Molecular Dynamics Simulations of Carbon-Supported Ni Clusters

adsorption on the basal plane and for binding to a hydrogen terminated graphite edge. 1. Introduction. Carbon-supported clusters of a few to a few hun...
6 downloads 0 Views 2MB Size
J. Phys. Chem. C 2008, 112, 12663–12668

12663

Molecular Dynamics Simulations of Carbon-Supported Ni Clusters Using the Reax Reactive Force Field Carlos F. Sanz-Navarro,† Per-Olof Åstrand,† De Chen,‡ Magnus Rønning,‡ Adri C. T. van Duin,§ Jonathan E. Mueller,§ and William A. Goddard III§ Department of Chemistry, Norwegian UniVersity of Science and Technology (NTNU), 7491 Trondheim, Norway, Department of Chemical Engineering, Norwegian UniVersity of Science and Technology (NTNU), 7491 Trondheim, Norway, and Materials and Process Simulation Center, DiVision of Chemistry and Chemical Engineering, California Institute of Technology, Pasadena, California 91125 ReceiVed: December 17, 2007; ReVised Manuscript ReceiVed: April 17, 2008

Molecular dynamics simulations have been performed using a Reax force field for C/H/Ni systems to study the structural changes of an Ni100 cluster adsorbed on a carbon platelet. Three different edges of a carbon platelet are considered. We find a complete restructuring of the initial structure of the Ni100 clusters adsorbed on the armchair and zigzag edges. Nonetheless, the mean Ni-Ni bond length hardly changes. Several preferential sites on each of the graphite edges are identified. Diffusion of the entire cluster is found both for adsorption on the basal plane and for binding to a hydrogen terminated graphite edge. 1. Introduction Carbon-supported clusters of a few to a few hundred metal atoms have recently attracted a great deal of attention as catalytic sites since they present a high surface area relative to their volume. Moreover, the catalytic activity of a metal can be enhanced by selecting a suitable support material. For instance, while gold has the reputation of being one of the most inactive metals, gold supported on the surface of certain oxides has been shown to be an active and selective catalyst.1,2 Therefore, the search for a suitable support for metal clusters is crucial for the design of highly effective catalysts. In particular, improvements in the cluster-driven electrocatalysis of water will reduce the high cost of hydrogen production for their use in proton exchange membrane (PEM) fuel cells.3,4 Information derived from experimental observation is limited by the inherent nanosize of the systems, leaving gaps in the overall picture of the cluster-support interaction. However, atomistic simulations can fill this gap. On one hand, electronic structure methods often provide an accurate description,5–7 while they are limited to systems of only around a hundred atoms and time scales of a few picoseconds. On the other hand, molecular modeling based on empirical force fields can deal with systems of a few million atoms and time scales beyond a nanosecond, but at the cost of a less accurate description of the interatomic interactions. Linking the two approaches is the main goal of reactive force fields, which attempts to model the manybody effect of the interatomic interactions with a much lower computational effort than required for quantum-mechanics calculational. Reactive force fields are normally based on a bond-order approach, which defines a bond-order term in the force field to account for the nature of the bond. Several schemes have been proposed to include the bond order in interatomic interactions such as the Tersoff-Brenner force field,8–19 Pettifor’s bond-order potentials (BOP),20–25 and the Reax force field (ReaxFF). The latter is based on a bond-order model in * To whom correspondence should be addressed. † Department of Chemistry, NTNU. ‡ Department of Chemical Engineering, NTNU. § California Institute of Technology.

