A publishing partnership

Articles

A MOLECULAR DYNAMICS STUDY ON SLOW ION INTERACTIONS WITH THE POLYCYCLIC AROMATIC HYDROCARBON MOLECULE ANTHRACENE

, , , and

Published 2014 February 13 © 2014. The American Astronomical Society. All rights reserved.
, , Citation J. Postma et al 2014 ApJ 783 61 DOI 10.1088/0004-637X/783/1/61

0004-637X/783/1/61

ABSTRACT

Atomic collisions with polycyclic aromatic hydrocarbon (PAH) molecules are astrophysically particularly relevant for collision energies of less than 1 keV. In this regime, the interaction dynamics are dominated by elastic interactions. We have employed a molecular dynamics simulation based on analytical interaction potentials to model the interaction of low energy hydrogen and helium projectiles with isolated anthracene (C14H10) molecules. This approach allows for a very detailed investigation of the elastic interaction dynamics on an event by event basis. From the simulation data the threshold projectile kinetic energies above which direct C atom knock out sets in were determined. Anthracene differential energy transfer cross sections and total (dissociation) cross sections were computed for a wide range of projectile kinetic energies. The obtained results are interpreted in the context of PAH destruction in astrophysical environments.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Polycyclic aromatic hydrocarbon molecules, or PAHs, are thought to be a common molecular constituent of the interstellar medium. They are found in all directions of observation and seem to be distributed over great distance scales. The interstellar medium affects these PAHs through the interstellar radiation field, through stellar winds, shock waves, and hot ionized gas. Since PAHs are observed over great distances, it is of interest to determine their interaction dynamics in these processing mechanisms. This may shed some light on whether PAHs might survive the harsh environments of interstellar space or if they would have to be protectively incorporated into larger (supra-molecular) structures in order to survive.

The origin and further (ion-)chemistry of PAHs is not fully understood. One pathway of formation might be that PAHs are formed in regions of high C content through chemical reaction mechanisms as might occur in cool stellar winds (Frenklach & Feigelson 1989). Another proposed mechanism is the erosion of dust grains in shock waves originating for example from supernova explosions (Tielens et al. 1994). PAHs present in these regions are also processed again by the hot post shock gas (Micelotta et al. 2010b). Shock velocities of this kind are relatively low (50–200 km s−1) and since hydrogen and helium are by far the most abundant particles present in the gaseous state (90.8% and 9.1% by number, respectively; Ferriere 2001) in the interstellar medium this suggests an investigation of interactions of ions and PAH molecules in the range of 10–500 eV projectile ion energies.

Over the last two years a number of research groups have started to experimentally study ion interactions with PAHs and PAH clusters. Impact of keV ions such as He2 + on anthracene (Postma et al. 2010), pyrene and coronene (Lawicki et al. 2011) was found to efficiently cause multiple ionization followed by extensive fragmentation. PAH ionization and dissociation energies have been determined comparing fragmentation spectra obtained in keV Xeq + collisions with pyrene and fluoranthene with density functional theory calculations (Seitz et al. 2011). Reitsma et al. have experimentally and theoretically determined activation energies (Reitsma et al. 2012) and kinetic energy releases (Reitsma et al. 2013) of anthracene and naphthalene dications, respectively, after ion collisions.

These pioneering studies shed light on the fundamental physics of ion–PAH interactions, however the collision energies under investigation were exceeding the most astrophysically relevant range.

In a recent article Micelotta et al. (2010b) have investigated PAH processing in shocks with velocities between 50 and 200 km s−1 using a framework of binary collision approximations (BCA) for the interactions of atomic projectiles and the PAH constituent carbon atoms. It was found that the molecular structure of PAHs is strongly affected by shock processing in the interstellar medium. It is suggested that nuclear interaction could be a pathway to formation of N-containing PAHs. The appeal of the BCA approximation is its straightforward analytical nature. The molecular properties of the target, however, are entirely neglected. A fully realistic description of ion–PAH interactions that includes all molecular properties would require to solve the time-dependent Schrödinger equation for the entire system. Computationally this approach is far too demanding for molecular systems that are more complex than, e.g., H2. As a feasible alternative, Kunert & Schmidt (2001) have modeled ion collisions with C60 molecules using non-adiabatic quantum molecular dynamics (MD) which combines time-dependent density functional theory (TD-DFT) with a classical MD description of the nuclear motion. A similar approach was recently employed by Wang et al. (2011) to study the dynamics of ethylene molecules in intense laser fields. At present the combination of TD-DFT with classical MD is still computationally challenging and very time consuming and thus not an ideal choice for systematic studies of ion–PAH collisions.

We have thus opted for an entirely classical MD approach to study the interactions of hyperthermal to sub-keV H and He atoms with PAHs. As with many other hydrocarbon molecules, the intramolecular forces in PAHs can be modeled realistically using Brenner's analytical bond-order potential (Brenner 1990). In an early study we have shown that this potential is suitable for studies on ion collisions with free C60 molecules (Schlathölter et al. 1999). Note that charge exchange processes are neglected in this study because neutral instead of ionic projectiles are treated.

In this article we present an MD case study on H and He collisions with the PAH anthracene (C14H10). After a brief introduction of the interatomic potentials and numerical techniques used, the MD code is validated. In the following, ion induced direct fragmentation and heating of anthracene are investigated and the respective absolute cross sections are derived. The findings are discussed within an astrophysical perspective.

