Nuclear Instruments and Methods in Physics Research B

Similar documents
Spontaneous Pattern Formation from Focused and Unfocused Ion Beam Irradiation

One and Two-Dimensional Pattern Formation on Ion Sputtered Silicon

Accepted Manuscript. Authors: Charbel S. Madi, Michael J. Aziz S (11) Reference: APSUSC 22249

Ion-induced pattern formation on Co surfaces: An x-ray scattering and kinetic Monte Carlo study

A Comparative Study on the Differences in the Evolutions of Thin Film Morphologies of Co-Al Binary System: Molecular Dynamics Study

Quantifying the order of spontaneous ripple patterns on ion-irradiated Si(111)

In-situ probing of near and below sputter-threshold ion-induced nanopatterning on GaSb(1 0 0)

Quantifying the order of spontaneous ripple patterns on ion-irradiated Si(111)

Molecular Dynamics Study of Plasma Surface Interactions for Mixed Materials

Molecular dynamics simulation of sputtering from a cylindrical track: EAM versus pair potentials

Wavelength tunability of ion-bombardment-induced ripples on sapphire

Medium-energy ion irradiation of Si and Ge wafers: studies of surface nanopatterning and

A comparison of molecular dynamic simulations and experimental observations: the sputtering of gold {1 0 0} by 20 kev argon

Ordered nanocrystals on argon ion sputtered polymer film

Multiple bifurcation types and the linear dynamics. of ion sputtered surfaces

Damage of low-energy ion irradiation on copper nanowire: molecular dynamics simulation

"Enhanced Layer Coverage of Thin Films by Oblique Angle Deposition"

Comparison of Molecular Dynamics and Binary Collision Approximation Simulations for Atom Displacement Analysis

PHYSICAL SELF-ASSEMBLY AND NANO-PATTERNING*

Making Functional Surfaces and Thin Films: Where are the Atoms?

Kinetic Monte Carlo: from transition probabilities to transition rates

Scaling during shadowing growth of isolated nanocolumns

Morphological instability of Cu nanolines induced by Ga+-ion bombardment: In situ scanning electron microscopy and theoretical model

Low Energy Nuclear Fusion Reactions in Solids

COMPUTATIONAL INVESTIGATION OF THE EFFECT OF CLUSTER IMPACT ENERGY ON THE MICROSTRUCTURE OF FILMS GROWN BY CLUSTER DEPOSITION

Molecular dynamics simulations of C 2,C 2 H, C 2 H 2,C 2 H 3,C 2 H 4, C 2 H 5, and C 2 H 6 bombardment of diamond (111) surfaces

Supporting Information Tuning Local Electronic Structure of Single Layer MoS2 through Defect Engineering

Nanopore sculpting with noble gas ions

Effect of Spiral Microwave Antenna Configuration on the Production of Nano-crystalline Film by Chemical Sputtering in ECR Plasma

Nanopatterning of silicon surfaces by low-energy ion-beam sputtering: dependence on the angle of ion incidence

Wang, Appl. Phys. Lett., vol.102, no. 5, 2013] is studied using molecular dynamic simulation.

Swift chemical sputtering of covalently bonded materials*

Nuclear Instruments and Methods in Physics Research B

Depth Distribution Functions of Secondary Electron Production and Emission

Electrostatic charging e ects in fast H interactions with thin Ar

Ion-irradiation-induced defects in bundles of carbon nanotubes

How ripples turn into dots: modeling ion-beam erosion under oblique incidence arxiv:cond-mat/ v2 [cond-mat.mtrl-sci] 12 May 2006.

Computer-Simulation Studies of Plasma- Surface Interactions

Dynamic Monte-Carlo modeling of hydrogen isotope reactive diffusive transport in porous graphite

31704 Dynamic Monte Carlo modeling of hydrogen isotope. reactive-diffusive transport in porous graphite

Stopping power for MeV 12 C ions in solids

Structural investigation of kev Ar-ion-induced surface ripples in Si by cross-sectional transmission electron microscopy

Sputtering Yield of Noble Gas Irradiation onto Tungsten Surface

Growth of carbon nanotubes by chemical vapor deposition

The effect of native oxide on ion-sputteringinduced nanostructure formation on GaSb surfaces

