Influence of Si deposition on the electromigration induced step bunching. instability on Si(111)

Similar documents
Universality of step bunching behavior in systems with non-conserved dynamics

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

ELECTROMIGRATION INDUCED STEP INSTABILITIES ON SILICON SURFACES DISSERTATION

ENERGETICALLY AND KINETICALLY DRIVEN STEP FORMATION AND EVOLUTION ON SILICON SURFACES DISSERTATION

Crystalline Surfaces for Laser Metrology

Scanning Tunneling Microscopy Studies of the Ge(111) Surface

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

Spontaneous Pattern Formation from Focused and Unfocused Ion Beam Irradiation

Optimizing Graphene Morphology on SiC(0001)

Surface Morphology of GaN Surfaces during Molecular Beam Epitaxy Abstract Introduction

Nanostructure Fabrication Using Selective Growth on Nanosize Patterns Drawn by a Scanning Probe Microscope

Spatially resolving density-dependent screening around a single charged atom in graphene

1 Corresponding author:

Scaling during shadowing growth of isolated nanocolumns

Dynamics of terraces on a silicon surface due to the combined action of strain and electric current

Role of Si adatoms in the Si 111 -Au 5 2 quasi-one-dimensional system

Distinguishing step relaxation mechanisms via pair correlation functions

One and Two-Dimensional Pattern Formation on Ion Sputtered Silicon

DEPOSITION OF THIN TiO 2 FILMS BY DC MAGNETRON SPUTTERING METHOD

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

ABSTRACT INTRODUCTION

Approved for Public Release _.UUUU/lufnll Distribution Unlimited x «-wvvwi-lf \JJ\J

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.

HYBRID MODELS OF STEP BUNCHING

SEMICONDUCTOR GROWTH TECHNIQUES. Introduction to growth techniques (bulk, epitaxy) Basic concepts in epitaxy (MBE, MOCVD)

SUPPLEMENTARY NOTES Supplementary Note 1: Fabrication of Scanning Thermal Microscopy Probes

PHYSICAL SELF-ASSEMBLY AND NANO-PATTERNING*

Two-dimensional facet nucleation and growth on Si 111

Quasi-periodic nanostructures grown by oblique angle deposition

Supplemental material: Transient thermal characterization of suspended monolayer MoS 2

NUCLEAR TRANSMUTATION IN DEUTERED PD FILMS IRRADIATED BY AN UV LASER

Imaging Methods: Scanning Force Microscopy (SFM / AFM)

Spin-resolved photoelectron spectroscopy

Uniform and ordered self-assembled Ge dots on patterned Si substrates with selectively epitaxial growth technique

GRAPHENE ON THE Si-FACE OF SILICON CARBIDE USER MANUAL

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

Chapter 3. Step Structures and Epitaxy on Semiconductor Surfaces

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

Hysteresis-free reactive high power impulse magnetron sputtering

In situ studies on dynamic properties of carbon nanotubes with metal clusters

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

b imaging by a double tip potential

MICRO-FOUR-POINT PROBES IN A UHV SCANNING ELECTRON MICROSCOPE FOR IN-SITU SURFACE-CONDUCTIVITY MEASUREMENTS

ABSTRACT ON VICINAL SURFACES

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

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

Analyses of LiNbO 3 wafer surface etched by ECR plasma of CHF 3 & CF 4

Length Scales in Step-Bunch Self- Organization during Annealing of Patterned Vicinal Si(111) Surfaces

EE C245 ME C218 Introduction to MEMS Design Fall 2007

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

SUPPLEMENTARY INFORMATION

Supporting Information. by Hexagonal Boron Nitride

Self-Assembled InAs Quantum Dots on Patterned InP Substrates

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

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

JARA FIT Ferienprakticum Nanoelektronik Experiment: Resonant tunneling in quantum structures

Bending of nanoscale ultrathin substrates by growth of strained thin films and islands