2. SIMULATION

To track the molecular dynamics of the ion–PAH system in time, the classical equations of motion need to be solved numerically for all constituents of the entire collision system. The forces acting on the atomic particles need to be derived from the respective interaction potentials. To obtain optimal results it is crucial to choose a realistic analytical potential, capable of reproducing PAH properties as well as properties of the various possible fragments.

2.1. Intra-molecular Potential

For the intra-molecular potential we use the Brenner reactive bond-order potential, which is based on the pioneering work on (reactive) bond-order potentials by Abell (1985) and Tersoff (1986). Brenner (1990) devised a functional form for the potential, which properly describes radical structures and conjugation and which gives excellent results for a great number of hydrocarbon structures. The potential energy is defined as the sum of an attractive pair potential VA(rij) and a repulsive pair potential VR(rij) between the atoms i and j, where rij is the internuclear distance:

Equation (1)

The many-body coupling between the atoms i and j, depending on the local environment of the bond, is included in the bond-order function $\overline{B}_{ij}$. $\overline{B}_{ij}$ explicitly depends on the angles between the ij bond and neighboring bonds ik and jl. Its functional form can be found in Brenner (1990). The pair potentials are defined as

Equation (2)

Equation (3)

For Sij = 2 the sum of these terms equals the Morse-potential (Morse 1929) with a potential well depth $D_{ij}^{e}$, an equilibrium internuclear distance $R_{ij}^{e}$ and a force constant βij. fij is a smooth cutoff function to limit the range to nearest neighbors only. The parameters employed here were determined by Brenner (1990) by fitting the free parameters to experimental data.

2.2. Screened Coulomb Interactions

The interaction between fast atomic particles, i.e., between the impinging ion and a PAH constituent atom, can be approximated by a repulsive screened-Coulomb potential:

Equation (4)

where Z1 and Z2 are the atomic numbers of the collision partners, e is the elemental charge and r the internuclear separation of the particles. The screening function Φ accounts for the action of the electron distributions of the collision partners. The screening length a sets the length scale for this effect. We have chosen the functional form for Φ that was suggested by Ziegler et al. (1985) and that is most widely used for the description of energetic collisions of atomic particles:

Equation (5)

with

Equation (6)

a0 = 0.0592 nm is the Bohr radius. The coefficients in the Ziegler–Biersack–Littmark (ZBL) potential are given in Table 1.

Table 1. ZBL Potential Coefficients ai and bi Appearing in the Screening Function

i a b
1 0.1818 3.2
2 0.5099 0.9423
3 0.2802 0.4029
4 0.02813 0.2016

Download table as:  ASCIITypeset image

2.3. Integration and Validation

To numerically integrate the classical equations of motion, we have employed the particularly accurate Beeman algorithm (Beeman 1976):

Equation (7)

Equation (8)

Here v and a are the velocity and acceleration in three dimensions of the particle under study. The algorithm is not self-starting, so to initialize the integration, the velocity-Verlet algorithm (Verlet 1967) was used. The projectile is initialized at a cutoff distance where interaction is negligible and the interaction is tracked until the projectile reaches the cutoff distance again. For perpendicular impact on anthracene, the cutoff distance was chosen to be 35 AU from the molecular plane.

The code was validated by reproduction of the atomization energies calculated by Brenner (1990). The results are shown in Appendix A. A deviation is only found for ethynylbenzene which is probably due to a typo in Brenner's work. Our value of 68.4 eV is in agreement with more recent studies (Che et al. 1999).

The projectile–target interaction was validated by simulating head-on binary collisions for which an analytical solution for the energy transfer between projectile and target is known (Ziegler et al. 1985):

Equation (9)

where Mp and Mt are the projectile and target masses respectively, E0 is the initial projectile kinetic energy and Tm is the maximum transferred energy of the projectile to the target particle. Excellent agreement was found. The simulation geometry used in the following consists of an anthracene molecule and a projectile atomic particle as displayed in Figure 1.

Figure 1.

Figure 1. Simulation geometry.

Standard image High-resolution image

3. RESULTS AND DISCUSSION

3.1. H and He Collisions with Anthracene: Direct Fragmentation

Figures 7 and 8 (Appendix B) show a series of frames of a He projectile colliding head-on with a C atom in the anthracene molecule. The frames are separated in time by 1000 AU = 2.42 · 10−14 s. The He kinetic energies for these trajectories are Ekin = 40 eV (Figure 7) and 50 eV (Figure 8), respectively. Clearly at 40 eV, the molecule is left vibrationally excited, but intact whereas at 50 eV the direct hit to a C atom leads to its ejection.

Figure 2 displays the atom numbering scheme for the anthracene molecule used in the following. For a better insight into the interaction process, first a number of characteristic collision geometries are investigated. Table 2 shows results for the situation sketched in Figure 1. The molecule is oriented in the yz-plane and a 100 eV He projectile initially moves toward the molecule in x-direction. The time step was 10 AU = 0.24 fs. Relevant quantities are the projectile kinetic energy loss ΔEkin, the kinetic energy of the knocked-out fragment(s) Efrag and the vibrational energy of the remaining molecule Evib. The latter is defined as Evib = ΔEkinEfragEtrans, with Etrans being the translational energy of the remaining molecule after the collision (usually on the order of 1–2 eV). For symmetry reasons only results for impact on C3, C4, C4a, C10, H3, H4, and H10 and at the midpoint of the respective bonds are presented. For each impact site, Table 2 lists ΔEkin (Column 2), Evib (Column 3), Efrag (Column 4) and the outcome of the collision (Column 5). Clearly, a substantial amount of the projectile energy is carried away by the knocked-out fragment.