conjunction with a charge-equilibration scheme,26 and it was first developed for hydrocarbons.27 Subsequently, the ReaxFF has been extended to many other elements and compounds comprised of Si/SiO,28 Al/AlO,29 Ni/Cu/Co/C,30 Mg/MgH,31 Li/ LiC,32 and BiMoOx.33 Recently, we studied the interaction of a Pt100 cluster with a graphite platelet34 using a ReaxFF for C/Pt/H systems. We showed that an initial cluster-substrate mismatch can determine whether the cluster atoms have to rearrange itself to be accommodated in some preferential sites on the graphitic edge. Here we extend our study to a Ni100 cluster. First the methodology is described, both for the parametrization of the ReaxFF and for the simulations. Then the methodology is applied to the study of the interaction of Ni100 clusters with the different edges of a graphite platelet. Finally, the armchair edge is hydrogen terminated and the effect of the termination on the binding between cluster and substrate is investigated. 2. Methodology 2.1. Hydrocarbon Fragment Binding Energies. A ReaxFF has been constructed to model the interatomic interactions in the C/Ni/H systems studied. Detailed information on the functional form of a general ReaxFF can be found in ref 27 and Supporting Information of ref.30 All ab initio hydrocarbon fragment/Ni [111] calculations were performed with the SeqQuest periodic density functional theory code and utilized the PBE density functional and pseudopotentials.35 A p(2 × 2) cell with three nickel layers was used for all calculations. The four nickel atoms comprising the bottom layer were fixed at the experimental nickel lattice distance (2.49 Å) and the eight nickel atoms in the middle and top layers were allowed to relax in the geometry optimizations. All calculations were performed in a 2-D periodic unit cell with periodic sides of length 4.98 Å each and an angle of 120° between them. The vertical dimension normal to the periodicity was 21 Å resulting in a minimum vacuum region of 6 Å both above and below the slab and adsorbent molecule. An energy minimized geometry was used in all cases. Forces were relaxed within 0.0005 Rydbergs/Bohr. Strict spin conservation rules were not taken into account based

10.1021/jp711825a CCC: $40.75  2008 American Chemical Society Published on Web 07/30/2008

12664 J. Phys. Chem. C, Vol. 112, No. 33, 2008 on the assumption that a large nickel crystal can act as a “spin bath” and absorb or donate spin to a surface site without changing the intensive spin properties of the whole crystal. Since the energies of various spin projections form an approximately harmonic well centered on the lowest energy spin projection, the ground-state spin projection was found to the nearest halfintegral Ms spin projection by finding the Ms value for which the energy values for the Ms states 1/2 Ms higher and lower both yield higher energies. The energies of these three states were typically within 2 kcal/mol of each other, so we expect our energy to be within 1 kcal/mol of the bottom of the Ms energy well. 2.2. Simulation Details. For comparison with previous results for a Pt100 cluster,34 a Ni100 cluster was selected. This cluster size lies halfway between two highly stable clusters for fcc metals: Ni55 and Ni14736 and its diameter, ∼2 nm, is within the size of clusters in experiments for fuel cell catalysis applications (i.e., 1-5 nm).37,38 Regarding the cluster structure, Ni atoms were initially arranged by using a genetic algorithm.39 It was checked that this cluster arrangement remains stable over a nanosecond simulation at 600 K and also, after energy minimization, the energy per atom, 3.86 eV, coincides with the value obtained by using an embedded-atom-method (EAM) potential.40 An MD simulation was carried out for each of the three edges of graphite platelet, i.e., the armchair and zigzag edges as well as the graphite basal plane. Three different carbon platelets were built regarding the specific cluster-nanosheet binding edge. In all of them, the distance between first neighbor atom was 1.42 Å as experimentally found for graphite.41 An equilibrium distance of 3.260 Å between graphitic layers was found by miminization of the ReaxFF potential energy. This value is quite close to that of graphite (i.e., 3.41 Å)41 and so it was considered as interlayer distance for the MD simulations. The graphite planes were perpendicular to the y axis and periodic boundary conditions were applied along the x-and y-directions. Details about the three graphite lattice are as follows: (1) A 10-layer lattice with dimensions of 29.826 Å × 29.340 Å × 28.290 Å was employed in the ReaxFF simulation of the binding of a Ni100 cluster to the armchair edge. Each graphite layer contained 336 carbon atoms. The armchair edge was perpendicular to the z-axis. This lattice was located within a simulation box with dimensions 29.826 Å × 32.60 Å × 50.0 Å. (2) A 10-layer graphite lattice with dimensions of 29.52 Å × 29.340 Å × 29.83 Å was employed for the simulations of a Ni100 cluster adsorbed onto the zigzag edge of a carbon platelet, where the zigzag edge was perpendicular to the z direction. Each graphite layer contained 360 atoms. This graphite lattice was introduced in a simulation box with dimensions 29.52 Å × 32.60 Å × 50.0 Å. (3) A 4-layer thin graphite layer with dimensions of 45.510 Å × 9.78 Å × 44.029 Å was chosen for the interaction between the Ni100 cluster and the graphite surface. Each graphite layer contained 836 atoms. This graphite lattice was located inside a simulation box with dimensions of 45.510 Å × 50.0 Å × 44.029 Å. For each platelet edge, a Ni100 cluster was located at a distance of 2.0 Å away from the substrate surface and then the overall energy of the system was minimized without constraints. Afterward, the system was subjected to an initial period of thermal equilibration at 600 K for 25 ps. First, a Berendsen thermostat42 was used to achieve a fast convergence of the system temperature to the target value. A Nose´-Hoover chain thermostat43 was used subsequently to obtain a proper NVT ensemble over a period of 250 ps.

