N + SO −→ NO + S (R 7)

N + NS −→ N2 + S. (R 8)

5.3. Chemical evolution along different trajectories

The previous results clearly demonstrate mixing processes doinfluence the chemical evolution. Of course, the local changein the composition of species due to diffusion may have an im-pact on the chemical evolution globally since the disk mate-rial is transported by advection. Given that we have a reaction-diffusion model to determine the kinetics, we now focus on howthe advective mass transport acts on the chemical evolution ofthe disk.

Because of the steady-state assumption, the dynamics of aprotoplanetary accretion disk are represented only by α and M.Therefore, discussions of the disk dynamics become only rea-sonable if we compare different steady-state solutions in termsof the parameters M and α. The trajectories as well as the localdisk structures referring to three different steady-state solutionsare shown in Fig. 3.

Page 11: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows 653

Fig. 5. Chemical evolution of HCS+, O, S, NH3, CSads., and HCO+ (clockwise from top left) along the trajectory from 10 to 1 AU; mod1ELISwithout vertical mixing. The contour lines refer to abundance values relative to the total amount of hydrogen of 10−20,−21,−22 (HCS+), 10−13.5 (O),10−15,−16 (HCO+), 10−15,−17 (S), 10−15,−16,−18 (CSads), and 10−5,−6 (NH3).

We now compare the models mod1ELIS and mod3ELIS (seeFig. 8). They can be classified by different trajectories, but quitesimilar disk structures. We find no significant differences inthe chemical evolution of CS. Although the vertical distribu-tion of CS is slightly larger in mod3ELIS than in mod1ELIS,this reflects the fact that the vertical extent of the former diskis larger than that of the latter. In particular, the main forma-tion and destruction processes along the trajectories are identi-cal. However, one obtains a gradual shift in the abundances of,e.g., SO, SO2, and S. Although the evolution of these moleculesfollows the same chronological sequence in both simulations,there is still a change in the abundances in relative high temper-ature regions as well as in regions with lower temperatures. The

change of the position within the contour lines does not followthe physical conditions such as, e.g., the temperature profile.Therefore, different aspects have to be taken into account suchas, e.g., minor differences in the temperature structure as wellas different transport velocities in general.

Let us consider, for instance, the contour line of SO refer-ring to the value of 10−13 relative to the total amount of hy-drogen. Compared to the result of mod1ELIS, the contour lineof mod3ELIS is shifted to larger radii (Figs. 6 and 8). The lo-cal variation in the abundances can be best explained by timedependent chemical effects rather than by minor differences inthe efficiency of adsorption and desorption processes due to

Page 12: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

654 M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows

Fig. 6. Chemical evolution of CS, SO, SO2, and S (clockwise from top left) along the trajectory from 10 to 1 AU; mod1ELIS incl. verticalmixing. The contour lines refer to abundance values relative to the total amount of hydrogen of 10−13,−16 and 10−10 for SO2 additionally.

slightly different temperature profiles3. That is, the time neededfor transport between 10 AU and 8.5 AU along the trajectoryof mod1ELIS is about 20 000 yrs while the radial transport ofmod3ELIS operates on a timescale roughly twice as long, cf.Fig. 3. Therefore, in the case of mod1ELIS less SO mantle ma-terial can be ejected into the gas phase.

In the outer regions of the computational domain, a smallvariation within the temperature profiles can only account fordifferences in the abundances if the thermal desorption is af-fected sensitively by local changes in the temperature. This, ofcourse, depends on the species-dependent value of binding en-ergies, and therefore species with high binding energies, suchas SO2, are expected to show temperature-dependent variationsin their abundances in those regions. Additionally, in terms ofthe SO2 sublimation temperature of 90 K (Wutz et al. 1992) theshift within the contour lines, e.g., of 10−10, cf. Figs. 6 and 8,respectively, follows exactly the variation of the temperatureprofile.