Figure 2.

Figure 2. Numbering scheme of atoms in the anthracene molecule.

Standard image High-resolution image

Table 2. 100 eV He on Anthracene: Perpendicular Head-on Collisions

Target Projectile Molecule Efrag Effects
ΔEkin (eV) Evib (eV) (eV)
C atoms
C3 70.60 16.25 54.35 CH loss
C4 70.57 15.90 54.67 CH loss
C4a 69.77 16.85 52.92 C loss
C10 70.52 15.55 54.97 CH loss
H atoms
H3 59.48 5.21 54.27 H loss
H4 59.38 5.17 54.21 H loss
H10 59.27 5.15 54.12 H loss
CC bonds
C3-C2 12.18 12.18 ... Bond scission
C4-C3 12.15 12.15 ... Bond scission
C4a-C4 11.44 11.44 ... Bond scission
C4a-C9a 10.59 10.59 ... Bond scission
C4a-C10 11.55 11.55 ... Bond scission
CH bonds
C3-H3 25.77 12.30 13.47 H loss
C4-H4 25.74 12.28 13.46 H loss
C10-H10 25.71 12.23 13.48 H loss

Note. The first column contains the target, the second column contains projectile kinetic energy losses ΔEkin, and the third and fourth columns contain the molecular vibrational excitation Evib and the kinetic energy of the fragment(s) Efrag, respectively.

Download table as:  ASCIITypeset image

The results for 1 keV He projectiles are shown in Table 3 (time step 1 AU = 0.024 fs). In collisions with such energetic projectiles, much less of the projectile kinetic energy is transferred elastically to the remaining molecule. Instead, in head-on collisions an energetic recoil atom is produced. In collisions with a C atom in the molecule, for example, it is obvious that the C atom is knocked out with a substantial amount of energy, while the hydrogen that was attached to it is more or less stationary and stays in the vicinity of the molecule.

Table 3. Same Setup as Table 2, But for 1 keV He on Anthracene: Perpendicular Head-on Collisions

Target Projectile Molecule Efrag Effects
ΔEkin (eV) Evib (eV) (eV)
C atoms
C3 744.82 16.64 C:728.00  H:0.18 C,H loss
C4 744.81 15.34 C:729.30  H:0.17 C,H loss
C4a 744.38 16.09 C:728.28 C loss
C10 744.77 15.02 C:729.58  H:0.17 C,H loss
H atoms
H3 635.32 4.95 630.37 H loss
H4 635.25 4.92 630.33 H loss
H10 635.13 4.89 630.24 H loss
CC bonds
C3-C2 1.05 1.05 ... ...
C4-C3 1.05 1.05 ... ...
C4a-C4 0.99 0.99 ... ...
C4a-C9a 0.92 0.92 ... ...
C4a-C10 1.00 1.00 ... ...
CH bonds
C3-H3 2.78 2.78 ... ...
C4-H4 2.78 2.78 ... ...
C10-H10 2.77 2.77 ... ...

Download table as:  ASCIITypeset image

Tables 4 and 5 display the results for 100 eV and 1 keV H impact, respectively (time step 1 AU = 0.024 fs). For through-bond trajectories, no bond scission is observed in either case. Head-on collisions always lead to knock out of the respective atom.

Table 4. 100 eV H on Anthracene: Perpendicular Head-on Collisions

Target Projectile Molecule Efrag Effects
ΔEkin (eV) Evib (eV) (eV)
C atoms
C3 27.37 18.68 8.69 CH loss
C4 27.36 17.99 9.37 CH loss
C4a 27.24 21.31 5.93 C loss
C10 27.35 17.28 9.97 CH loss
H atoms
H3 90.32 4.95 85.37 H loss
H4 90.15 4.92 85.23 H loss
H10 89.96 4.88 85.08 H loss
CC bonds
C3-C2 0.95 0.95 ... ...
C4-C3 0.95 0.95 ... ...
C4a-C4 0.90 0.90 ... ...
C4a-C9a 0.83 0.83 ... ...
C4a-C10 0.90 0.90 ... ...
CH bonds
C3-H3 2.48 2.48 ... ...
C4-H4 2.48 2.48 ... ...
C10-H10 2.47 2.47 ... ...

Note. The first column contains the target, the second column contains projectile kinetic energy losses, and the third and fourth columns contain the molecular vibrational excitation and the energies of the fragments, respectively.