Sanz-Navarro et al.

Figure 1. QM (DFT) and ReaxFF hydrocarbon fragment binding energies to a Ni [111] slab.

We have investigated the changes in the local structure surrounding a cluster atom by means of a common-neighboranalysis (CNA) approach44 as well as the bond-length distribution. We have also calculated the strain in the bond length expressed as

strain )

dstrained - dunstrained dunstrained

(1)

where dstrained and dunstrained denote the strained and unstrained bond-length, respectively. Both Ni-Ni and C-Ni bonds have been taken into account in the analysis. 3. Results and Discussion 3.1. Parameterization of ReaxFF. The ReaxFF Ni/C/H parameters were optimized against a training set consisting of hydrocarbon fragment binding energies, obtained from DFT calculations, to the Ni[111] surface. These hydrocarbon fragments range from under-coordinated (single C atom, CH fragment) to fully coordinated species (acetylene, ethylene, and graphene) thus spanning the entire range of carbon binding options. Figure 1 shows the ReaxFF and DFT-results for these hydrocarbon fragment binding energies, demonstrating that ReaxFF gives an adequate description of the Ni-surface/CxHy interactions. In addition to these data we also included the full ReaxFF hydrocarbon,27 all-carbon27 and Ni-atom/hydrocarbon30 training set in the ReaxFF parametrization. The Ni-Ni ReaxFF parameters were obtained by training against DFT-derived equations of state for the fcc, bcc, a15, simple cubic, and diamond Ni-metal lattices and against DFT and experimental data for Ni-surface, cohesive and vacancy energies. The full set of ReaxFF parameters are available in the Supporting Information. 3.2. Structural Changes of Adsorbed Ni Clusters. As inferred from Figure 2, the number of adsorbed Ni atoms increased considerably for both the cluster bound to the armchair and zigzag edges during the equilbration stage. In contrast, the number of adsorbed Ni atoms of the cluster deposited on top of the basal plane remained oscillating slightly around its initial value. The increase in the number of adsorbed cluster atoms

Molecular Carbon-Supported Ni Clusters

J. Phys. Chem. C, Vol. 112, No. 33, 2008 12665

Figure 4. Snapshots taken during the production stage of binding to (a) armchair, (b) zigzag, and (c) basal plane carbon edges.

Figure 2. Number of adsorbed Ni atoms as a function of time during the thermal equilibration stage.