It was already pointed out that the chemical evolutionbased on the simulations mod1ELIS and mod3ELIS can bedescribed by the same sequence of main formation and de-struction processes. However, important differences within the

3 Here, with respect to the binding energy of SO ∆H=2000/kB[K]temperature-dependent effects are negligible.

relative abundances can still be found, especially in regionswith r ≤ 5 AU. The different values in the abundances mainlyoriginate from different timescales which the radial transportoperates on. As long as one compares the chemical evolutionwithin the same temper ature region, the evolution is quite sim-iliar. For example, the destruction of SO which occurs in bothsimulations for T ≥ 300 K is illustrative of this particular be-haviour.

The chemical evolution of SO is intimately connected tothe oxygen chemistry and, in particular, with the distributionof reactive oxygen. The “destruction” of SO for r ≤ 3 AU ismainly caused by missing production routes such as

S + OH −→ SO + H (R 9)

and

S + O2 −→ SO + O. (R 10)

While the abundance of atomic sulphur remains approximatelyconstant, most of the hydroxyl will be channeled into water via

OH + H2 −→ H2O + H (R 11)

at higher temperatures. For the temperature region T ≥ 300 K,the formation of OH is mainly due to the reaction

O + H2 −→ OH + H. (R 12)

Page 13: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows 655

Fig. 7. Chemical evolution of CSads, O, HCS+, and NS (clockwise from top left) along the trajectory from 10 to 1 AU; mod1ELIS incl. verticalmixing. The contour lines refer to abundance values relative to the total amount of hydrogen of 10−15,−16 (CSads), 10−13.5 (O), 10−22.5 (NS), and10−20,−21,−22 (HCS+).

Consequently, significant production of molecular oxygen inthe gas phase via the reaction

O + OH −→ O2 + H. (R 13)

is inhibited since the amount of O and OH is limited at increas-ingly higher temperatures.

Although the kinetics interacts with advective and diffusivemass transport processes the chemical evolution may be gov-erned by local properties like the temperature structure. Theexplanation for the destruction of SO given above holds forboth simulations mod1ELIS and mod3ELIS. The quite similiardistributions of O, O2, and OH along the trajectories, cf. Fig. 9,may be explained in terms of the same chemical signature de-termined mainly by the temperature structure. Therefore, in ad-dition to the importance of global transport processes in diskson the chemical evolution, we should never underestimate theimportance of the parameters of the kinetics.

Continuing the discussion, we now refer to the simulationsmod2ELIS and mod3ELISwhich have quite different local diskstructure but with approximately similar transport by advection(cf. Figs. 8 and 10). Because of the minor differences betweenthe time scales of advective transport below t ≤ 70 000 yrs,cf. Fig. 3, the differences within the local disk structures ofmod2ELIS and mod3ELIS beyond r≥7 AU account for the dif-ferent chemical evolution. This region is of special interest asreactions in the gas phase begin to dominate the chemical evo-lution in mod3ELISwhile the gas-grain interactions still remain

the driving source in mod2ELIS. However, peaks in the abun-dances of species in the gas phase, signifying the contributionsdue to thermal desorption processes, are linked to temperatureregions where these processes become effective.

It is remarkable that with respect to the late stages, that isin the inner disk, the chemical evolution based on mod2ELISand mod3ELIS, respectively, is quite similar and no additionalreactions have to be taken into account to explain the chemicalevolution.

6. Conclusion

We have investigated the chemical evolution of a protoplane-tary disk. For this study, we used the 1+1-D-approximation ofcylindrical steady disk models. Using a Lagrangian description,the change in the molecular composition was evaluated alongdifferent trajectories. Besides advective transport we also con-sidered transport processes by diffusion.

As these models have been prepared for the study of a re-acting flow system, we analysed the underlying physical trans-port properties in detail in order to estimate where coupling ofmass transport processes and kinetics is needed. Modeling thevertical mixing requires a reaction-diffusion model while in thecase of the advective transport decoupling seems to be an ap-propriate approach to treat both processes.

