Surface morphologies associated with thermal desorption: Scanning tunneling microscopy studies of Br GaAs(110)

Similar documents
Atomic processes during Cl supersaturation etching of Si(100)-(2Ã 1)

Equilibrium morphologies for Cl-roughened Si 100 at K: Dependence on Cl concentration

Author(s) Okuyama, H; Aruga, T; Nishijima, M. Citation PHYSICAL REVIEW LETTERS (2003), 91(

SUPPLEMENTARY INFORMATION

Role of surface stoichiometry in the Cl 2 /GaAs(001) reaction

Supporting Information

Surface Morphology of GaN Surfaces during Molecular Beam Epitaxy Abstract Introduction

Scanning Tunneling Microscopy Studies of the Ge(111) Surface

Reflection high energy electron diffraction and scanning tunneling microscopy study of InP(001) surface reconstructions

Supplementary Information:

Evidence for structure sensitivity in the high pressure CO NO reaction over Pd(111) and Pd(100)

Supporting Online Material for

Si/XeF2 etching: Temperature dependence

Crystalline Surfaces for Laser Metrology

Acidic Water Monolayer on Ruthenium(0001)

Surface atoms/molecules of a material act as an interface to its surrounding environment;

Supporting Information

CHAPTER 5 ATOMIC HYDROGEN

STM spectroscopy (STS)

Site seectivity in the initial oxidation of the Si 111-7=7 surface

desorption (ESD) of the O,/Si( 111) surface K. Sakamoto *, K. Nakatsuji, H. Daimon, T. Yonezawa, S. Suga

Selectivity in the initial C-H bond cleavage of n-butane on PdO(101)

Molecular dynamics simulations of fluorine molecules interacting with a Siˆ100 (2 1) surface at 1000 K

Deuterium and fluorine radical reaction kinetics on photoresist*

Evidence for partial dissociation of water on flat MgO(1 0 0) surfaces

Indium incorporation and surface segregation during InGaN growth by molecular beam epitaxy: experiment and theory

Optimizing Graphene Morphology on SiC(0001)

Adsorption, desorption, and diffusion on surfaces. Joachim Schnadt Divsion of Synchrotron Radiation Research Department of Physics

Applied Surface Science CREST, Japan Science and Technology Corporation JST, Japan

SUPPLEMENTARY INFORMATION

Mater. Res. Soc. Symp. Proc. Vol Materials Research Society

Surface Physics Surface Diffusion. Assistant: Dr. Enrico Gnecco NCCR Nanoscale Science

Hydrogen termination following Cu deposition on Si(001)

LOW-TEMPERATURE Si (111) HOMOEPITAXY AND DOPING MEDIATED BY A MONOLAYER OF Pb

The Transition State for Surface-Catalyzed Dehalogenation: C I Cleavage on Pd(111)

Kinetic Monte Carlo simulation of nucleation on patterned substrates

Appearance Potential Spectroscopy

Oligomer desorption during heterogeneous catalytic synthesis of polymers

SUPPLEMENTARY INFORMATION

1 Corresponding author:

Direct observation of a Ga adlayer on a GaN(0001) surface by LEED Patterson inversion. Xu, SH; Wu, H; Dai, XQ; Lau, WP; Zheng, LX; Xie, MH; Tong, SY

5.8 Auger Electron Spectroscopy (AES)

The first three categories are considered a bottom-up approach while lithography is a topdown

arxiv:chem-ph/ v2 11 May 1995

Supporting information for. Direct imaging of kinetic pathways of atomic diffusion in. monolayer molybdenum disulfide

Curriculum Vitae December 2006

Spectroscopy at nanometer scale

Solid Surfaces, Interfaces and Thin Films

SUPPLEMENTARY INFORMATION

Molecular Dynamics on the Angstrom Scale

Magic numbers in Ga clusters on GaAs (0 0 1) surface

Observation of Bulk Defects by Scanning Tunneling Microscopy and Spectroscopy: Arsenic Antisite Defects in GaAs

Scanning tunneling microscopy of monoatomic gold chains on vicinal Si(335) surface: experimental and theoretical study