Download table as:  ASCIITypeset image

Table 5. Same Setup as Table 4, But for 1 keV H on Anthracene: Perpendicular Head-on Collisions

Target Projectile Molecule Efrag Effects
ΔEkin (eV) Evib (eV) (eV)
C atoms
C3 284.27 15.71 C:268.16  H:0.40 C,H loss
C4 284.27 15.39 C:268.48  H:0.40 C,H loss
C4a 284.26 16.17 C:268.09 C loss
C10 284.27 15.07 C:268.80  H:0.40 C,H loss
H atoms
H3 969.85 4.90 964.95 H loss
H4 971.63 4.87 966.76 H loss
H10 971.99 4.84 967.15 H loss
CC bonds
C3-C2 0.081 0.081 ... ...
C4-C3 0.081 0.081 ... ...
C4a-C4 0.077 0.077 ... ...
C4a-C9a 0.071 0.071 ... ...
C4a-C10 0.078 0.078 ... ...
CH bonds
C3-H3 0.22 0.22 ... ...
C4-H4 0.22 0.22 ... ...
C10-H10 0.22 0.22 ... ...

Download table as:  ASCIITypeset image

In this context it is of interest to investigate the collision dynamics as a function of projectile atom energy, in particular in the energy range where direct atom knock-out sets in. The threshold kinetic energy transfer $T_0=\Delta E_{{\rm kin}}^{{\rm threshold}}$ is reached when the hitting atom obtains sufficient momentum to get liberated from the anthracene molecule. The quantity T0 is often referred to as the vacancy formation energy. T0 can be determined for a given projectile and impact site by simulation of the respective collision as a function of the projectile kinetic energy.

In Figure 3 the projectile kinetic energy dependence of ΔEkin, Evib, Efrag and Etrans for collisions of H (blue) and He (red) with the anthracene carbon atom number 4a is depicted. The dashed vertical lines indicate the threshold kinetic energy for direct knock-out $E_{{\rm kin}}^{{\rm threshold}}$. Note that prompt bond scission already sets in at projectile energies a few eV below $E_{{\rm kin}}^{{\rm threshold}}$. ΔEkin (Figure 3, top) exhibits an almost linear increase with Ekin as expected from classical mechanics. For low projectile kinetic energies, ΔEkin is almost entirely transferred into Evib (≈95% for H, ≈90% for He), which accordingly exhibits an almost linear dependence on Ekin as well. Because of the much larger mass of the anthracene molecule as compared to the projectile atom, molecular translational energies Etrans stay below 2.5 eV and 2 eV for He and H respectively and increase almost linear with Ekin.

Figure 3.

Figure 3. Energetics of H (blue) and He (red) atoms colliding with C4a of anthracene. From top to bottom: total energy loss ΔEkin, anthracene vibrational energy Evib, kinetic energy of the knocked out atom(s) Efrag, and anthracene translational energy Etrans (in eV).

Standard image High-resolution image

Molecular excitation exhibits a clear peak at $E_{{\rm kin}} \approx E_{{\rm kin}}^{{\rm threshold}}$, the threshold projectile kinetic energy for carbon knock-out. These values (He: 43 eV, H: 99 eV) are strongly dependent on the projectile mass. Independent from the projectile mass is the quantity $\Delta E_{{\rm kin}}^{{\rm threshold}} = T_0$ which amounts to approximately 27 eV (see horizontal line in Figure 3, top). For solid targets T0 is often referred to as "threshold displacement energy"). It depends not only on the site but also on the orientation and therefore is a rather ill-defined quantity, still experimentally undetermined for free PAHs. Experimental and theoretical determinations of T0 on solid carbon have yielded numbers ranging from 5 eV for amorphous graphite (Cosslett 1978), over 7.6–15.7 eV for carbon nanotubes (Füller & Banhart 1996), to 15–20 eV for graphitic nanostructures (Banhart 1997). In their BCA study, Micelotta et al. chose a conservative value of T0 = 7.5 eV for most of their calculations, which is in line with the nanotube data. Very recently, however, from first principles density functional theory MD calculations on electron collisions with graphene, a displacement energy T0 = 22.03 eV was obtained (Kotakoski et al. 2010). Calculations using the tight binding approximation found T0 = 23 eV (Zobelli et al. 2007), whereas classical MD simulations of ion bombardment on graphene with interactions based on the Brenner potential yielded T0 = 22.2 eV (Lehtinen et al. 2010). Using the first principles value T0 = 22.03 eV, experimentally observed cross sections for electron induced knock-out from graphene sheets could be very well reproduced (Meyer et al. 2012).

Tables 6 and 7 display T0 for perpendicular head-on collisions of H and He projectiles on the different C sites in anthracene obtained from our MD simulation. For comparison, the values for coronene (C24H12) and graphene are displayed as well (graphene was simulated as a sheet of 200 C atoms with the impact site in the center). Clearly, very similar values are observed for the two PAHs and for graphene, i.e., the recently obtained threshold displacement energies for graphene can be used as a reference for our data. Our results for graphene are markedly higher than the experimentally determined T0 = 22.03 eV. To some extent, this discrepancy is due to the long range nature of the interaction potentials, which implies an interaction between the incoming ion/atom and typically more than one C atom in the target. Energy is thus also transferred to neighboring C atoms and not only to the knock-out atom. Furthermore, the minimum value for T0 depends also on the orientation of the system. Direct initialization of the knock out atom with appropriate momentum in the MD code used here gives Td ≈ 21.3 eV, which is close to the reference data. It can thus be concluded, that the T0 values for PAHs displayed in Tables 6 and 7 are realistic.