Figure 3. Evolution of the structure of the cluster over the thermal equilibration stage.

on the armchair and zigzag edges led to a flattening of the cluster shape as observed in Figure 3. A similar flattening of Pt clusters adsorbed on carbon nanosheets has already been observed in both experiments and MD simulations.34 In contrast, this phenomenon seems to be more pronounced for an Ni cluster. The energy to take an atom out from a 100-atom cluster can be approximated by the energy per atom, Eatom, in the cluster since such a quantity hardly changes in the reaction

Ni100(Ecluster ) 100Eatom) f Ni99(Ecluster ≈ 99Eatom) + Ni (2) Therefore, the fact that less energy is required to take out an atom from an Ni100 cluster (ReaxFF energy per atom ) 89 kcal/ mol) than from a Pt100 cluster (ReaxFF energy per atom ) 113 kcal/mol) seems to be the most plausible explanation for the notable difference in the ultimate number of adsorbed atoms. In Figure 4, three snapshots taken during the production stage of the binding of Ni100 clusters to each of the platelet edges are displayed. The rearrangement of the Ni atoms of the cluster adsorbed on the zigzag and armchair led to a layered structure. Each individual layer resembled a distorted fcc(111) surface. The distance between two adjacent layers was approximately 2.5 Å in the region near the cluster-graphite interface, but as the distance from the platelet increased, the interlayer distance in the cluster gradually converged to 2.0 Å, which coincides with the distance between two adjacent Ni(111) layers derived from the experimental fcc lattice parameter of 3.52 Å45 at room

temperature. As for the cluster deposited on the graphite basal plane, its structure evolves toward a more icosahedral-like shape as seen in Figure 4c. A CNA search found only a few atoms with a fcc-like signature (Figure 5). Interestingly, the cluster adsorbed on the armchair edge had more atoms with a hcp-like local structure (g signature) and some icosahedral-surface-edge atoms (isignature), while the cluster adsorbed on the zigzag edge had a bigger number of atoms in a truncated octahedral structure (f signature). The CNA analysis of the cluster deposited on top of the basal plane confirmed a transformation to a more icosahedral-like structure; an average of 35 atoms had a (a-l) signature, while almost no atoms had a (a-f) signature. In fact, the number of atoms with h signature (icosahedral spine or atoms on the line connecting the center of an icosahedron to a vertex atom) was considerably higher than that of the isolated cluster. In Figure 6, the bond length distribution of an isolated Ni100 cluster is compared with that of an Ni100 cluster adsorbed to each of the three distinct platelet edges. The mean Ni-Ni bond length of the isolated cluster is 2.603 Å ( 0.002. Upon adsorption, it is shown that there is an increase in the probability of bond lengths shorter than 2.45 Å as well as larger than 2.8 Å. The increase in the probability of both short and long Ni-Ni bond lengths (as compared to the mean bond-length for the isolated cluster) results in a very tiny change in the mean bond length distribution for the adsorbed clusters: 2.606 ( 0.004 Å (strain ) 1.2 × 10-3) for the cluster bound to the armchair, 2.600 ( 0.004 Å (strain ) -1.2 × 10-3) for the cluster bound to the zigzag edge and 2.613 ( 0.003 Å (strain ) -3.8 × 10-3) for the cluster bound to the basal plane. Therefore, a significant change in the bond length cannot be concluded by only looking at the mean bond length over the entire cluster. In contrast, a much more clear variation in the bond length is revealed when the mean bond length is represented as a function of the distance to the platelet surface. Such a distribution for each of the three platelet edges is represented in Figure 7. Clearly the bond length increases as the distance to the platelet edge does. Since the overall mean bond length remains unchanged after adsorption in the MD simulations, the bond length distribution in the region of the cluster away from the cluster must therefore increase after adsorption. Thus the increase in the bond-length probability for values larger than 2.8 Å in Figure 6 mainly corresponds to atoms away from the platelet, while the increase in probability for values lower than 2.45 Å is primarily connected to atoms in the region close to the platelet-cluster interface. Thus it is expected to see different catalytic behavior between both regions. Regarding to the number of atoms belonging to each of such regions, Figure 8 displays the atom distribution as a function of the distance to the platelet edge. This figure along with Figure 7 suggests that the number of atoms of the cluster is equally distributed in regions with negative and positive strain. Moreover, according to Figure 7, the change in the bond-length with respect to the average value (∼2.6Å) goes from -0.075

12666 J. Phys. Chem. C, Vol. 112, No. 33, 2008

Sanz-Navarro et al.