Electronic excitation in atomic collision cascades

Emergence of non-linear effects in nanocluster collision cascades in amorphous silicon

Quasi-periodic nanostructures grown by oblique angle deposition

Swift chemical sputtering of covalently bonded materials

Sputtering by Particle Bombardment

dynamics simulation of cluster beam deposition (1 0 0) substrate

Accurate detection of interface between SiO 2 film and Si substrate

Chapter 4. Surface defects created by kev Xe ion irradiation on Ge

Fabrication of InP nano pillars by ECR Ar ion irradiation

Multiscale modelling of D trapping in W

Effect of surface morphology on the sputtering yields. II. Ion sputtering from rippled surfaces

Surface Science 603 (2009) Contents lists available at ScienceDirect. Surface Science. journal homepage:

Molecular Dynamics Simulation of Chemical Sputtering of Hydrogen Atom on Layer Structured Graphite

Molecular dynamics simulations of plasma interaction with berylliumbased fusion reactor materials

Multiscale modelling of electrical breakdown at high electric field

A Measurement of Monoenergetic Neutrons from 9 Be(p,n) 9 B

Radiation Damage Modeling of Fused Silica in Fusion Systems

The Monte Carlo Simulation of Secondary Electrons Excitation in the Resist PMMA

Effects of substrate rotation in oblique-incidence metal(100) epitaxial growth

The role of electronic friction of low-energy recoils in atomic collision cascades

Table of Content. Mechanical Removing Techniques. Ultrasonic Machining (USM) Sputtering and Focused Ion Beam Milling (FIB)

Implantation Energy Dependence on Deuterium Retention Behaviors for the Carbon Implanted Tungsten

Fundamental science and synergy of multi-species surface interactions in high-plasma--flux environments

Irradiation-assisted substitution of carbon atoms with nitrogen and boron in single-walled carbon nanotubes

Comparison of tungsten fuzz growth in Alcator C-Mod and linear plasma devices!

The limits of volume reflection in bent crystals

Interface roughness evolution in sputtered WSi 2 /Si multilayers

Steady-State Statistical Sputtering Model for Extracting Depth Profiles from Molecular Dynamics Simulations of Dynamic SIMS

The use of MIM tunnel junctions to investigate kinetic electron excitation in atomic collision cascades

Computer simulation of the single crystal surface modification and analysis at grazing low-energy ion bombardment

ANNUAL REPORT INTERNATIONAL RESEARCH COLLABORATION AND SCIENTIFIC PUBLICATION

Molecular Dynamics Simulation of Hydrocarbon Reflection and Dissociation Coefficients from Fusion-Relevant Carbon Surfaces

Evolution of Cu nanoclusters on Cu(100)

Simulation of the cathode surface damages in a HOPFED during ion bombardment

A. Burenkov, P. Pichler, J. Lorenz, Y. Spiegel, J. Duchaine, F. Torregrosa

Supplementary Figure 1 Experimental setup for crystal growth. Schematic drawing of the experimental setup for C 8 -BTBT crystal growth.

Analysis of Ion Implantation Profiles for Accurate Process/Device Simulation: Analysis Based on Quasi-Crystal Extended LSS Theory

Filigree with a Sledge Hammer

S1. X-ray photoelectron spectroscopy (XPS) survey spectrum of

7 Wettability of 60 kev Ar-ion irradiated rippled-si surfaces

Chemical Sputtering of Carbon Materials due to Combined Bombardment by Ions and Atomic Hydrogen

Helium effects on Tungsten surface morphology and Deuterium retention

Interaction of ion beams with matter

Molecular dynamics simulations of the clustering and dislocation loop punching behaviors of noble gas atoms in tungsten

Two simple lattice models of the equilibrium shape and the surface morphology of supported 3D crystallites

Nuclear Instruments and Methods in Physics Research B 255 (2007)

Mechanisms of pattern formation in grazing-incidence ion bombardment of Pt(111)

Molecular dynamics simulation of deposition of nickel nanocluster on copper surface

Fluence Effects in C 60 Cluster Bombardment of Silicon

Monte Carlo simulation of thin-film growth on a surface with a triangular lattice

Atomistics of the Lithiation of Oxidized Silicon. Dynamics Simulations

Comparisons of DFT-MD, TB- MD and classical MD calculations of radiation damage and plasmawallinteractions

