THEOfi (01H 91¢”? This is to certify that the dissertation entitled STRUCTURAL AND ELECTRONIC PROPERTIES OF RARE EARTH SILICIDES ON Sl(001): NANOWIRES AND 30 ISLANDS presented by Chigusa Ohbuchi has been accepted towards fulfillment of the requirements for the Ph.D. degree in Physics and Astronomy ,Bug AD Major Prcfiessor’s Signature g/IQ/Okr, Date MSU is an Affirmative Action/Equal Opportunity Institution digs-.. Ac-Q-.-._.-.-.-.-.-._.-I-._.-.-- .—-_.-4 .-.—-—.—.-.._. —.-._.-o-n-l-a-o-.--o-.--_-—. _ LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 6/01 cJCIRC/DatoDuopGS-p. 1 5 STRUCTURAL AND ELECTRONIC PROPERTIES OF RARE EARTH SILICIDES ON SI(OOl): NANOWIRES AND 3D ISLANDS By Chi gusa Ohbuchi A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Physics and Astronomy 2004 ABSTRACT STRUCTURALAND ELECTRONIC PROPERTIES OF RARE EARTH SILICIDES ON SI(OOl): NANOWIRES AND 3D ISLANDS By Chi gusa Ohbuchi Deposition of rare earth (RE) metals on the silicon(001) surface at elevated temperature results in the formation of RE silicide islands coexisting with a reconstructed substrate surface. This thesis presents the growth behavior of holmium, samarium, and dysprosium on Si(OOl) and the electronic properties of the resulting surface structures. The growth of Ho and Sm on Si(001) at 600°C was studied using scanning tunneling microscopy (STM) and low energy electron diffraction. Ho grows in a Stranski-Krastanov mode with highly elongated nanowires (W8) and a 2x4 reconstructed substrate at low metal coverage. As the coverage increases, three-dimensional (3D) compact silicide islands coexist with the W3. A 2x7 two-dimensional (2D) phase appears in the medium coverage range, and it always coexists with the 2x4 phase. Scanning tunneling spectroscopy (STS) data show that HoSiz NWs are more metallic than the 3D silicide islands or the reconstructed substrate. Srn/Si(001) induces two different surface reconstructions: a lower coverage 2x3 phase and higher coverage 3x2/c(6x2) phases. In the high coverage regime, large 3D silicide islands are observed instead of elongated NW 3. The topographical evolution of 3D Dy silicide structures and the 2D reconstructed surface in the temperature range of 600 to 750°C was investigated by STM and low energy electron microscopy (LEEM). The initial surfaces were covered in the 2x7 superstructure plus NWs of uniform width defined by the 2x7 unit cell. Post—growth annealing at 700°C increases the number of 3D islands and the average island size. At the same time, the number of NWs decreases and surface reconstruction is swept away. Therefore, 3D silicide islands are a stable phase whereas NWs are a metastable phase at 700°C. Ex-situ transport properties of the grown RE silicide films were correlated with film morphology as observed by STM. A 1 nm thick film of interconnected DySiz islands shows surface resistivity at 4.2 K similar to the reported bulk silicide value. A sample with sparse, disconnected islands shows higher surface resistivity than the 1 nm film but lower than that of the clean Si surface. ACKNOWLEDGMENTS First of all, I would like to thank my major advisor Professor Jun Nogami for having given me an opportunity to work in his group. I decided to do my research at Michigan State University because Dr. Nogami was there. He generously accepted me as his student even though I did not have enough background of surface physics. Without his offer, I could not even come to the US. and my Ph.D. life would never have stated. I remember the days when he explained to me how STM works in JAPANESE because my English comprehension level was poor at that time. I was so grateful to him for his patience and generosity. I also thank him for giving me his time, support, and valuable advise, and feedback on my Ph.D. work. I would like to thank my committee members, Professor Norman Birge, Professor Jack Baldwin, Professor Phillip Duxbury, Professor Thomas Glasmacher, and the previous committee, Professor Carl Bromberg who is currently on the Sabbatical, for their helpful comments and suggestions. Especially, I thank Dr. Birge for providing me technical assistance and knowledgeable advice on transport measurements. There are several other people whom I would like to thank for helping me with transport measurements. Dr. Reza Loloee, a SQUID (superconducting quantum interference devices) specialist in Physics, made pre-deposited Ta and Ti contacts in the sputtering system. Michael Crosser and Ion Moraru, graduate students in Dr. Birge’s lab, helped me with preparations for transport measurements in the clean room. They always gave me a hand whenever I needed them, even though they are busy with their own research. iv I would like to express my appreciation to Dr. Wacek Swiech who is a LEEM specialist in the Frederick Seitz Materials Research Laboratory at the University of Illinois for his LEEM operation and image acquisition. I really enjoyed doing new experiments with him although his lab was extremely cold. I thank former and current coworkers, Dr. Bangzhi Liu, Michael Katkov, Gangfeng Ye, and Jiaming Zhang for their kind support and friendship. Bangzhi and Micheal taught me how to do the STM experiments. Besides research, we used to have a cookout on summer weekend for fun and relaxing. I am really thankful to Gangfeng and Jiaming for encouraging me throughout the writing of this thesis. I also thank Professor Evangelyn Alocilja in Biosystems and Agricultural Engineering for her continuous encouragement, prayer, and support for my successful research progress and future proposals. Sean Davis and Beth Czischke are my best American friends and also my best English teachers. I learned a lot about American cultures and traditions from them. I am grateful to them for making my life more fulfilling and happier. Finally, I would like to thank my fiancé Yuzuru Honda and my family in Japan for their loving and caring about me all the time. My parents sent me my favorite Japanese food for my energy source from time to time. Yuzuru has been waiting for the completion of my Ph.D. since I started the program, which becomes a powerful motivation for me to work hard and finish my Ph.D. within five years. I have been always feeling his overwhelming love and support although we are far apart. My gratitude to him is beyond words. I thank all the other friends in Japan and the US. who cheered me up. N- . :, TABLE OF CONTENTS LIST OF TABLES ................................................................................ viii LIST OF FIGURES ................................................................................. ix ABBREVIATIONS ............. '. .................................................................. xvi CHAPTER 1 INTRODUCTION ............................................................. 1 1.1 Rare earth (RE) silicides and self-assembled RE disilicide NWs ................. 1 1.1.1 The rare earths ...................................................................... 2 1.1.1.1 RE silicides ..................................................................... 4 1.1.2 Crystal structures of RESiz.x ...................................................... 5 1.1.2.1 Epitaxial growth of RE silicides on Si substrate .......................... 10 1.1.2.1.1 Epitaxial growth of RE silicides on Si(lll) ........................... 10 1.1.2.1.2 Epitaxial growth of RE silicides on Si(001) .......................... 12 1.1.2.2 Schottky barrier height of RE silicides on Si .............................. 16 1.1.2.3 Electrical resistivity of RE silicides ........................................ 17 1.2 Organization of the thesis ............................................................... 22 CHAPT ER 2 EXPERIMENTAL ............................................................ 23 2.1 Experimental and research methods ................................................... 23 2.2 Scanning tunneling microscopy (STM) ................................................ 24 2.2.1 Basic principles of STM ............................................................ 24 2.2.2 STM data acquisition and image processing .................................... 27 2.2.3 Scanning Tunneling Spectroscopy (STS) ....................................... 28 2.3 Low energy electron microscopy (LEEM) ............................................ 30 2.3.1 Basic physics of the low-energy electron interaction with surface: Elastic scattering and inelastic scattering ................................................. 32 2.3.2 LEEM instrumentation and image acquisition .................................. 33 2.4 Si (001) surfaces .......................................................................... 35 CHAPTER 3 HOLMIUM GROWTH ON SI(001): SURFACE RECONSTRUCTIONS AND NANOWIRE FORMATION......... 40 3.1 Results and discussion ................................................................... 40 3.1.1 Coverage dependence ............................................................... 40 3.1.2 Nanowires and 3D islands .......................................................... 42 3.1.3 Two-dimensional substrate surface reconstructions ............................ 47 3.1.4 Ho content of the 2x4 substrate reconstruction and the nanowires. . . . . . .....51 3.1.5 Scanning tunneling spectroscopy ................................................. 53 3.2 Conclusions ................................................................................ 55 vi CHAPTER 4 SAMARIUM INDUCED SURFACE RECONSTRUCTIONSS OF SI(001) ............................................................................ 56 4.1 Results and discussion .................................................................. 56 4.1.1 Coverage dependence ............................................................ 56 4.1.2 General surface structure evolution with metal coverage .................. 57 4.1.3 The low coverage 2x3 and zigzag chain structures .......................... 61 4.1.4 The medium coverage regime 3x2 and C(6x2) phases ...................... 66 4.2 Conclusions .............................................................................. 70 CHAPTER 5 TOPOGRAPHIC EVOLUTION OF DY SILICIDE ON SI(001): SHAPE TRANSITIONS FROM NANOWIRES TO 3D ISLANDS .................................................................................... 71 5.1 Results and discussion .................................................................. 71 5.1.1 STM experiments .................................................................. 71 5.1.2 LEEM experiments ................................................................ 78 5.2 Discussion ................................................................................ 85 5.3 Conclusions .............................................................................. 89 CHAPTER 6 TRANSPORT MEASUREMENTS ........................................ 91 6.1 Sample preparation ...................................................................... 91 6.1.1 Pre-deposited Ta (or Ti) silicide contacts ....................................... 92 6.1.2 In-situ Au contacts ................................................................. 95 6.1.3 Amorphous Ge passivation ....................................................... 97 6.2 Transport measurements: Results and discussion ................................... 98 CHAPTER 7 FUTURE WORK ............................................................. 107 7.1 Future work on growth kinetics ...................................................... 107 7.2 Reliability and reproducibility of electrical contacts for transport measurements ............................................................................................ 108 7.3 Single nanowire measurements ...................................................... 109 7.4 Temperature dependence and magnetic-field dependence of magnetic, structural, and transport properties of low-dimensional rare earth silicides... 110 APPENDICES ..................................................................................... 117 REFERENCES ..................................................................................... 128 vii Table 1.1 Table 1.2 Table 1.3 Table 1.4 Table 6.1 Table 6.2 Table 7.1 Table 7.2 Table A1 LIST OF TABLES Periodic table of the elements ........................................................ 3 RESiz lattice parameters and mismatches with Si substrate ..................... 8 Schottky barrier heights, (DB, of RE metals and silicides on n- and p-type Si substrates. ............................................................................... 19 Resistivity of RE silicides. ............................................................ 20 Dy and Ho film conductivity at 4.2 K vs. metal coverages. All the values correspond to those in Figure 6.12. There are no values for the conductivity of bulk Ho silicide in the literature, but the conductivity of the RE silicides lie in the range 4.5x1o3 to 6.7x105 (sz-cm)‘l at 4.2 K as in Table 1.4. ........... 104 Summary of transport samples measured. Data for the shaded samples are collected in Figure 6.12 and Table 6.1. ........................................... 105 Bulk properties of RE elements. ................................................... 113 Magnetic properties of RE disilicides. ............................................. 114 Structure of the H-saturated Si(001) surface and H exposure conditions. 1 19 viii LIST OF FIGURES CHAPTER 1 Figure 1.1 Spatial models for (a) the hexagonal Ale structure, and (b) tetragonal ThSiz structure or orthorhombic GdSiz structure. (c) Projection of Ale structure along the chex-axis. ((1) Projection of ThSizlGdSiz along the b-axis; Introduction of shear planes in Ale generates the ThSiZ/GdSiz structure. .. 7 Figure 1.2 Structural model for Si(111)1xl-RE: (a) top view and (b) side view. 12 Figure 1.3 Structural model of Si(001)1xl-RE: (a) top view and (b) side view. The orientation relationship between hexagonal Ale and NW structures are shown in (c). .......................................................................... 13 CHAPTER 2 Figure 2.1 Wave function I” for an electron with kinetic energy E tunneling a thin potential barrier of height V0 and width s. The wave function is oscillatory in the left and right regions of the barrier. Within the barrier, the wave function decays exponentially. ..................................................... 25 Figure 2.2 Energy diagram for tunneling between a tip and a sample. Tunneling is possible only when there are empty states on the right. Such empty states are created when eV is applied to lower the Fermi level on the right. ......................................................................................... 26 Figure 2.3 Schematic diagram of an STM. A metallic tip is mounted on a piezoelectric tripod actuator which can move the tip in the x—y plane of the sample surface, and in the z direction normal to the surface, with atomic resolution. ............................................................................. 28 Figure 2.4 Schematic of an STS experiment. An ST M tip is held above the sample, and the current is measured while the voltage is ramped. This measures the conductivity in the surface normal direction. ................................ 30 Figure 2.5 Diffraction contrast in LEEM (a) Bright field, (b) tilted bright field, (c) dark field. ................................................................................... 31 Figure 2.6 Schematic of LEEM instrument (IBM Type I). ................................. 34 Figure 2.7 Schematic ray diagrams for LEEM. .............................................. 34 Figure 2.8 (a) Diamond structure of silicon, and (b), (c) top view of model of the dimer reconstruction on Si(001). .......................................................... 35 ix Figure 2.9 Figure 2.10 Figure 2.11 (a) filled- and (b) empty-state STM images of Si(001) surface with a side view of Si dimer model. 2x1 unit cells corresponding to the dimer position are marked. ........................................................................... 36 Schematic drawing of Si(001) 2X1 reconstructed surface consisting of three terraces. Si atoms on top, second, third, and fourth layers are shown with black, dark gray, light gray, and white colors, respectively. The height of the terrace lowers from left to right. The directions of surface stress are marked with black arrows. ......................................................... 37 LEED pattern of a Si(001) surface with 2x1 superstructure. (%, 0) and (g— , O) diffraction spots and (0, %) and (0, %) diffraction spots are indicated with white and gray circles, respectively. 2x1 unit cell is marked with a white rectangle as well. ..................................................... 38 Figure 2.12 Dark-field LEEM image of Si(001) surface with opposite contrast. Images (a) and (b) are acquired using one of the diffraction spots with (a) white and (b) gray circles, respectively. ...................................................... 39 CHAPTER 3 Figure 3.1 Coverage dependence of LEED patterns and topography for Ho growth on Si(001) at 600°C. .................................................................... 41 Figure 3.2 Two 400x400 nm2 images of HoSiz W8 and 3D islands. (a) NWs formed by 0.36 ML of Ho. Each NW grows on a single terrace. (b) As the coverage is increased (0.90 ML), 3D compact silicide islands are found. .. 42 Figure 3.3 Relationship between coverage and the percentage of surface covered with Ho surface reconstruction, NW 8, and 3D islands. .............................. 43 Figure 3.4 (a) Triangular (left) and rectangular (right) NWs at 0.3 ML (200x200 nmz). Steps on the substrate flow to accommodate the NWs. Lines A-B and C-D show the locations of the cross sections shown in (b) and (c). ................ 44 Figure 3.5 (a) Bundled NWs with second layer growth at 0.36 ML (90x90 nmz). (b) Atomic resolution on bundled NW 3 (12x14 nmz). A single c(2x2) unit cell is marked with a white square. ..................................................... 45 Figure 3.6 The distributions of (a) the width and (b) the height of single NW 8 and 3D islands. ................................................................................ 46 Figure 3.7 (a) Empty- and (b) filled-state images of slightly different areas of the same surface with 0.55 ML Ho showing both 2x4 and 2x7 structures (30x30 nmz). The position of gray circle on the NW in (a) corresponds to that in (b). ..... 47 Figure 3.8 Figure 3.9 Figure 3.10 Figure 3.11 (a) Empty-state (1.76 V) and (b) filled-state (-1.76 V) images of the 2x4 phase in the reconstructed surface shown in Figure 3.7 (10x10 nmz). (a) and (b) correspond to the square areas numbered by 1 and 3 in Figure 3.7, respectively. (c) Line drawing showing the empty-state (gray circles) and filled-state (dark circles) maxima of the 2x4 structure. Clean 2x1 Si dimer positions are shown as a reference on the topmost row. ....................... 48 (a) Empty-state (1.76 V) and (b) filled-state (-l.76 V) images of the 2x7 phase extracted from Figure 3.7. (a) and (b) correspond to the square areas numbered by 2 and 4, respectively. A 2x7 unit cell consists of three types of maxima in rows labeled A and B in both images. (c) Line drawing showing the empty-state (gray circles and oblongs) and filled-state (dark circles) maxima of the 2x7 structure. 2x1 Si dimer positions from the clean surface are shown as a reference on the topmost row. .................. 50 Comparison of experimental coverage with calculated coverage under different assumptions of Ho content (see text). ................................. 52 (a) The STS I-V curves and normalized spectra for a NW, 3D island, and the reconstructed substrate. The surface topography is shown in (b) (150x150 nmz). The image (c) is a partial image of (b). Each STS spectrum was averaged from the STS data acquired in the areas shown by solid or dashed lines in the STM image shown in (c) (40x40 nmz). (d) Magnification of the STS I-V curves at near-zero bias voltage. .................................... 54 CHAPTER 4 Figure 4.1 Figure 4.2 Figure 4.3 Figure 4.4 Figure 4.5 Coverage dependence of LEED patterns and topography for Sat/Si(001) at 600°C. ................................................................................. 56 LEED pattern of Sm/Si(001) at 0.54 ML shows 2x3 (black rectangle) and c(6x2) (black arrows) periodicities. The LEED pattern was obtained from the surface shown in Figures 4.8 (c) and 4.9 (a). The electron beam energy is 63 eV. ............................................................................... 57 (a) Empty-state (1.90 V) and (b) filled-state (-1.91 V) STM images of the same area showing 2x3 phase and the zigzag chain structures at 0.067 ML (15x15 nmz). Cross-sectional profile along the line A is shown on the bottom of each image. ............................................................... 58 15x15 nm2 image of a surface at 0.59 ML. Both a 3x2 phase and a highly defected Si 2x1 surface are seen. The bias voltage is —1.77 V. ............ 60 Sm silicide island at 1.46 ML (300x300 nmz). .................................. 