Pb thin films on Si(111): Local density of states and defects

Dielectric constant measurement of P3HT, polystyrene, and polyethylene

Supplementary Information. for. Controlled Scalable Synthesis of Uniform, High-Quality Monolayer and Fewlayer

Oxidation of hydrogenated crystalline silicon as an alternative approach for ultrathin SiO 2 growth

Highly Efficient and Anomalous Charge Transfer in van der Waals Trilayer Semiconductors

Supplementary Figure S1 Scheme and description of the experimental setup. UHV-EC experiment includes the sample preparation and STM measurement in

Supplementary Information. In colloidal drop drying processes, multi-ring depositions are formed due to the stick-slip

Temperature Dependence of the Diffusion. Coefficient of PCBM in Poly(3-hexylthiophene)

Supplementary Information. Characterization of nanoscale temperature fields during electromigration of nanowires

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

2 M. N. POPESCU, F. FAMILY, AND J. G. AMAR a set of deterministic, coupled reaction-diffusion equations describing the time (coverage) dependence of a

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

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

Extreme band bending at MBE-grown InAs(0 0 1) surfaces induced by in situ sulphur passivation

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

Energy-level alignment at interfaces between metals and the organic semiconductor 4,4 -N,N -dicarbazolyl-biphenyl

New evidence for the influence of step morphology on the formation of Au atomic chains on vicinal Si(111) surfaces

Removal of Cu Impurities on a Si Substrate by Using (H 2 O 2 +HF) and (UV/O 3 +HF)

SUPPLEMENTARY INFORMATION

Graphene: Plane and Simple Electrical Metrology?

Microwave Absorption by Light-induced Free Carriers in Silicon

Hydrogen termination following Cu deposition on Si(001)

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

Surface morphology of GaAs during molecular beam epitaxy growth: Comparison of experimental data with simulations based on continuum growth equations

Nanometer-Scale Investigation of Metal-SiC Interfaces using Ballistic Electron Emission Microscopy

Level set approach to reversible epitaxial growth

Segregated chemistry and structure on (001) and (100) surfaces of

Spin pumping in Ferromagnet-Topological Insulator-Ferromagnet Heterostructures Supplementary Information.

In situ electron-beam processing for III-V semiconductor nanostructure fabrication

Nanosphere Lithography

Metal-coated carbon nanotube tips for Magnetic Force Microscopy

Precise control of size and density of self-assembled Ge dot on Si(1 0 0) by carbon-induced strain-engineering

Anisotropic spin splitting in InGaAs wire structures

Outline Scanning Probe Microscope (SPM)

Supporting Online Material for

Vacuum Pumps. Two general classes exist: Gas transfer physical removal of matter. Mechanical, diffusion, turbomolecular

Available online at Physics Procedia 15 (2011) Stability and Rupture of Alloyed Atomic Terraces on Epitaxial Interfaces

Repeatability of Spectral Intensity Using an Auger Electron Spectroscopy Instrument Equipped with a Cylindrical Mirror Analyzer

Supplementary Figure 1 Detailed illustration on the fabrication process of templatestripped

Supporting Data. The University of Texas at Dallas, 800 West Campbell Road, Richardson, Texas 75080, United

Coulomb crystal extraction from an ion trap for application to nano-beam source"

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

Growth of Graphene Architectures on SiC

Transcription:

Influence of Si deposition on the electromigration induced step bunching instability on Si(111) B.J. Gibbons, J. Noffsinger a), and J.P. Pelz b) Department of Physics, The Ohio State University, Columbus, OH 43210 Abstract Effects of Si deposition on electromigration induced step bunching on the Si(111) (1x1) surface were studied for Temperature Regimes I (~850 950 C) and II (~1040 1190 C) on dimpled samples that have a range of initial surface miscut angles (±0.5 ). We find that a stepdown electric current is required to induce bunching under both net sublimation and depositions conditions in temperature Regime I, in agreement with previous reports. However, for temperature Regime II we observe that step-up current is required to induce step bunching for both net deposition and net sublimation conditions, in contradiction with the report of Métois et al. [Surf. Sci. 440 (1999) 407] and suggested step permeability model of Stoyanov [Surf. Sci. 416 (1998) 200]. We further observe a strong reduction in the number of crossing steps on the wide terraces for near equilibrium Si flux conditions. We also report a systematic, nearly linear dependence of the step bunching rate on the initial sample miscut angle in both Regimes I and II, which is independent of net deposition/sublimation conditions. a) Permanent address: Department of Physics & Astronomy, The University of Kansas, Lawrence, KS 66045. b) Corresponding author. Tel.: 614-292-8388; fax: 614-292-7557. E-mail address: jpelz@mps.ohio-state.edu (J.P. Pelz) 1

Keywords: Atomic Force Microscopy, Diffusions and migration, Evaporation and sublimation, Growth, Step formation and bunching, Surface diffusion, Surface structure, morphology, roughness, and topography. Introduction Understanding the processes that govern the motion of vicinal surface steps has been a longstanding problem of great fundamental interest in surface science. For semiconductors, understanding the behavior of surface steps is technologically critical for the growth of epitaxial overlayers and for device processing. Recently there has been great interest in processes that lead to spontaneous rearrangement of steps into surface structures of size ranging from a few nm up to many µm [1 5]. Since 1988 [6] it has been known that heating a slightly-miscut Si(111) with direct current (DC) in an ultra-high vacuum (UHV) environment leads to large scale changes in surface morphology such as the formation of step bunches. The earliest reports of these surface electromigration phenomena by Latyshev and coworkers [6] showed that step bunching on Si(111) only results for one direction of the current relative to the vicinal step-up or stepdown direction (see Fig. 1(a)), but also that this required current direction reverses multiple times with increasing temperature. In temperature Regime I (~850 950 C) and Regime III (~1200 1300 C) bunching occurs only for step-down heating current, while in Regime II (~1040 1190 C) and Regime IV (>1320 C) bunching occurs only for step-up heating 2

current. In all temperature regimes the opposite current direction maintains an initially-vicinal surface and accelerates relaxation of an initially step-bunched surface [7] 1. Still controversial is the physical origin of these temperature-dependent reversals of the current direction required for step bunching. It is generally accepted that the diffusing surface species (thought to be individual Si adatoms for Si(111)) each have an effective charge q eff that causes them to drift (or flow) either parallel (for q eff > 0) or antiparallel (for q eff < 0) to the applied electric field. It is also generally agreed that a step-down adatom flow will cause a bunching instability provided steps have a sufficiently small attachment probability and are sufficiently impermeable (i.e., a diffusing surface atom must incorporate into a step before it can cross onto the adjacent terrace [8]). One general model proposed that q eff changes sign as the temperature is increased so that adatom flow is parallel to the applied electric current in Regime I and Regime III, and anti-parallel in Regime II and Regime IV [9, 10]. In this case there will be step-down adatom flow (and hence bunching) only for step-down (step-up) electric current in Regimes I and III (Regimes II and IV). Another general view is that adatom flow is always parallel to the applied electric field (i.e., q eff is always positive) but that temperature-dependent changes to the step permeability [11, 12] or to the relative adatom diffusivity close to a step edge [13] lead to bunching even for step-up adatom flow. Here we focus on the model proposed by Stoyanov [11] and supporting experimental evidence [12] which hold that significantly increased step permeability in Regime II causes step bunching for a step-up adatom flow in that regime. A key prediction of this model is that permeable steps will bunch for step-up adatom flow only under net sublimation conditions (when the Si sublimation rate R sub is larger than an applied Si deposition rate R dep ) while a step-down flow is required to produce bunching under net 1 As addressed later, so called in-phase step wandering has sometimes been observed for step-down current in regime II. 3