Mechanisms of inert gas impact induced interlayer mixing in metal multilayers grown by sputter deposition

, MgAl 2. and MgO irradiated with high energy heavy ions O 3

STM characterization of Cu thin films grown by direct ion deposition

Transcription:

Nuclear Instruments and Methods in Physics Research B 269 (2011) 2605 2609 Contents lists available at ScienceDirect Nuclear Instruments and Methods in Physics Research B journal homepage: www.elsevier.com/locate/nimb Anisotropic rearrangement of the substrate atoms during Ar bombardment on Pd(0 0 1) surface Sang-Pil Kim a,1, Byung-Hyun Kim a,b, Haeri Kim a,c, Kwang-Ryeol Lee a,, Yong-Chae Chung b, Jikeun Seo d, Jae-Sung Kim e a Computational Science Center, Interdisciplinary Fusion Technology Division, Korea Institute of Science and Technology, Seoul 136-791, Republic of Korea b Department of Materials Science and Engineering, Hanyang University, Seoul 133-791, Republic of Korea c Department of Physics, Ewha Womans University, Seoul 120-750, Republic of Korea d Department of Ophthalmic Optics, Chodang University, Muan 534-701, Republic of Korea e Department of Physics, Sookmyung Women s University, Seoul 140-742, Republic of Korea article info abstract Article history: Received 7 March 2011 Received in revised form 19 July 2011 Available online 26 July 2011 Keywords: Cascade induced rearrangement MD simulation Pd(0 0 1) surface Surface nanostructuring Using a three-dimensional molecular dynamics (MD) simulation, we investigated the atomic scale rearrangement that occurs on a Pd(0 0 1) surface after energetic bombardment by Ar at room temperature. High energy Ar bombardment provoked the significant rearrangement of Pd atoms in a ballistic manner with a fourfold symmetric lateral distribution aligned along the h1 1 0i direction. The MD simulation of uniform Ar bombardment at normal incidence on a Pd surface reproduced the experimentally observed fourfold symmetric nano-scale surface structure. The present result supports that the ballistic rearrangement of the substrate atoms plays an important role in the ion induced surface structure evolution. Ó 2011 Elsevier B.V. All rights reserved. 1. Introduction Self-organized nanostructures of either dots, holes or ripples produced by bombardment of energetic ions have drawn much attention, since the ordered structures of nanoscale can be produced by relatively simple and thus cheap processes [1 6]. Conventional understanding of these phenomena is based on the Bradley Harper model [7]. In this model, the pattern formation is mainly governed by two competing processes: the erosion of the substrate atoms which is proportional to the total energy deposited at each position as prescribed by the Sigmund theory of sputter erosion [8] and the structure relaxation via surface diffusion that can be enhanced by the ion bombardment. Recently, it was suggested from experimental observation and molecular dynamics (MD) simulations [9 14] that diverse surface reactions other than the sputter erosion and surface diffusion are provoked by ion bombardment, and significantly contribute to the morphological evolution of the bombarded surface. An example is the MD simulations of Ar bombardment on Si surface reported by Kalyanasundaram et al. [13]. They revealed that atoms in the bombarded region are not simply sputtered away, but major part of them are either displaced or redeposited around the region, thus forming Corresponding author. Tel.: +82 103 720 5491. E-mail address: krlee@kist.re.kr (K.-R. Lee). 1 Presently at School of Engineering, Brown University, Providence, RI 02912, USA. large craters at the point of impact and a heap of atoms nearby. This cascade induced rearrangement of the substrate atoms can be distinguished from the thermally activated surface diffusion in long time scale since it increases the surface area of the bombarded region and it occurs in a ballistic manner that can be simulated in typical MD time scale (a few tens ps). The MD simulations prompt the development of a new multi-scale kinetic model, which uses the crater function obtained by the simulation of the ion bombardment [15 17]. These models excluded the assumption that surface height change is proportional to the net energy deposited by the ion, but considered both the sputter erosion and the rearrangement by the ion bombardment. It was reported that the new kinetic model explains key experimental observations such as the stability of smooth surface during ion bombardment and the model quantitatively agrees with the ripple amplitude saturation [16,17]. The formalism also reproduced the classical Bradley Harper results under the appropriate simplifying assumptions [16]. Analysis of the rearranged substrate atoms such as its spatial distribution and quantitative yield per ion would be thus significant to understand the surface structure evolution by the energetic ion bombardment. In the present work, we focused on the rearrangement behavior of the substrate atoms when energetic Ar bombards a Pd(1 0 0) surface at room temperature using an empirical MD simulation technique. One of the motivations of the present work was the experimental observations by Kim et al. showing that 500 ev Ar bombardment on a Pd(1 0 0) surface resulted in a 0168-583X/$ - see front matter Ó 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.nimb.2011.07.093