61 xi Figure 4.6 Figure 4.7 Figure 4.8 Figure 4.9 (a) Empty-state (1.57 V) and (b) filled-state (-l.57 V) STM images in the same area showing 2x3 phase at 0.067 ML (10x10 nmz). There are out-of-phase boundaries shown with dashed lines. Two 2x3 unit cells are marked with white rectangles. (c) Line drawing showing the empty-state (white circles and ovals) and filled—state (gray circles and ovals) maxima of the 2x3 structure. .................................................................... 62 (a) Filled-state (-1.82 V) image of zigzag chain structures with buckled Si dimers below 0.1 ML (10x10 nmz). The 2x3 unit cell is also shown as a pair of two maxima on the lower right area. (b) Line drawing illustrating the zigzag chain structures with surrounding buckled Si dimers. This dimer buckling causes c(4x2) periodicity. A c(4x2) unit cell is marked with a black rectangle. ............................................................. 65 (a) Large-scale STM image of a surface consisting of two types of step edges. Straight step edges (SA step) run along the dimer row direction on terrace A, and terraces B have jigsaw-shaped step edges (83 step). (b) Both types of terraces consist of 3x2 and c(6x2) phases and Si 2x1 dimer with zigzag defects shown with white arrows. (c) Close-up image of zigzag defects on Si dimer rows. The ends of four zigzag defect structures are marked with black and white arrows. All three images were at different Sm coverages: (a) 0.93 ML, (b) 0.59 ML, and (c) 0.54 ML. The image sizes are (a) 300x300 nmz, (b) 50x50 nmz, and (0) 25x25 nmz. Bias voltages are (a) —2.31 V, (b) -l.77 V, and (c) 1.28 V. ........................ 67 10x10 nm2 images in (a) empty states (1.28 V) and (b) filled states (-l.77 V) showing 3x2 and c(6x2) phases at different coverages: (a) 0.54 ML and (b) 0.59 ML. Both 3x2 and c(6x2) unit cells are marked with white rectangles on each image. (c) Line drawing showing the empty-state (gray circles) and filled-state (dark circles) maxima of the 3x2 and c(6x2) structures. Si dimers are shown on the leftmost row as a reference. .......................... 68 CHAPTER 5 Figure 5.1 STM images of DySiz NWs at 0.58 ML (a) before and (b) after annealing at 600°C for 8 minutes. (a) Dy was deposited at 600°C. Almost all the 2D surface is covered with 2x7 stripe structure and W3 have comparatively uniform width distribution. 2x7 phase generally disappears after 1 min annealing at 600°C. (b) After annealing, 2x4 superstructure coexists with bare silicon. More bundled NWs are observed and the width of NWs becomes wider. There is also more second layer growth on top of the NWs. ......................................................................................... 72 xii Figure 5.2 Figure 5.3 Figure 5.4 Figure 5.5 Figure5.6 Figure 5.7 Figure 5.8 Figure 5.9 Small-scale images of (a) and an) in Figure 5.1 showing (a) 2x7 (-1.80 V) and (b) 2x4 (+1.04 V) reconstructed surface around NWs. Two unit cells are marked with white rectangles in each image. ............................... 73 Annealing behavior of Dy grown on Si(001) at 0.58 ML. 3D structures of DySiz transit from large 3D islands by annealing. 2x7 superstructure dominates on the 2D surface after rapid deposition of Dy. After 1 min annealing at 600°C, 2x7 superstructure is replaced by 2x4 superstructure and bare Si. Less 2x4 structure appears with higher temperature annealing or longer annealing time. ........................................................... 74 Transition from NWs to large 3D islands of DySiz by annealing. (a) 0.58 ML of Dy was deposited at 600°C, (b) 1St annealing at 600°C for 8 min, (c) 2nd annealing at 600°C for 20 min, ((1) 6h annealing at 700°C for 10 min, and2 (e) 7‘h annealing at 720°C for 10 min. All image sizes are 400x400 nm . ................................................................................... 77 (a) The LEED pattern of Dy grown on Si(001) at ~600°C (0.78 ML). 1x7 phase and 1/2 order streaks are shown. The electron beam voltage is 60 V. (b) Dark-field LEEM image (2 eV) taken from 2/7 spot shown with the white arrow in (a). The field of view is 4 urn. (c) Bright-field LEEM image (9 eV) in the same area of (b) showing the stripe structure. Contours of terraces are drawn with white lines. The stripes are oriented at an angle of 90 degrees on each terrace, which correspond to the NW 3. ............... 78 Sequence of LEEM images showing the growth of DySiz islands (0.78 ML). The field of view is 4 pm. The white arrows show the same island. Small grain structures start appearing at ~700°C and grow as islands. BF: Bright field, TBF: Tilted bright field. ............................................ 80 Annealing duration vs. temperature. Annealing temperature was gradually increased from RT (00’00”) to 750°C. The black arrow shows the time range during which the islands are observed as shown in Figures 5.8 and 5.9. ......................................................................................... 81 Time evolution of the length and the width of 3D islands and the surrounding Si depletion areas in different growth types during annealing. ......................................................................................... 83 Change in the island size by annealing (0.78 ML of Dy). Black, gray, and white marks indicated large, medium, and small islands, respectively. The islands numbered from 1 to 13 correspond to the islands with the same numbers in Figure 5.6 (f). The labels (a)~(f) are marked for the time at which the LEEM images in Figure 5.6 were taken. ............................ 84 xiii CHAPTER 6 Figure 6.1 Figure 6.2 Figure 6.3 Figure 6.4 Figure 6.5 Figure 6.6 Figure 6.7 Figure 6.8 Figure 6.9 Figure 6.10 Sample preparation procedure for transport measurements. Si and Ta (or Ti) depositions are done in the high vacuum chamber (Step 1 and 2). Sample flashing, deposition of RE metals, Au contacts, and amorphous Ge are performed in the UHV chamber (Step 4~8). ................................ 92 (a) Atomic Force Microscopy (AFM) image of the boundary between TaSiz and Si substrate after annealing 100 nm thick Ta on silicon. (b) Cross-sectional profile traced along a white line in (a) showing a deep trench. (Width = 23 um, depth = 800 nm) ................................... 94 Figure 6.3 Scanning Electron Microscopy (SEM) images of (a) TaSiz and (b) TiSiz films. For (a), 200 nm of Si + 100 nm of Ta films are annealed after deposition. For (b), 100 nm of Si + 50 nm of Ti films are annealed after deposition. ...................................................................... 94 (a) Sputtering mask for Ta (or Ti) deposition. (b) Shadow mask for Au deposition. ............................................................................ 95 (a), (b) Optical microscope images of Au contacts covering beneath of TiSiz contacts. (b) Close-up image of the area marked with a white rectangle in (a). (0) 3D STM image of the Au contact. The formation of many small Au islands is observed. ................................................................. 96 Schematic diagram of Au evaporation with a shadow mask. ................. 97 STM images of the Au boundaries (200x200 nmz). The density of Au increases from (a) to (c). DySiz NWs can be seen underneath of Au. ..... 97 (a) Before Ge deposition: DySiz NWs with interconnected 3D islands. (180x180 nmz) Coverage is unknown. (b) After 5 ML of Ge deposition on top of NWs. (200x200 nmz) ................................................. 98 (a) The STM image of 3 ML of interconnected Dy silicide islands. (b) Experimental setup for transport measurements. Surface resistances were measured at liquid helium temperature (4.2 K). (c) In order to measure the change in surface resistance of the sample (a), the sample was scratched in the horizontal and vertical directions. The numbers labeled on the bottom 0, l, and 2 indicate ‘before scratch’, ‘scratched in the horizontal direction’, and ‘scratched in the both directions’, respectively. ........................... 99 (a) Change in 4-terminal surface resistance of interconnected Dy silicide islands due to the sample scratching. A diagram shows the resistances in the vertical (open circles) and horizontal (solid circles) directions, respectively. Dashed lines on top are the surface resistance of clean Si surface. (b) Schematic diagrams of each step of sample scratching and corresponding current flow. ...................................................... 100 xiv Figure 6.11 STM images of (a) DySiz NWs at 0.22 ML and (b) interconnected Dy silicide islands at 3 ML. ........................................................... 102 Figure 6.12 Dy and Ho film conductivity at 4.2 K vs. metal coverages. Solid (open) circles show the surface conductivity of Dy (Ho) on Si(001). .............. 104 CHAPTER 7 Figure 7.1 Figure 7.2 Post growth deposited contacts Step I: Grow sparse network of NWs. Step 2: Deposit alignment marks: Use AFM to determine the NW/mark registration. Step 3: Deposit fine Au contacts. .............................. 108 Magnetic structures of heavy RE metals [after Koehler (1972)]. ........... 111 APPENDICES Figure A1 Figure A2 Figure A3 Schematic of the various possible hydrogenated structures on Si(001) surface. .............................................................................. 117 (a) Schematic diagram of a Si dimer ‘before’ and ‘after’ H-passivation. A half-filled 7t. anti-bonding orbital is completely filled due to H-desorption. (b) Empty-state (1.74 V) and (c) filled-state (-1.82 V) images of H—passivated Si(001) surface. The left half of the images is the surface after it was passivated by applying high tip voltage (20x30 nmz). Cross-sectional profile in the horizontal direction is shown next to each STM image. ........................................................................ 120 STM image (90x120 nmz) of a H—passivated Si(001) surface. Brighter areas correspond to an STM-induced H desorption. By operating the tip position and the voltage, we can write a “MSU” logo! ..................... 121 Figure A4 (a) Before H-passivation: DySiz NWs on Si(001) at 0.80 ML (180x180 nmz). Figure Bl Figure Cl Figure C2 Figure C3 (b) After H-passivation: NWs still survive, although there are small gaps on the NWs shown by white arrows (50x50 nmz). The change in the electrical properties of NWs due to the small gaps is unknown. ....................... 122 Schematic diagram of current passing through a thin film. .................. 123 Constant-current circuit using a ballast resistor. .............................. 124 (a) Circuit showing a two-terminal resistance measurement. (b) Equivalent circuit for (a). ...................................................................... 124 (a) Circuit showing a four-terminal resistance measurement. (b) Equivalent circuit for (a). ......................................................... 125 XV ABBREVIATIONS The following abbreviations are used in this thesis. 1D 2D 3D AFM BF DF LEED LEEM RT SEM STM STS TEM one-dimensional two-dimensional three-dimensional Atomic Force Microscopy bright field dark field Low Energy Electron Diffraction Low Energy Electron Microscopy monolayer (see p.23 for the definition) nanowire rare earth room temperature Scanning Electron Microscopy Scanning Tunneling Microscopy Scanning Tunneling Spectroscopy tilted bright field Transmission Electron Microscopy ultra high vacuum xvi Chapter 1 Introduction Our research objective is motivated by the potential for application of self-assembled rare earth (RE) silicide nanowires (NWs) for electronic and photonic devices in the semiconductor industry and to explore the structural and electrical properties of those W5 and their growth kinetic and dynamics. In this chapter, the background and basic ideas of rare earth (RE) silicides and their growth behavior and electric properties are introduced. The organization of this thesis will be described in the end of this chapter. 1.1 Rare earth (RE) silicides and self-assembled RE disilicide NWs RE disilicide NWs were firstly discovered by Preinesberger et al. in 1998 and they showed that Dy deposited on Si(001) forms NWs under certain conditions [1]. Unfortunately, this fascinating phenomenon was not noticed for more than 1 year. On November 1, 1999, the New York Times carried an article reporting that a Hewlett-Packard research group had produced conductive wires about 10 atoms wide [2]. These NW 3 were actually made of RE silicide [3]. This result was described as part of an impending revolution in molecular electronics and nanotechnology. Since then, epitaxial RE silicides grown on silicon have attracted particular interest. In the study of RE silicides grown on Si(001), it has been reported that so far eight types of RE silicides can form self-assembled NWs: YSig [4], ScSiz [5], SmSiz [6], GdSiz [5, 7, 8], DySiz [1, 5, 6, 9-12], HoSiz [9, 13, 14], ErSiz [3, 5, 6, 15, 16], and YbSiz [17]. For the other RE silicide/Si substrate systems, NW formation has been recently found on Dy/Si(111) [18], Gd/Si(111) [19], and Dy/Si(110) [20]. The formation of W3 is ascribed to small lattice mismatch in one direction and large lattice mismatch in the perpendicular direction on Si(001) substrate. The mechanics of nanowire formation will be explained on page 15. 1.1.1 The rare earths The rare earths are a group of 15 metallic elements: La (atomic number 57) through Lu (71) which appear in the extended sixth row of the periodic table as shown in Table 1.1. Along with Sc (21) and Y (39), they constitute the “RE elements”. RE elements can be classified into two groups: the light RE elements (La to Eu) and the heavy RE elements (Gd to Lu, Sc and Y). The RE metals are typically trivalent, which are distinct by the progressive filling of the 4f localized subshell, while the external electrons are shared between 63 and 5S subshells which determine the conduction band maximum with a total charge of between two and three electrons under normal conditions [21]. The 5d and 6s subshells mix in the solid state, and RE metals have a conduction band of hybrid s-d character; the extent of hybridization varies along the series in a rather irregular way. In contrast to most of the trivalent RE metals, Sm, Eu, and Yb are divalent. The divalence of Eu and Yb metals is well explained in terms of very stable configuration of half filled (4 f 7) and completely filled 4f band (4f ‘4 ), respectively [21]. Sm forms divalent or trivalent compounds depending on the chemical environment, because two valence states have rather similar energies [22, 23]. EBB. Eon 5E» uoEmemE 8m $52 E8 :3 £033 2808 mm BEEPQQ Be 232: mm :8. oZ~E 32:: Enos amen L08 SE; EUR :23 3&3 m2? D: «as Free o<$ =4: .53 Saree ammo ems hnee £3 as. am? Ema. Ems uZS 5% 00mm 3% 35820 53m 2am ESE 85820 mm Em: mic—10x 25m: E £52. m: 639.: m: 253 255 55o: HE§ mime :mu: mmee DDS. “m3. «m3 is cm: “<9“ omen 5% 9mg F; wig :5; EM: h: woes ox? 3: «H? Ema :er amen 6% oxen Fm £er am; amen 5%. BO: m<$ ems. am: am: out? 02? AZ; HNe. >3 amen am: gem cm: omen 342 00m «0: :NS =03 2% 00: our” 523 .53 >2 F2 em: «08 M2 5.2 5: m2 m: a: 2: 322 «Z: 02.: ma Ow Z» De mm one Em 0mm F 3.5820 05 no 058 Ever—om fl“ 29¢. 1.1.1.1 REsiIicides The investigation of the intrinsic character of RE silicide systems started to develop in the early 1980s; interface and thin film formation, electronic structure, and transport and magnetic properties [24-29]. Since then, studies of RE compound layers on semiconductors have become increasingly popular. Some of their advantages are the low growth temperatures, the good thermal stability, the low Schottky barrier on n-type Si substrates [24], and their compatibility to the Si technology. Moreover, these silicides are characterized by a small lattice mismatch relative to the Si(lll) surface, ranging from zero to 2.5 %, thereby allowing epitaxial growth with chemically sharp interfaces and a high degree of crystallinity and structural perfection. Another interest is that of magnetic properties of RE compounds, associated with a large value of the magnetic moment of 4f layer of RE elements. These properties are unusual and can be studied by electrical transport methods, because the magnetic ions interact strongly with conduction electrons (see Chapter 7). The RE silicides occur in three major stoichiometries: in metal rich form as RE5313 and RE5$i4 silicides, as monosilicides RESi, and as disilicides of varying composition, silicon rich RESiz.x with 09:50.33 (from RE3Si5 to RESiz) [30-34]. From the present viewpoint, the disilicides are of most interest, since this type of stoichiometry develops upon annealing of RE overlayers on Si. 1.1.2 Crystal structures of RESih The crystallographic properties have been described earlier in several papers [30, 35, 36]. As shown in Figure 1.1, three main crystallographic structures are encountered, depending on the RE metal and on the silicon content: the hexagonal Ale structure (p6/mmm space group) [Figure 1.1 (a)], the tetragonal ThSiz structure (I4/amd space group), and its distorted orthorhombic variant GdSiz (Imma space group) [Figure 1.1 (b)]. The structures and lattice constants are listed on Table 1.2. The first three RE metals (La, Ce, and Pr) crystallize in the tetragonal ThSiz or orthorhombic GdSiz structure. The next RE metals (Nd to Gd) can crystallize all the three structures depending on the RE elements. The hexagonal AIB2 structure is found mainly in heavy RE compounds like YbSiz.x and ScSiz.x for a stoichiometry close to R3Si5 [37, 38], and corresponds to the structure of epitaxial films grown on Si(lll). The type of structure adopted depends on the nature of the RE, on the value of x in RESiz.x and on the temperature. ‘Ale structure hosts the metal atoms in hexagonal planes along the c-axis ([0001] direction) [Figure 1.1 (a)]. The hexagonal RE planes alternate with Si atoms in a honeycomb arrangement. In this lattice, Si atoms occupy the interstitial sites between the hexagonal layers of RE atoms. However, the growth of epitaxially ordered RE3Si5 layers is observed and this bulk-like silicide forms a defected hexagonal Ale structure, consisting of stacked hexagonal RE planes and graphite-like Si planes with an ordered arrangement of vacancies at every sixth Si lattice site [39]. Therefore, the actual compositions of Si atoms vary between 1.67 and 2.00 per RE atom although the stoichiometry for perfect Ale-type RE silicides would be RESiz. Tetragonal ThSi2 structure [Figure 1.1 (b)] is closely related to the MB; structure. Starting from the Ale structure [Figure 1.1 (c)] and introducing shear planes parallel to (10I0)hex with a shear vector (b.m + cued/2 and a periodicity of flab“, the ThSiz structure is derived [Figure 1.1 (d)]. As a consequence, the lattice parameters of the two structures are related: a = b = (ahex + chcx)/2 and c z 2 J3 am The orthorhombic GdSiz structure is a deformation of the tetragonal ThSiz structure with a rt b. Raga} _I 9% m ((1) Figure 1.1 Spatial models for (a) the hexagonal Ale structure, and (b) tetragonal ThSiz structure or orthorhombic GdSiz structure. (c) Projection of Ale structure along the aux-axis. ((1) Projection of ThSizlGdSiz along the b-axis; Introduction of shear planes in Ale generates the ThSi2/GdSi2 structure. Table 1.2 RESiz lattice parameters and mismatches with Si substrate. (as, = 3.840 A) Silicide Structure Lattice parameter and mismatch Ref a (A) (%) b (A) (%) c (A) (%> SCSI” hexagonal Ale 3.66 (-4.69) 3.87 (+0.78) [40] tetragonal ThSi2 orthorhombic GdSiz YSiz hexagonal Ale 3.842 (+0.05) 4.144 (+7.92) [36] tetragonal ThSiz 4.04 (+5.21) 13.42 [40] orthorhombic GdSi2 4.05 (+5.52) 3.954 (+2.97) 13.36 [40] LaSiz hexagonal Ale tetragonal ThSiz 4.31 (+12.2) 13.80 [41] orthorhombic GdSiz 4.271 (+11.2) 4.182 (+8.91) 14.035 [42] CeSiz hexagonal Ale tetragonal ThSiz 4.192 (+9.17) 13.86 [42] orthorhombic GdSiz 4.19 (+9.11) 4.13 (+7.55) 13.92 [42] PrSi; hexagonal A132 tetragonal ThSiz 4.22 (+9.90) 13.71 [42] orthorhombic GdSiz 4.20 (+9.38) 4.16 (+8.33) 13.76 [42] NdSiz hexagonal Ale 4.12 (+7.29) 4.44 (+156) [43] tetragonal ThSiz 4.111 (+7.06) 13.56 [42] orthorhombic GdSiz 4.155 (+8.20) 4.125 (+7.42) 13.67 [42] PmSiz hexagonal Ale tetragonal ThSiz orthorhombic GdSiz Sm3Si5 hexagonal Ale 3.90 (+1.64) 4.21 (+9.64) [44] SmSiz tetragonal ThSiz 4.08 (+6.25) 13.51 [421 orthorhombic GdSi2 4.105 (+6.90) 4.035 (+5.08) 13.46 [421 EuSiz hexagonal Ale tetragonal ThSiz 4.29 (+11.7) 13.66 [421 orthorhombic GdSiz Table 1.2 (cont’d) Silicide Structure Lattice parameter and mismatch Ref a (A) (%) b (A) (%) c (A) <%) GdSiz hexagonal Ale 3.877 (+0.96) 4.172 (+8.65) [36] tetragonal ThSiz 4.10 (+6.77) 13.61 [42] orthorhombic GdSiz 4.09 (+6.51) 13.44 13.44 [42] TbSiz hexagonal Ale 3.847 (+0.18) 4.146 (+7.97) [36] tetragonal ThSiz orthorhombic GdSiz 4.057 (+5.65) 3.97 (+3.39) 13.375 [42] DySiz hexagonal A1B2 3.831 (-O.23) 4.121 (+7.32) [36] tetragonal ThSiz 4.03 (+4.95) 13.38 [42] orthorhombic GdSiz 4.04 (+5.21) 3.95 (+2.86) 13.34 [421 HoSiz hexagonal AlB2 3.816 (-0.63) 4.107 (+6.95) [36] tetragonal ThSiz 3.964 (+3.23) 13.297 [45] orthorhombic GdSiz 4.03 (+4.95) I 3.944 (+2.71) 13.30 [42] ErSiz.x hexagonal Ale 3.79 (-1.30) 4.09 (+6.51) [46] (“‘05) tetragonal ThSiz 3.96 (+3.13) 13.26 (+245) [46] orthorhombic GdSi; 5.6 (+458) | 13.26 (+245) 5.6 [461 5.6 (+453) | 13.26 (+245) 11.2 [461 TmSiz hexagonal Ale 3.768 (-l.88) 4.070 (+5.99) [36] tetragonal ThSiz orthorhombic GdSiz YbSiz hexagonal Ale 3.784 (-l.46) 4.098 (+6.71) [36] tetragonal ThSiz orthorhombic GdSiz LuSiz hexagonal Ale 3.747 (-2.42) 4.046 (+5.36) [36] tetragonal ThSiz orthorhombic GdSiz 1.1.2.1 Epitaxial growth of RE silicides on Si substrate Since RE silicides are extremely reactive with oxygen, which hindered early attempts at fabrication, both growth technique and epitaxial conditions are very important for improving epitaxial quality in silicides. Continued in-situ work in UHV systems is necessary for the production of RE silicide films of adequate quality. RE silicides can be grown on Si substrates by codeposition of RE and Si at elevated temperatures or by solid-phase epitaxy, i.e. by RE deposition on Si and subsequent annealing [25, 26, 36, 47-49]. In the latter case, the diffusion processes necessary for silicide formation give rise to a complex interplay of energetic and kinetic processes, requiring a detailed control of the process parameters. A Si(lll) substrate favors the epitaxial growth on it of the RE silicide phase with the Ale structure, because it contains three equivalent directions (the three <110> directions) related with a three-fold symmetry [Figure 1.2 (a)]. In the case of a Si(001) substrate, there are only two <110> directions related with a two-fold symmetry [Figure 1.3 (a)]and thus a tetragonal phase is favored for the formation of thin films. 