deposition conditions (R sub < R dep ). Métois et al. [12] reported observations that bunching in Regime II indeed follows this predicted pattern, with step-up current required for bunching under net sublimation, and step-down current required under net deposition conditions. This would appear to rule out competing models [9, 10, 13] which predict that the required current direction for step bunching should be the same under either net sublimation or net deposition conditions. Here we report a series of measurements made under clean UHV conditions, which clearly show that changing from net sublimation to net deposition conditions has no effect on the required current direction for step bunching in Regime II, in disagreement with the step permeability model [11] and supporting experiments [12]. Our measurements utilize a Si(111) wafer with a large-radius spherical dimple [14 19] so that behavior for step-up and step-down current flow could be observed on different parts of a wafer annealed under identical experimental conditions. We do observe that the net deposition or sublimation has a strong effect on the density of crossing steps on the terraces between the step bunches, in agreement with basic considerations and the observations of Stoyanov et al. [20]. We also report a systematic, nearly linear dependence of the rate of step bunching on the initial sample miscut angle (in both Regime I and Regime II) which may be useful to test different models. Experiment Samples were cut from an n-type Si(111) wafer (0.075 0.2 Ω-cm) into 25 x 6 x 0.4 mm 3 rectangular strips with the long edge parallel to 112. Shallow spherical dimples of radius of ~76 mm were ground into the sample surface providing a ±0.5 range of surface miscuts angles and cleaned as described elsewhere [14 16]. To minimize contamination during annealing, unmounted wafer samples where introduced through a load-lock into our UHV annealing chamber (base pressure <1 x 10-10 Torr) and inserted in-situ into a well-outgassed sample holder 4

[22] consisting of two ~2.5 cm long Mo heating supports held far from anything else in the system. Samples were outgassed in UHV for ~2 hours at 780 C by passing 60 Hz alternatingcurrent (AC) through the sample, and then cleaned by flashing repeatedly at ~1220 C (each flash lasting ~5 10s) for a total time ~2.5 5 minutes. Samples were then annealed for 3.0 h at 940 C or 1090 C under various deposition conditions. Si deposition (up to ~0.25 Å/s) was done using a commercial rod-fed electron beam evaporator (Tectra GmbH e-flux evaporator) held ~10 cm from the Si wafer. The Si deposition rate was monitored by a quartz crystal that was shielded from the hot sample to ensure it only received Si flux from the deposition source. Temperature was monitored using an optical pyrometer. The pressure remained below ~4x10-10 Torr during long anneals in the absence of Si flux, and below ~1x10-9 Torr for the cases of highest Si deposition flux. After annealing under a given sublimation/deposition condition, samples were transferred out of the UHV and characterized ex-situ with atomic force microscopy (AFM), as shown in Figs. 1 and 2. On samples annealed with the electron beam evaporator operating, we did observe widely-spaced pinning sites [21, 22] but these only affected step bunching within 5 10 µm of the pinning site (see Fig. 3(a)). As done previously [10, 21] we used the shape of pinned steps close to pinning sites (see Fig. 3(b)) and 3(c)) to accurately determine whether a given annealing condition produced net sublimation or net deposition conditions. We could estimate the sublimation rate at a given temperature by measuring the deposition flux that produced near zero net sublimation/deposition rate R net (R dep R sub ). Results and Discussion Fig. 1(a) shows a schematic sample cross-section near the dimple bottom, with the applied electric field (and hence the applied current direction) directed from right to left. All 5