Table 6. Threshold Kinetic Energies and Kinetic Energy Transfer for Vacancy Formation by H Projectiles in eV

Target Projectile Ekin Projectile ΔEkin Efrag Effect
(T0)
Anthracene
C3 91 −24.8 6.9 CH loss
C4 90 −24.5 5.8 CH loss
C4a 99 −27.0 5.0 C loss
C10 89 −24.2 4.1 CH loss
Coronene
C 94 −25.5 4.0 C loss
Graphene
C 95 −25.8 4.7 C loss

Note. First column: target site; second column: threshold kinetic energy; third column: threshold kinetic energy transfer to the molecule (T0); fourth column: fragment kinetic energy; fifth column: effects of the collision. For the graphene calculation, a sheet consisting of 200 C atoms was used.

Download table as:  ASCIITypeset image

Table 7. Threshold Kinetic Energies and Kinetic Energy Transfer for Vacancy Formation by He Projectiles in eV (See Table 6 for Details)

Target Projectile Ekin Projectile ΔEkin Efrag Effect
(T0)
Anthracene
C3 42 −27.7 7.1 CH loss
C4 41 −26.9 5.1 CH loss
C4a 43 −27.5 5.4 C loss
C10 41 −26.8 7.8 CH loss
Coronene
C 42 −26.7 6.1 C loss
Graphene
C 42 −26.7 6.4 C loss

Download table as:  ASCIITypeset image

For projectile kinetic energies exceeding the knock-out threshold the excitation of the molecule decreases asymptotically to the vacancy energy Evac. The vacancy energy is a quantity that weakly depends on the C-atom location in the molecule and on the size of the PAH. For C4a in anthracene, Evac = 16.0 eV whereas for a coronene inner ring atom, Evac = 15.3 eV. In the Evib plot in Figure 3, the vacancy energy (Evac) is indicated by a horizontal line.

It is obvious that—particularly close to threshold—the removal of a carbon atom from a PAH molecule implies Evib > Evac. The implication of these findings is that the Evib of a PAH after a direct C knock out depends on Ekin. It starts at the threshold at T0 and decreases asymptotically with Ekin to reach Evac.

3.2. Monte Carlo Simulations: Cross Sections

Differential cross sections for elastic energy transfer from a projectile atom to an anthracene molecule are obtained by averaging over many collision events where the target molecule is randomly oriented and where the impact parameter is randomly chosen. To this end Monte Carlo simulations of H and He projectiles interacting with the anthracene molecule were performed. These Monte Carlo simulations were performed by implementing a parallelized version of the code on the "Millipede" computer cluster of the Center for High Performance Computing and Visualization of the University of Groningen.

The molecule was positioned at the origin of the simulation geometry with a fixed orientation or with random orientations. The ion started a fixed distance of 35 AU away from the origin to ensure negligible interaction. A great many ion trajectories with random impact parameters were generated in a window of 40 AU by 40 AU. For every ion trajectory the kinetic energy transfer to the molecule was recorded. This allowed for the determination of differential cross sections for energy transfer for a number of projectile kinetic energies of astrophysical interest.

Figure 4 displays differential cross sections for energy transfer (projectile kinetic energy loss) of He atoms of various kinetic energies interacting with anthracene for ion trajectories perpendicular (top panel) to the plane of the molecule and for random orientations (bottom panel), respectively. The differential cross section for energy transfer is defined as

Equation (10)

where N(E) is the number of events with energy transfer E, $\mathcal {L}$ is the (time-)integrated luminosity (number of particles per surface area), and ΔE the bin width. The data were binned using bins of 0.1 eV width.

Figure 4.

Figure 4. Differential cross sections for energy transfer of He projectiles of various kinetic energies to the anthracene molecule for upright (i.e., projectile trajectories perpendicular to the plane of the molecule) orientation (top panel) and random orientations (bottom panel).

Standard image High-resolution image

Total cross sections for classes of collisions with energy transfers exceeding a defined threshold can now be computed easily. The values are computed by integrating the differential cross sections using an appropriately chosen lower cut-off value Tco. For a better comparison to the BCA results (Micelotta et al. 2010b), it is instructive to choose their knockout energy value of Tco = 7.5 eV. As already shown, this value falls short of the more realistic knock-out energy, obtained from the MD simulation. However, it is in the range of activation energies for anthracene dissociation. Figure 5 displays absolute cross sections for He impact perpendicular to the PAH plane (top panel) and for random orientations (bottom panel) as a function of projectile energy. In addition, cross sections for Tco = 4.6 eV are displayed, as this value is the activation energy for the important C2H2 loss process. Also shown are the respective data for anthracene targets from which the H atoms were removed, as well as for 14 single non-interacting C atoms, mimicking the anthracene carbon skeleton. The latter data are compared to the BCA results (solid line), representing 14 times the cross section for the above threshold (7.5 eV) energy transfer to an isolated C atom target.

Figure 5.

Figure 5. Total cross sections for above threshold energy transfer of He projectiles of various kinetic energies to the anthracene molecule for upright (i.e., projectile trajectories perpendicular to the plane of the molecule) orientation (top panel) and random orientations (bottom panel). The BCA results are added for comparison. For random orientations, the BCA cross section is corrected for orientational averaging to 50% (Micelotta et al. 2010b).