Figure 5. CNA signature for an Ni100 cluster without adsorption (a) and adsorbed on the armchair (b), zigzag (c), and basal (d) planes. A full explanation of the notation employed to represent local structure by CNA signatures can be found in reference.47 Briefly, a ) fcc bulk, b ) fcc (100) surface, c ) fcc (111) surface, d ) fcc (111)-(100) edge, e ) fcc(111)-(111) edge, f ) truncated octahedron, g ) icosahedral internal twinning plane, h ) icosahedral spine, i ) icosahedral surface edge, j ) icosahedral central atom, k ) icosahedral surface vertex or decahedral notch vertex, l ) truncated icosahedral vertex or decahedral notch vertex, m ) decahedral notch edge, and n ) tetrahedral edge. The figures only show atoms, whose local structure has been successfully determined.

Figure 6. Comparison of the Ni-Ni bond-length distribution of an isolated Ni100 cluster and that of an adsorbed Ni100 cluster to the three distinct edges of a carbon platelets.

Å to 0.075 Å for the armchair and zigzag edges and, therefore, the strain derived from the bond length runs from -3 × 10-2 to 3 × 10-2. Assuming an equal number of cluster atoms in regions with negative and positive strain and a linear variation in the bond length (which seems to be a good approximation according to Figure 7 and 8), the mean of the absolute value is therefore on the order of 10-2 for the armchair and zigzag edges. This order of magnitud in the bond-length strain matches quite well with that of the experimental value (∼2.5 × 10-2) of Ni clusters adsorbed in platelets.46 3.3. Preferential Binding Sites. Several preferential binding sites on the carbon substrate have been identified. Such binding sites are dependent on the edge as illustrated in Figure 9. For the armchair surface, each adsorbed Ni atom is covalently bound to two carbon atoms as shown in Figure 9a. Conversely, in our

Figure 7. Mean bond-length distribution for an Ni100 cluster as a function of the distance to the platelet edge.

previous study,34 we showed that a Pt100 cluster has preference for sites located in the gap between graphite layers. The reason for the preference of distinct binding sites seems to be connected with the ReaxFF binding energy of an adatom located in each of these preferential sites. Figure 10 shows the ReaxFF binding energy of an Ni and Pt adatom bound to three stable sites on the armchair edge. Visualization of the MD data reveals a certain preference for the Ni atoms to be found in a β site, whereas site γ was found to be the preferential site for Pt atoms in previosu MD simulations of an Pt100 cluster. As displayed in Figure 10, the energy difference between sites β and γ is large for a Pt adatom, highly favoring the γ site. In contrast, the energy difference between these two sites is much smaller for an Ni adatom, but still a γ site for an Ni adatom has lower energy than a β site. This is, however, just a rough simplification of the more complex situation when the cluster interacts with the

Molecular Carbon-Supported Ni Clusters

Figure 8. Normal atom distribution for an Ni100 cluster as a function of the distance to the platelet edge.

J. Phys. Chem. C, Vol. 112, No. 33, 2008 12667

Figure 10. Comparison of the ReaxFF energy between three stable binding sites for a Pt and Ni adatom bound to an armchair edge.

Figure 11. The cluster slides along the direction perpendicular to the graphite basal plane during the equilibration stage.