images shown here correspond to this geometry, so the left (right) side of the dimple corresponds to step-up (step-down) current. Figs. 1(b) and 1(c) show step-up and step-down areas of a sample annealed for 3 hours at 940 C (in regime I) in the absence of the Si deposition flux (free sublimation with R sub 0.002 Å/s), while Figs. 1(d) and 1(e) show step-up and step-down regions of a different sample annealed at the same temperature but under a Si flux R dep 0.012 Å/s, and so R net +0.01 Å/s. Step bunching occurs only for step-down current under both net sublimation and net deposition conditions, consistent with past reports for Regime I [10, 12]. We also observe in Fig. 1(c) (free sublimation) that the lowest step in each bunch has separated slightly from the bunch, as observed by Homma et al [23], who suggest this is due to step-flow growth when free adatoms on the wide terraces attach to the bounding step edges during the sample quench. Fig. 1(e) shows slightly different behavior under net deposition, where now a number of crossing steps [24] can be observed on the wide terraces between bunches. During annealing, the net adatom deposition flux on the terraces causes significant step-flow growth of the lowest step of each bunch, which continuously break away from the bottom of one bunch and cross over to the top edge of the adjacent bunch. As discussed later, crossing steps should in general exist under either net sublimation or net deposition conditions [24], but not close to balanced deposition/sublimation conditions. Figure 2 shows the bunching behavior at 1090 C (well into Regime II) where under free sublimation conditions step bunching is known to exist only for step-up current [6, 10, 12, 23, 25, 26, 27]. Figs. 2(a) and 2(b) show that we also observe bunching only for step up current under free sublimation (R sub 0.16 Å/s), consistent with these previous reports. Figs. 2(c) and 2(d) show bunching behavior at 1090 C near balanced sublimation/deposition conditions (R dep R sub 0.16 Å/s), where again bunching is only observed for step-up current. In this near- 6

balanced case however, we do observe a dramatic reduction in the number of crossing steps on the bunched surface, consistent with the report of Stoyanov et al. [26]. We will return to this point later. Figs. 2(e) and 2(f) show bunching behavior at 1090 C under net-deposition conditions, with R dep 0.25 Å/s and hence R net 0.09 Å/s. Here again bunching is clearly observed only for step-up current, in direct disagreement with the report [12] that net deposition conditions in Regime II cause bunching to occur only for step-down current. We have repeated this measurement more than 5 times under a variety of net deposition/sublimation conditions, and so far have never observed step bunching for step-down current in Regime II. Our observations are inconsistent with the step-permeability model [11] which predicts a reversal of the required current direction for step bunching in Regime II. Our observations are consistent with the model which assumes sign reversal of the effective charge in Regime 2 [9, 10], and the recent model which assumes diffusion near step edges becomes very fast in Regime II [13]. Both of these models predict that net deposition/sublimation have no effect on the required current direction for bunching. Fig. 2(c) shows a marked reduction in the density of crossing steps under near-balanced conditions (R dep R sub ), consistent with the report of Stoyanov et al. [26]. We can understand this within the simultaneous bunching/debunching instability model of Kandel and Weeks [24] in the following way. The applied electric field causes the adatom concentration to vary across a wide terrace between bunches as shown schematically in Fig. 4, assuming here that surface electromigration causes the adatoms pile-up on the down-hill side of the wide terrace. In strong net sublimation [Fig. 4(b)], the adatom concentration is below the equilibrium value c eq on both sides of the large terrace. This will result a net loss of adatoms from the top step in the right 7

bunch, causing it to break away from the right bunch and cross over to the left bunch. In strong net deposition [Fig. 4(d)], the adatom concentration is above c eq on both sides of the large terrace, so now the bottom step of the left bunch is unstable against crossing over. However, for a certain range of near-balanced conditions the concentration will be above c eq on the right side and below c eq on the left side, causing stability for the bounding steps on both sides of the wide terrace. We are currently exploring the possibility of using the dependence of the crossing step density on R net to quantify the adatom pile-up during DC heating. Lastly, we observe a clear trend in the bunching rate with sample miscut, by inspecting regions within the dimple which have different miscut. Figure 5 shows a plot of the average bunch height as a function of local miscut angle for several samples annealed for 3.0 h at 1090 C and 940 C under various deposition conditions. Each data point is the calculated average of numerous bunches at a given miscut, and the error bars represent the standard deviation of the mean. It is clear that the average bunch height increases systematically (and approximately linearly) with sample miscut. As a result, the average size of the large terrace between bunches is similar for all miscuts in this range. Interestingly, the bunching rate does not depend strongly on the deposition conditions, consistent with the images shown in Fig. 2. This measured dependence of bunching rate on initial miscut angle may prove useful in comparing different step-bunching models. Finally, we note that in-phase step wandering (IPSW) has been reported in regime II for a step-down directed heating current after long heating times [18]. In our study, we observed IPSW on two samples out of the eight we studied. The IPSW occurred on those two samples at regions of step-down current and fairly high local miscut ( 0.3 ), at near balanced conditions (R net -0.015 and 0.005 Å/s). We note that the ~3 h annealing time in the present study was 8