Supporting Information for Ultra-narrow metallic armchair graphene nanoribbons

Scanning tunneling spectroscopy of C 60 adsorbed on Si(100)-(2 1)

Cross-Section Scanning Tunneling Microscopy of InAs/GaSb Superlattices

G. L. KELLOGG. Sandia National Laboratories Albuquerque, NM USA. 1. Abstract

The effectiveness of HCl and HF cleaning of Si 0.85 Ge 0.15 surface. Stanford Synchrotron Radiation Lab, Menlo Park, CA 94025

Self-Assembly of Two-Dimensional Organic Networks Containing Heavy Metals (Pb, Bi) and Preparation of Spin-Polarized Scanning Tunneling Microscope

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

Cross-sectional scanning tunneling microscopy of InAsSb/InAsP superlattices

Dynamics of step fluctuations on a chemically heterogeneous surface of AlÕSi )Ã)

Characterization of metal clusters (Pd and Au) supported on various metal oxide surfaces (MgO and TiO 2 )

Photo-Reactivity. Jerusalem, Israel. Israel

1. Robust hexagonal rings on Cu(111) Figure S1 2. Details of Monte Carlo simulations

Simultaneous Etching and Oxidation of Vicinal Si(100) Surfaces: Atomistic Lattice-Gas Modeling of Morphological Evolution

Outlines 3/12/2011. Vacuum Chamber. Inside the sample chamber. Nano-manipulator. Focused ion beam instrument. 1. Other components of FIB instrument

C. D. Lee and R. M. Feenstra Dept. Physics, Carnegie Mellon University, Pittsburgh, PA 15213

SUPPLEMENTARY FIGURES

Vapor-Phase Cutting of Carbon Nanotubes Using a Nanomanipulator Platform

Hydrogen-Induced Changes in Palladium Surface Structure: Field Ion Microscopy Study

Model for nucleation in GaAs homoepitaxy derived from first principles

Gregory S. Rohrer,* Victor E. Henrich, Dawn A. Bonnell

Effects of methanol on crystallization of water in the deeply super cooled region

Vacancy migration, adatom motion, a.nd atomic bistability on the GaAs(110) surface studied by scanning tunneling microscopy

Growth and Morphology of Pb Phases on Ge(111) Y. Sato 1 and S. Chiang 2,3. Department of Physics. University of California Davis. 1 Shields Ave.

J. Günster, J. Stultz, S. Krischok, and D. W. Goodman a),b) Department of Chemistry, Texas A&M University, College Station, Texas

Supporting Information. for. Angew. Chem. Int. Ed. Z Wiley-VCH 2003

Supplementary Materials for

b imaging by a double tip potential

Physical and chemical properties of high density atomic oxygen overlayers under ultrahigh vacuum conditions: 1 1 -O/Rh 111

Surface physics, Bravais lattice

Supplementary Information for Topological phase transition and quantum spin Hall edge states of antimony few layers

Lecture 10 Thin Film Growth

Pt-induced nanowires on Ge(001): A density functional theory study

Structure and Energetics of P-rich GaP(001) Surfaces

Catalytic Activity of IrO 2 (110) Surface: A DFT study

Reconstruction and intermixing in thin Ge layers on Si 001

Scanning Tunneling Microscopy. how does STM work? the quantum mechanical picture example of images how can we understand what we see?

Table 1: Residence time (τ) in seconds for adsorbed molecules

Supplementary Information for Solution-Synthesized Chevron Graphene Nanoribbons Exfoliated onto H:Si(100)

1 IMEM-CNR, U.O.S. Genova, Via Dodecaneso 33, Genova, IT. 2 Dipartimento di Fisica, Università di Genova, Via Dodecaneso 33, Genova, IT

= 6 (1/ nm) So what is probability of finding electron tunneled into a barrier 3 ev high?