Standard image High-resolution image

For perpendicular orientation, BCA and MD give almost identical results for the case of 14 C atoms and Tco = 7.5 eV. When the MD code is run for a target molecule with the H atoms removed, i.e., with a molecular carbon skeleton, the total cross sections already exceed the BCA data by about 50% at 55 eV collision energy and Tco = 7.5 eV. This indicates that the results of the BCA framework cannot immediately and straightforwardly be extrapolated to molecular structures. MD results for the entire anthracene molecule peak at He collision energies of about 20 eV while the BCA data peaks around 70 eV. At Tco = 7.5 eV, these MD calculations exceed the BCA data by more than a factor of two. For Tco = 4.6 eV, the discrepancy amounts to a factor of 3.

Clearly, cross sections drop, when going from a perpendicular collision geometry to random orientations, mainly because of the smaller geometric cross section. In the extreme case of in-plane collisions, projectile interactions only directly affect PAH constituents facing the projectile, while the others are shadowed. This effect is obviously absent in the BCA and MD cases for atomic C. Micelotta et al. therefore corrected BCA cross sections by a factor 0.5 to account for orientational averaging. Including this correction, the discrepancy between MD for the entire molecule and BCA for single atoms would be almost identical in both panels.

In the next section, the cross sections will be used to estimate PAH dissociation cross sections due to excitation by the projectile.

3.3. Dissociation of PAHs

In recent collision studies on anthracene (and other PAH) cations with neutral He at center of mass energies of 110 eV, Stockett et al. observed single C atom loss for all PAHs under study (M. H. Stockett et al. 2014, private communication), which clearly is the experimental signature of the direct knock-out process investigated here. From this experimental study as well as from the MD results presented here, it is however obvious, that direct C knock out is not the main pathway leading to PAH destruction. In photoionization (Jochims et al. 1994), low energy electron impact (Deng et al. 2006), or collision induced dissociation (Arakawa et al. 2000) studies, typically H, H2, and C2H2 loss from singly charged PAH cations are identified as the most important channels and typically no signatures of single C atom loss are observed. These channels and their respective energetics are appropriate for the studies here, as under astrophysically relevant conditions ions rather than neutrals impact on the PAHs, typically leaving the latter positively charged. It is straightforward to assume a statistical process which leads to a scenario in which an excited molecule is subject to competition between two relevant de-excitation mechanisms: infrared (IR) photon emission and dissociation of the molecule.

Once the energy transferred to the molecule (Evib) is known, it is possible to determine a dissociation probability for the molecule if the rates for dissociation and IR-emission are known (Tielens 2005). The dissociation rate can be determined using an Arrhenius relation,

Equation (11)

where Teff is the effective temperature of the system, NC the number of carbon atoms in the molecule, and E0 the binding energy in eV of the fragment produced in the dissociation process. kB is Boltzmann's constant. Evib is the excitation energy in eV.

Dissociation competes with successive emission of IR photons of energy epsilon (typically 0.16 eV for a C-C vibrational mode) at a rate kIR, which in turn reduces the molecular excitation energy. This implies that the dissociation probability changes with each IR photon emission. The probability for the excited molecule to dissociate between the nth and (n + 1)th IR photon emission equals

Equation (12)

where the probability for dissociation at each step is pi. The unnormalized probability for dissociation at every step pi equals

Equation (13)

i.e., the ratio of the rate constants for both processes. The total (unnormalized) probability for dissociation of the molecule at the end of the emission chain is the sum of the probabilities for dissociation at each photon emission step. Micelotta et al. assume a constant, average pi = pav and write the unnormalized total dissociation probability after a maximum of nmax photon emissions as (Micelotta et al. 2010a)

Equation (14)

The average temperature Tav associated with an average probability pav is taken as the geometric mean $\sqrt{T_{{\rm in}}\times T_{n_{{\rm max}}}}$, where Tin is the effective temperature immediately after excitation of the molecule by the ion and $T_{n_{{\rm max}}}$ is the effective temperature after nmax IR photon emissions corresponding to an internal energy of (TEnmaxepsilon). In line with Micelotta et al. (2010a), for anthracene we furthermore assume nmax ≈ 3, kIR = 100 s−1 (Jochims et al. 1994) and k0 = 1.4 × 1016 s−1 (Ling & Lifshitz 1998). Of crucial importance is the determination of the Arrhenius energy E0. By fitting the Arrhenius rate to experimental data from Jochims et al. (1994), Micelotta et al. (2010a) find E0 = 3.65 eV, which is lower than the ≈4.2 eV C2H2 binding energy determined for small PAHs (Ling & Lifshitz 1998). For better agreement with astrophysical data, they determine an alternative value of E0 = 4.6 eV (Micelotta et al. 2010a). Note that the determination of E0 is fundamentally difficult because C2H2 loss competes with H and H2 loss, which have very similar activation energies.

The Monte Carlo calculations yield the distributions of energy Evib transferred to the anthracene molecule, and thus the initial effective temperature Tin, for every single collision. Together with E0 = 4.6 eV and setting nmax = 3 this in turn allows one to determine the anthracene dissociation probability on an event-by-event basis. The dissociation probability is used to calculate total dissociation cross sections by integrating the differential cross sections (for energy transfer of the ion to the molecule) multiplied by the dissociation probability for each energy transfer:

Equation (15)

The dissociation probability is relatively constant over the range of energy transfers at approximately 0.6.