Figure 9. Preferential binding sites for the three edges.

in Figure 11, moving from graphite layer to graphite layer. This seems to be connected to the fact that, as pointed out for a bare platelet edge in the previous section, the Ni atoms have preference for a 2-fold site in the plane of a graphite layer,which are already occupied by H atoms. Conversely, for a Pt100 cluster,34 it has been observed that hydrogen atoms may be removed from the graphite surface and replaced by platinum atoms, and second, the platinum atoms can still find a stable location between graphite layers even though the graphite surface is hydrogen terminated. On the time scales in our simulations, we did not observe any breaking of C-H bonds by interaction with an Ni cluster. Over the first 50 ps, the cluster moves over the support, until it finds a more stable position. Then it remains immobilized over a nanosecond time scale. This suggests that a Ni100 cluster may diffuse over a hydrogen-terminated graphite surface until finally it finds a more stable location with respect to the support. Therefore, a higher dispersion of the Ni clusters, as compared to that of Pt clusters, is expected when the cluster is adsorbed to a hydrogen-terminated carbon edge. 4. Conclusions

graphite sheets at 600 K since also entropic and lattice distortion effects as well as many-body effects due to neighbor Pt atoms are added. For the zigzag edge, each adsorbed Ni atom was found to be bound to only one C atom (Figure 9b). Finally, the Ni atoms of the cluster in contact with the basal plane were positioned on top of a hollow site (hexagon’s center) as schematically represented in Figure 9c. Interestingly, the distance between two adjacent preferential sites on graphite for both the zigzag and basal planes is very similar to the Ni-Ni first neighbor distance (i.e., ∼2.5 Å), thereby the Ni atoms can be easily accommodated in adjacent preferential sites without a large price in the energy of the Ni-Ni bond. 3.4. Adsorption of Ni Clusters on Hydrogen-Terminated Graphite Edges. Unlike previous simulations, the cluster slides over the armchair edge during the equilibration stage, as shown

The adsorption of a Ni100 cluster adsorbed to both bare and hydrogen-terminated platelets have been simulated. It is found that the cluster atoms settle down on favorable sites, leading to an overall rearrangement of the cluster structure. Such a restructuring is less appreciable for adsorption on the graphite basal plane. A negative strain effect in the Ni-Ni bond is found at the platelet-cluster interface, whereas a positive strain effect is seen in regions of the cluster far away from the platelet. These opposite effects cancel each other, and therefore, the mean Ni-Ni bond length hardly changes upon adsorption. Furthermore, a different catalytic behavior between the two identified regions is expected. The hydrogen termination of the edges of the platelets decreases the strength of the cluster-substrate binding in such a way that the cluster slides on the carbon surface before finding a more stable site. Thus the dispersion of a cluster on the support