1.1.2.1.1 Epitaxial growth of RE silicides on Si(lll) Small lattice mismatch to Si(lll) substrates makes epitaxial layer growth possible and ensures an ideal interface structure with minimized defect states [36]. For the growth of trivalent RE metals (except for Sm, Eu, and Yb) on Si(lll) at elevated temperatures, a p(1x1) interface is formed in the submonolayer range and consists of a bulk-like Si substrate with a single RE T4 site (second-layer substrate atom) with a rotated Si bilayer on top on the RE [50, 51] as shown in Figure 1.2. The growth of divalent RE metals 10 (Sm, Eu, and Yb) on Si(lll) behaves differently. Low energy electron diffraction (LEED) and scanning tunneling microscopy (STM) studies show 3x1, 3x2, 5x1, and 7x1 periodicities for Sm [52] and Yb [52-54], and 3x1, 3x3, and 5x5 for Eu [55]. As the coverage increases, J3 x J3R30° structure has been mainly observed for most of the RE metals [51, 56, 57] although two-dimensional (2D) Eu or Yb layer on Si(lll) has 2x1 reconstruction [58, 59]. The J3 xJ3R30° structure is assigned to an ordered superlattice of Si vacancies producing a «[3 xJ3 mesh in the silicide film [60]. This silicide film is identified to be hexagonal RESiz.x with Ale structure where the Si atoms are arranged in a hexagonal plane normal to the silicide c-axis, which has a good matching of the lattice parameters (as,= 0.384 nm) of the Si(lll) surface. This allows interface configurations with lattice mismatches of less than 2 %, resulting in high quality epitaxial growth. The orientation relationships of hexagonal RESiz.x on Si(lll) [46, 61-63] are determined to be: RESi2-x(0001)// Si(lll) and RESi2-x[1 100] // suizi] as shown in Figure 1.2. 11 00 9,. 00 00 [11201 QREmetar o 0 ._ 0 O 0 - ,/ ‘ _ 0 1st SI layer .0 0‘; .0 .0 .0 hex [110°] g 2nd St layer 0 O L 0 0 0 [101151 9 D '0/ O i _ _ (a) e o o o O 0 0 “2115‘ 0.384nm [11115i Chex [1100] [1120] [111]sr (b) [I2I]Si nails: Figure 1.2 Structural model for Si(lll)1x1-RE: (a) top view and (b) side view. 1.1.2.1.2 Epitaxial growth of RE silicides on Si(001) Many studies of thin RE silicide film growth on Si(lll) have been reported due to the easy epitaxial growth in this crystallographic orientation. However, until the discovery of self-assembled RE disilicide NWs [1, 3, 13], only a few studies were performed on Si(001) [61, 64-66] which exhibits a square unit cell as shown in Figure 1.3 (a). Most prior thin film growth studies looked at film thicknesses between 25 and 200 nm, where the most interesting features were the crystallographic phases present, and the epitaxial relationship with respect to the substrate. 0 {.13 RE metal Si 0 [0001] Chex 9 [1120] [1i00] (a) é—p [11in] [0001] Cher [001]Si "a651,: e'e % e [110181 [11019 (b) ahex Figure 1.3 Structural model of Si(001)1x1-RE: (a) top view and (b) side view. The orientation relationship between hexagonal Ale and NW structures are shown in (c). As already mentioned before, a Si(001) substrate contains only two <110> directions and thus the growth mode is different from Si(lll). In the case of ErSiz thin films, the lattice parameters of hexagonal Ale—type ErSiz (ahex = 0.379 nm and cm = 0.409 nm) are relatively close to the lattice parameter of Si(001) (as, = 0.384 nm) (see Table 1.2). 13 Epitaxy with the ahex- and chcx-axes parallel to the two mutually perpendicular <110>51 directions of the substrate has two obvious orientation variants, with the ahex- and chex-axes interchanging directions in the two variants. Although there are several epitaxial modes as observed in hexagonal ErSiz.x [67], the orientation relationships of hexagonal RESiz.x (h-RESi2-x) on Si(001) [46, 61-64] are generally determined to be: h—RESi2-x(1I00) // Si(001) and 114115312.x [0001] // Si[lIO] or h-RESi2-x(1 100) // Si(001) and h-RESiz.x [1150] // Si[lIO] (rotated by 90°) The vacancy ordering superstructure of unit cell (ch/3 aJ3 2c) is also found in epitaxial hexagonal YSi2-,,, TbSi2-x, DySi2-x, and ErSiz.x thin films with Ale structure on Si(001) [61]. The tetragonal ThSi2 and orthorhombic GdSiz structures are grown in the similar way. Several RESiz.x thin films such as DySiz.x [68], EI‘SIz-x [46, 62], and LuSiz.x [69] on Si(001) form tetragonal or orthorhombic structures and the orientation relationships of tetragonal / orthorhombic RESiz_x (t-RESiz-x) on Si(001) are determined to be: t-RESi2-x(010) // Si(001), t-RESiM [100] // Si[110], and t-RESiz.x [001] // Si[ 1 I0] For the Gd/Si(001) system, the coexistence of epitaxial h-GdSiz.x and o-GdSiz.x , and polycrystalline o-GdSiz is observed depending on the growth conditions [48, 63]. The formation of an epitaxial H0812 layer by solid phase reaction is reported, but no structural detail is given [70]. When the amount of metal deposited onto the Si surface is insufficient to create a continuous thin film, then the RE silicide grows in what can be described as the Stranski-Krastanov mode. In reality, the RE grows as a 2D metal induced surface reconstruction involving at most 1 ML of metal, and 3D silicide nanostructures that can 14 be described as either NWs or islands [1, 3-9, 11-14, 17]. Various types of periodicities due to the surface reconstruction are observed at low metal coverages: 2x3 and 2x4 for Nd [71], 2x4 and 2x7 for Gd [72], Dy [10, 72], and Ho [14], 2x3, c(6x2) and 2x1 for Sm [73], 2x3 and c(2x4) for Br [74], 1x5 [17]or 2x3, 2x4, 2x6, and c(6x2) [71, 75] for Yb. As mentioned previously, in the initial stages of 3D silicide growth, several RE metals form silicide NWs. The formation of these NWs arises from the anisotropic lattice mismatch between the RE silicide and the Si(001) substrate. The RE disilicide NW has a hexagonal Ale structure. The orientation relationships of hexagonal RE812 on Si(001) are determined to be: h-RESi2(1IOO) // Si(001), h-RESi2[0001]//Si[1IO], and h-RESi2[112O] //Si[110] as shown in Figures 1.3 (a) and 1.3 (b). The atomic resolution on top of NW shows p(1x1) [12, 17] and c(2x2) [8, 9, 11, 12, 14] which are consistent with a hexagonal Ale structure. Comparing the lattice mismatch in a and c directions of the hexagonal structure with Si(001) 1x1 unit cell, the lattice mismatch in a direction is very small (within it 1.64 % for Y, Sm, Gd, Dy, Ho, Br, and Yb) as shown in Table 1.2. On the other hand, the lattice mismatch in c direction is large (over 6.51 %). Because of those anisotropic lattice-mismatch strains, hexagonal RESiz grows faster in a direction than in c direction. Therefore, a and c directions of hexagonal structure correspond to a length and width of NW, respectively, as shown in Figure 1.3 (c). For ScSiz, the direction of large and small lattice mismatches is opposite, and thus the growth direction of ScSiz NW is perpendicular to that of other RESiz NWs. Also, YbSiz 15 NWs grow in a wide variety of orientations [17]. YbSiz NWs grow along the [001] direction which is 45 degrees rotated with respect to the other RESi,, NW-growth direction. They also merge into the other NWs running in the different orientation. This behavior might be controlled by larger lattice mismatch (-1.46 %) in a direction, and thus there is a competition between different orientations of YbSiz NWs on Si(001). When the lattice mismatch is large in both a and c directions such as NdSiz, three-dimensional (3D) compact silicide islands form instead of elongated NWs [71]. Those 3D islands are also observed in most of the RESi7/Si(001) system [1, 7, 12, 14, 16, 76] depending on the annealing temperatures and annealing durations. 1.1.2.2 Schottky barrier height of RE silicides on Si Almost all metals can form silicides, but most of them result in high Schottky barrier heights on n-type silicon and low barrier heights on p-type silicon. Unlike most other silicides, RE silicides produce a high Schottky barrier height of 0.7~0.8 eV on p-type Si and a low barrier height of 0.3~0.4 eV on n-type Si [24, 25, 49, 56, 60, 77-79] as listed in Table 1.3. This behavior makes them in particular interesting for applications as Ohmic contacts for n-type Si [25]. Furthermore, the corresponding high barrier height on p-type Si substrates [24] is an interesting property for infrared detectors or photovoltaic applications. In case of ErSiz-x/Si diodes, excellent rectifying behavior is observed for p-type Si, whereas the electrical characteristics of the film to n-type Si are relatively Ohmic at RT and rectifying at low temperatures [49]. Most of the values in Table 1.3 are quite close to one another, which reflects the great physical and chemical similarities of the different RE metals (including Sc and Y). 16 1.1.2.3 Electrical resistivity of RE silicides Our knowledge of the electrical properties of RE silicides is very limited and fragmentary. There are only a few reported works on electrical properties of RE silicide thin films. The values of resistivity of RE silicides are summarized in Table 1.4. RE silicide thin layers exhibit an almost metallic behavior. The resistivity decreases linearly with decreasing temperature and tends to a limiting residual resistivity, p0, at very low temperatures. The abrupt decrease of p(T) at very low temperatures is due to the antiferromagnetic ordering arising from the incomplete 4f shell of the RE3+ ion, which is operative below the Néel temperature TN. La [80], Gd [81-86], Th [83, 87], Dy [68, 83], Er [29, 49, 83, 84, 86, 88, 89], Tm [29, 62], and Yb [29] have a significant resistivity drop, but Y [29, 88] and Lu [69] do not. In case of LaSiz.x (OSXSOA), there is a superconducting transition at 2.5 K [80]. The resistivity p(T) is determined using Matthiessen’s rule which includes the various independent scattering mechanisms: ptT) = p0 + and) + pesto) where ,00 and pm(T) are the residual resistivity and the magnetic contribution to the resistivity, respectively. p0 arises due to imperfections in the crystalline lattice such as defects and impurities. pm(T) is the magnetic contribution to the resistivity above the Néel temperature TN and generated by scattering from randomly oriented spins [90]. The temperature dependent term pe-ph(T) is the contribution of the electron-phonon scattering mechanism expressed by the Bloch-Gruneisen formula [91]: 5 4 I T D/T x ._ T = T— dx p" M ) p (90] f (ex-lxl-e’x) 17 where OD is the Debye temperature, p’ is the high temperature limit of paph. At very low temperatures (GD/T > 20), the Bloch-Gruneisen formula can be expressed as pph(T) ~ T5 and at higher temperatures (T ~ 91)), the relation is approximated by the linear formula: pph(T) ~ T. 18 69:2: :oamEooaonm ”So—E 348:3 Ewen H2329: ”mm: EoEmomoc omecm.come> ”On; Eonfiomgo 38.405 ME. 32 sea 86 H “no 94> a :E >4 .35 me o> 54 >4 who my 5 .3 >4 wad E. SE Sosa 86 H moo 94> 4mm. >4 86 H 93 my NEE 4a >4 3o H So we Nwe: 3o H moo 94> 4E >4 86 H So my "55 >4 mod $3 925 $.o 49 8a eeeo 3o H moo 94> 4E Amine >4 8d H 43 me as >4 3o H two me :25 Re .5 22a 3 H 3 we Em 5: Sega So we o0 >4 who $3 223 Ed E 25> >4 woo :fi 92.4 mod > :2 >4 3 me :2 >4 to me _mom 3 ufiQEP—flmflo—fl 0:52:93 ,uflUEP—flmgnfl 0:62-88“ 023—? mu— aom 4.. 32482 93 aae 8:858 255 com me $282 $3 gee =3...er 255 be .32: mm .83593 E 099$ use -= so 330:8 can mEQE mm we .40 .mfimfic SE43 Esosom m4 033. 19 8:35.488 05 3209. 0240358943 ”m 64324.3 oEEoEozto AOV .Ecomaboa CC .3835: 95 £35.25 EE 6 FM 3 53560.4 065.55 HQ 53362 34553“ noQ .3 8m NEE 2.3 EN 04 o: co Harm Ba 8+ ".mou 44$ 98 04 o: om £30 043 v4 m 64 R 64 em 8a a. 2 "a 3.3wa 83 cm m4 m4 :3 co 6 am E cm :3 cm 043 v4 m £53 4me mm 2&3 8: ea 53 44a SN 04 c: we £23 :35 SN em 8E an £3 4me we S am 4me mm 25> as S4 3 4m; SE E E» 844 84 $3 mam 8 4m; 52 mm 844 e9. 35% an 4m 2 844 on _mom 8: a: flaw «om A65 c ES Gave. AEo Giveq Mom 945 o A80 Givq A80 @1464 you A80 SEQ AEo G5 em 54:95:... 2845484...— o==m osoam .8225 mm o £24284 1 29¢ 20 82 cm Smtom: 82 ow 8254 E? as 8 2 £53 :5. mam 4: £25 043 vv m er 43 8m 94 4mm; 32 8 E ea E 3% 4mg 8 E m: E 3% as com 8 £495. 3: mm em 3 E on E 43 $502: :3 8:8 E 24 E34 :a 828 E mom E :2 5: mm no 35:4 as E a flea ocmi «.3544: 3a cam: 5.44%: 82 we E :4 E .9. 52 we E on E em Pa 828 E 34 E 9: :a 828 E E E 4.4m £55 Nine 53 8m «44 4mm; 5: we 4o 3% 3455 :3 8.3 E om E 8 :55 $3 a? cam ewe—ow :3 8m 04 c: mo :36 Ga mam "54.2 :3 ma as: m: 43.2 Lem A5544 25 G15 95 6.3164 mom 945$ Eu Give. E8 33164 Lem Eu ~4in ES @136 SEE—444,4 3:85:54..— aem oEoEm 2.4.258 v4 055. 21 1.2 Organization of the thesis The remainder of this thesis is organized as follows. In Chapter 2, experimental background will be provided. The principles of scanning tunneling microscopy (STM) and low energy electron microscopy (LEEM), and the basic structure of the of Si(001) surface are presented. From Chapter 3 to Chapter 6, the growth behavior and electrical properties of RE silicides on Si(001) are discussed. In Chapter 3 and Chapter 4, the growth behavior of two different RE silicides, Ho and Sm silicide NWs and/or islands on Si(001) is studied by STM. The results of scanning tunneling spectroscopy (STS) show that HoSiz W3 and 3D islands are metallic. Chapter 5 contains LEEM studies of growth kinetics of Dy silicide islands on longer length scales and at elevated temperatures that are not accessible by STM. The LEEM studies were carried out at the University of Illinois in collaboration with Dr. W. Swiech of the Frederick Seitz Materials Research Laboratory. In Chapter 6, transport measurements of Dy thin films are presented. The details of ultra-high vacuum (UHV) sample preparation for transport are also provided. In Chapter 7, suggestions for future work will be described, continuing the themes of this thesis in several different areas. The Appendices include: a summary of studies of hydrogen on Si(001), a definition of sheet resistance, the principles of two- and four-terminal measurements, a description of the van der Pauw’s method, and a list of publications. 22 Chapter 2 Experimental In this chapter, the experimental methods used, the basic principles and mechanisms of scanning tunneling microscope (STM) and low energy electron microscope (LEEM), and the basic structures of Si(001) substrate are introduced. 2.1 Experimental and research methods All experiments except transport measurements were performed in an ultra high vacuum (UHV) system with a base pressure of < 2x10'10 Torr. The chamber was equipped with an STM, a four-grid low energy electron diffraction (LEED) optics, and facilities for sample heating and metal deposition. The Si(001) samples were chemically cleaned, and then outgassed by heating to 975°C. After outgassing, they were flashed briefly at 1175°C to remove surface oxide, and held at 975°C for 10 min, and then cooled slowly to room temperature (RT). The temperatures of the samples were measured using an optical pyrometer. Metals were evaporated from a tungsten wire basket, and the evaporation rate was calibrated by a quartz-crystal thickness monitor. For holmium (Ho) deposition, typical evaporation rates were 0.05~0.09 monolayers (ML) per minute [1 ML = was,2 = 1/(0.384 mm)2 = 6.78x10“ atoms/cm2 for Si(001): see p.35 for details of silicon surface] and total coverages studied were between 0.06~0.9 ML. For dysprosium (Dy) and samarium (Sm) depositions, typical evaporation rates were 0.1~0.6 MIJmin and 0.03~0.09 Ml/min and total coverages were between 0.l6~3 ML and 0.05~1.5 ML, respectively. The chamber pressure remained below 1x10”8 Torr during evaporation, and immediately recovered to below 5x10‘lo Torr after the metal source was turned off. All growth was done at a substrate temperature of 600°C unless 23 specifically mentioned. All STM imaging and LEED observation were done at RT. 2.2 Scanning tunneling microscopy (STM) STM was invented in 1981 [100, 101] by G Binnig and H. Rohrer, IBM in Zurich, Switzerland, gaining them the 1986 Nobel prize in physics. Their first STM image of Si(111)-7x7 reconstructed surface with its display of individual atoms [102] was the stimulus that initiated many STM projects all over the world. STM is a powerful technique for studying the structural and electronic properties of surfaces with real-space atomic resolution. When a sharp metallic tip is brought close enough (~10 A) to a conducting surface and a bias voltage (mV~1OV) is applied, the tunneling current flows between the tip and sample. The lateral resolution is about 1 A whereas a vertical resolution up to 0.01 A can be achieved. In addition, STM can be used not only in UHV, but also under other environments such as in air or in solution. Recently, atoms and molecules have been able to be moved controllably on surfaces, either by sliding them by means of the tip or by lifting them off the surface by the tip and depositing them at a different location [103-116]. The atomic-scale manipulation of surfaces by STM-tip is a promising candidate for future nanoelectronic device applications. 2.2.1 Basic principles of STM The operating principle of the STM is based on the quantum mechanical phenomenon of tunneling. Electrons transmit through a vacuum barrier between two conductors, in this case the sample and the tip, when they are brought very close together. A simple 24 understanding of this principle follows from consideration to the solution of the Schrddinger equation for a one-dimensional square barrier of height V0, as shown in Figure 2.1. For a particle of energy E < Vo incident on this square barrier from the left, the solutions of the Schrbdinger equation have the form: leti’“ +Re“’“ (x<0) w(x) = < AK” + 36““ (0 S x S s) (1) TeI‘k" (s < x) L where hk =J2mE, fix: 2m(V0 —E), R and T are reflection and transmission coefficients, respectively, and A and B are constants. ’lflf\\nr UU Lit 0 s x Figure 2.1 Wave function [[1 for an electron with kinetic energy E tunneling a thin potential banier of height V0 and width s. The wave function is oscillatory in the left and right regions of the barrier. Within the barrier, the wave function decays exponentially. The wave function [(1 decays exponentially into the classically forbidden region 0 S x S s with a decay length of 1/ K. For an electron at the Fermi level of a conductor with a work function ¢, this decay length would be 1/ K = ‘ihz / 2m¢ (Vo — E is just the work function). Since most work functions are around 4~5 eV, the decay length is typically 1/ K ~ 1 A. When a barrier of width sis much thicker than the wave function decay length of UK, the transmission probability T, or the tunneling current J decays 25 exponentially as: J = fille cc e‘2’° = e—Zn/ZmWhZ (2) m Thus the tunneling current drops by an order of magnitude for every 1 A. For the tip-sample distance of ~10 A, the tunneling current is a few nA. For tunneling between two conductors with a voltage difference V across the gap, only the states within eV above or below the Fermi level can contribute to tunneling; electrons in states within eV below the Fermi level on the negative side tunnel into empty states within eV above the Fermi level on the positive side. In the case of Figure 2.2, the tunneling current flows from a tip to a sample, since the positive bias voltage is applied to the sample. When the sample is positively and negatively biased, it is called “empty” and “filled” states, respectively. Tip r».-. Sample FL,I‘“L, o. . . Energy levels from which tunneling can occur Figure 2.2 Energy diagram for tunneling between a tip and a sample. Tunneling is possible only when there are empty states on the right. Such empty states are created when eV is applied to lower the Fermi level on the right. 