The global chemical evolution was found to change inall models due to vertical mixing processes, compared to

Page 14: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

656 M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows

Fig. 8. Chemical evolution of CS, SO, S, and SO2 (clockwise from top left) along the trajectory from 10 AU to 1 AU; mod3ELIS includingvertical mixing. The contour lines refer to the abundance values relative to the total amount of hydrogen: 10−13,−16 and additionally 10−10

for SO2.

results without vertical mixing. Furthermore, the local changesin abundances due to diffusion will be “transported” by advec-tion. Diffusion processes appear to have only a limited effect onthe chemical evolution in regions where gas-grain proccessesdominate the kinetics.

In particular, models with different disks parameters M andα may show similarities in the chemical evolution of the diskgas. Thus, in addition to mass transport processes, informationabout the local conditions, which determine the kinetics, arestill needed in order to analyze the chemical evolution.

Appendix A: Boundary conditions determinedby radiation transport

For a one-dimensional planar atmosphere, the steady radiationtransfer equation is given by

cosΘ∂Iν∂τν= Iν(τν,Θ) − S ν(τν) (A.1)

τν =

z∫−∞κν(z) dz,

where Θ, Iν, S ν, τν, and κν refer to the photon propaga-tion angle and the frequency-dependent quantities: the specific

intensity of radiation, the source function, the optical depth,and the absorption coefficient. z denotes the geometrical depth.The emergent specific intensity in general and its value outsidethe atmosphere, i.e., τν=0, can be obtained by

Iν(τν,Θ) = secΘ

∞∫0

S ν(τ∗ν)e−(τ∗ν−τν) secΘ dτ∗ν (A.2)

Iν(0,Θ) = secΘ

∞∫0

S ν(τ∗ν)e−τ∗ν secΘ dτ∗ν.

In particular, the monochromatic radiation flux πFν can bespecified by

Fν(τν) = 2

π∫0

Iν(τν,Θ) cosΘ sinΘ dΘ (A.3)

Fν(0) = 2

∞∫0

S ν(τ∗ν)K2(τ∗ν) dτ∗ν,

Page 15: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows 657

Fig. 9. Chemical evolution of OH, O, and O2 along the trajectory within 4 AU; The figures on the left and right panel refer to results of thesimulations mod1ELIS and mod3ELIS, respectively. Vertical mixing were taken into account in both simulations.

where the last equation corresponds to the flux outside the at-mosphere. K2 denotes the exponential integral of the secondkind which belongs to the family

Kn(x) =

∞∫1

y−ne−xydy.

Replacing the source function S ν by its power expansioni.e.,

S ν(x) = S ν(x0)

+∂S ν∂x

∣∣∣∣∣x0

(x − x0) +12

∂2S ν∂x2

∣∣∣∣∣∣x0

(x − x0)2 + O(x3),

Page 16: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

658 M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows

Fig. 10. Chemical evolution of CS, SO, S, and SO2 (clockwise from top left) along the trajectory from 10 AU to 1 AU; mod2ELIS includingvertical mixing. The contour lines refer to the abundance values relative to the total amount of hydrogen: 10−13,−16 and additionally 10−10 forSO2.

substituting into Eq. (A.3), and taking the following expressioninto account

∞∫0

xsKn(x) dx =s!

s + n,

one obtains the final expression for the monochromatic radia-tion flux πFν(0) (Unsold 1968):

Fν(0) = S ν(τ∗ν) +

(23− τ∗ν

)∂S ν∂τν

∣∣∣∣∣∣τ∗ν

(A.4)

+12

59+

(23− τ∗ν

)2 ∂2S ν∂τ2ν

∣∣∣∣∣∣∣τ∗ν

+ O(τ3ν

).