short compared to the ~ 4 48 h times use in the previous reports of IPSW [18], making it hard to determine whether the net deposition/sublimation conditions affect the formation of IPSW. Conclusion In conclusion we have studied the bunching behavior of Si(111) in regimes I and II under various Si flux deposition conditions. In Regime 2, we observe no reversal in the required current direction for step bunching under net deposition conditions, which argues against the step permeability explanation [11, 12] of the bunching behavior in Regime II. We also observed a minimum in crossing step density for near-balanced deposition/sublimation conditions, and a systematic near-linear increase in the bunching rate on initial sample miscut. Acknowledgements We thank S. Schaepe for assistance and discussions. This work was supported by NSF Grant No. DMR-0074416. J. Noffsinger was supported by the NSF REU program through NSF Grant No. PHY-0242665. References [1] Y.,-W. Mo, D.E. Savage, B.S. Swartzentruber, and M.G. Lagally, Phys. Rev. Lett. 65 (1990) 1020. [2] D.E. Jones, J.P. Pelz, Y.Hong, E. Bauer, and I.S.T. Tsong, Phys. Rev. Lett. 77 (1996) 330. [3] G. Medeiros-Ribeiro, A.M. Bratkovski, T.I. Kamins, D.A.A. Ohlberg, and R.S. Williams, Science 279 (1998) 353. [4] F.M. Ross, R.M. Tromp, and M.C. Reuter, Science 286 (1999) 1931. [5] J.-F. Nielsen, J.P. Pelz, H. Hibino, C.-W. Hu, and I.S.T. Tsong, Phys. Rev. Lett. 87 (2001) 136103-1. [6] A.V. Latyshev, A.L. Aseev, A.B. Krasilnikov, and S.I. Stenin, Surf. Sci. 213 (1989) 157. 9

[7] E.S. Fu, D.-J Liu, M.D. Johnson, J.D. Weeks, and Ellen D. Williams, Surf. Sci. 385 (1997) 259. [8] S. Stoyanov and V. Tonchev, Phys. Rev. B 58 (1998) 1590. [9] D. Kandel, and E. Kaxiras, Phys. Rev. Lett. 76 (1996) 1114. [10] Y. N. Yang, E.S. Fu, and E.D. Williams, Surf. Sci. 356 (1996) 101. [11] S. Stoyanov, Surf. Sci. 416 (1998) 200. [12] J.J. Métois and S. Stoyanov, Surf. Sci. 440 (1999) 407. [13] T. Zhao, J.D. Weeks, and D. Kandel, Phys. Rev. B 70 (2004) 161303. [14] V. Brichzin and J.P. Pelz, Phys. Rev. B 59 (1999) 10138. [15] J. F. Nielsen, M.S. Pettersen, and J.P. Pelz, Surf. Sci. 480 (2001) 84. [16] J. F. Nielsen, M. S. Pettersen, and J. P. Pelz, Surface Review and Letters 7 (2000) 577. [17] M. Degawa, H. Nishimura, Y. Tanshiro, H. Minoda, and K. Yagi, Jpn. J. Appl. Phys. 38 (1999) L308. [18] K. Yagi, H. Minoda, and M. Degawa, Surf. Sci. Rep. 43 (2001) 45 [19] X. Tong and P.A. Bennett, Phys. Rev. Lett. 67 (1991) 101. [20] S. Stoyanov, J.J. Métois, and V. Tonchev, Surf. Sci. 465 (2000) 227. [21] H.C. Abbink, R.M. Broudy, and G.P. McCarthy, J. Appl. Phys. 39 (1968) 4673. [22] B.S. Swartzentruber, Y.-W. Mo, M.B. Webb, and M.G. Lagally, J. Vac. Sci. Technol. A 7 (1989) 2901 [23] Y. Homma and N. Aizawa, Phys. Rev. B 62 (2000) 8323. [24] D. Kandel and J.D. Weeks, Phys. Rev. Lett. 74 (1995) 3632. [25] E.D. Williams, E. Fu, Y. N, Yang, D. Kandel, and J.D. Weeks, Surf. Sci. 336 (1995) L746. [26] S. Stoyanov, J.J. Métois, and V. Tonchev, Surf. Sci. 465 (2000) 227. 10