26 2.2.2 STM data acquisition and image processing STM can be operated in several imaging modes: constant current mode, constant height mode, barrier height imaging, and scanning tunneling spectroscopy (STS). Most of the STM images presented in this thesis are acquired in the constant current mode unless specifically mentioned. In the constant current mode, the tip is scanned in the two lateral (x and y) dimensions, while a feedback circuit constantly adjusts the tip height to keep the current constant. The position of the tip is accurately controlled by the piezoelectric drivers. The tip position is recorded as a function of the x-y surface coordinates as shown in Figure 2.3. All STM bias voltages were applied to the sample with the tip at virtual ground. According to equation (2), the tunneling current reflects the surface topography only when the work function does not change during scanning. In addition, the tunneling current changes with the electronic state of the tip and sample. Therefore, essentially, STM image shows the electronic density of states at the sample surface instead of surface topography of the sample. The most common method of displaying STM data is to construct a grayscale image where the brightness represents the surface electronic density: Brighter areas are higher density of electrons at the surface. For easy understanding, brighter areas can be considered topographically higher than darker areas. In all of the STM images in this thesis, only minimal image processing has been done, consisting of planar background subtraction, drift correction, or contrast adjustment in order to extract the desired information. Image processing was done with either “Image SXM” which is the public domain image analysis software written by Steve Barrett [117], 27 or NOeSYS “Transform” software for the correction of image distortion. z Tunneling X Image ‘ Vz = Output Y x-y Raster Reference V0“ Current --—-h Piezoelectric 4 Tripod l I z v V Feedback \ A Position x Controller. .5 Tip ' . ; Vbias / Tunnel Current Sample Figure 2.3 Schematic diagram of an STM [118]. A metallic tip is mounted on a piezoelectric tripod actuator which can move the tip in the x-y plane of the sample surface, and in the z direction normal to the surface, with atomic resolution. 2.2.3 Scanning Tunneling Spectroscopy (STS) STS is used for studying local surface density of states with atomic resolution because the probe apex is very small consisting of a few atoms. The sensitivity of STS to 28 electronic structure can be a tremendous advantage over other spectroscopic methods such as ultraviolet photoelectron spectroscopy (UPS), X-ray photoelectron spectroscopy (XPS), Auger electron spectroscopy (ABS), and inverse photoelectron spectroscopy (IPES). A single tunneling spectrum can show surface states both above and below the Fermi level simultaneously, at the same spatial location. Figure 2.4 shows the schematic of an STS experiment. The sample-tip separation is fixed and the tunneling current I is recorded as a function of applied bias voltage V (I-V curve). Since I-V characteristics depend exponentially on the sample-tip separation, the or the normalized differential conductivity dI/dV , dan I /V logarithmic derivative which is a dimensionless quantity, is generally used to remove the influence of sample-tip separation. The approximate expression for the tunneling current [119] is given by: V I cc 1: ps(E)T(E)dE (3) where p3 is the surface density of states, T is the transmission probability in equation (1), and V is the applied sample voltage. After the first order approximation of E (or V), the normalized differential conductivity can be expressed as: dI/dV ~ p,(eV) I IV 1 V (4) W E p.(E)dE Therefore, is proportional to the normalized surface density of states. For the STS data plot and graphing, WaveMetrics “Igor Pro” software was used. 29 Figure 2.4 Schematic of an STS experiment. An STM tip is held above the sample, and the current is measured while the voltage is ramped. This measures the conductivity in the surface normal direction. 2.3 Low energy electron microscopy (LEEM) LEEM is a relatively new electron microscopy technique for in-situ real-time studies of surface dynamics, such as epitaxial growth, phase transitions, interface formations, strain relaxation phenomena, and chemisorption over a wide temperature range. LEEM was developed by E. Bauer and his student W. Telieps in 1985 [120] after its invention by E. Bauer in 1962 [121]. RM. Tromp and MC. Reuter designed a new LEEM that has higher resolution and wider capability of the instrument in 1991 [122]. Electron microscopes use beams of electrons instead of light, and magnets instead of lenses. However, the working principle of electron microscopes is more or less the same as in the optical microscope. In LEEM, the sample is illuminated with a well-collimated beam of low-energy electrons which diffract from the sample. The diffracted electrons are then used to form an image. In contrast to STM in which the probe tip is scanned back and forth across the surface to form images, all pixels are 30 imaged simultaneously from the illuminated area on the surface. At low energies of the order of 1~50 eV, in particular at very low energies (< 10 eV), high brightness and a good resolution can be obtained. Contrast is mainly produced by diffraction, surface t0pography, and local changes in electron work function and/or electron density. For the diffraction contrast, there are three imaging modes in LEEM as shown in Figure 2.5: bright-field (BF) imaging, tilted bright-field (TBF) imaging, and dark-field (DF) imaging. When electron beam is on the optical axis of objective lens, and the specularly reflected (0,0) beam is used to form a BF image. TBF image is also formed by using (0,0) beam, but with tilted incident angle. In DF imaging, non-specular diffracted beam (any beam other than (0,0) beam) is used. TBF or DF contrast must be used for the observation of azimuthally rotated domains, because domains are equivalent at normal incidence and they do not produce contrast in BF. For the interface contrast, surface steps are surrounded by stress and strain fields which cause local changes of the diffraction conditions. The resulting diffraction contrast may enhance or reduce the geometric phase contrast. (b) (c) specimen \ lens I (hk) (00) ' """k‘e‘ aperture - o (oo) (bk) 0 optical axis Figure 2.5 Diffraction contrast in LEEM (a) Bright field, (b) tilted bright field, (c) dark field. 31 The spatial resolution is determined mainly by the spherical and chromatic aberrations of the objective lens and by diffraction at the contrast aperture. Lateral resolution of LEEM is 5~10 nm on video-rate imaging [123, 124] , but in principle the best resolution of 3 nm is achievable [123]. Diffraction contrast provides depth resolution of a single atomic layer. 2.3.1 Basic physics of the low-energy electmn interaction with surface: Elastic scattering and inelastic scattering For the low energy electrons in the LEEM geometry, inelastic scattering and elastic backscattering are important (No forward scattering). At low energies, the elastic backscattering does not increase monotonically with nuclear charge because the incident electrons can ‘feel’ the details of potential distribution of neighbor atoms. Incident electrons can be reflected from surfaces with a high reflection coefficient within the first few atomic layers when their energy is in a band gap. (When electrons have wave vectors at the Brillouin zone at which the band gaps occur in the band structure, they fulfill the Laue condition k — k0 = 27:11 and are reflected.) Or they can penetrate deeply into the crystal when the incident energy E matches to E(k) states in the solid. The electron energy loss mechanism in solids is excitation of valence band electrons. For low energy electrons, the inelastic processes do not involve inner electron shells because of the insufficient energy but only outer electron shells. The electron density in the valence band (or Fermi energy EF) is nearly constant for most materials, and it gives the ‘universal curve’ of inelastic mean free path A =1/y , where M E) is the attenuation coefficient for inelastic scattering. ME) dominates I/(E), the attenuation coefficient for elastic scattering, above the 32 threshold ET for plasmon excitation. ET varies with E1: from 11 to 27 eV when E: changes from 5 to 15 eV [125]. At E>ET, inelastic scattering processes determine the penetration depth corresponding to a mean free path in the range of 0.3 to 1 nm. On the other hand, at E type direction. Each NW usually terminates at a step edge or at a perpendicular NW, but perpendicular NWs are on different terraces and do not cross. At higher coverages, rectangular islands can nucleate at the intersection of NWs. The NWs in Figure 3.2 (a) are comparatively long as the step density on this substrate is low. On substrates with higher step density, steps can flow to accommodate NWs, as shown in Figure 3.4 (a). In this image, the steps curve to follow the NWs so that each remains on a single terrace. Nevertheless, there is some limit to the number of steps that can flow to accommodate a NW, and so NW length is still limited by step density. 42 (%) (nm) 100 C] % of bare Si I % of Ho/Si surface 80 reconstruction - + - % of NWs and/or 3D 60 ' ' islands on surface _D_. average height of 40 NWs and/or 30 islands 20' (O N ('3 (O (D (D IO N O) 0. N. 0' <0. m, V. o' N. 0' coverage (ML) 0 O O O O O l V \ \ ’\ 7 30 islands Figure 3.3 Relationship between coverage and the percentage of surface covered with Ho surface reconstruction, NWs, and 3D islands. As shown in Figure 3.4 (a), there are two distinct NW morphologies; one is triangular in cross section, and the other is rectangular. Figures 3.4 (b) and 3.4 (c) are cross—sectional profiles of these NWs. At 02-03 ML, both types of NWs are seen at equal probability. As the coverage increases up to 0.4 ML, rectangular NWs are more often observed and at the same time NWs tend to form parallel bundles as shown in Figure 3.5 (a). While the narrow NWs in the edges of bundles are mostly triangular in cross section, those in the middle of the bundles are certainly rectangular. 43 [nm [NM] 1.6 1.6 1 2 1 2 0.8 0.8 A a c o 0.4 0.4 O 5 10 15 20[nm] 0 5 10 15 20 [nm] (b) triangular nanowire (c) rectangular nanowire Figure 3.4 (a) Triangular (left) and rectangular (right) NWs at 0.3 ML (200x200 nmz). Steps on the substrate flow to accommodate the NWs. Lines A—B and OD show the locations of the cross sections shown in (b) and (c). The structure of the rectangular NWs can be understood in terms of the bulk silicide HoSim, which has a hexagonal Ale type structure, with the [0001] axis parallel to the surface, and the [1120] direction along the long axis of the NW as shown in Figure 1.3 in Chapter 1. Atomic resolution on the top surface of rectangular NWs shows c(2x2) periodicity [Figure 3.5 (b)]. The same periodicity is seen on the surface of second layer NW growth. The height of the second layer NW is always 0.33 nm which corresponds to 1 bulk HoSiz unit cell height [9]. This height includes one metal layer and two Si layers. The apparent height of the first NW silicide layer varies because the electronic properties of the NW and substrate surface are different. The structure of the triangular 44 cross section NWs is not clear from the STM data, but it is likely that it has the same epitaxy characteristic as the rectangular NW with the [1120] direction lying along the long axis. It is possible that these triangular wires are simply a minimal width rectangular wire with perhaps a different step edge structure. Figure 3.5 (a) Bundled NWs with second layer growth at 0.36 ML (90x90 nmz). (b) Atomic resolution on bundled NWs (12X14 nmz). A single c(2x2) unit cell is marked with a white square. When the widths and the heights of silicide islands are measured, it is clear that the 3D compact silicide islands have widely varying dimensions whereas those of NWs are restricted to narrow ranges. Figures 3.6 (a) and 3.6 (b) are the distributions of the width and the height of single W3 and compact silicide islands, respectively. The top graphs show the details of the leftmost bar on the bottom graphs, and include data from both isolated NWs and NWs in bundles. Narrowcr widths shown by ‘A’ come from the edges of bundled NWs, and except for these data, there is no particular difference between the widths of bundled and isolated NWs. For isolated NWs, 45 rectangular NWs are wider than triangular ones. The most probable NW width is in the range in 2~3 nm, which corresponds to 5~9 silicide cells wide. The maximum NW width of ~5 nm can be related to the 6.8 % lattice mismatch between HoSiz and Si substrate along the c axis. This was also noted for DySiz NWs which have similar lattice mismatch and maximum width [9]. 20 40 a I o 17~o. 56ML b I 0.17~0.56ML g 15 ( ) (nanowires) .8 30 ( ) (nanowires) "E A E 8 1° H 8 2° N (U c 5 c B I I. I.. 3 1° #1: =11: O - L o l .. 0 2 3 4 5 8 7. “8 9 0 0.2 0.4 0.6 0.8 1.0 1.2 width lnml height [ml 3 1 -3 g - 0.17~0.56ML (nanowires) g - 0.17~O.56ML (nanowires) E m 0.40~0.90 ML (30 islands) E rat: 0.40~0.90 ML (30 islands) 0 , ._ . ..-. D I . (‘0 (0 a '5 l 2 2 i I S 2 ._ 0 40 80 120 160 200 ‘_ 0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 2 Width [mm ‘2 height [nm] Figure 3.6 The distributions of (a) the width and (b) the height of single NWs and 3D islands. The most probable NW height is 0.4~O.6 nm, which is 1~2 bulk HoSiz unit cell heights. Although we have statistics for comparatively few compact 3D islands, it is clear that both the width and the height of the compact silicide islands are spread over a much 46 wider range than those for the NWs. The fact that the 3D islands' sizes can exceed the maximum width of the NWs suggests that their growth is not coherent. It is possible that there are dislocations at the interface that do not propagate up to the surface of the islands. However, for most islands there is no clear indication of such dislocations from the STM images, unlike the dislocation networks observed for other epitaxial silicides grown on Si(lll) [148]. Further work with transmission electron microscopy (TEM) would be useful in clarifying the structure of the islands. 3.1.3 Two dimensional substrate surface reconstructions Figures 3.7 (a) and 3.7 (b) are empty- and filled-state images of slightly different areas of the same surface with 0.55 ML Ho. Coexistence of 2x4 and 2x7 structures on the 2D surface is observed. A NW runs diagonally across the image, and gray circles on both Figures 3.7 (a) and 3.7 (b) indicate the same physical position on the surface. Figure 3.7 (a) Empty- and (b) filled—state images of slightly different areas of the same surface with 0.55 ML Ho showing both 2x4 and 2x7 structures (30x30 nmz). The position of gray circle on the NW in (a) corresponds to that in (b). 47 Figures 3.8 (a) and 3.8 (b) show the empty-state (1.76 V) and filled-state (-I.76 V) images of the 2X4 structure corresponding to the square areas numbered by ‘1’ and ‘3’ in Figure 3.7, respectively. These images are of the same area of the surface. Both figures have three sets of 2X4 unit cells along the Si dimer row direction [llO] outlined in white. In the empty—state image [Figure 3.8 (a)], three maxima are visible in each ZIZIZIZIZIZIZIZIZIZIZIZIZIZ . . . . o . o 0'0 O 20400.....o o ....o 0 00'00 O 0"00 0 O O o (C) o oIo oIo oIo o 3 2 a a Figure 3.8 (a) Empty-state (1.76 V) and (b) filled-state (-l.76 V) images of the 2x4 phase in the reconstructed surface shown in Figure 3.7 (10X10 nmz). (a) and (b) correspond to the square areas numbered by 1 and 3 in Figure 3.7, respectively. (c) Line drawing showing the empty-state (gray circles) and filled-state (dark circles) maxima of the 2x4 structure. Clean 2x1 Si dimer positions are shown as a reference on the topmost row. 48 2x4 unit cell. There are two types of maxima that are lined up in one of two possible configurations; a darker maximum (small circle) between brighter maxima (big circles), or a brighter maximum between two darker maxima. Either combination is equally probable. In the filled state image [Figure 3.8 (b)], the contrast between the two types of maxima is reversed, and more pronounced. Bright oval maxima appear in the positions of the darker empty-state maxima. By comparison of both bias images with respect to the Si(001) substrate, the registration of empty and filled state of maxima of the 2x4 structure can be diagrammed as in Figure 3.8 (c). All the empty-state maxima are positioned in what would normally be Si dimer positions, and the filled-state maxima are located in the same positions as the darker empty-state maxima. The overall character of the 2x4 structure is the same as that seen for Dy on Si(001) [10]. Figures 3.9 (a) and 3.9 (b) show enlargements of the 2x7 structure corresponding to the square areas numbered by ‘2’ and ‘4’ in Figure 3.7, respectively. Two types of rows (row A and B) are observed and three types of maxima are visible in both biases. In the empty-state image [Figure 3.9 (a)], row A consists of oblong maxima elongated along the 7X direction. Row B consists of three maxima and the smaller maximum is located between two larger maxima. In the filled-state image, row A includes paired maxima, and row B has three maxima and the larger maximum is located between two smaller maxima. The registration of 2x7 phase can be obtained by comparison with positions of 2x4 phase from both bias images. The registration and approximate spatial extent of the empty- and filled-state maxima of the 2x7 structure are diagrammed in Figure 3.9 (c). This illustration contains five 49 rows (row ABABA) which are seen in Figure 3.9 (b), and the clean 2x1 Si dimer row positions are shown on top as a reference. The empty-state maxima of the bright features in row A and B are represented by gray oblongs and circles, respectively. The filled-state maxima are illustrated using dark circles in three sizes. In row A, the paired filled-state of maxima are located on the bridge site, and an oblong empty-state maximum is positioned off bridge site. In row B, in both states, the middle maximum is located on IIIIIIIIIIIIIIII I—l I'—'o‘ o'- 0.0 - o ”o “3.0' o o ° °o° are 0 I ’V o o I": o o o o o a - 0.0 0:0 0&0 o I IIIIIIII ‘ o o o o o” . o - T 0.0 0 ¢ OT (C) shifted row A and B by 1a Figure 3.9 (a) Empty-state (1.76 V) and (b) filled-state (-1.76 V) images of the 2x7 phase extracted from Figure 3.7. (a) and (b) correspond to the square areas numbered by 2 and 4, respectively. A 2x7 unit cell consists of three types of maxima in rows labeled A and B in both images. (c) Line drawing showing the empty-state (gray circles and oblongs) and filled-state (dark circles) maxima of the 2x7 structure. 2x1 Si dimer positions from the clean surface are shown as a reference on the topmost row. 50 the bridge site, and both edges of the maxima are positioned off bridge site. Row A or B in the both states often shifts by laSI in the perpendicular direction to the 2x1 Si dimer row ([110] direction). In this sense, the overall periodicity should be defined as the coexistence of 2x7 and c(2x14) unit cells. For simplicity, we denote this phase by 2x7. Since at a given coverage Ho metal can be distributed in NWs and both the 2x4 and the 2x7 reconstructed surfaces, it is difficult to decide the density of H0 in the 2x7 structure. Therefore we are not proposing an atomic structure for the 2x7 phase. 3.1.4 Ho content of the 2x4 substrate reconstruction and the NWs From the results in Figure 3.3, one can compare experimental coverage with calculated coverage under different assumptions for Ho content of both the 2D substrate reconstruction and the NWs as shown in Figure 3.10. The calculation was done on the basis of the first five coverages in Figure 3.3. The table on the right in Figure 3.10 shows two different assumptions of Ho content. The first assumption is the number of Ho atoms per 2x4 unit cell on the 2D surface: ‘A’, ‘B’, and ‘C’ indicate one, one and half, and two Ho atoms per 2x4 unit cell, respectively. The second assumption is the number of metal layers in a NW: ‘1’ and ‘2’ represent one and two metal layers in the minimum height NW, respectively. Here we also assumed that each metal layer in the NW has a density of 1 ML as implied by the bulk crystal structure as shown in Figure 1.3 in Chapter 1. The coverages shown in the graph on the left hand side where calculated under the six possible combinations of assumptions enumerated to the right, along with the measured percentage of the surface covered by bare surface, 2x4, and NWs as shown in Figure 3.3. The best model should come closest to the diagonal line. Error bars are 51 omitted for clarity. The error in coverage is less than 20 %, and the error in the vertical position of the symbols is less than 10 %. As can be seen in the figure, the best fit is for ‘BI ’ which implies 1.5 Ho atoms per 2x4 unit cell, and one layer of metal atoms in the minimum height NW. There is one layer of metal atoms in each bulk unit cell of the silicide, and so this is not a surprising result. Nevertheless, this coverage dependent analysis was necessary to deduce this fact since the geometric height of the first layer NWs as seen by STM was bias dependent. We cannot exclude the possibility that the minimum height NW may have more than two layers of Si atoms, but this depends in some sense on the definition of where the substrate ends and the NW begins. The 1.5 Ho atoms per 2x4 unit cell is somewhat more surprising. The simplest interpretation is that one of the two types of filled state maxima is associated with a single Ho atom. As shown in Figure 3.8, two possible configurations of 2x4 unit cells alternate on the surface, one with a bright-dim-bright arrangement of maxima, and the 0.7 3' 0.6- 8, 0.5 g 04 > - ” 8 0 ° . . '0 Q3 d’ -— Ideal km: A: 1HO/2X4 2 B X 2; 3:1.5 Ho/2x4 .53 0,2 - 0 0 - 31 C: 2Ho/2x4 2 D 82 , . . 