Neglecting the nonlinear terms in (A.4), we findFν(0) = S ν (τ∗ν = 2/3), which is well-known as theEddington-Barbier relation. The emergent monochromaticradiation flux is equal to the source function at optical depthτ∗ν=2/3.

With respect to the vertical disk model, the transport ofradiation energy has to be modified within the frequency-integrated variables I, S , and τ. With exception of grey ma-terial, i.e., κ ≡ κν, the transfer equation obtained serves only as

an approximation. Providing a frequency-averaged absorptioncoefficient in terms of a single mean opacity κ= κ, one obtainsthe zeroth and first angular moment of the transfer Eq. (A.1):

1κ(ν)

∂zH(ν) = J(ν) − S (ν) (A.5)

1κ(ν)

∂zK(ν) = H(ν), (A.6)

determinded by the frequency-integrated variables S , J, H, andK. The angular moments of the (monochromatic) specific in-tensity are given by

(Jν,Hν,Kν

)=

∫Iν(1, cosΘ, cos2Θ

) dω4π·

In addition, the indices within the parentheses refer to the exact(frequency-dependent) description. Based on this set of equa-tions one can derive a single mean opacity κ which guaranteesthe correct total flux of radiation energy, i.e.

∫Hνdν = H,

∞∫0

1κν

∂Kν∂z

dν =1κ

∞∫0

∂Kν∂z

dν. (A.7)

The assumption that the specific intensity Iν is close to itsblackbody value holds as long as one considers regions of

Page 17: Astronomy c ESO 2004 Astrophysics - star.pst.qub.ac.uktjm/papers/ilgner04.pdf · We briefly summarise the basic concepts of accretion disk ... there are “dead zones” where other

M. Ilgner et al.: Transport processes and chemical evolution in steady accretion disk flows 659

a non-negligible amount of optical depth. Because of theisotropic radiation field, the cos2Θ term can be replaced byits mean value 1/3 which finally results in Eddington’s approx-imation K(ν) = 1/3 J(ν) and the Rosseland mean opacity κR,respectively. This kind of mean opacity was used to approxi-mate the radiative energy transport in Eq. (7). Correspondingto the relationship (A.7) the boundary condition for τ∗ν = 2/3can be replaced by τ∗=2/3.

Finally, we have to specify the relation between the phys-ical temperature T and the effective temperature Teff used tomeasure the emitted flux from a black body, see Eq. (5). Byrequiring radiative equilibrium, Eqs. (A.5–A.6) yield (Unsold1968)

J = 3H

(23+ τ

), (A.8)

where Eddington’s approximation was used. Applying theStefan-Boltzmann law one obtains the simple relation

T 4 =34

(23+ τ

)T 4

eff . (A.9)

With respect to the chosen boundary value of τ∗ one can easilyspecify T (z∗)=Teff .

Acknowledgements. M.I. thanks Edwin Bergin for useful discussions.We thank P. D’Alessio for providing data of her models. M.I. wassupported by DFG grant He 1935/17-1. Research in Astrophysics atUMIST is supported by PPARC. The computations were performed,using the equipment of the Department of Mathematics, Friedrich-Schiller-Universitat Jena.

References

Aikawa, Y., Umebayashi, T., Nakano, T., & Miyama, S. 1999, ApJ,519, 705

Aikawa, Y., & Herbst, E. 1999, A&A, 351, 233Aikawa, Y., & Herbst, E. 2001, A&A, 317, 1107Aikawa, Y., van Zadelhoff, G., van Dishoeck, E., & Herbst, E. 2002,

A&A, 386, 622Balbus, S. A., & Hawley, J. F. 1991, ApJ, 376, 214Bauer, I., Finocchi, F., Duschl, W., Gail, H.-P., & Schloder, J. 1997,

A&A, 317, 273Bell, K. R., Cassen, P., Klahr, H. H., & Henning, Th. 1997, ApJ, 486,