Hydrogen adsorption by graphite intercalation compounds

Low-temperature Scanning Tunneling Spectroscopy of Semiconductor Surfaces

Consequences of Surface Oxophilicity of Ni, Ni-Co, and Co Clusters on Methane. Activation

Enantiospecific Desorption of Chiral Compounds from Chiral Cu(643) and Achiral Cu(111) Surfaces

Scanning tunneling microscopy of Ir adsorbed on a Ge(111) surface

Soft x-ray photoelectron spectroscopy study of the reaction of XeF 2 with GaAs

CO Adsorption on Co(0001)-Supported Pt Overlayers

Transcription:

Surface morphologies associated with thermal desorption: Scanning tunneling microscopy studies of Br GaAs(110) C. Y. Cha, J. Brake, B. Y. Han, D. W. Owens, and J. H. Weaver a) Department of Materials Science and Chemical Engineering, University of Minnesota, Minneapolis, Minnesota 55455 Received 16 October 1996; accepted 31 January 1997 Scanning tunneling microscopy was used to characterize the developing surface morphology found during typical temperature programmed desorption experiments for halogen GaAs. Surfaces exposed to Br 2 at 300 K were heated to temperatures between 450 and 675 K, followed by scanning at room temperature. This made it possible to relate the temperature-dependent gas phase etch product distribution to the surface structure and thereby examine atomic-level surface processes associated with the evolution of volatile products. We associate the desorption of GaBr 3 around 500 K with the initiation of single-layer-deep terrace pits. Desorption of GaBr and As 2 above 600 K accounts for the lateral enlargement of the pits. 1997 American Vacuum Society. S0734-211X 97 00203-5 I. INTRODUCTION The interaction of halogens with semiconductor surfaces is of technological importance because chlorine- and bromine-containing reactants are used in dry-etching processes. Accordingly, much effort has been devoted to define and understand the relevant surface processes. Pivotal understanding of the underlying chemistry was gained by temperature programmed desorption TPD experiments that identified the volatile reaction products. Detailed work by Tyrrell et al., 1 Su et al., 2 Ludviksson et al., 3 and French et al. 4 demonstrated that the distribution of these products will depend on the initial halogen exposure initial surface concentration and on the surface temperature. In work with the Cl GaAs 100 system, they showed the desorption of GaCl, GaCl 3, and As 2 As 4, with rates that varied with processing conditions. Etching reactions are more complex than elemental desorption processes because they require the formation of volatile products involving substrate atoms. Since these formation and desorption processes have activation energies that depend on the particular surface site, the relative importance of the various reaction pathways should vary with the local morphology. In turn, the local morphology changes as material is removed. Unfortunately, there have been few studies that provide snapshots of the surface structure so as to complement gas-phase product analysis. Scanning tunneling microscopy STM makes it possible to observe surface structures with atomic resolution. Previously, it was used to characterize GaAs 110 surface etching at elevated temperature during continuous exposure to molecular halogen beams. 5 In those experiments, the steady state surface concentration of halogens reflected the balance between halogen adsorption and the desorption of the various volatile reaction products. It follows that the concentration depends on temperature because of the different desorption energies for the products. a Electronic mail: weave001@maroon.tc.umn.edu In the present work, we used STM to investigate Brexposed GaAs 110 after heating to successively higher temperatures between 450 and 675 K. The initial submonolayer surface concentration was selected by exposure at a temperature below the desorption threshold. The images, obtained at room temperature after each heating cycle, reflect conditions representative of surfaces realized in thermal desorption investigations where the concentration diminishes and reflects the activation of different processes. We found evidence for two distinct stages of surface modification associated with two stages of etching identified previously in desorption studies. 1 4 The first is associated with the creation of singlelayer-deep etch pits on terraces. This sets in at about 500 K where TPD of Cl GaAs has shown GaCl 3 desorption. The second stage, activated above 600 K, involves lateral enlargement of the etch pits. This stage can be related to etching via GaCl loss and the desorption of As 2 As 4. The TPDlike etching studied here produces quite different surface morphologies than those reached with steady state etching 5 because of the dependence of the reaction channels on surface halogen concentration. II. EXPERIMENTAL PROCEDURE Our experiments were conducted in an ultrahigh vacuum system housing a Park Scientific Instruments STM. Clean GaAs 110 surfaces were obtained by cleaving p-type posts Zn doped at 2 10 18 cm 3. Electrochemically etched tungsten wire was used for the tips of the STM. These tips were cleaned by electron bombardment prior to use. STM topographs were acquired in the constant current mode, with the sample bias between 1.5 and 3.5 V. A solid state AgBr CdBr 2 electrochemical cell was used as a source for Br 2. The Br exposure is expressed in ma s, corresponding to operation of the cell with a current of 5 A for a defined duration. For reference, an exposure of 1 ma s at 300 K results in about 0.5 monolayer ML surface coverage, determined by counting the adsorbate features. After exposure to Br 2 at 300 K, the sample was heated by tungsten filaments 605 J. Vac. Sci. Technol. B 15(3), May/Jun 1997 0734-211X/97/15(3)/605/5/$10.00 1997 American Vacuum Society 605