8 0.1 — + 01 1. 1 metal layer In nanowwe O 02 2:2metal layers in nanowire 0 O 0.1 0.2 0.3 0.4 0.5 0.6 0.7 Experimental coverage (ML) Figure 3.10 Comparison of experimental coverage with calculated coverage under different assumptions of Ho content (see text). 52 other with a dim-bright-dim arrangement. There are equal populations of both types of unit cells, meaning that the average density of either type of maxima is 1.5 per 2x4 unit cell. This is not to say necessarily that the maxima seen in either bias is a Ho atom. In the case of the Dy 2x4, there appears to be three metal atoms per unit cell and yet bright and dim maxima are seen as well. Determination of the atomic structure of the 2x4 surface and the origin of the differences between the Ho and Dy induced structures will require complementary information from other surface analytic techniques. 3.1.5 Scanning rltrnneling Spectroscopy Figure 3.11 (a) shows the STS I-V curves and normalized spectra of a NW, 3D island, and the adjacent areas of reconstructed substrate at 0.56 ML. Since the 2D surface is reconstructed by Ho, the spectrum of the substrate is different from that of clean Si. The upper three curves are the actual I-V data, and the lower three curves are the normalized I/dV differential conductivity STS data were acquired at every pixel of the image shown in Figure 3.11 (c) which is a partial image of Figure 3.11 (b). Each STS curve is an average of the data taken in the area enclosed by the solid or dashed lines. The normalized spectra show that the NW has non-zero conductivity at near-zero bias. Figure 3.11 (d) is the STS I-V curves at near-zero bias voltage. It is clear that the I-V curve of the NW is steeper than that of 3D island and the reconstructed substrate. Although both NW and 3D island are metallic, NW has greater surface conductivity. Bulk HoSiz is metallic with a conductivity of approximate 4.4)(103 [9 cm]'1 at RT [99]. 53 (a) -— nanowire ’0'5 - - - SD island ----- substrate - 0.4 (NI)/(Ap/Ip) _ nano re — - - 30 island ..... substrate -o.2 -03 0'0 o.'1 o.'2 V(Vo|ts) Figure 3.11 (a) The STS I-V curves and normalized spectra for a NW, 3D island, and the reconstructed substrate. The surface topography is shown in (b) (150x150 nmz). The image (c) is a partial image of (b). Each STS spectrum was averaged from the STS data acquired in the areas shown by solid or dashed lines in the STM image shown in (c) (40x40 nmz). (d) Magnification of the STS I-V curves at near—zero bias voltage. 54 3.2 Conclusions We have studied the initial stages of Ho growth on the Si(001) surface. Metal coverages between 0.06~0.9 ML were deposited onto substrates heated to 600°C, and the samples were then cooled to RT for study by STM. Under these growth conditions, Ho forms either highly elongated silicide NWs or compact 3D silicide islands, together with several possible Ho induced 2D substrate surface reconstructions. STS spectra reveal that the NWs are more metallic than either the large 3D islands or the reconstructed substrate. 55 Chapter 4 Samarium induced surface reconstructions of Si(001) In this chapter, the growth behavior of Sm on Si(001) is described, including coverage dependence, 2D substrate surface reconstructions, and 3D island formation. Also we provide the first report that there are two different surface reconstructions: a lower coverage 2x3 phase and a higher coverage 3x2 phase. 4.1 Results and Discussion 4.1.1 Coverage dependence Figure 4.1 shows the coverage dependence of Sm growth on Si(001) at 600°C. The horizontal axis indicates the metal coverage in monolayers (ML), and both LEED patterns and topography as seen by STM are shown. Gray circles are the metal coverages at which the experiments were performed. The error in coverage is less than 20 %. Generally, a 2x3 phase dominates the 2D reconstructed surface for the coverages aaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaaa . . I D . . 0‘ 2D reconstructed surface only 3D island 0 0.5 1.0 1.5 [ 1 Figure 4.1 Coverage dependence of LEED patterns and topography for Sm/Si(001) at 600°C. 56 below 1 ML. At very low coverages (below 0.07 ML), only a 2x1 LEED pattern is observed. A c(6x2) phase coexists with 2x3 phase above 0.54 ML as in the LEED pattern shown in Figure 4.2. Figure 4.2 LEED pattern of Sm/Si(001) at 0.54 ML shows 2x3 (black rectangle) and c(6x2) (black arrows) periodicities. The LEED pattern was obtained from the surface shown in Figures 4.8 (c) and 4.9 (a). The electron beam energy is 63 eV. 4.1.2 General surface structure evolution with metal coverage An overall description of all of the ordered surface reconstructions will be presented, followed by a more detailed discussion of each surface structure. At low coverages, 0 < 0.2 ML, there are two Sm associated surface structures: a 2x3 ordered phase and a more sparse zigzag chain structure. Figure 4.3 shows a pair of images (15x15 nmz) in (a) empty and (b) filled states at 0.067 ML. Two 2x3 unit cells on the different terraces are shown with white and black rectangles on each image. The rectangles are slightly distorted as a result of drift while acquiring the images. Also shown are line traces taken from both images along the dashed lines A shown in Figures 57 4.3 (a) and 4.3 (b). A small terrace of clean 2x1 Si dimers is seen in the central part of the image. Zigzag chains of brighter features are seen on the upper part of the terrace, and there is a small patch of brighter ordered 2x3 reconstruction on top of this terrace as well. Surrounding the central terrace are larger ordered areas of 2x3 reconstruction that are one atomic height step lower than the brighter 2x3 patch in the central lower part of the image. 0.20 - - 0.24 - - 0.16 -A - 02° ' ' I ' 0.16 -A - 0.12- - _ _ L l l l l l I l 0'12 -l I I l I I I r- 02468101214nm 02468101214nm Figure 4.3 (a) Empty-state (1.90 V) and (b) filled-state (-1.91 V) STM images of the same area showing 2x3 phase and the zigzag chain structures at 0.067 ML (15x15 nm‘). Cross-sectional profile along the line A is shown on the bottom of each image. 58 The trace along A in both images shows that the large areas of 2x3 are lower than the central 2x1 terrace at both bias polarities [0.019 (0.039) nm lower in empty (filled) states]. This makes it reasonable to assume that the 2x3 is geometrically lower than the 2x1. The height of the zigzag chains is similar to that of the 2X3 area on top of the central terrace [about 0.116 (0.065) nm in empty (filled) states], and the upper 2x3 structure is about 0.12 nm higher than the lower 2x3, which is about the single atomic height step (0.125 nm) on the Si(001) surface. This supports the view that the 2x3 structure on both terraces is in fact the same. One can describe this 2x3 structure as being built up on top of a Si(001) terrace, or embedded in a terrace that is one level lower. For the purposes of this thesis, we will describe the 2x3 as being embedded into the level of the central terrace. This means that the upper 2x3 structure is effectively embedded in the terrace above the one that is not visible in this image. With this definition, the ordered reconstruction is denoted as a 2x3 structure embedded in a 2x1 clean Si terrace, implying that the 2x periodicity is along the same direction as the Si dimerization direction of the surrounding clean Si areas. In the medium coverage regime (0.5 < 0 < 1 ML) the surface looks at first glance to be similar to the lower coverages. Figure 4.4 shows an image of a surface at 0.59 ML, which is covered in a combination of 2x1 Si and a 2x3 reconstruction. However, closer examination reveals several differences. Firstly, the uncovered Si areas have a large concentration of dark missing dimer defects that are arranged in zigzag patterns, rather than bright protrusions. Secondly, the 3x periodicity of the “2x3” is perpendicular to the Si dimer rows in the 2X1 terrace at a similar height in the center of this image. This 59 is perpendicular to the 2x3 structure seen at the lower coverages, and so this structure is denoted as “3x2”. It should be noted that anti—phase defects in this 3X2 structure can also produce a c(6x2) periodicity that can also be seen in the LEED patterns as shown in Figure 4.2. Figure 4.4 15X15 nm2 image of a surface at 0.59 ML. Both a 3x2 phase and a highly defected Si 2x1 surface are seen. The bias voltage is —1.77 V. In the high coverage regime (0 > 1 ML), 3D silicide islands are seen in coexistence with a 2x1 reconstructed substrate. An STM image of a 3D island is shown in Figure 4.5. It is known that Sm-Si system forms Sm38i5 and SmSiz compounds [42, 44]. Although no elongated structure is observed for 3D Sm silicide compounds in this work, it has been reported that SmSiz.x forms NWs on vicinal Si(001) substrate at the same growth temperature [6]. This structural difference in the formation of silicide is likely due to the metal deposition rate, sample annealing temperature, and substrate type. 60 Figure 4.5 Sm silicide island at 1.46 ML (300x300 nmz). 4.1.3 The low coverage 2x3 and zigzag chain structures Figures 4.6 (a) and 4.6 (b) show empty-state (1.57 V) and filled-state (-l.57 V) images of the same area at 0.067 ML, respectively. No 2x3 unit cells are marked with white rectangles on each image. An out—of—phase boundary is also indicated with a dashed line. This type of phase shift in 2X3 structures can be frequently observed. Si dimer rows running diagonally are shown in the lower right region. In both empty and filled states, two round maxima can be observed with different size and contrast within each 2x3 unit cell. Here, the bright and dim maxima are represented by circles and ovals in the 2X3 unit cell, respectively, on each image. By comparing both bias images of Figures 4.6 (a) and 4.6 (b) with respect to the Si(001) substrate, the registration of empty (white) and filled (gray) states of maxima of the 2x3 structure can be diagramed as in Figure 4.6 (c). The 2X3 maxima are in registry with the centers of Si dimers on any line that is drawn along the Si dimer row directions and 3x periodicities are aligned along the [ITO] direction, as is readily apparent from these images. The 61 registry in the perpendicular direction has been determined from other similar images that are not shown here. This registry of features is different from Ragan et al.’s result of Sm on vicinal Si(001) where Sm atoms are positioned on the trenches between the Si dimer rows [6]. (c) 00 1.57 V :31 -1.57 V H Si dimer 0 1st Si layer 0 2nd Si layer Si [110] Figure 4.6 (a) Empty-state (1.57 V) and (b) filled-state (-1.57 V) STM images of the same area showing 2x3 phase at 0.067 NIL (10x10 nmz). There are out-of-phase boundaries shown with dashed lines. Two 2x3 unit cells are marked with white rectangles. (c) Line drawing showing the empty-state (white circles and ovals) and filled-state (gray circles and ovals) maxima of the 2X3 structure. On the line drawing, in empty states, the maxima (white circles and ovals) are situated almost in the center of an underlying Si square lattice consisting of four Si atoms in each 62 corner, but they are slightly off center. The filled-state maxima (gray circles and ovals) lie between two Si atoms, which is shifted by half as] (as, = 0.384 nm) from the underlying Si square lattice. The dim maxima in empty states (white ovals) are positioned almost between the filled-states maxima, while the bright maxima in empty states (white circles) positioned on the dark region of filled states. In fact, the number and the appearance of the maxima appearing in STM images are bias dependent. In empty states, below 2.56 V, two maxima can be seen with different contrast in the 2x3 unit cell [Figures 4.3 (a) and 4.6 (a)]. Above 2.74 V, only one maximum is visible (not shown). In filled states, below —1.85 V, there are two maxima in the 2x3 unit cell [Figure 4.6 (b)], while only one maximum can be seen above -1.90 V [Figure 4.3 (b)]. The bias can also affect the relative height of the 2x3 structure and the 2x1 clean Si surface. Zigzag chain structures are also observed perpendicular to the Si 2x1 dimer rows as shown in Figures 4.3 (a) and 4.3 (b). The maxima of the zigzag chain structures in empty and filled states are located over the trenches between the Si dimer rows. The features of maxima in 2x3 structure and zigzag chain structures appear the same, although they look different due to a tip effect in Figure 4.3 (a). Figure 4.7 (a) shows the more detailed features of zigzag chain structures with surrounding buckled Si dimers. An area of 2x3 phase can be also seen in the lower right area. In this bias voltage (-1.82 V), the maxima of dimerlike pairs can be observed to have an oblong shape. Comparing the two filled-state images {—1.91 V for Figure 4.3 (b), and -1.82 V for Figure 4.7 (a)], the paired maxima in Figure 4.7 (a) are positioned over the large maximum in Figure 4.3 (b). The dimerlike pairs are 2aSI wide in [110] direction and shifted by lag, in [110] direction 63 resulting zigzag chain structures. Since these dimerlike pairs have a width of 203,, there is not enough space to align straight on the trenches with a 2as; periodicity, and thus zigzag chains are generated. However, they can align in a row with a 3a5I periodicity forming 2x3 structure as indicated with a rectangle A in Figure 4.7 (a). The maxima in the zigzag chain structure can move along the dimer rows under repeated scanning with STM, and so the pattern of zigzag chain structure slightly changes from image to image. In fact, the movement of zigzag chain causes noise in STM images as appears in the oval area B in Figure 4.7 (a). However, the arrangement of 2x3 reconstruction in the small patch does not change even after 10 scans. In addition, Sm zigzag chain structures induce buckling of the Si dimer rows with c(4x2) arrangement in which the pattern of buckling is out of phase as illustrated in Figure 4.7 (b). A c(4x2) unit cell is marked with a black rectangle. The Si dimer buckling extends possibly more than 20 dimers along a row without decay. However, the buckling does not affect the adjacent row: if there is no Sm atom in a certain row, the buckling does not occur in this row even though there are Sm atoms forming zigzag chain structures in the adjacent rows. This implies that the interaction between Srn and Si atoms in the same row is very strong, whereas the Sm-Si interaction in the perpendicular to the dimer rows is ignorable. 83?: 1’23: 32?: O 0 ,:.°.:’ $080.0 ,:::8 a? 2:00 0008 3: 3: 3. 3° 8 o 23 12? 0.000 O O O 0 0‘0 0 O O ' .3. 0’0 0 O O O 0 O O O O 0 0:0 0 O 0 0‘0 0 0:0 0 O 0 Figure 4.7 (a) Filled-state (-1.82 V) image of zigzag chain structures with buckled Si dimers below 0.1 ML (10x10 nm‘). The 2x3 unit cell is also shown as a pair of two maxima on the lower right area. (b) Line drawing illustrating the zigzag chain structures with surrounding buckled Si dimers. This dimer buckling causes c(4x2) periodicity. A c(4x2) unit cell is marked with a black rectangle. The comparison of the experimental coverage with the percentage of the 2D surface covered with Sm provides a result that there is one Sm atom in 2x3 unit cell. This is again different from Ragan et al.’s result which suggests two Sm atoms in a 2x3 unit cell [6]. It is not possible to propose a model for this phase based on the limited amount of information that we can deduce from the STM data. However, we note that the appearance, bias dependence, and registration of STM maxima are all very similar to that observed for a Ba induced 2X3 phase on Si(001) [149]. Ba is divalent with an ionic radius of 0.135~0.l63 nm and so it might be expected to behave similarly to divalent Sm (ionic radius 0.113 nm). However, the Ba model proposed in ref [149] requires 2 metal atoms per unit cell. 65 4.1.4 The medium coverage regime 3x2 and c(6x2) phases As noted previously in the coverage regime (0.5 < 0 < 1 ML), the surface is covered in a combination of highly defective Si 2x1, and Sm reconstructed 3x2 and c(6x2). A large-scale image is shown in Figure 4.8 (a). The unusual step structure is a signature of large scale Si adatom displacement during the formation of a reconstructed surface. Similar step structures have been seen in the case of B on Si(001) [150, 151] and Ge on Si(001) [152, 153]. Terrace A has a straight step edges (SA step) which run along the dimer row direction, and terrace B has jigsaw-shaped step edges (83 step). Both terraces are covered with the same surface structures: the mixture of 3x2 and c(6x2) phases, and Si dimer rows with zigzag defect structures. In Figure 4.8 (b), the area of the front edge of the jigsaw shape on terrace B shows the alternate terraces A and B. 3x2 and c(6x2) phases can be seen in the four corners of the image and Si dimer rows with diagonal rows of missing dimer defects appear in the middle. Close-up image of zigzag defects is shown in Figure 4.8 (c). The ends of zigzag defects are indicated with white or black arrows in Figures 4.8 (b) and 4.8 (c). The angles between Si dimer rows and the zigzag defects (0° S B S 90°) vary from 41 to 76 degrees depending on the surface area. These missing dimer defects relieve strain due to a low density of Sm atoms substituting into the Si surface. The Sm atoms do not show up directly in the empty state images, but they make the filled state images of the 2x1 structure appear disordered. Figures 4.4 and 4.8 (b) shows the coexistence of Si 2x1 and the 3x2 phase. From this image, the 3x2 phase is measured to be about 0.079 (0.078) nm below the surrounding Si 2x1 areas in empty (filled) states. This measured height is slightly different from that of the lower coverage 2x3 structure, and once again, the reconstruction is rotated by 90°. 66 Figure 4.8 (a) Large—scale STM image of a surface consisting of two types of step edges. Straight step edges (S A step) run along the dimer row direction on terrace A, and terraces B have jigsaw-shaped step edges (SB step). (b) Both types of terraces consist of 3x2 and c(6x2) phases and Si 2x1 dimer with zigzag defects shown with white arrows. (c) Close-up image of zigzag defects on Si dimer rows. The ends of four zigzag defect structures are marked with black and white arrows. All three images were at different Sm coverages: (a) 0.93 ML, (b) 0.59 NIL, and (c) 0.54 ML. The image sizes are (a) 300x300 nmz, (b) 50x50 nmz, and (0) 25x25 nmz. Bias voltages are (a) —2.31 V, (b) —1.77 V, and (c) 1.28 V. 67 Figure 4.9 10x10 nrn2 images in (a) empty states (1.28 V) and (b) filled states (-1.77 V) showing 3x2 and c(6x2) phases at different coverages: (a) 0.54 ML and (b) 0.59 ML. Both 3x2 and c(6x2) unit cells are marked with white rectangles on each image. (c) Line drawing showing the empty-state (gray circles) and filled-state (dark circles) maxima of the 3x2 and c(6x2) structures. Si dimers are shown on the leftmost row as a reference. Figures 4.9 (a) and 4.9 (b) show the empty- and filled-state images of 3x2 and c(6x2) structures, respectively. Both 3x2 and c(6x2) unit cells are marked with white rectangles on each image. The STM images were acquired from the surfaces at different coverages. In empty states [Figure 4.9 (a)], two types of round maxima are observed with slightly different contrast. Bright and dark maxima appear alternatively with a 1.5a5I periodicity in the perpendicular to the dimer rows (along the [110] direction). The 68 bright and dark maxima each form rows along the dimer row direction ([1 I0] direction) with a 2aSI periodicity. In filled states [Figure 4.9 (b)], only one type of maximum is visible. In the similar way as empty states, the filled-state maxima lie along the [1T0] direction with a 2a3i periodicity forming rows. These rows appear along [110] direction with a 3am periodicity. Figure 4.9 (c) shows the registration of the maxima in the 3x2 and c(6x2) phases with respect to the Si(001) substrate. The registration was decided by the same surface area which is not shown here. In empty states, the bright and dark maxima are distinguished with large and small gray circles, respectively. The filled-state maxima are marked with dark circles. All the maxima are located on the bridge site in both biases. For the 3x2 structure in empty states, bright maxima are situated in the center of underlying 1x1 Si lattice, while dark maxima locating between bright maxima lie on the two underlying Si atoms. The filled-state maxima are positioned between the empty-state bright maxima on the same rows. When the positions of the maxima are shifted by lag, in the dimer row direction, c(6x2) structure can be obtained. Thus 3x2 and c(6x2) phases have basically the same structure. The number of Sm atoms per 3x2 unit cell is definitely more than 2 according to the experimental coverages (0.54 and 0.59 ML). However, this coverage can be affected by the possibility that 3D silicide compounds sparsely exist somewhere on the surface which STM did not find. When the coverage increases to 0.93 ML, more than 90 % of the surface is covered with Si 2x1 with zigzag defects and few 3x2 phase can be observed, although the LEED pattern shows 2x3, c(6x2), and 2x1 phases. 