2606 S.-P. Kim et al. / Nuclear Instruments and Methods in Physics Research B 269 (2011) 2605 2609 fourfold symmetric surface nanostructure that cannot be described by the conventional kinetic models [18]. The statistical analysis of 1000 single bombardment events showed that the distribution of the rearranged atom was found to be along h110i directions, resulting in the square-like distribution around the region of the bombardment. Randomly superposing the ballistic rearrangement generated the fourfold symmetric surface structures in agreement with the experimental observation. 2. Simulation procedure We simulated the bombardment of neutral energetic Ar on Pd(0 0 1) at 300 K for a variety of angles of incidence. Even if ionic Ar was used in most experiments, we assumed that the ionic Ar can be treated as neutral energetic Ar in MD simulation because free electrons on the Pd surface will significantly diminish the ionic contribution of the reaction. The kinetic energy of the Ar atoms ranged from 0.5 to 2.0 kev. Embedded atom method (EAM) potential was employed for Pd after benchmarking the potential parameters [19,20]. Strong repulsion between Pd atoms can be induced by the energetic Ar bombardment. We thus combined the EAM potential with the Ziegler, Biersack and Littmark (ZBL) potential [21] using a switching function [20]. The interaction between Ar and Pd was described by the ZBL potential. For Ar Ar interactions, the Born-Mayer type potential was employed [22]. Fig. 1 shows the geometry of the Ar-beam and the substrate orientation. The direction of the Ar-beam is defined by the polar angle h with respect to the surface normal, and the azimuthal angle u between the projection of the ion-beam on the surface and the h100i direction of the substrate. Polar angle ranged from 0 to 75 with two azimuthal angles, 45 and 22.5. The size of the substrate was 30 30 30a 3 o, where a o is the lattice constant of Pd (0.389 nm). A periodic boundary condition was adopted in both the x and y directions. The atomic positions of the bottom three layers were fixed to simulate a thick substrate. The temperature of the bottom half of the remaining layers was fixed at 300 K to play the role of a thermal bath during the Ar bombardment. All the other atoms were completely unconstrained. We used very short time step, 0.01 fs, for the first 3 ps after the Ar bombardment, because the substrate atoms would move considerably fast in the early period after the bombardment. The time step was 0.1 fs for the next 10 ps and 1.0 fs for the final 10 ps. This approach produces a sufficiently accurate description of the dynamic behavior with efficiency in the simulation. Large-scale Atomic/Molecular Massively Parallelized Simulator (LAMMPS) was used for all the simulations [23]. Fig. 2. (a) A snapshot of the Pd(0 0 1) surface at 2.3 ps after the commencement of normal incidence bombardment of 2.0 kev Ar at 300 K (color online). (b) Ratio of the rearrangement yield to the sputter erosion yield for various values of polar angle and the kinetic energy of Ar. 3. Results and discussion Fig. 2(a) shows the typical morphology of the Pd(0 0 1) substrate after bombardment by 2.0 kev Ar at normal incidence. Color Fig. 3. The total energy change of the system and mean square displacement of the rearranged Pd atoms during the simulation of 0.5 kev Ar bombardment on the Pd(0 0 1) surface. Fig. 1. The geometry of the Ar beam and the substrate orientation. of the atom represents the height of the atomic position. As the energetic Ar dissipates its energy to the substrate, local melting followed by micro-explosion and rapid relaxation of the lattice atoms occurs. This surface reaction results in both the sputter-erosion and the rearrangement of the substrate atoms that generates a crater at the bombarding position and an atomic heap around