606 Cha et al.: Morphologies associated with thermal desorption 606 mounted on the holder. 6 Upon reaching the designated temperature, 450 T 675 K as measured with a pyrometer, the sample was cooled to ambient and was characterized by STM. The process was then repeated for the next higher temperature. Thermally activated reactions can be described by the Arrhenius form k exp( E A /kt) where E A is the activation energy and is the probing frequency of the reaction. With a given heating time, only surface reactions that are fast on the experimental time scale will lead to measureable changes of the surface morphology. Thus, after each heating cycle the surface reflects a configuration where reactions that are fast have saturated at the maximum temperature. Extended annealing at the maximum temperature would lead to gradual changes of the surface structure, but would not initiate etching pathways with significantly higher activation energies. We verified this by annealing samples for 20 min at maximum temperature after first reaching this temperature during desorption cycles. The fact that the maximum temperature and not the time at that temperature determined the morphology justifies our approach of simulating TPD-related surface structures with a series of heating cycles. III. RESULTS A. Previous TPD results In TPD, a surface is initially exposed to a specified amount of gas phase reactant at a fixed temperature. It is then heated in a controlled fashion while the desorbing species are analyzed with a mass spectrometer. 7 The energetics and order of surface reactions can be deduced from the desorption temperatures, the intensities of particular masses, and the dependencies on initial coverage and ramping rate. The published TPD results for Cl/GaAs(100) agree that the primary desorbing species is GaCl 3 below 600 K, but that desorption of GaCl, As 2, and As 4 dominates above 600 K. 1 4 The intensity of the trihalide contribution is enhanced by higher initial coverages up to saturation at 300 K. While AsCl 3 is not observed following exposure at 300 K, this species can be formed when multilayers of Cl 2 are condensed on GaAs at low temperature. In agreement with this, photoemission experiments have shown AsBr 3 on the surface at low temperature and its desorption at about 150 K. 8 B. STM results Figure 1 a is a STM image of a GaAs 110 surface that was exposed to 1.5 ma s of Br 2 at 300 K. This surface is nearly saturated with Br adatoms. Figures 1 b 1 e are images of the same surface after heating to 450, 500, 600, and 675 K and with no further Br exposure. As shown by Patrin and Weaver, 5 Br atoms bond to Ga and As sites, forming adsorption islands in which every Ga atom and every second As atom is Br terminated. The maximum Br coverage that can be attained with this structure is 0.75 ML where 1 ML is the atom density of GaAs 110. The bright objects in Figs. 1 a 1 c correspond to Br atoms chemisorbed on As sites, producing mixed (2 1) and c(2 2) symmetry see Fig. 3 of Ref. 5 for models of the structures. The Br atoms on Ga sites are lower in height and are less visible, appearing as the darker background. 5 Island formation reflects an attractive interaction among Br adsorption sites. This picture is supported by photoemission studies by Stepniak et al. 9 and by first-principles calculations by Khoo and Ong 10 and by Corkill and Chelikowsky. 11 In the low coverage regime, the Br atoms form narrow chains elongated along 001. The underlying adsorbate interaction stretches the adsorption islands along 001, as is rationalized by the calculations by Corkill and Chelikowsky. 12 These chains coalesce at higher coverage. The image of the surface heated to 450 K in Fig. 1 b shows no significant etching, and the surface is almost saturated with Br structures. There is, however, an increased occurrence of (2 1) structures compared to Fig. 1 a. This indicates that the (2 1) island structure is energetically favored relative to the c(2 2) islands. Although they have the same adsorbate density, the (2 1) structure has As rows free of Br along 001. It is possible that (2 1) enables a rowwise buckling with 2 periodicity along 110 that allows better relaxation than c(2 2), where the vacant As sites are staggered. Heating to 500 K significantly reduces the density of bright features attributed to Br on As sites, Fig. 1 c. The remnant Br appears in narrow chains. By analogy to the low coverage results of Patrin and Weaver, 5 we associate these chains to elongated Br As chains with double rows of Br Ga features at lower height. The zigzag row structure of exposed GaAs 110 is visible in some regions. We associate the dark patches with the onset of etching labeled II. Most of the surface remains covered with poorly ordered adsorption features, and it is impossible to quantify the amount of Br that has desorbed by etching. Heating to 600 K results in the surface structure shown in Fig. 1 d. Here, the zigzag row structure of GaAs 110 is clearly visible, with lighter areas marking the original top layer labeled I and darker ones revealing the second layer labeled II. The second layer remains intact with almost no pits in it. In contrast to Figs. 1 a 1 c, the regular Br features have vanished in Fig. 1 d. While the top layer is decorated with irregular structures, their density is much reduced. We attribute them to Br bonded to Ga and to As released during etching. As 2 was observed to desorb at 600 K in the TPD spectra from Cl/GaAs 100. 1 4 The fact that GaCl 3 desorbs at lower temperature implies that the surface will be As rich between the onset of etching and the onset of As 2 desorption. Such As enrichment was observed by Simpson et al. 13 in photoemission of Cl/GaAs 110. Thermodynamically, As 2 formation is favored by 382 kj/mol compared to atomic As. 14 The accumulation of As 2 on the GaAs 110 surface would help to explain the absence of ordered Br islands in Figs. 1 c and 1 d because surface As 2 would hinder the diffusion of Br. The nearly exposed lower terraces remain free of As 2 features, perhaps because As atoms ejected onto the lower terrace are accommodated at steps. J. Vac. Sci. Technol. B, Vol. 15, No. 3, May/Jun 1997