69 4.2 Conclusions In the initial stage of Sm deposition on Si(001) at 600°C, the surface is reconstructed with a 2x3 phase. Sm-induced zigzag chain structures coexist with a 2x3 structure. STM observations show that the appearance of the 2x3 structure is bias-dependent. As the coverage is increased, a second 3x2 phase appears with a co-existing c(6x2) phase. The 3x2 phase is different in appearance from the lower coverage 2x3 phase. At higher coverages, large 3D Sm silicide islands are found. This research was partially supported by NSF grant DMR-0305472. 70 Chapter 5 Topographic evolution of Dy silicide on Si(001): Shape transitions from nanowires to 3D islands In this chapter, we show that DySiz NWs can be formed with a uniform width by controlling the deposition rate and the deposition duration, which will play an important role for nanodevice assembly. The shape transition from NWs to 3D islands is studied by scanning tunneling microscopy (STM). Low energy electron microscopy (LEEM) is used to study Dy growth on longer length scales that are not accessible by STM. Dynamic behavior and growth of 3D Dy disilicide islands at elevated temperatures are reported. 5.1 Results and Discussion 5.1.1 STM experiments Figure 5.1 shows STM images of DySiz NWs at 0.58 ML (a) before and (b) after annealing at 600°C for 8 minutes. As shown in Figure 5.1 (3), almost the whole surface is covered with a 2x7 substrate reconstruction. There are also some NWs with a comparatively uniform width of 7 times the lattice constant of silicon as, (= 0.384 nm). Formation of NWs with uniform widths with My may be attributed to coexistence with the well-ordered 2x7 substrate reconstruction. A more detailed image of the 2X7 superstructure is shown in Figure 5.2 (a). Two unit cells are marked with white rectangles. The detailed atomic structure of Dy 2x7 superstructure on Si(001) surface is reported by 31. Liu et al. [72] Rapid deposition provides only 2x7 superstructure on the 2D surface regardless of the metal coverage. Six different metal coverages from 0.19 to 0.84 ML were obtained by 71 adjusting the deposition rates and keeping the deposition time fixed for rapid deposition experiments. Surface morphology at those coverages only consists of 2X7 superstructure on 2D surface and NWs for 3D structure. More NWs are observed at higher coverages and NWs tend to bundle above ~0.8 ML. On the other hand, 2x7 superstructure dominates on 2D surface at any coverage between 0.19 and 0.84 ML. After 8 min annealing at 600°C, bundled NWs are more often observed and there are more second layer growth on top of the NWs as shown in Figure 5.1 (b). In comparison with ‘before’ annealing, the NW width distribution ‘after’ annealing is broader because of the presence of bundled NWs. However, the overall percentage of the surface occupied by NWs does not change significantly ‘before’ (15.1 %) and ‘after’ (18.5 %) annealing. Figure 5.1 STM images of DySiz NWs at 0.58 ML (a) before and (b) after annealing at 600°C for 8 minutes. (a) Dy was deposited at 600°C. Almost all the 2D surface is covered with 2x7 stripe structure and NWs have comparatively uniform width distribution. 2x7 phase generally disappears after 1 min annealing at 600°C. (b) After annealing, 2x4 superstructure coexists with bare silicon. More bundled NWs are observed and the width of NWs becomes wider. There is also more second layer growth on top of the NWs. 72 The 2x7 superstructure is swept away and a disordered 2x4 superstructure appears with bare silicon on the 2D surface as shown in Figure 5.2 (b). Two 2x4 unit cells are marked with white rectangles. The 2x4 superstructure consists of one or two round bright maxima in filled states. The detailed atomic structure of Dy 2X4 phase on Si(001) is explained by B.Z. Liu et al. [10] 30 X 30 nm2 Figure 5.2 Small-scale images of (a) and (b) in Figure 5.1 showing (a) 2x7 (-1.80 V) and (b) 2x4 (+1.04 V) reconstructed surface around NWs. Two unit cells are marked with white rectangles in each image. Generally, more than 1 min annealing at 600°C replaces the whole 2x7 superstructure by 2x4 superstructure coexisting with bare silicon surface. Immediate disappearance of the 2x7 phase can be seen by LEED when the sample heating starts. 2x1 and x4 LEED patterns are observed during annealing at 600°C after the 2x7 phase disappears. The extreme sensitivity of the 2x7 superstructure to annealing is characteristic of the overall ease with which the 2X7 can be converted to the 2x4 structure. It can be further noted that the balance between 2x7 and 2X4 on the substrate is also sensitive to 73 deposition rate. In earlier work where the Dy deposition was slower, 2x7 and 2x4 coexist on the substrate over a range of metal coverages. The present results show that fast deposition with post deposition quenching results in a pure 2x7 substrate phase which can be converted rapidly to 2X4 by annealing. Figure 5.3 shows the evolution of the 0.58 ML sample shown in Figures 5.1 and 5.2, after repeated annealing cycling. The percentage of the surface covered with 2D substrate reconstructions (2x7, 2x4, or 2x1) and 3D structure (NWs or islands) is expressed in the vertical axis. Annealing processes are shown in the horizontal axis. The sample was annealed seven times in total: 1St annealing (600°C 8min), 2"‘1 (600°C 20 (%l 5’ Around the NWs Around the Islands (a) Before (b) 1st (c) 2nd (d) 6th 531 7th Annealing Annealing Annealing Annealing neallng 600°C 600°C 700°C 720°C 8 min 20 min 10 min 10 min Figure 5.3 Annealing behavior of Dy grown on Si(001) at 0.58 ML. 3D structures of DySiz transit from large 3D islands by annealing. 2x7 superstructure dominates on the 2D surface after rapid deposition of Dy. After 1 min annealing at 600°C, 2x7 superstructure is replaced by 2X4 superstructure and bare Si. Less 2x4 structure appears with higher temperature annealing or longer annealing time. 74 min), 3rd (635°C 12 min), 4” (655°C 10 nrin), 5th (678°C 10 rrrin), 6th (700°C 10 min), and 7th (720°C 10 min). STM images were taken after each annealing step. The large-scale STM images of changes in surface morphology (a)~(e) in Figure 5.4 correspond to the bar graphs labeled with the same letters in Figure 5.3. The measured areas were averaged over more than five images after each annealing step. The results of 3rd ~ 5th annealing are omitted for simplicity because the surface morphologies after 3“1 ~ 5th annealing are very similar to that after 2“d annealing. Elongated NWs with uniform width and 2x7 superstructure dominate on the surface after rapid deposition without post-annealing [Figure 5.3 (a)] as already shown in Figures 5.1 (a) and 5.2 (a). After 1St annealing, the substrate surface reconstruction completely changes to 2x4 superstructure while the percentage covered by NWs remains almost the same. The fraction of the substrate covered with 2x4 superstructure [Figure 5.3 (b)] is 66 i 8 % and the remainder is bare Si. STM images indicate that 2x4 superstructure does not unifonnly distribute on the 2D surface. Furthermore, there seems to be no clear influence of either steps or NWs on the distribution of 2x4 superstructure on the surface. The metal content on the surface is conserved before and after the 1" annealing step at 600°C. The calculation below indicates that the increased quantity of Dy atoms in the NW 3 is equal to the decreased quantity of Dy atoms on the 2D surface. Firstly, the calculation for the metal content in NWs is performed. Dy metal density in the DySiz is poy = 19.1 atoms/nm3, where the lattice parameters of hexagonal DySiz structure are a = 0.3831 nm and c = 0.4121 nm [36]. If we assume that the minimum height NWs include one bulk DySiz unit cell high (= one Dy layer plus two silicon layers = 0.33 nm 75 high), 0.93 ML (= 19.1 atoms/nm3 X 0.33 nm = 6.30x1014 atoms/cm‘) of Dy covers the whole surface as DySiz NWs. Because there is ~20% of second layer growth on top of NWs after annealing, the percentage difference of metal content in NWs before and after annealing is: A(NWs) = 18.5 % x 1.2 (after annealing) — 15.1 % (before annealing) = 7.1 %. Therefore, 0.066 ML (= 7.1% of 0.93 ML) of Dy atoms is expected to have moved from the 2D reconstructed surface to the NWs by annealing. The second part is the calculation of the metal content on 2D surface. If we assume that there are five (three) Dy atoms per 2x7 (2x4) unit cell [72], the difference in metal content on the 2D surface before and after annealing is: A(2D) = 0.357 ML x 85 % (before annealing: 2x7) — 0.375 ML x 66 i 8 % (after annealing: 2x4) = 0.055 i 0.030 ML. Therefore, there are less Dy atoms on the 2D surface after annealing. This coverage range of 0.055 i 0.030 ML for 2D surface also agrees with the change in coverage of 0.066 ML from the calculation for NW 5. After the 2“d annealing, 3D compact silicide islands start appearing, although NWs are still dominant on the surface [Figure 5.4 (c)]. This phenomenon supports the results reported by 32. Liu et al. [12] The longer side of small 3D islands tends to be oriented along the direction of NWs. More 2x4 superstructure can be seen around NWs than 3D islands [Figure 5.3 (c)]. The percentage of the 2D surface covered with 2x4 superstructure is 36 i 13 % (21 i 4 %) around NWs (3D islands). This result implies that 3D islands are a more favorable state than NW 5 since there are less Dy atoms around 3D islands than NWs. Additional annealings increase the number of 3D islands and the island size. At the same time, the number of NWs decreases and the surface reconstruction is gradually swept away. After the 7th annealing at 720°C, various sizes 76 of 3D islands are visible and very few NWs are observed on the surface [Figure 5.4 (c)]. There is no obvious metal induced reconstruction left on the surface, but surface defects are present on approximately 10 percent of the 2x1 bare Si(001) substrate. It is difficult to confirm the mass conservation of Dy in this limit since the 3D islands are very sparse and the size range of the 3D islands is very wide. Also, Transmission Electron Microscopy (TEM) study reveals that large 3D silicide islands extend underneath the plate of the surface [154], making it difficult to accurately calculate their volume from what is seen in the STM data. Figure 5.4 Transition from NWs to large 3D islands of DySiz by annealing. (a) 0.58 ML of Dy was deposited at 600°C, (b) 1St annealing at 600°C for 8 min, (c) 2“Cl annealing at 600°C for 20 min, (d) 6‘11 annealing at 700°C for 10 min, and (e) 7Lh annealing at 720°C for 10 min. All image sizes are 400x400 nmz. 77 5.1.2 LEEM experiments Figure 5.5 shows (a) the LEED pattern, (b) a dark-field LEEM image, and (c) a bright-field LEEM image of 0.78 ML of Dy grown on Si(001) at ~600°C. Under these conditions, the surface should be covered in the 2X7 reconstruction plus NWs as in Figures 5.1 (a) and 5.4 (a). All three images were taken at RT after deposition. The Figure 5.5 (a) The LEED pattern of Dy grown on Si(001) at ~600°C (0.78 ML). 1x7 phase and 1/2 order streaks are shown. The electron beam voltage is 60 V. (b) . Dark-field LEEM image (2 eV) taken from 2/7 spot shown with the white arrow in (a). The field of view is 4 pm. (c) Bright-field LEEM image (9 eV) in the same area of (b) showing the stripe structure. Contours of terraces are drawn with white lines. The stripes are oriented at an angle of 90 degrees on each terrace, which correspond to the NWs. 78 electron beam voltage is 60 V in Figure 5.5 (a). The diffraction pattern with 7x periodicity and 1/2-order streaks are similar to typical LEED pattern Observed from 2x7 reconstructed surface of Gd, Dy, or Ho grown on Si(001) at 600°C [72]. Figure 5.5 (b) shows the dark-field LEEM image (2 eV) taken from the 2/7-order diffraction spot denoted with the white arrow in Figure 5.5 (a). The field of view is 4 um. A clear bright-dark contrast is observed on alternate terraces. Each terrace shows a cross-hatched texture, as well as a few small isolated islands. The spacing of the cross-hatching is about 50 nm, and it is oriented along <1I0> directions but it is not a direct image of the NW structures. Figure 5.5 (c) is a bright-field image (9 eV) in the same area of Figure 5.5 (b) showing the striped structure. The contours of terraces are drawn with white lines. There is a rough alteration in stripe direction across the terraces, which is consistent with NW related features, but and again, the apparent size and spacing are too large to be the NWs. Figure 5.6 shows the sequence of LEEM images presenting the growth of Dy disilicide islands (0.78 NH.) during annealing. Annealing duration time and the temperature are indicated on the lower left side, and imaging mode (BF: bright field, TBF: tilted bright field) and the incident electron energy are indicated on the lower right side of each image. As already mentioned in Section 5.1.1, the 2x7 superstructure is replaced by a 2x4 superstructure after 1 min annealing. Also, for LEED pattern in Figure 5.5 (a), the position of 2/7-order diffraction spot is very close to that of 2x4-diffraction spot. Since the aperture used is not small enough to distinguish those two spots, 2/7-order and/or the 2x4-diffraction spot are used to form the dark-field image during annealing. The shape of the terraces in the dark-field image remains the same up to ~620°C. This clarifies 79 51 '30" ' 56'31" 742°C 745°C Figure 5.6 Sequence of LEEM images showing the growth of DySiz islands (0.78 ML). The field of view is 4 pm. The white arrows show the same island. Small grain structures start appearing at ~700°C and grow as islands. BF: Bright field, TBF: Tilted bright field. 80 that silicon atoms does not move on a macroscopic scale when the 2x7 to 2x4 phase transition occurs. The terrace contrast becomes weaker above ~700°C, although the terraces still have the original profile. The annealing temperature was gradually increased from RT to ~750°C for 60 minutes. The relationship between annealing duration and temperature change above ~600°C is shown in Figure 5.7. In Figure 5.6, the white arrow points out the same island and its surrounding area in each image. Different islands labeled with numbers in Figure 5 .6 (f) are classified as follows: (#1-5) large islands, (#6-10) medium islands, and (#11-13) small islands. The incident electron energies of 2~3 eV, 4.5 eV, and 11~12 eV give the best contrast for surface terrace in the dark-field image [Figure 5.5 (b)], step phase in the bright-field image [Figure 5.6 (f)], and stripe structures in the bright-field image [Figures 5.5 (c) and 5.6 (b)], respectively. Tilted bright field (TBF) image provides the information of both island growth and surface terrace morphology. 750 9 fl 9. o 0 9 :3 700 ,9. ...................................... 92 99 3 o *‘ o e 650 .............. 3 ............................................................. 3 . .O M 600 5’ . . - - - - . 30'00" 40'00" 50'00" 60'00" Annealing duration Figure 5.7 Annealing duration vs. temperature. Annealing temperature was gradually increased from RT (00’00”) to 750°C. The black arrow shows the time range during which the islands are observed as shown in Figures 5.8 and 5.9. 81 The nucleation of small structures that eventually grow into islands starts at ~700°C [Figure 5.6 (a)]. However, the initial stage of the nucleation process cannot be seen because of the limited resolution, and small islands are hardly distinguishable from the darker contrast in the stripe structure. Islands become clearly visible 15 seconds after nucleation. In most cases, the shape of the islands is a rounded rectangle, but a few square (#4) and elongated (#5) shapes can be also found [Figure 5.6 (1)]. During annealing process, three types of island growth are observed as shown in Figure 5.8. Time evolution of the length and the width of three different islands [#4, 5, and 6 in Figure 5.6 (f)] are plotted. The length and the width of islands are marked with black squares and white triangles, respectively. Note that the vertical axes in Figure 5.8 (and Figure 5.9) show the length and the width of 3D islands including the area of surrounding trenches which are clearly visible in STM [Figure 5.4 (c)] but are not resolvable from the actual islands in the LEEM. Therefore, the actual island dimensions are slightly smaller than those shown in the graphs. However, it is clear enough to understand the growth behavior of islands since the surrounding Si depletion areas approximate the shape of islands. The first type of island growth is square growth [Figure 5.8 (a)] in which the width and the length of the island increase at the same rate. The square island (#4) grows in this manner. The second type of growth is elongated growth [Figure 5.8 (b)] in which the length increases while the width remains constant. The elongated island (#5) is this case. The third type of growth is the most common growth, rectangular growth [Figure 5.8 (c)]. Both width and length increase uniformly but the growth rate of length is faster than that of width. Here, island #6 was illustrated by an example in Figure 5.8 (c) although the 82 other islands (#1-3 and #7-13) grow in the similar way. Except for the elongated island (#5), the island width and length grow uniformly in all the directions while maintaining the same aspect ratio. The aspect ratios of square (#4) and elongated (#5) islands are 1.2:1 and 7.321, respectively, while those of rectangular islands are in the range of 1.8:1 ~ 4.6:1. ’é‘ (a) Square growth: #4 5 200 I I % 150~-----------°-'---2.r-2"Z'é ---------- 'fi """""" E 100» ------ fig.“- -------------------------------- g I 5 sou-13L ------------------------------------------ g) 0 . . . 3 45 50 55 60 Annealing time (min) A (b) Elongated growth: #5 g 00 ll""I I I i 1:: 400 -------- 'J' ------------------------------------ E 300+ ----------------------------------------------- I ‘2 200 ------------------------------------------------ m 100 ------------------------------------------------ g» 0 Mat: A A - A A - 3 45 50 55 50 Annealing time (min) A c Rectan ular rowth:#6 E 50( ) 9 9 V ..................................... 1.1 ..... g 200 . I I I '5 150» ------- #9- --------------------------------- g 100 ----- .‘i'x """"" a “8A """""" A‘A """ g 50052214:1M9 -------------------------------------- C a: 0 ‘ ‘ ‘ -‘ 45 50 55 60 Annealing time (min) Figure 5.8 Time evolution of the length and the width of 3D islands and the surrounding Si depletion areas in different growth types during annealing. 83 Figure 5.9 shows the change in the size of the islands by annealing. Black, gray, and white marks indicate large, medium, and small islands, respectively. Numbers from 1 to 13 correspond to the islands labeled with the same numbers in Figure 5.6 (f). The labels (a)~(f) indicate times at which the LEEM images in Figure 5.6 were captured. The island growth starts and ceases in the same time regardless of the island size. This implies that the growth rate of larger islands is faster than that of smaller islands. After 5 minutes of island growth, the island size remains constant up to ~750°C. In general, the island sizes do not change up to ~760°C and Dy metals start leaving the surface by annealing at above ~780°C according to the experiments under similar conditions. Large islands 3 __ 10 2A 39 g 4. 5— E E Medium Islands E C 532;. =;:;. a :32 9 .312: 10- a) V 2 Small islands 5 t0 11012A130 "o 9 c to to c a) o ‘U '4: 1 c a) 1.“ a .‘2 g 8 '0') 0 40 45 Annealing time (min) Figure 5.9 Change in the island size by annealing (0.78 ML of Dy). Black, gray, and white marks indicated large, medium, and small islands, respectively. The islands numbered from 1 to 13 correspond to the islands with the same numbers in Figure 5.6 (f). The labels (a)~(f) are marked for the time at which the LEEM images in Figure 5.6 were taken. 84 Silicon terraces start breaking up during the island growth [Figure 5.6 (c)]. After the island growth ceases, terraces start rearranging themselves to form regular Si(001) surface [Figure 5.6 (e)] with 2x1 periodicity of the LEED pattern. TBF images of Figures 5.6 (c) and 5.6 (e) have an opposite contrast due to the different incident electron energy and the different beam tilt in the azimuthal angle. As the traces of streaked structure shown in Figure 5.6 ((1), small dots are left on the terraces with weak contrast. This might be a reflection of surface defect areas that are generally observed on STM images as explained in Section 5.1.1. 