S.-P. Kim et al. / Nuclear Instruments and Methods in Physics Research B 269 (2011) 2605 2609 2607 Fig. 4. The lateral distribution of the rearranged and eroded Pd atoms for Ar ions of 0.5 kev bombarding a Pd(0 0 1) surface: the lateral distribution of (a) the rearrangement for normal incidence of Ar and (b) the eroded atom position in the crater for normal incidence of Ar. Fig. 5. The lateral distribution of the rearranged Pd atoms for Ar ions of 0.5 kev bombarding a Pd(0 0 1) surface: the lateral distribution of the rearrangement for (a) h = 45 and u =45, (b) h =60 and u =45, (c) h =45 and u = 22.5 and (d) h =60 and u = 22.5. the crater as indicated by an arrow in Fig. 2(a). The present observation agrees with the previous MD simulations that reported the details of the surface reaction during high-energy ion bombardment [10 13]. In this paper, we defined the rearranged atoms as those located above the initial surface after finishing the surface reaction or after 23 ps from the Ar bombardment as will be discussed in Fig. 3. Fig. 2(b) shows the ratio of the rearrangement yield to the sputtering yield for various polar angles and incidence energies of Ar. The ratio was almost independent of the incidence energy within uncertainty of the calculated yields but gradually decreased from about 3.0 to 2.0 as the polar angle increased. This result shows that the rearrangement of the substrate occurs two or three times more in abundance than the sputter erosion. Fig. 3 shows the change in specimen temperature and the mean square displacement of the rearranged atoms when 500 ev Ar bombarded on Pd(0 0 1) surface. By depositing kinetic energy to the substrate, high temperature spike dissipated rapidly (for less than 1 ps) followed by the gradual decrease in the temperature of the specimen. The mean square displacement of the rearranged atoms showed that the rearrangement finished within 2.5 ps. Hence, it can be inferred that most of the surface processes were completed within about 2.5 ps after the Ar bombardment. After

2608 S.-P. Kim et al. / Nuclear Instruments and Methods in Physics Research B 269 (2011) 2605 2609 the MD simulation for 23 ps, kinetic Monte Carlo (kmc) simulation was applied to investigate the surface diffusion in long time scale. It was hardly observed for the surface diffusion to occur at 300 K after the ballistic rearrangement. Fig. 4(a) shows the lateral distribution of the rearranged atoms for normal incidence Ar of kinetic energy of 0.5 kev bombarding a Pd(0 0 1) surface. The distribution was obtained by collecting the distributions of 1000 independent bombardment events. The intensity of each data points is proportional to the number of the rearranged atoms. It is noted that the rearranged atoms tend to align along the h1 1 0i directions, resulting in a fourfold symmetric distribution. This behavior is in contrast to the rearrangement on Si (0 0 1) surface that shows an isotropic lateral distribution without resemblance to the symmetry of the substrate [13]. The difference between Pd and Si would originate from the fact that the region around the impact point of Si substrate becomes amorphous, while the metallic substrate swiftly recovers its crystallinity after each bombardment of energetic particle. It tells that the ballistic rearrangement in Pd is guided by the crystallinity of the substrate. Fig. 4(b) shows the collected distribution of the initial position of eroded atoms. In contrast to the distribution of the rearranged atoms, that of eroded atoms seems less regular, suggesting that the bond breaking and thus the distribution of the eroded atoms did not reflect the symmetry of the