[27] K. Fujita, M. Ichikawa, and S. Stoyanov, Phys. Rev. B 60 (1999) 16006. 11

Fig. 1. (a) Schematic view of vicinal stepped surface close to the bottom of a spherical dimple. For right-to-left applied electric field (and conventional current) direction, the left side (right side) has step-up (step-down) current. (b) and (c): Derivative-mode AFM images (which appear as if illuminated from the left) of the surface annealed for 3.0 h at 940 C under free sublimation conditions (Rnet -0.002 Å/s) for (b) step-up and (c) step-down current. (d) and (e): surfaces annealed at 940 C under net deposition of 0.01 Å/s for (d) step-up and (e) step-down current conditions. Bunching is only seen for step-down current for both net-sublimation and netdeposition conditions. 12

Fig. 2. Derivative mode AFM images for Si(111) samples DC annealed for 3.0 h at 1090 C. The DC is oriented from right to left in all images as in Fig. 1. (a) and (b): annealed under free sublimation conditions (with Rnet -0.16 Å/S), (c) and (d): annealed under near-balanced deposition conditions (Rnet 0 Å/s), and (e)) and (f): annealed under net deposition conditions (Rnet +0.09 Å/S). In all cases the step train remains uniform for step-down current (right side), but bunch under step-up current (left side). 13

Figure 3: (a) Typical large area scan of bunched surface (annealed at T=1090 C) with a high Si deposition flux (Rdep 0.17 Å/S Rnet 0.01 Å/S). Close up views of pinning site under slight (b) net sublimation and (c) net deposition conditions. In net sublimation (net deposition) the steps retract from (advance past) the pinning site. 14

Fig. 4. (a) Schematic of a bunched surface, assuming that the applied field causes a higher adatom concentration on the down-hill edge of a wide terrace between bunches. (b)-(d) the corresponding adatom concentration across the terrace (solid line) relative to the equilibrium concentration (dashed line) for (b) net sublimation, (c) balanced, and (d) net deposition conditions. For balanced conditions, the bottom (top) step of the bunch on the left (right) side of the terrace will both be stable against breaking away from the bunch. 15

Heigth (nm) 25 20 15 10 5 a) T=1090 C Height (nm) 0 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 7 6 5 4 3 2 1 0 Initial miscut (degrees) b) T=940 C 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 Initial miscut (degrees) Fig. 5. Plot of the average bunch height after a 3.0 anneal as a function of sample miscut for Regimes I and II. (a) 1090 C anneal for net conditions: ( ) R net -0.16 Å/s, ( ) -0.045 Å/s, ( ) - 0.015 Å/s, ( ) -0.005 Å/s, ( ) 0.005 Å/s, ( ) 0.025 Å/s, and ( ) 0.085 Å/s. (b) 940 C anneal for net conditions ( )R net -0.002 Å/s and ( ) 0.01 Å/s. For reference, the solid lines are power law fits, with best-fit exponents 0.92 and 0.93 for (a) and (b) respectively. There exists a clear near-linear increase in bunch height with initial surface miscut angle, but with little dependence on net deposition/sublimation conditions. 16