372Blake, G., van Dishoeck, E., & Sargent, A. 1992, ApJ, 391, L99Brown, P., Byrne, G., & Hindmarsh, A. 1988, LLNL Report UCRL-

98412

Charnley, S. 1997, ApJ, 481, 396Chiang, E., & Goldreich, P. 1997, ApJ, 490, 368D’Alessio, P., Canto, J., Calvet, N., & Lizano, S. 1998, ApJ, 500, 411Doty, S., van Dishoeck, E., van der Tak, F., & Boonman, A. 2002,

A&A, 389, 446Duschl, W., Gail, H.-P., & Tscharnuter, W. 1996, A&A, 312, 624Dutrey, A., Guilloteau, S., & Guelin, M. 1997, A&A, 317, L55Finocchi, F., Gail, H.-P., & Duschl, W. 1997, A&A, 325, 1264Finocchi, F., & Gail, H.-P. 1997, A&A, 327, 825Gail, H.-P. 1998, A&A, 332, 1099Gail, H.-P. 2001, A&A, 378, 192Gail, H.-P. 2002, A&A, 390, 253Hasegawa, T., Herbst, E., & Leung, C. 1992, ApJ, 82, 167Hasegawa, T., & Herbst, E. 1993, MNRAS, 263, 589Hayashi, C., Kakazawa, N., & Nakagawa, Y. 1985 in Protostars and

Planets II , ed. D. Black & M. Matthews (Arizona Press), 1100Kley, W., & Lin, D. N. C. 1992, ApJ, 397, 600Kley, W., Papaloizou, J. C. B., & Lin, D. N. C. 1993, ApJ, 416, 679Langer, W., van Dishoeck, E., Bergin, E., et al. 2000, in Protostars and

Planets IV, ed. V. Mannings & A. Boss, 29–58 (Tucson: Universityof Arizona)

Launhardt, R., Evans, N., Wang, Y., et al. 1998, ApJS, 119, 59Levermore, C. D., & Pomraning, G. C. 1981, ApJ, 248, 321Lynden-Bell, D., & Pringle, J. E. 1974, MNRAS, 168, 603Markwick, A., Ilgner, M., Millar, T, & Henning, Th. 2002, A&A, 385,

632Pollack, J. B., Hollenbach, D., Beckwith, S., et al. 1994, ApJ, 421, 615Pringle, J. E. 1981, A&ARv, 19, 137Qi, C. 2000, Ph.D. Thesis, California Institute of Technology,

PasadenaRawlings, J., & Hartquist, T. 1997, ApJ, 487, 672Semenov, D., Henning, Th., Helling, Ch., Ilgner, M., & Sedlmayr, E.

2003, A&A, 410, 611Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337Snyder, L., Veal, J., Woodney, L., et al. 2001, ApJ, 121, 1147Spalding, D. B. 1971, Chem. Eng. Sci., 26, 95Strehmel, K. & Weiner, R. 1995, Numerik gewohnlicher

Differentialgleichungen (Stuttgart: Teubner)Toomre, A. 1964, ApJ, 139, 1217Taylor, S., Morata, O., & Williams, D. 1996, A&A, 313, 269Umebayashi, T., & Nakano, T. 1980, Publ. Astron. Soc. Jpn, 32, 405Umebayashi, T., & Nakano, T. 1981, Publ. Astron. Soc. Jpn, 33, 617Unsold, A. 1968, Physik der Sternatmospharen (Berlin: Springer)Wehrstedt, M., & Gail, H.-P. 2002, A&A, 385, 181Willacy, K., Klahr, H., Millar, T., & Henning, Th. 1998, A&A, 388,

995Willacy, K., & Langer, W. 2000, ApJ, 544, 903Wutz, M., Adam, H., & Walcher, W. 1992, Theorie und Praxis der

Vakuumtechnik, Vieweg, BraunschweigXie, T., Allen, M., & Langer, W. 1995, ApJ, 440, 674