607 Cha et al.: Morphologies associated with thermal desorption 607 FIG. 1. a STM image of GaAs 110 exposed to 1.5 ma s of Br 2 at 300 K to produce a nearly saturated surface. Also shown are images of the same surface heated to b 450, c 500, d 600, and e 675 K. Areas marked with I and II are portions of the first and second layers, respectively. The images are 280 280 Å 2. Pitting is evident in c and the pits are seen to expand laterally at 600 and 675 K. When the sample that was represented by Fig. 1 d was heated to 675 K, there was significant removal of surface material. The fact that the top layer was reduced to small islands with only a few isolated adsorption features demonstrates that there was appreciable residual Br at 600 K. Pitting was minimal. These STM results show that the morphology exhibits two distinct etching stages. The first is characterized by development of small single-layer-deep terrace pits at 500 K. The rest of the terrace is covered with features derived from Br and As 2. The second stage at T 600 K is characterized by the lateral enlargement of the pits and desorption of the features that decorate the terraces. Most of the substrate material is removed in stage two. IV. DISCUSSION Examination of TPD results 1 4 in light of the STM data suggests that the evolution of GaBr 3 at lower temperature corresponds to terrace pit initiation, whereas GaBr and JVST B - Microelectronics and Nanometer Structures

608 Cha et al.: Morphologies associated with thermal desorption 608 FIG. 2. Diagram that illustrates the increase in the GaAs system energy with physical atomic-layer removal as Ga and As atoms are placed at infinity. Such atom removal by retreat of a single step leads to a linear, minimal energy increase that corresponds to the cohesive energy times the number of atoms of the surface dashed line. In this case, there is no change in surface energy. Terrace pitting increases the surface energy due to the step and kink energy components solid line. The surface contribution decreases as the pits coalesce. This diagram neglects thermochemical gains associated with exothermic etching of GaAs by Br. As 2 As 4 desorption at higher temperature reflects etch pit enlargement. On an atomic scale, the evolving surface reflects the details of bonding, with terrace atoms being more tightly bound than those at steps or kinks. Thermodynamically, the removal pathway that keeps the surface energy at a minimum is the one where atoms desorb from step kinks so that steps retreat. Such step retreat is depicted in Fig. 2 where the dashed line in the plot represents the linear increase in the total energy, scaling with atom removal as atoms are taken to infinity and the remaining crystal stays in its thermodynamic ground state. In contrast, atom removal from a terrace site pitting increases the surface energy because it creates steps and kinks. This contributes to an increase in the total energy, as shown in Fig. 2, although greatly exaggerated, since the step energy may be a few tenths of an ev per atom compared to a cohesive energy of 6.5 ev per Ga As pair. As the pits coalesce, they reduce the surface energy because the total step length and kink density diminish. In addition to these considerations of the physical removal of atoms to infinity, etching involves chemical bonding with an extrinsic species. Thermochemically, the monobromide reaction 3 GaAs s 3/2 Br 2 g 3 GaBr g 3/2 As 2 g is exothermic by 270 kj/mol, whereas the tribromide evolution GaAs s 3/2 Br 2 g GaBr 3 g 1/2 As 2 g is exothermic by 426 kj/mol Ref. 15. Clearly the desorption branch that produces GaBr 3 is thermochemically favored over the GaBr branch. When added to Fig. 2, this would result in