substrate. Instead, the rearranged atoms, approaching equilibrium to the substrate and losing their initial high kinetic energies, would diffuse preferentially along the close packed directions where the diffusion barrier is minimal, before they are stabilized to display fourfold symmetric distribution with edges along h1 1 0i. This behavior would be also determined by the excess energy of the surface atoms. Since the atoms are mostly packed along h1 1 0i directions, the rearranged atoms aligned along h1 1 0i directions would have the lowest excess energy. It must be noted that local incidence angle to the surface can be different from the macroscopic incidence angle as the rough surface in nanometer scale evolves. It is thus required to investigate the lateral distribution with various incidence conditions. Figs. 5(a) and (b) are the lateral distributions of the rearranged atoms when Ar bombards the surface at h =45 and 60 respectively with u =45 (parallel to the h1 0 0i direction). The arrows show the projected beam directions on the surface and the y axis of the graph is in h010i direction. Similar fourfold symmetry was observed for h =45, although more atoms are displaced to the front of the ion beam. As h increased to be 60 or with the more glancing incidence, a large fraction of the rearranged atoms were located alongside the beam direction while the number on the front and the rear sides of the incidence beam direction decreased (see Fig. 5(b)). High grazing incidence along h110i direction resulted in the distribution proximate to twofold symmetry, although the characteristics of square-like distribution still remained. Figs. 5(c) and (d) show the distribution when the azimuthal angle of the incident Ar beam is in the less symmetric direction, h210i which corresponds to u = 22.5. The projection of the Ar beam trajectory on the surface is indicated by arrows in the figures. The distribution of the rearranged atoms preferentially aligned along the h110i direction, giving square-like patterns of the distribution as shown in Fig. 5(c) and (d). However, the rearrangement density is enhanced, most notably on the front edge of square pattern, and the maximum of the distribution is rotated from h110i to the incidence direction. Especially when h =60, the distribution with u = 22.5 (Fig. 5(d)) is closer to the square-like one than that with u =45 (Fig. 5(b)). This enhancement seems to be caused by the fact that the incident Ar along h2 1 0i has shorter channeling distance than that along h1 1 0i, and more energy dissipation of the incident Ar occurs near surface region. Similar behaviors were observed for the higher kinetic energies of the incidence Ar. Fig. 6. (a) A snapshot of the Pd(0 0 1) surface after 4200 Ar atoms of 0.5 kev were bombarded at normal incidence at 300 K (color online). (b) The autocorrelation function of the surface shown in (a). The statistical analysis of the simulations showed that the lateral distribution of the rearranged Pd atoms has a strong tendency to be aligned along the h1 1 0i direction, irrespective of the incident Ar direction. This result means that the fourfold symmetric lateral rearrangement will be retained even if the energetic Ar at macroscopically normal incidence has a different local incidence angle on a roughened surface. It should be thus interesting to examine if the local fourfold symmetric distribution of the rearranged atoms can lead to the fourfold symmetric surface nanostructures as observed in the experiment [18]. To answer the question, we simulated the morphological evolution of a Pd(0 0 1) surface at 300 K by bombarding 4200 Ar atoms of 0.5 kev at normal incidence.