The total dissociation cross section can now be determined for every collision energy. To do so, trajectories leading to direct knock out are not counted separately for two reasons. First of all, direct knock out is only a quantitatively small dissociation channel and second, this channel is associated with relatively high excitation energies, leading to subsequent dissociation processes. Figure 6 shows the total dissociation cross sections for perpendicular impact (top panel) and for random molecular orientations (bottom panel). Fortunately, these total dissociation cross sections are not very sensitive to the precise value of the integration lower cut-off of 4.6 eV or 7.5 eV, making the choice of the cut-off value less critical. The reason is the fact that the dissociation probability is close to zero near this threshold.

Figure 6.

Figure 6. Total anthracene dissociation cross sections after He collisions (squares). The solid line gives the total BCA cross section for collisions leading to excitation energies exceeding Tco = 7.5 eV. For random orientations, the BCA cross section is corrected for orientational averaging to 50% (Micelotta et al. 2010b).

Standard image High-resolution image

Micelotta et al. (2010b) have determined PAH dissociation cross sections using the BCA and assuming ion induced C atom knock-out as the underlying mechanism. Knock out was assumed to occur if the transferred energy exceeded a displacement threshold of 7.5 eV. We have already shown that MD calculations reveal a significantly larger displacement energy, but how do the resulting fragmentation cross sections compare?

The knock-out results of BCA calculations for the 7.5 eV threshold from Figure 5 are displayed as a solid line in Figure 6 for comparison. Once again, these values have to be corrected by 0.5 for orientational averaging in the case of randomly oriented targets. It can be seen that for perpendicular impact the MD cross sections exceed the BCA data by almost a factor of 2 for 40 eV projectile energies whereas the curves almost converge at high projectile velocities. For random orientation, cross sections are again markedly lower and the maximum is now observed for projectile energies around 100 eV. Here, the dissociation cross section is more than twice as high as the BCA data (corrected for orientational averaging). Differences are smaller for lower and higher projectile energies.

In Table 8 we have made a comparison between the average energy transfer of He projectiles to the anthracene molecule as obtained from our simulations and the BCA of Micelotta et al. (2010b). The average energies obtained in this work are calculated using the lower limit of integration or cut-off of Tco = 4.6 eV instead of the somewhat artificially chosen lower limit of Tco = 7.5 eV for direct single carbon knock out, but as the range of energy transfers grows larger this difference is felt less and less. At projectile energies of 55 eV for He the average energy transfers are already very similar.

Table 8. Average Energy Transfers for He Projectiles at Various Kinetic Energies Colliding on Anthracene

Projectile TE (eV) TE (eV)
Ekin (eV) This Work (Micelotta et al. 2010b)
(Tco = 4.6) Tco = 7.5
10 5.4 ...
25 9.8 12.0
40 14.2 15.2
55 17.6 17.9
70 20.2 20.2
85 22.2 22.3
100 23.9 24.1
250 33.7 36.1
500 42.1 46.1

Note. The second column shows our average energy transfers. The third column gives average energy transfers from the binary collision approximation of Micelotta et al. (2010b).

Download table as:  ASCIITypeset image

The astrophysical implications of ion induced PAH destruction in interstellar shocks have been investigated in great detail by Micelotta et al. (2010b). Based on BCA calculations assuming a displacement energy of 7.5 eV, they show that for a PAH containing 50 C atoms, destruction only sets in at shock velocities of about 100 km s−1. For a displacement energy of 15 eV, this value shifts to about 125 km s−1. In the light of the MD calculations, it is clear that the BCA systematically underestimates cross sections for above threshold energy transfer. However, even the displacement energy of 15 eV falls short of the ≈22 eV obtained by MD calculations. We conclude that the results presented here will not significantly change the shock velocity onset for PAH destruction. Electron collisions on the other hand are already very important for shock velocities of 75 km s−1 (Micelotta et al. 2010b). The main astrophysical implication of the present study thus is the "outcome" of the collision. Due to the high displacement energies, direct C atom knockout is not a very strong destruction pathway. Destruction proceeds mainly through C2H2 loss from the periphery of the PAH. If the PAH is not fully destroyed, C2H2 loss can initiate isomerization and cage formation. The small relevance of knock-out collisions on the other hand leaves little room for PAHs with internal vacancies. N atom incorporation at such reaction sites is thus not an obvious pathway toward nitrogenated PAHs.

4. CONCLUSIONS

Using MD simulations, we have studied the interactions of hydrogen and helium atoms with anthracene as prototypical systems for particle collisions with PAH molecules.

Two main conclusions can be drawn from this work.

First, the threshold kinetic energy for a knock-out collision is much larger than estimated previously. With the MD approach, C knock out in a head on collision with the respective C-atom occurs for kinetic energies above 43 eV (He) and 99 eV (H) while BCA calculations find threshold kinetic energies of 10 eV and 26 eV, respectively when assuming a knock-out energy of 7.5 eV (Micelotta et al. 2010b). The reason for the discrepancy is in part due to the fact, that BCA approaches neglect the molecular nature of the PAH. Furthermore, only recently a precise experimental determination of the C threshold displacement energy in graphene bas been obtained (T0 − 22.03 eV (Meyer et al. 2012)). Our MD approach reproduces this knock out energy. For anthracene, we find appreciable numbers of knock-out events for energy transfer of at least ≈27. Previous studies had to rely on older and much less consistent data, hinting at much lower threshold displacement energies as used in the BCA study. As a consequence, direct C-loss is a weak channel and accordingly this process appears not to provide a route toward incorporation of N atoms into the PAH structure. However, the recent first experimental observation of direct C atom knock out in PAH collisions (M. H. Stockett et al. 2014, private communication) is an important first step toward a detailed understanding of collision induced vacancy production in PAH molecules. Therefore it is now feasible to experimentally test the presented MD results in detail in the region where knock out sets in.