12668 J. Phys. Chem. C, Vol. 112, No. 33, 2008 can be enhanced by pretreating the support in a hydrogen-rich atmosphere, whereas hydrogen-free surfaces are required for applications in which a low dispersion of the nickel clusters is interesting to, for instance, optimize the surface-volume ratio of the catalyst material. Along with previous work,34 this study demonstrates that platinum and nickel metal catalysts behave differently upon adsorption on a carbon-nanostructured surface. Simulations have the advantage over experiments that various types of atoms, in this work nickel atoms close and far away from the carbon surface respect to it, can be labeled. In this work this has been fruitful to interpret the experimental bond length distribution. Although the work performed so far gives indications about mechanisms at the atomic scale, it is still not sufficient to lead to conclusions about how to enhance the catalyst performance of the clusters. However, the mechanisms found here and in our previous work can help to understand future experimental observations. Further work will examine how the curvature radius of the support may help to control the catalytic activity. Since the cluster-support binding is very different between platinum and nickel cluster, it is expected that the radius of curvature will have different effect depending on the catalyst material. Acknowledgment. C.F.S.-N., P.-O.Å., D.C., and M.R. have received support from the Norwegian Research Council through a Nanomat program “FUNMAT: Materials for Hydrogen Technology”, Project No. 158516/S10. C.F.S.-N. and P.-O.Å. have received a grant of computer time from the Norwegian Council and NTNU. Supporting Information Available: Full set of ReaxFF parameters. This material is available free of charge via the Internet at http://pubs.acs.org. References and Notes (1) Hutchings, G. J.; Siddiqui, M. R. H.; Burrows, A.; Kiely, C. J.; Whyman, R. J. Chem. Soc., Faraday Trans. 1997, 93, 187–188. (2) Guzman, J.; Gates, B. C. J. Phys. Chem. B 2002, 106, 7659–7665. (3) Marshall, A.; Børresena, B.; Hagen, G.; Tsypkin, M.; Tunold, R. Energy 2007, 32, 431–436. (4) Grigoriev, S. A.; Porembsky, V. I.; Fateev, V. N. Int. J. Hydrogen Energy 2006, 31, 171–175. (5) Nørskov, J. K.; Bligaard, T.; Logadottir, A.; Bahn, S.; Hansen, L. B.; Bollinger, M.; Bengaard, H.; Hammer, B.; Sljivancanin, Z.; Mavrikakis, M.; Xu, Y.; Dahl, S.; Jacobsen, C. J. H. J. Catal. 2002, 209, 275–278. (6) Nilsson, A.; Pettersson, L.; Hammer, B.; Bligaard, T.; Christensen, C.; Nørskov, J. K. Catal. Lett. 2005, 100, 111–114. (7) Rossmeisl, J.; Logadottir, A.; Nørskov, J. K. Chem. Phys. 2005, 319, 178–184. (8) Tersoff, J. Phys. ReV. Lett. 1988, 61, 2879–2882. (9) Tersoff, J. Phys. ReV. B 1989, 39, 5566–5568. (10) Brenner, D. W. Phys. ReV. B 1990, 42, 9458–9471. (11) Beardmore, K.; Smith, R. Phil. Mag. A 1996, 74, 1439–1466. (12) Stuart, S. J.; Tutein, A. B.; Harrison, J. A. J. Phys. Chem. 2000, 112, 6472–6486.