5.2 Discussion A DySiz NW has a hexagonal Ale structure. The formation of DySiz NWs arises from the anisotropic lattice mismatch between DySiz and the Si(001) substrate. The orientation relationships of hexagonal DySiz on Si(001) are determined to be [5, 9, 12]: DySi2(lTOO)// Si(001), DySi2[OOOl] //Si[1'1‘0], and DySi2[11§O] //Si[110]. Comparing the lattice mismatch in a and c directions of the hexagonal structure with Si(001) 1x1 unit cell, the lattice mismatch in the a direction is very small (-0.23 %). On the other hand, the lattice mismatch in the c direction is large (+7.32 %). Because of those anisotropic lattice-mismatch strains, hexagonal DySiz is favored to grow in the a direction than the c direction. Therefore, a and c directions of hexagonal structure correspond to a length and width of NW, respectively. When NWs are grown under conditions where the substrate is at least partially covered in the 2x4 phase, then the widths of NW5 are relatively broadly distributed between about 2as, and 16a3i. The limit that one would expect for coherent growth is about 14am, 85 given the lattice mismatch in the direction of the NW width. However, as shown in this work, if the substrate is dominated by the 2x7 phase, most of the NW widths are 7as, which suggests that the presence of the 2x7 surface reconstruction can affect how the NWs nucleate and grown. The stronger influence of the 2x7 phase may be due to two factors. Firstly, even in the case of the 2x4 surface, most NWs are wider than 3 or 4a3i which suggests that NWs prefer to grow wider than the width of the 2x4 unit cell. Secondly, it is clear in images of surfaces with a mixture of 2x4 and 2x7 that the 2x7 phase is much more highly ordered. As can be seen in Figure 5.1 (a), the bright rows of the 2x7 phase are ordered over a similar length scale as adjacent NWs. Annealing at 600°C [Figure 5.1 (b)] converts some of the 2x7 into 2x4, and at the same time, NWs appear to have moved and reconfigured since they are much more likely to be seen in bundles, and they are much more likely to show second layer growth. Figure 5.1 illustrates that the NWs are a metastable phase, and that the structural order and uniformity of the NWs are a very sensitive function of growth conditions. Islands grow in the temperature regime where conversion from NWs to 3D islands is seen in the STM images shown in Figure 5.4. Specifically, significant conversion of NW to 3D islands is not seen even after extended annealing at 6005C [Figure 5.4 (c)], whereas the annealing above 700°C causes the significant conversion of surface morphology [Figure 5 .4 (d)]: transition from NWs to 3D islands. When the NW width reaches a certain critical width during annealing, where the maximum width mismatch between DySiz and Si(001) substrate reaches lag, a dislocation is generated in order to release the misfit strain. Cross-sectional TEM observation shows two types of misfit dislocations at the interface between 3D islands 86 and Si substrate [154]. In a Type I dislocation, there are several tilted misfit dislocations along the interface between 3D islands and Si substrate. In a Type H dislocation, the whole island is tilted at a small angle with respect to the Si substrate, but there is no partial misfit dislocation. When the Type I dislocation occurs, the DySiz can readily grow in both directions resulting in the formation of rectangular islands. For the elongated island [island #5 in Figure 5.6 (0], there is a larger misfit strain along the width direction, since the width of an elongated island is narrower than that of rectangular islands. If the misfit dislocation along the width direction is suppressed during annealing for some reason, the misfit strain relaxation occurs in the length direction and the island length increases. The formation of the elongated island might be related to a Type H dislocation which has a tilted crystal structure. After all, it is possible that 3D islands can be formed in different growth modes. DySiz NW 8 grow in the Stranski-Krastanov (SK) mode. As 3D islands start growing, the growth mode changes from SK mode to Volmer-Weber (VW) mode [Figure 5.3]. When the island nucleation occurs, Dy atoms are desorbed from the 2D surface and attach along the island edges resulting the linear increase of the island area. The island-size distribution is broadened. Once Dy atoms attach to the 3D islands, they preferentially remain on the island regardless of the island size. It seems that the attachment probability of Dy atoms to 3D islands is equal, so that larger islands can attract more atoms because of longer circumferences. Therefore, the island growth rate is size-dependent: Larger and smaller islands grow with faster and slower growth rates, respectively [Figure 5.9]. The relative size differences of large and small islands do not change during the island growth. 87 Even though the general behavior of islands in all size ranges is similar, Figure 5.8 clearly shows that islands can exhibit both a variety of morphologies and associated growth behaviors. The most common island is rectangular, and although there is a distribution of aspect ratios seen, each individual island is seen to maintain the same aspect ratio during growth. Square growth is a variation of this behavior. Finally, a few islands show elongated growth, where the width of the island is fixed as the island lengthens. This type of 1D growth has been seen in the case of other silicides, including ErSiz, and it has been related to various theories. However, the diversity in growth behaviors and island morphologies seen in the LEEM data in Figure 5.6 make it clear that at least in this system, one cannot make any generalizations about growth in this system. Furthermore, it is dangerous to make a case for growth behavior base on just one type of island. For the RE silicides, it is possible that some of these different behaviors could be related to the presence of different silicide structures. DySiz can crystallize into hexagonal, orthorhombic and tetragonal phases, all of which have different lattice constants. It is tempting on geometrical grounds to associate square, rectangular, and elongated growths with the tetragonal, orthorhombic, and hexagonal phases, respectively. In fact, TEM studies of 3D islands grown in this system have shown that islands can be either hexagonal or orthorhombic / tetragonal, at least at a growth temperature of 600°C. Correlation of island morphology and silicide crystal structure requires further plan view TEM work. The LEEM data relates island morphology to growth behavior. All NWs are gone at 720°C, which is roughly the temperature at which the 3D islands stop growing in the LEEM measurement. Essentially, growth stops at the point where 88 there is no longer a source of Dy available, from either the substrate reconstruction, or the NWs. After island growth stops, there is no further evolution of the islands, even after extended annealing at 750°C. No ripening or coarsening is seen. This is consistent with the behavior reported for Er silicide islands on Si(001) annealed at 800°C [76]. One other interesting point is that the Si steps are observed to wander after island growth stops, whereas they are largely static during the conversion of NW to islands. Given the strong interaction between NW and single height step configuration seen in the STM images, it is possible that the substrate steps are pinned in place by a few remaining NW, until the conversion to 3D silicide islands is complete. 5.3 Conclusions After rapid Dy deposition, almost all 2D surface is covered with 2x7 superstructure, and NW 5 have uniform width with 7a3i which is 7 times as wide as the lattice constant of silicon. Surface reconstruction phase immediately changes from 2x7 to 2x4 superstructure by annealing, while the percentage of NWs remains almost the same. Therefore, this fact insists that 2x7 superstructure is the metastable state. The metal content on the surface also conserves before and after annealing at 600°C. Annealing at 700°C increases the number of 3D islands and the average island size. At the same time, the number of NWs decreases and surface reconstruction is swept away. This fact insists that 3D islands are the stable state whereas NWs are a metastable state at 700°C. When the sample is heated, the silicon terraces have the same contours up to ~620°C. Silicon terraces start breaking up and the terrace contrast weakens at ~700°C as the 89 nucleation of islands starts. The island growth starts and ceases in the same time regardless of the island size. Larger 3D islands grow at a faster growth rate than smaller islands. 90 Chapter 6 Transport Measurements Deposition of a RE metal on the Si(001) surface at an elevated temperature results in the formation of silicide islands and W3 coexisting with a reconstructed substrate surface. Although it is important to understand the electrical properties of RE silicide NWs and islands in air for practical applications, no one so far has done the transport measurements of ultra high vacuum (UHV)—prepared samples in air. We describe our approach for making transport measurements on nanometer scale thickness films grown on atomically clean silicon in UHV. A combination of ex-situ Ti silicide and in-situ Au deposition is used for contacts. Samples are passivated with amorphous germanium (Ge) before being removed from UHV for transport measurements at 4.2 K. The transport properties are correlated with film morphology as observed by STM. In this chapter, we will present the results of ongoing measurements on macrosc0pic networks of the Dy silicide islands. 6.1 Sample preparation The procedure needs several steps for the sample preparation as shown in Figure 6.1. After a sample is chemically cleaned (Step 1), Si and Ta (or Ti) depositions are carried out in a dc-magnetron triode sputtering system with a base pressure of 3x10’8 Torr (Step 2). The sputtering Ar pressure is 2X10'3 Torr. Ar and Ta (or Ti) purities are 99.999 % and 99.99 %, respectively. The sample with pre-deposited Ta (or Ti) contacts is then transferred into the UHV chamber and flashed briefly at 1175°C to remove surface oxide (Step 4). At the same time, Ta (or Ti) forms silicides which serve as contacts for 91 transport measurements (Step 5). After a RE silicide thin film is made by the same experimental procedure as Section 2.1 (Step 6), gold (Au) is deposited through a shadow mask on top of TaSiz (or TiSiz) (Step 7). Since Au partially covers the RE silicide thin film, there is a direct connection between Au contacts and the RE silicide thin film. In the end, the sample is passivated with amorphous Ge at RT to avoid surface oxidation (Step 8), and then removed from UHV chamber for transport measurements. Ta or Ti Oxidizaiion - Silt. - s - — Si subsira-le Flashing! .Nm.m_fi-wa___..u l i 5 b '7 8 Passivafion i TaSiz or Tifiz Nanowires " l g :drfi .1‘1;.‘-:~’Ly,’:‘, 3:: , . .r, Figure 6.1 Sample preparation procedure for transport measurements. Si and Ta (or Ti) depositions are done in the high vacuum chamber (Step 1 and 2). Sample flashing, deposition of RE metals, Au contacts, and amorphous Ge are performed in the UHV chamber (Step 4~8). 6.1.1 Pre-deposited Ta (or Ti) silicide contacts Refractory metal disilicides have been of special interest due to their temperature stability and relatively low resistivity [155-165]. TaSiz and TiSiz are the candidates for pre-growth deposited contacts for transport measurements, since these silicides can 92 survive the high temperature flash necessary to clean the Si(001) surface [166]. The contact resistances are in the range of 6.8~52 S2 for TaSiz and 5~15 $2 for TiSiz at 4.2 K [155]. TaSiz is a hexagonal structure (C40) with a c/a ratio of about 1.38 [155]. The sheet resistance of the TaSiz thin film decreases with an increase in film thickness and annealing temperature [156, 157]. On the other hand, TiSiz exists in two phases: a metastable C49 base-centered orthorhombic phase that is formed at lower temperature and the stable face-centered orthorhombic C54 phase formed at higher temperature [158]. The C54 phase is more desirable for device applications because the C54 phase has lower resistivity than the C49 phase. But the resistivity of TiSiz increases due to the failure of the C49 to C54 phase conversion as the lateral dimension of TiSi2 shrinks to less than 1 urn [159]. However, the type of TiSiz phases will not affect our macroscopic measurements since our contact dimension and thickness are not in the submicron range. Si diffuses into Ti layers from 500°C and then TiSi2 is formed at 600°C. TaSiz is also formed from 600°C [160], but the crystallization of TaSiz occurs mainly between 800°C and 900°C [157]. The initial attempts at making Ta silicide contacts involved depositing just Ta onto the substrate, and then annealing in UHV. As shown in Figure 6.2, when the sample with a pre-deposited Ta is annealed, a deep trench is created on the boundary between TaSiz and Si surface because Si around Ta is consumed to form TaSi2_ The same phenomenon can be seen for TiSiz. Since these trenches might interfere with connections to a RE thin film to TaSiz contacts, the deposition of Si with Ta is necessary beforehand as shown in Figure 6.1. The ratio of metal to deposited Si thickness is 122.23 (1:227) for Ta (Ti) in order for the stoichiometry of the deposited material to be 93 MSiz. When this is done, the trenches around the contacts are largely gone, and also the rough boundary region around the contacts is narrower. Figures 6.3 (a) and 6.3 (b) show the surface morphology of TaSiz and TiSiz films, respectively. Although the films do not appear smooth, they are continuous. Figure 6.2 (a) Atomic Force Microscopy (AFM) image of the boundary between TaSiz and Si substrate after annealing 100 nm thick Ta on silicon. (b) Cross-sectional profile traced along a white line in (a) showing a deep trench. (Width = 23 um, depth = 800 nm) Figure 6.3 Scanning Electron Microscopy (SEM) images of (a) TaSiz and (b) TiSiz films. For (a), 200 nm of Si + 100 nm of Ta films are annealed after deposition. For (b), 100 nm of Si + 50 nm of Ti films are annealed after deposition. 94 For the deposition of Si and Ta (or Ti) in Step 2 in Figure 6.1, sputtering masks are used as shown in Figure 6.4 (a). Three samples are fitted on one mask. A rectangle in Figure 6.4 (a) indicates the position of the sample sitting in the middle of the sputtering mask. Other two samples can be put on the top and bottom of the middle sample. After Si and Ta (or Ti) deposition, four contacts are made on each sample as shown in Figure 6.5 (a). Since the maximum of 24 samples can be loaded on the sputtering system, the possibility of making many pre-deposited contacts at once offers a great advantage. ll sample size = 1.23m Figure 6.4 (a) Sputtering mask for Ta (or Ti) deposition. (b) Shadow mask for Au deposition. 6.1.2 In-situ Au contacts After the RE silicide growth, Au contacts are deposited at RT by covering a sample with the middle of a shadow mask as shown in Figure 6.4 (b) (Step 7 in Figure 6.1). About 50 ML (~6 nm) of Au is deposited on top of TaSiz (or TiSiz) contacts [Figure 6.5 (a)]. The Au thin-film contacts are composed of many small islands which are connected to each other [Figure 6.5 (c)]. Figure 6.5 (b) shows a magnified view of the edge of the Ti silicide contact overlaid with Au. As mentioned previously, the codeposition of Si and Ti before annealing eliminates any trenching around the silicide. At the same time, it is clear that the edges of the Au contact are not sharp. This is 95 because of a combination of the distance between the shadow mask and the sample (which is about 0.5 mm), the distance between the shadow mask and Au evaporation source (about 5 mm), the dimension of the shadow mask (2.5 mm), and the finite size of the Au evaporation source (about 2 mm diameter). Given the geometry of the source relative to the mask and the sample, one expects that the boundary of the Au deposition would be spread over about 0.1 mm as shown in Figure 6.6. Figures 6.7 (a)~(e) show the STM images of the boundaries on the Au contact. The distance between the areas shown in these images is about 1 um. The density of Au increases from image (a) to (0), meaning that the edge of the Au boundary is not microscopically sharp. DySiz NWs can be also seen underneath the Au. 1‘ ”f“ i ‘ )1 __ \u - lilrrm 'l‘iSi: Figure 6.5 (a), (b) Optical microscope images of Au contacts covering beneath TiSiz contacts. (b) Close-up image of the area marked with a white rectangle in (a). (c) 3D STM image of the Au contact. The formation of many small Au islands is observed. 96 sample 0.1 mm 25 um Figure 6.6 Schematic diagram of Au evaporation with a shadow mask. Figure 6.7 STM images of the Au boundaries (200x200 nmz). The density of Au increases from (a) to (c). DySiz NWs can be seen underneath of Au. 6.1.3 Amorphous Ge passivation Since RE elements and RE silicides are very reactive with oxygen, it is important to protect the sample surface from oxidation in air. The materials used for surface passivation should meet the following two conditions. Firstly, it must not affect the conductance of the sample. Hence, it should be nonconductive at 4.2 K at which transport measurements are performed. Secondly, it should not be reactive with RE 97 silicide and Si at RT so that surface morphology underneath does not change. Although we first used hydrogen for the surface passivation for historical reasons (see Appendix A), Ge seems to work better in terms of more reliable transport measurements. Figures 6.8 (a) and 6.8 (b) show the STM images of ‘before’ and ‘after’ 5 ML of Ge deposition on DySiz W5 and interconnected 3D islands, respectively. NWs still survive underneath the amorphous Ge in Figure 6.8 (b). Figure 6.8 (a) Before Ge deposition: DySiz NWs with interconnected 3D islands. (180x180 nmz) Coverage is unknown. (b) After 5 ML of Ge deposition on top of NWs. (200x200 nm2) 6.2 'II-ansport‘ measurements: Results and discussion All transport measurements are done at liquid helium temperature (4.2 K) as shown in Figure 6.9 (b). Four-terminal surface resistances are measured by a Keithley digital multimeter. Note that four-terminal measurements can exclude the contact resistance and provide only sample resistance (see Appendix C). From the two-terminal measurements, contact resistances are a few 9 at RT and >1 M9 at 4.2 K. 98 (b) Liquid He Figure 6.9 (a) The STM image of 3 ML of interconnected Dy silicide islands. (b) Experimental setup for transport measurements. Surface resistances were measured at liquid helium temperature (4.2 K). (c) In order to measure the change in surface resistance of the sample (a), the sample was scratched in the horizontal and vertical directions. The numbers labeled on the bottom 0, l, and 2 indicate ‘before scratch’, ‘scratched in the horizontal direction’, and ‘scratched in the both directions’, respectively. In the first part of the experiments, changes in surface resistance due to sample scratching was investigated. The surface topography of the sample with 3 ML of Dy is shown in Figure 6.9 (a). The surface is covered with interconnected Dy silicide islands. The surface resistances were measured in three steps: without scratch (Step 0), with a scratch in the horizontal direction (Step 1), and both directions (Step 2). The numbers labeled in Figures 6.9 (c) and 6.10, indicate ‘before scratch (0)’, ‘scratched in the 99 mag , - horizontal direction (1)’, and ‘scratched in the both directions (2)’. Scratches were drawn with a diamond tipped pencil. AFM images show that the width and depth of the scratches are ~6i2 um and ~200i100 nm, respectively. Figure 6.10 (a) shows the change in four-terminal surface resistance of interconnected Dy silicide islands. The left (right) diagram represents the surface resistance in the vertical (horizontal) direction. The resistance values are averaged between left and right (top and bottom) sides of contacts for the vertical (horizontal) direction, and the resistances in three steps of sample scratching are marked with open (solid) circles. As a reference, the surface resistance of clean Si surface is indicated with a dashed line on top of each diagram. Corresponding current flows are schematically diagramed in Figure 6.10 (b). (a) 4-terminal Measurements (b) 3 8 Rvertical Rhorizontal 7 7 6 6 6.46 E a s 5 E 5 § 4 g 4 f u 3 3 E 3 m 2 m 2 1 1 1.2x1o-4 6.6x1o-4 o o e L 0 1 2 0 1 2 Figure 6.10 (a) Change in 4-terminal surface resistance of interconnected Dy silicide islands due to the sample scratching. A diagram shows the resistances in the vertical (open circles) and horizontal (solid circles) directions, respectively. Dashed lines on top are the surface resistance of clean Si surface. (b) Schematic diagrams of each step of sample scratching and corresponding current flow. 100 Before the sample scratching (Step 0), surface resistances in both vertical and horizontal directions are ~1.2x102 52. For this particular sample, the distance between two contacts in the vertical (horizontal) direction is 1.45 mm (1.75 mm). In case of the surface resistance in the vertical direction (open circles), the resistance increases dramatically to 7.2 MQ after scratching in the horizontal direction (Step 1), because there is a large barrier between contacts and the current cannot pass though easily [Rvertical ‘1’ in Figure 6.10 (b)]. After scratching in the vertical direction (Step 2), the resistance drops, because the vertical barrier tends to confine the flowing current to the left side [vamcal ‘2’ in Figure 6.10 (b)] thereby decreasing the voltage drop measured across the right hand side contacts. In the similar way, the change in the surface resistance in the horizontal direction can be explained. After scratching in the horizontal direction (Step 1), the resistance slightly increases, but the resistance value remains the same order of magnitude as that before scratching. After scratching in the vertical direction (Step 2), the resistance jumps up to 6.4 M52 due to the barrier between the contacts. It should be also mentioned that the surface resistance values of interconnected Dy silicide islands are smaller than that of clean Si surface which is drawn by dashed lines in Figure 6.10 even after scratching the sample in both directions. This is a reasonable result because there is small electrical contribution to the surface conductivity from Dy metals on Si substrate, although there are deep boundaries in the middle of the sample. In the conclusion for the first part of our experiments, the destruction of interconnected Dy silicide islands increases the surface resistance. 