a reduction in overall energy. The likelihood of probing the GaBr 3 etching pathway is determined by the local Br concentration. While islands of the 2 1/c(2 2) structure are stable at 300 K, they are likely to be lost at etching temperatures as Br moves onto the terrace. It follows then that, at low initial concentration, Br would be dispersed and this would make attainment of the precursor state unlikely kinetically limited. For high initial concentrations, the local concentration is maintained and the lateral motion of Br atoms is inhibited. GaBr 3 formation is limited by a barrier related to Br insertion into the backbonds and the associated disordering of the structure. Once GaBr 3 is formed on the surface, it is readily desorbed. This was indicated by molecular beam scattering experiments by DeLouise on GaAs 110 that showed that GaBr 3 desorption involves almost no activation barrier. 16 A possible surface precursor state is one in which both As sites are occupied next to a Br-terminated Ga site. That configuration with 1 1 symmetry would also tend to weaken the Ga backbonds. While not a stable structure at 300 K, it could be populated at high temperature within 2 1/c(2 2) islands via next-neighbor jumps of Br atoms on As sites. Overall, the barrier associated with such GaBr 3 formation must be less than the desorption barrier of GaBr because GaBr 3 is the lower temperature desorption product. The halogen concentration will decrease as a consequence of etching during the temperature rampup for a nearly saturated starting configuration. At the start of that process, the concentration is high and the temperature is low. Hence, etching is biased toward GaBr 3 evolution and creation of small etch pits Fig. 1 c. This relates to the GaBr 3 peaks observed around 500 K in TPD. As the Br concentration decreases and the temperature increases, etching via step retreat is favored, and the pits formed at lower temperature provide ample sites for such etching. That process relates to GaBr evolution. The absence of appreciable pitting of the second layer supports the notion that the Br concentration on this layer never reaches the density needed for trihalide formation. Instead, Br is consumed by step etching reactions. The net contributions of the GaBr 3 and GaBr desorption pathways depend on the initial coverage, with the former gaining weight at higher coverage, in agreement with TPD results. The coverage dependent ratio of these pathways manifests itself in the etch yield, i.e., the number of atoms removed per Br atom. With GaBr 3 etching, three Br atoms participate in the removal of one Ga atom, whereas GaBr etching requires only one. An increased contribution of GaBr 3 evolution would reduce the yield. This has been observed in studies where we varied the initial coverage and evaluated the total number of surface atoms removed after annealing to 700 K. As discussed in Ref. 17, the yield was close to 2 at very low initial Br concentration each Br atom removes one Ga atom and As 2 desorbed spontaneously. The yield fell to 1.6 for an initial coverage near saturation. The relative contribution of the two etching pathways was estimated by fitting the measured coverage dependence of the yield to a reaction kinetic model. For starting coverages below about 0.1 ML, the GaBr 3 contribution was lower than 1%. For an initial coverage of 0.5 ML, it rose to 12%. As a result, it was possible to quantitatively describe the competition between a first-order and a third-order bromide desorption pathway. V. CONCLUSIONS This work has characterized the evolution of Br-exposed GaAs 110 as the temperature was ramped from 300 K. With J. Vac. Sci. Technol. B, Vol. 15, No. 3, May/Jun 1997