S.-P. Kim et al. / Nuclear Instruments and Methods in Physics Research B 269 (2011) 2605 2609 2609 The surface area of the substrate composed of 48,000 Pd atoms was 40 40a 2 o. We adopted an acceleration method to realize the simulation because ion bombardment on the surface of area 40 40a 2 o is very rare in MD time scale under the real experimental condition. Simulation of the surface reaction at the experimental flux is thus impractical in terms of computation time. Each Ar atom was thus consecutively incident to the surface every 23 ps around which we found no further mass transport near the surface as shown in Fig. 3. Because no more significant surface reaction is expected after 23 ps from Ar bombardment, the present MD simulation would mimic the phenomena in real situation. Under this acceleration scheme, the flux was equivalent to 1.80 10 22 /cm 2 s and the fluence was set to 1.74 10 15 /cm 2. We rescaled the temperature of the system to 300 K prior to the projection of each Ar atom. Therefore, surface diffusion over long time scale was not considered in this simulation as tested by the kinetic MC simulation at room temperature. The (x, y) position of Ar incidence was randomly selected to mimic uniform Ar irradiation. Fig. 6(a) shows the Pd surface after the bombardment of 4200 Ar atoms. Color of the surface atom represents the height of the position according to the color index bar. The snapshot of Fig. 6(a) showed that many Pd atoms are placed above the initial surface even after 4200 Ar bombardment (indicated by white arrows). The only way for the Pd atoms to be placed above the initial surface is the ballistic rearrangement caused by the energetic Ar bombardment. In the early stage of the surface structure evolution, surface nano dots are formed by the (red colored) Pd atoms agglomerated above the initial surface surrounded by the eroded region. Therefore, the present snapshot shows that the ballistic rearrangement plays a significant role in the surface structure evolution. Fig. 6(b) shows the autocorrelation function of the surface shown in Fig. 6(a). Despite the statistical noise due to the small substrate size and the low fluence, the surface features are aligned along the h110i direction with fourfold symmetry. This correlation function is comparable with the experimental results of Kim et al. who observed a fourfold symmetric nano-dot pattern on an Ar + ion-sputtered Pd(0 0 1) surface for normal incidence at 300 K [18], where the dots are aligned along the h110i direction. 4. Conclusions We investigated the atomic scale rearrangement that occurs on a Pd(0 0 1) surface during the energetic Ar bombardment at room temperature using molecular dynamics simulation. It was observed in wide range of the incidence angles of energetic Ar beam that the rearrangement of Pd atoms occurs in a ballistic manner with a fourfold symmetric distribution aligned along the h1 1 0i direction. This result shows that the fourfold symmetry of the rearrangement can be maintained even when the local incidence angle varies with the surface structure evolution. The simulated Pd surface structure evolved by the uniform Ar beam bombardment of normal incidence was comparable with that of the experimentally observed fourfold symmetric nano-dots, which exposed the important contribution of the ballistic rearrangement to the surface structure evolution. Acknowledgment This research was supported by the Converging Research Center Program through the Ministry of Education, Science and Technology (2010K000992) and the KIST Core Capability Enhancement Program (2E21580). References [1] S. Facsko, T. Dekorsy, C. Koerdt, C. Trappe, H. Kurz, A. Vogt, H.L. Hartnagel, Science 285 (1999) 1551. [2] W.L. Chan, E. Chason, J. Appl. Phys. 101 (2007) 121301. [3] U. Valbusa, C. Boragno, F.B. de Mongeot, J. Phys.: Condens. Matter 14 (2002) 8153. [4] M. Kalf, G. Comsa, T. Michely, Surf. Sci. 486 (2001) 103. [5] O. Malis, J.D. Brock, R.L. Headrick, M.-S. Yi, J.M. Pomeroy, Phys. Rev. B 66 (2002) 035408. [6] M.V.R. Murty, T. Curcic, A. Judy, B.H. Cooper, A.R. Woll, J.D. Brock, S. Kycia, R.L. Headrick, Phys. Rev. Lett. 80 (1998) 4713. [7] R.M. Bradley, J.M.E. Harper, J. Vac. Sci. Technol. A 6 (1988) 2390. [8] P. Sigmund, J. Mater. Sci. 8 (1973) 1545. [9] C.M. McQuaw, E.J. Smiley, B.J. Garrison, N. Winograd, Appl. Surf. Sci. 231 232 (2004) 39. [10] K. Nordlund, Phys. World 14 (2001) 22. [11] E.M. Bringa, K. Nordlund, J. Keinonen, Phys. Rev. B 64 (2001) 235426. [12] H. Gades, H.M. Urbassek, Phys. Rev. B 50 (1994) 11167. [13] N. Kalyanasundaram, M. Ghazisaeidi, J.B. Freund, H.T. Johnson, Appl. Phys. Lett. 92 (2008) 131909. [14] M. Moseler, P. Gumbsch, C. Casiraghi, A.C. Ferrari, J. Robertson, Science 309 (2005) 1545. [15] B. Davidovitch, M.J. Aziz, M.P. Brenner, Phys. Rev. B 76 (2007) 205420. [16] S. Norris, M.P. Brenner, M.J. Aziz, J. Phys.: Condens. Matter 21 (2009) 224017. [17] N. Kalyanasundaram, J.B. Freund, H.T. Johnson, J. Phys.: Condens. Matter 21 (2009) 224018. [18] T.C. Kim, C.-M. Ghim, H.J. Kim, D.Y. Noh, N.D. Kim, J.W. Chung, J.S. Yang, Y.J. Chang, T.W. Noh, B. Kahng, J.-S. Kim, Phys. Rev. Lett. 92 (2006) 246104. [19] S.M. Foiles, M.I. Baskes, M.S. Daw, Phys. Rev. B 33 (1986) 7983. [20] S.-P. Kim, S.-J. Kim, D.-Y. Kim, Y.-C. Chung, K.-R. Lee, J. Kor. Vac. Soc. 17 (2008) 81. [21] J.F. Ziegler, J.P. Biersack, U. Littmark, The Stopping and Range of Ions in Solids, Pergamon, New York, 1985. [22] A.A. Abrahamson, Phys. Rev. 178 (1969) 76. [23] S.J. Plimpton, J. Comput. Phys. 117 (1995) 1.