Because direct knock out of C atoms from PAHs did not prove to be an efficient channel toward PAH destruction, we considered the dissociation of PAHs through vibrational excitation, as previously done by Micelotta et al., for electron collisions (Micelotta et al. 2010a). This dissociation channel competes with de-excitation through IR photon emission. The dissociation probability (≈0.6) appears to be quite constant over the range of studied projectile energies. Calculating total dissociation cross sections by weighing the integral over the differential cross section for energy transfer with the dissociation probability for that particular energy transfer gives results that are clearly larger than the results obtained from BCA calculations and a C knock-out model. Total dissociation cross sections do not appear to be very sensitive to the precise lower cut-off value. To summarize, for small PAHs as anthracene, ion processing primarily induces loss of C2H2 (or other fragments) rather than C-atom knockout. The corresponding dissociation cross sections exceed those determined from a knock out model and imply a larger role of ionic particles than previously thought. Furthermore, dissociation hints at subsequent isomerization and cage formation, rather than formation of reactive "inner" vacancies, which may act as active sites for PAH nitrogenation.

APPENDIX A: TABLE OF ATOMIZATION ENERGIES

To validate the MD code, atomization energies were computed for all the hydrocarbon compounds that were studied in Brenner's original work (Brenner 1990). Table 9 compares the respective data obtained here with the numbers determined by Brenner and with experimental data. The agreement is excellent with the exception of ethynylbenzene. For this particular compound, Brenner's result is likely to be incorrect, as our value agrees well with more recent studies (Chen et al. 1999).

Table 9. Energies of Atomization (eV) of a Number of Molecular Structures

Structure This Work Brenner Experiment
Alkanes
methane 17.57 17.6 17.6
ethane 29.72 29.7 29.7
propane 41.99 42.0 42.0
n-butane 54.26 54.3 54.3
i-butane 54.27 54.3 54.4
n-pentane 66.54 66.5 66.6
isopentane 66.54 66.5 66.6
neopentane 66.79 66.8 66.7
cyclopropane 35.51 35.5 35.8
cyclobutane 48.65 48.7 48.2
cyclopentane 61.35 61.4 61.4
cyclohexane 73.63 73.6 73.6
Alkenes
ethylene 23.63 23.6 23.6
propene 36.23 36.2 36.0
1-butene 48.50 48.5 48.5
cis-butene 48.83 48.8 48.6
isobutene 48.39 48.4 48.7
(CH3)2C=C(CH3)2 73.15 73.2 73.4
cyclopropene 28.19 28.2 28.8
cyclobutene 42.41 42.4 42.4
cyclopentene 55.75 55.7 55.6
1,4-pentadiene 55.00 55.0 54.8
CH2=CHCH=CH2 41.84 41.8 42.6
CH3CH=C=CH2 40.42 40.4 42.1
H2C=C=CH2 27.82 27.8 29.6
Alkynes
acetylene 17.15 17.1 17.1
propyne 29.42 29.4 29.7
1-butyne 41.69 41.7 42.0
2-butyne 41.69 41.7 42.2
Aromatics
benzene 57.47 57.5 57.5
toluene 69.63 69.6 70.1
1,4-dimethylbenzene 81.79 81.8 82.6
ethylbenzene 81.90 81.9 82.5
ethenylbenzene 76.19 76.2 76.5
ethynylbenzene 68.39 69.8 69.9
naphthalene 91.39 91.4 91.2
anthracene 125.30 ... ...
Radicals
CH2 7.77 7.8 7.8
CH3 12.71 12.7 12.7
H3C2H2 25.67 25.7 25.5
H2C2H 18.88 18.9 18.9
C2H 12.24 12.2 12.2
CH2CCH 24.45 24.5 25.8
n-C3H7 37.94 37.9 37.8
i-C3H7 38.25 38.3 38.0
t-C4H9 50.47 50.5 50.5
phenyl 72.71 52.7 72.7

Note. The columns "Brenner" and "Experiment" are adopted from Brenner (1990).

Download table as:  ASCIITypeset image

APPENDIX B: COLLISION FRAMES

Figures 7 and 8 illustrate the time evolution of a head-on collision process for the example of a He projectile impinging onto the anthracene C9a atom with an initial projectile momentum perpendicular to the anthracene plane. The frames are separated in time by 1000 AU = 2.42 · 10−14 s. The He kinetic energies for these trajectories are Ekin = 40 eV (Figure 7) and 50 eV (Figure 8), respectively. Clearly at 40 eV, the molecule is left vibrationally excited but intact, whereas at 50 eV the direct hit to a C atom leads to its ejection.

Figure 7.

Figure 7. 40 eV He on anthracene: The molecule remains intact.

Standard image High-resolution image
Figure 8.

Figure 8. 50 eV He on anthracene. A C atom is knocked out of the molecule. The impact parameter is identical to the one in Figure 7.

Standard image High-resolution image
Please wait… references are loading.
10.1088/0004-637X/783/1/61