Sanz-Navarro et al. (13) Brenner, D.; Shenderova, O.; Harrison, J.; Stuart, S.; Ni, B.; Sinnott, S. J. Phys.: Condens. Matter 2002, 14, 783–802. (14) Albe, K.; Nordlund, K.; Averback, R. S. Phys. ReV. B 2002, 65, 195124. (15) Nord, J.; Albe, K.; Erhart, P.; Nordlund, K. J. Phys.: Condens. Matter 2003, 15, 5649–5662. (16) Ni, B.; Lee, K.-H.; Sinnott, S. B. J. Phys.: Condens. Matter 2004, 16, 7261–7275. (17) Juslin, N.; Erhart, P.; Tra¨skelin, P.; Nord, J.; Henriksson, K. O. E.; Nordlund, K.; Salonen, E.; Albe, K. J. Appl. Phys. 2005, 98, 123520. (18) Erhart, P.; Albe, K. Phys. ReV. B 2005, 71, 035211. (19) Erhart, P.; Juslin, N.; Goy, O.; Nordlund, K.; Muller, R.; Albe, K. J. Phys.: Condens. Matter 2006, 18, 6585–6605. (20) Pettiford, D. G. Phys. ReV. Lett. 1989, 63, 2480–2483. (21) Drautz, R.; Murdick, D. A.; Nguyen-Manh, D.; Zhou, X. W.; Wadley, H. N. G.; Pettifor, D. G. Phys. ReV. B 2005, 72, 144105. (22) Drautz, R.; Pettifor, D. G. Phys. ReV. B 2006, 74, 174117. (23) Aoki, M.; Nguyen-Manh, D.; Pettifor, D. G.; Vitek, V. Progr. Mater. Sci. 2007, 52, 154–195. (24) Drautz, R.; Zhou, X. W.; Murdick, D. A.; Gillespie, B. A.; Wadley, H. N. G.; Pettifor, D. G. Progr. Mater. Sci. 2007, 52, 196–229. (25) Gillespie, B. A.; Zhou, X. W.; Murdick, D. A.; Wadley, H. N. G.; Drautz, R.; Pettifor, D. G. Phys. ReV. B 2007, 75, 155207. (26) Rappe´, A. K.; Goddard III, W. A. J. Phys. Chem. 1991, 95, 3358– 3363. (27) van Duin, A. C. T.; Dasgupta, S.; Lorant, F.; Goddard, W. A., III. J. Phys. Chem. A 2001, 105, 9396–9409. (28) van Duin, A. C. T.; Strachan, A.; Stewman, S.; Zhang, Q.; Xu, X.; Goddard, W. A., III. J. Phys. Chem. A 2003, 107, 3803–3811. (29) Zhang, Q.; C¸agˇın, T.; van Duin, A.; Goddard, W. A.; Qi, Y.; Hector, L. G., Jr Phys. ReV. B 2004, 69, 045423. (30) Nielson, K. D.; van Duin, A. C. T.; Oxgaard, J.; Deng, W.-Q.; Goddard, W. A., III J. Phys. Chem. A 2005, 109, 493–499. (31) Cheung, S.; Deng, W.-Q.; van Duin, A. C. T.; Goddard, W. A., III. J. Phys. Chem. A 2005, 109, 851–859. (32) Han, S. S.; van Duin, A. C. T.; Goddard, W. A., III; Lee, H. M. J. Phys. Chem. A 2005, 109, 4575–4582. (33) Goddard, W. A., III; van Duin, A. C. T.; Chenoweth, K.; Cheng, M.-J.; Pudar, S.; Oxgaard, J.; Merinov, B.; Jang, Y. H.; Persson, P. Top. Catal. 2006, 38, 93–103. (34) Sanz-Navarro, C. F.; Åstrand, P-O.; Rønning, M.; Chen, D.; Jacob, T. J. Phys. Chem. A 2008, 112, 1392–1402. (35) Schultz, P. A. Unpublished. A description of the method is in Feibelman, P. J. Phys. ReV. B 1987, 35, 2626. (36) Paulus, P. M.; Goossens, A.; Thiel, R. C.; van der Kraan, A. M.; Schmid, G.; de Jongh, L. J. Phys. ReV. B 2001, 64, 205418. (37) Lee, K.; Zhang, J.; Wang, H.; Wilkinson, D. P. J. Appl. Electrochem. 2006, 36, 507–522. (38) Kvande, I.; Briskeby, S. T.; Tsypkin, M.; Hammer, N.; Rønning, M.; Sunde, S.; Chen, D.; Tunold, R. Top. Catal. 2007, 45, 81–85. (39) Hobday, S.; Smith, R. J. Chem. Soc., Faraday Trans. 1997, 93, 3919–3926. (40) Grigoryan, V. G.; Springborg, M. Phys. Chem. Chem. Phys. 2001, 3, 5135–5139. (41) Chemistry: The Central Science, 10th ed.; Prentice Hall: Upper Saddle River, NY, 2005. (42) Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; DiNola, A.; Haak, J. R. J. Chem. Phys. 1984, 81, 3684–3690. (43) Martyna, G. J.; Klein, M. L.; Tuckerman, M. J. Chem. Phys. 1992, 97, 2635–2643. (44) Faken, D.; Jo´nsson, H. Comput. Mater. Sci. 1994, 2, 279–286. (45) CRC Handbook of Chemistry and Physics, 83rd ed.; CRC: New York, 2002. (46) Ochoa-Ferna´ndez, E.; Chen, D.; Yu, Z.; Tøtdal, B.; Rønning, M.; Holmen, A. Surf. Sci. 2004, 554, L107-L112. (47) Cleveland, C. L.; Luedtke, W. D.; Landman, U. Phys. ReV. B 1999, 60, 5065–5077.

JP711825A