609 Cha et al.: Morphologies associated with thermal desorption 609 STM, we gained morphological information about processes often found in TPD. 1 4 By considering atom-level reaction mechanisms, we correlated gas phase reaction products and surface processes. Heating a saturated surface to 500 K results in the formation of scattered single-layer etch pits that we associate with GaBr 3 evolution. Such desorption is the low barrier/high concentration channel. The initiation of etch pits creates high energy step sites at which GaBr desorption can occur when the temperature is increased. This is the high barrier/low concentration channel. The change of surface structure reflects the removal of atoms from the perimeters of first-layer islands with little pitting in the exposed second layer. ACKNOWLEDGMENTS The authors thank C. J. Palmstro m, J. R. Chelikowsky, H. C. Kim, and G. S. Khoo for useful discussions. This work was supported by the U.S. Army Research Office. 1 G. S. Tyrrell, D. Marshal, J. Beckman, and R. B. Jackman, J. Phys., Condens. Matter 3, 179 1991. 2 C. Su, H.-Q. Hou, G. H. Lee, Z.-G. Dai, W. Luo, M. F. Vernon, and B. E. Bent, J. Vac. Sci. Technol. B 11, 1222 1993. 3 A. Ludviskson, M. Xu, and R. M. Martin, Surf. Sci. 222, 282 1992. 4 C. L. French, W. S. Balch, and J. S. Foord, J. Phys., Condens. Matter 3, S351 1991. 5 J. C. Patrin and J. H. Weaver, Phys. Rev. B 48, 17913 1993 ; J. C. Patrin, Y. Z. Li, M. Chander, and J. H. Weaver, Appl. Phys. Lett. 62, 1277 1993. 6 The ramp-up time was typically 10 min. Since we set a constant filament current at the onset of the cycle, the approach to the maximum temperature was not linear rapid heating initially but a leveling off toward the end. When the current was turned off, the sample cooled to near room temperature within 5 min. 7 See, for example, J. T. Yates, Methods in Experimental Physics, edited by R. L. Park and M. G. Lagally Academic, Orlando, 1985, Vol. 22, pp. 425 64; P. A. Redhead, Vacuum 12, 203 1962. 8 C. Gu, Y. Chen, T. R. Ohno, and J. H. Weaver, Phys. Rev. B 40, 10 197 1992. 9 F. Stepniak, D. Rioux, and J. H. Weaver, Phys. Rev. B 50, 1929 1994. 10 G. S. Khoo and C. K. Ong, Phys. Rev. B 50, 1924 1994. 11 J. L. Corkill and J. R. Chelikowsky, Phys. Rev. B 50, 10 796 1994. 12 J. L. Corkill and J. R. Chelikowsky, Phys. Rev. B 53, 12 605 1996. 13 W. C. Simpson, W. M. Tong, C. B. Weare, D. K. Shuh, and J. A. Yarmoff, J. Chem. Phys. 104, 320 1996. 14 K. P. Huber and G. Herzberg, Molecular Spectra and Molecular Structure IV van Nostrand Reinhold, New York, 1979. 15 CRC Handbook of Chemistry and Physics Chemical Rubber, Boca Raton, 1993. 16 L. A. DeLouise, J. Vac. Sci. Technol. A 9, 1723 1991. 17 J. Brake, C. Y. Cha, B. Y. Han, D. W. Owens, and J. H. Weaver, J. Vac. Sci. Technol. B 15, 670 1997. JVST B - Microelectronics and Nanometer Structures