101 ,3.) 450x450 nm2 . ,5 180x180nm2 Figure 6.11 STM images of (a) DySiz NWs at 0.22 ML and (b) interconnected Dy silicide islands at 3 ML. The second part of our transport measurements is the comparison of surface conductivity at different metal coverages. Figure 6.11 shows the STM images of (a) DySiz NWs (450x450 nmz) and (b) interconnected Dy silicide islands (180x180 nmz). A few NWs are seen with 2x4 reconstructed surface at 0.22 ML [Figure 6.11 (a)], while the varying shape of 3D islands are connected and bare Si is not visible on the surface at 3 ML (=1 nm) [Figure 6.11 (b)]. Figure 6.12 shows Dy and Ho film conductivity at 4.2 K vs. metal coverages. Solid (open) circles show the surface conductivity of Dy (Ho) on Si(001). The conductivity of Dy thick film with ~180 ML (60 nm) is added as a reference [68]. The values of conductivity are indicated in Table 6.1. A 3 ML (zl nm) thick film of interconnected DySiz islands shows surface conductivity at 4.2 K similar to the reported bulk silicide value. A sample with sparse and disconnected NWs (0.16 ML) shows lower surface conductivity than the 3 ML film but higher than that of the clean Si surface, because the 102 2D surface is reconstructed with Dy metal. The conductivity of a sample with a DySiz NW network (0.87 ML) is lower than that of the sample at lower coverage. One possible explanation for lower surface conductivity on the high-coverage sample is the contact type: instead of pre-deposited Ti and in-situ Au contacts, ex-situ Al contacts were used (see sample #12 on Table 6.2). For surface conductivities of Ho on Si(001) with ex-situ A1 contacts, a sample with a HoSiz NW network shows higher conductivity than that with 2x4 reconstructed surface. Overall, the surface conductivity rises sharply somewhere in the coverage range of 1~3’ NIL. Many variations in sample preparation were made to improve the reproducibility: changing the choice of contact materials, the thickness of the contacts, and sample cleaning procedures. The methods of sample preparation and the results of transport measurements of each sample are summarized in Table 6.2. The samples from which at least some resistance values could be obtained at 4.2 K, regardless of the adequacy of the values, are gray shaded on the background of the leftmost column. There is no consistency among the methods of sample preparation and the results of measurements. Although both two- and four-terminal surface resistances are all measurable at RT, low-temperature measurements lose the stability due to the background noise. Unfortunately, the reliability of this experiments were not achieved from three-year efforts. Besides instability of low-temperature measurements, the limited amount of metal source and the multiple failures of the in-situ Au evaporator (Section 6.2.2) made the measurements very difficult. This work is supported by NSF ECS-0303801. 103 105 z’cL .... 104 O 3‘ .[68] E 3ML“ 60nm o 103 8%— C; E: 102 8% IE 101_¢0.16ML n 8 (C E 100 90.46 ML (,(L 5 10-100.18ML n 00.87 ML “ 10'2 1 r r ()(l L 0 l 2 3 180 Metal coverage (ML) Figure 6.12 Dy and Ho film conductivity at 4.2 K vs. metal coverages. Solid (open) circles show the surface conductivity of Dy (Ho) on Si(001). At 4.2 K Metal layer Surface topography Conductivity Sample # coverage (ML) (9 cm)’1 on Table 6.2 Dy 0.16 2x4 1.0x10l # 20 Dy 0.87 NW networks 3.9x10'2 # 12 Dy 3 Interconnected islands 1.9x104 # l6 Dy 180 (60 nm) Thin film 2.9x104 Ref [68] Ho 0.18 2x4 3.7x10" # 9 Ho 0.46 NW networks 1.3 # 6 Table 6.1 Dy and Ho film conductivity at 4.2 K vs. metal coverages. All the values correspond to those in Figure 6.12. There are no values for the conductivity of bulk Ho silicide in the literature, but the conductivity of the RE silicides lie in the range 4.5> o: > N 3 0m83> 8:30 5:5 53:28: 00850 ”90: >~Q 03805 20:22 Om 32 dev 2Q CR :0 0H E: om + mm E: 83 050:. 3:20:0805 a m _ 0380:: 032 32 meg >9 CM 80 0H E: mm + «m E: om: 050p. 082800-08 a E a so: u .2. 032 02858:: 32 one 5 ES 40 :o: 0002. a 028:5 Ea 02V 7.. m 2 E93 3 n >SQ E: mmd m 000563: >92 . Eod cm H >~Q VXN + 032 32 $9 29 :E On :8 000910 “020050 AE: 38 _< m .2 0380:: va 32 $9 on :E cm :8 Doomv :0 0200::0 AE: ©va 10. 52 -Rm a . 3 0302:: VXN + 032 38 0 32 mmdv 03 43200 to 0200: 2 w< 55 808:8 w< 5.2-5 A 9 E93 0.3 u >SQ E: 2.0 m va 3 00053038 Eve :0 u :5 4x0 32 2.8 on 5:: 00 a: 0002. a 028:5 :5: 02V 2 5:40 a 0 E93 m.SXm.m n A.8058; 0:: E08 00:50 35:03:09 90 0Q 5553: 32 5.8 2Q :E mm :8 Doomv :0 32055 AE: 83 _< 52.35 3:96 No.8 2 m 0380:: 32 02 Jim :2 8.8 oz EE mm :8 Uooflq 00 020050 AE: as :0. 5.2.06 m 0 Sea 8.: u >05 E: mmd m 30533: 32 80d 8.: u :5 8:950: 32 32 84.8 on :E mm :8 0°09. :0 00.0050 AE: SS 2 52-2% m o, I 032 “00:00:55: 32 3.8 cm E96 8.8 = m 8.53:0: 32 32 00.9 on Eye 58 a 0 l 20:82 0:0 032 00:09: 32 $9 on 3:93 8.8 = m 0380:: :03E030 >15 E 00:00: at .0303 E _< m N .— 2 S. a E: :0 8 0505050002 85589 00825 099 88:00 0&0 30:35 .33 030 H 6:0 2.0 05wE E 8:800:00 0:0 00350 8:05 05 :8 80D 50526:: 00352 8:80:08 .3 50855 N6 0308 105 E: 83 3. 0:: E: 0 :- 53: + 53:: 083:0: 32 :52 T: E 60:: a a. E: m: + 2 E: 0:: N22 80:20:00-2: E90 :20 E 00 E: 8023 :2 5.0.0... +32 ... : 0 =< 4.5-E + | SE 0.2 00 + 3:2 000 :0 609. a 2 E: m: + 2 E: :3 22 80808-2: : :0 £2 a 3. 5.0-0.. + bEcE 32 3.0 00 609. a F E: m: + 2 E: 0:: N22 0008:0002 0: mm E90: 2 u 0 32 21:0 3. 5.2-E + a: 8012 u ...: SE 0.40 00 609. :0 F E: m: + 2 E: 20 ~22 02:88-2: .: 00 SE 32 8on 3. 5.2.5 + 608:. 0300:: 3.0 00 + 0:850: 32 32 t :0 a 2 E: 0: + 2 E: 8: N22 00:82:00-2: : :0 50:26.: M Q E: :0 m 0x0 .6 20:02:: 32 we 3. 5.2.5 + 6000' 0.2 20.: u ...: SE 3.0 00 + vxm 32 2.8 :0 a F E: m: + 2 E: 2: 22 09088.8: a :0 32 :2va :< 5.2.3 + A008: 58:: 32 0.2 00 + 32 0.8 :0 :0 F E: on + 2 E: :0 N2:. 00:00:00-2: : 2 32 20$ 3. 5.0-0.: + 6000. 2:32: 32 0.8 00 a 2 E: m: + 2 E: 80 ~22 00:00:00-8: : 2 E90: 8.: u 0 3:2 :2 :< :02: + ,, 02 8.: n 5: :53 600' :0 a. E: 0: + 2 E: 0:: N2:. 02:88-2: : 2 E901 2 n 0 E: : m 20:00:: :23 32 0.0 00 32 oavv 3. 5.2.3 + . 0 0: n ...: + 5:02 028::85E a 02 o :0 6.01:0 F E: m: + 2 E: :5 ~22 02:88-2: : . 2 M N4: :0 099 _ 255050003 53050.80: 008-5m 0%: 88:00 80:23; 3.308 N6 0309 Chapter 7 Future work In this chapter, suggestions for future work as a continuation of this thesis will be described: growth kinetics of RE silicide NWs and 3D islands, reproducibility of electrical contacts for transport measurements, single NW measurements, and temperature dependence and magnetic-field dependence of magnetic, structural and transport properties. 7.1 Further work on growth kinetics In Chapter 5, we reported the LEEM observation of the evolution of Dy silicide islands during annealing. Our original purpose was to investigate the growth kinetics of RE silicide NWs such as NW nucleation, size and shape evolution, and Si step motion. Unfortunately, the width of RE silicide NWs (the width of single NW is typically < 5 nm) was below the lateral resolution limit of IBM Type I LEEM. The investigation of growth dynamics of RE silicide NWs, therefore, requires high-resolution LEEM/PEEM, or other alternative real-time surface analysis methods. Recently, a few studies of the growth evolution of elongated Er silicide islands and Ti silicide NWs have been done by photoemission electron microscope (PEEM) or LEEM [76, 167, 168]. Since their dimensions are about 50 nm wide and several um long, which is one order of magnitude larger than those of RE silicide NWs, they are resolvable in LEEM/PEEM. Moreover, it is important to interpret a correlation of the growth behavior of the width and length (lateral directions) and the height (vertical direction) of NWs and 3D islands so that the surface diffusion mechanism, surface stress and strain relaxation can be better 107 understood. Combining these results with TEM data would provide stronger support, since TEM data offers information about the crystal structure (hexagonal, tetragonal, or orthorhombic structure) of the 3D islands and about the interface defects between the 3D islands and the Si substrate [154]. 7.2 Reliability and reproducibility of electrical contacts for transport measurements Normally, Si substrates with a resistivity of p ~2.5 Q-cm at RT are used for our transport measurements so that the substrates do not contribute to the sample resistance in measurements at 4.2 K [169]. When the resistance of a highly-doped Si substrate which has low resistivity at 4.2 K (p ~10’2 Q-cm at RT) was measured, resistance values were very stable and did not fluctuate during the measurements. Therefore, one can assume that the Si substrate and contacts, and the contacts and leads are well contacted. However, the reproducibility of our measurements on RE silicides on Si substrates is rather poor as mentioned in Section 6.3. This implies that there is a possibility of an unstable connection between the RE silicide thin film (or NW network) and the contacts. To improve this situation, deposition of thicker Au layer (more than 10 nm) will be helpful in Step 7 in Figure 6.1. Additional Au deposition after Ge passivation (Step 8 in Figure 6.1) will be worth trying as well. Ex-situ deposited contacts (e.g. evaporation of Snm of Ti and 200 nm of Au in the clean room) will be an alternate method, although this has the disadvantage that samples are no longer totally UHV-prepared. Electron-beam (e-beam) lithography is another alternative method for contact deposition. In this case, very fine electrons can be defined so that measurements can be done on much smaller scales. 108 The use of intrinsic Si wafers allows transport measurements at RT, because the intrinsic resistivity of Si (05, ~2.4x105 [Q-cm] at RT) is nine orders of magnitude larger than that of RE silicide thin films (plasma-dc ~104 [Q-cm] at RT) [170]. However, STM on such high resistivity wafers is problematic. 7.3 Single NW measurements Single NW measurements are even more challenging, because the measurement dimension is at the nanoscale. Determination of the precise positioning of contacts on a NW is a key for this measurement. Figure 7.1 shows a schematic diagram of post growth deposited Au contacts for single NW measurements. In Step 1, a sparse network of NWs are grown in UHV chamber. Ex-situ alignment marks are fabricated by e-beam lithography (Step 2). After the registration of NWs and marks is determined by AFM, appropriately configured fine Au contacts (~50 nm) will be deposited by e-beam lithography. When sparse, parallel NWs are formed on a vicinal Si substrate [7], deposition of four contacts on a single NW would be very possible, because the length of these NWs can be microns in length. 9‘2: ~':.-‘. =“ "t i. -.. c K 1 ‘ _ ~ ~:j {w .. w res ...... Figure 7.1 Post growth deposited contacts Step I: Grow sparse network of NWs. Step 2: Deposit alignment marks: Use AFM to determine the NW/mark registration. Step 3: Deposit fine Au contacts. 109 7.4 Temperature dependence and magnetic-field dependence of magnetic, structural and transport properties of low-dimensional RE silicides The investigation of magnetic properties of RE disilicide NWs and 3D islands at low temperature or in high magnetic field would be interesting topics, since the magnetism of one-dimensional (1D) systems is different from that of 2D or 3D systems. In case of Co atomic chains on a Pt substrate, long-range ferromagnetic order is stabilized by large anisotropy energy barriers which arise from large localized orbital moments [171, 172]. RE elements and bulk RE silicides have a variety of unusual magnetic properties. For example, several magnetic structures can be observed in heavy RE metals below RT: sinusoidal structure [Figure 7.2 (b)], circular cone structure [Figure 7.2 (d)], spiral antiferromagnetic structure [Figure 7.2 (c)], and ferromagnetic phase in the basal plane [Figure 7.2 (f)] [42]. Intermediate mixed phases, where the c axis component is modulated along the c axis (CAM structure) are often seen [Figures 7.2 (a)~(c)]. The same metal can have several of these structures at different temperatures. Since the 4f shell, which is responsible for almost all of the magnetism of RE ions, is localized deep inside the atoms and shielded by the filled outer 5s and 5p shells, it is not perturbed by the crystal field. Therefore, spin-orbit coupling (LS-coupling) dominantly contributes to the magnetic moment. Experimental results are fairly consistent with theoretical calculations as shown in Table 7.1. Although Sc, Y, La, Yb, and Lu ions are diamagnetic because of the absence of a partially filled 4f shell (no unpaired electrons), they generally show paramagnetic behavior. 110 m» a a «9 @- 9 I, I ’l I —-‘~—-—-—‘-* a 6m \ .1 0&000 “a. a @— (9+9) 000000000 000000000 000000000 0 036-6} 6:?) Br, Tm Er Ho, Er Tb, Dy, Ho Gd, Tb, Dy (3) (b) (C) (d) (e) (0 Figure 7.2 Magnetic structures of heavy RE metals [after Koehler (1972)] [42]. The magnetic properties of RE disilicides are also summarized in Table 7.2. There is a large variety of ordering: no magnetic ordering in CeSiz, ferromagnetism in light RE silicides, PrSiz and NdSiz, antiferromagnetism in heavy RE silicides, GdSiz, TbSiz, DySiz, HoSiz, and ErSiz. Pierre et al. mentioned that these behaviors might be explained by the magnetocrystalline anisotropy, which is governed by the three different crystallographic structures (see Figure 1.1) in the series [173]. Before then, Sekizawa et al. described that the effective radial extent of 4f wavefunction of RE metal may influence the sign of the exchange interaction J (J>O for 111 ferromagnetic, and J1xl (LEED) Additional dose at 400 K [230] 3x1-92x1 (LEED) A mild annealing at 575 K [185] 2x1—91x1—)2xl (LEED) Additional dose at RT, [185] then annealing at 600 K Figure A2 (a) shows the schematic diagram of a Si dimer ‘before’ and ‘after’ H-passivation. For the 2x1 M structure, one H atom can attach to each Si dangling bond to fill the empty states. Therefore, in STM images, the brightness of the maxima ‘before’ and ‘after’ H-passivation appears different. The empty- and filled-state images of H-passivated Si(001) surface with lO-min H exposure are shown in Figures A2 (b) and A2 (c), respectively. Small white spots are contamination coming from the H source. In both images, the half left side is the depassivated area produced by STM-induced H desorption. H atoms are desorbed by applying high tip voltage. Here, a sample bias voltage of +8 V is used. In empty states [Figure A2 (b)], the depassivated surface appears brighter than the H-passivated surface because it corresponds to bare Si surface. In filled states [Figure A2 (c)], the H-passivated surface appears brighter than the depassivated surface because the electron density is higher when H atom is absorbed on the dangling bond. 119 Depassivated surface Hydrogen-passivated by high tip voltage Si(001) surface IIIIIrI 'IIIIIIIIJ' O 10 20 nm lllllllllll IIIIlIIllLJ O 10 20 nm Figure A2 (a) Schematic diagram of a Si dimer ‘before’ and ‘after’ H-passivation. A half-filled 1t. anti-bonding orbital is completely filled due to H-desorption. (b) Empty-state (1.74 V) and (c) filled-state (-l.82 V) images of H-passivated Si(001) surface. The left half of the images is the surface after it was passivated by applying high tip voltage (20x30 nmz). Cross-sectional profile in the horizontal direction is shown next to each STM image. This H desorption mechanism can be used for nanoscale patterning as shown in Figure A3. By manipulating the tip position and the voltage, the letters “MSU” have been 120 patterned with a bias voltage of +8 V. The average linewidth is ~10 nm. It is known that the desorption yield depends on the sample bias voltage [212]. A wider linewidth is obtained as the sample bias voltage increases. Figure A3 STM image (90x120 nmz) of a H-passivated Si(001) surface. Brighter areas correspond to an STM-induced H desorption. By operating the tip position and the voltage, we can write a “MSU” logo! So far no study of a H-saturated RE silicide film on Si(001) substrate has been done, although there are a few studies of H adsorption on Er silicide films on Si(lll) [231-237]. LEED studies show that at low H exposures on ErSi1_7(0001), H is absorbed in the same manner as on Si(lll), confirming that Er silicide is terminated by a buckled Si plane without Si vacancies [231, 234]. Moreover, H is not only adsorbed on but also absorbed in the layer. It is suggested that this form of chemisorbed H corresponds to a saturation of the dangling bonds related to the Si vacancies in the silicide bulk. H-passivation of both external and internal dangling bonds strongly stabilizes the films [233]. Here, our attempt of H passivation of DySiz NWs on Si(001) substrate is briefly reported. Figures A4 (a) and A4 (b) show the ‘before’ H-passivation of DySiz NWs on 121 Si(001) at 0.80 ML and ‘after’ H—passivation of the same surface, respectively. Although there are small gaps on the NWs shown by white arrows [Figure A4 (b)], the structures of NWs still survive after H-passivation. The change in the electrical properties of NWs due to the small gaps is unknown. There are no changes in the H-induced features on the NW s after an attempt to depassivate at +8V on the tip. This fact is very different from the depassivated Si(001) surface as shown in Figures A2 (b) and A2 (c). This implies that the bonding between the H atom and the RE silicide is stronger than that between H and Si. Figure A4 (a) Before H-passivation: DySiz NWs on Si(001) at 0.80 ML (180x180 nmz). (b) After H-passivation: NWs still survive, although there are small gaps on the NWs shown by white arrows (50x50 nmz). The change in the electrical properties of NWs due to the small gaps is unknown. 122 APPENDIX B Sheet Resistance [238, 239] The resistance R of a thin film with a length 1, width w, thickness t, and resistivity p can be written as Rzpist L [Q] R 5 wt w 5 Nlb [Q/square] where R3 is the sheet resistance of a layer. Here, we assume that the current density is uniform throughout the film. I Ni it 1‘ all? Figure B1 Schematic diagram of current passing through a thin film. t Sheet resistance is a resistance of a film, material, or layer measured by a four-point probe. The resistance represents the parallel resistance of an infinite number of infinitely thin parallel sheets. Strictly speaking, the unit for sheet resistance is the ohm (since W is dimensionless). However, to avoid confusion between R and Rs, sheet resistance is expressed as ohms per square area which is equal to the average resistivity of a layer divided by the thickness of the layer. FYI p : i : n; [9.cm] 0' net where m, n, e, and r are the electron mass, electron density, electron charge, and relaxation time, respectively. 123 f APPENDIX C Two- and four-terminal measurements [240] To measure a sample resistance R, the easiest way is to pass a constant current through the sample as shown in Figure C1. A known voltage V0 is connected to a series circuit consisting of a sample resistor R and a ballast resistor R3. R3 is chosen to be RB >> R so that the current is I = Vo/(R+RB) ~ VOIRB, which is independent of change in R. R3 Vo Q Figure Cl Constant-current circuit using a ballast resistor. (a) Rs (b) RB RC1 R RC2 0— V0 Figure C2 (a) Circuit showing a two-terminal resistance measurement. (b) Equivalent circuit for (a). Figures C2 (a) and C2 (b) show a circuit for a two—terminal resistance measurement and its equivalent circuit, respectively. When the sample is connected to the circuit as in Figure C2 (a), the measured voltage drop V is V: 1(RC1 +R + RC2) as in Figure C2 (b), 124 where RC] and RC2 are the contact resistances. Therefore, we are not measuring the sample resistance R itself but a sum of the sample resistance R and contact resistances RC] and R02. To eliminate the contact resistances from the total resistance, four-terminal connections are used as shown in Figure C3 (3). In four-terminal measurement, current leads are separated from voltage leads. Figure C3 (b) shows the equivalent circuit of Figure C3 (3). The current and voltage contacts are labeled as RC1 ~ RC4. Since the contact resistances RC; and RC; << R3, the current I is still I = Vo/(RB + RC1 + R + RC2) ~ Vo/RB. A sample voltage is measured between the other two contacts with a voltmeter which has very large input impedance so that it draws very little current. Since no current will flow through the contact resistances RC3 and RC4, the voltage drop across the sample can be measured as V = IR. Unless the sample resistance is comparable with the internal resistance of the voltmeter, contact resistances RC3 and RC4 have no effect on the measurements. (3) RB (b) RB RC1 RC3 ll <‘. 0 II 7:: V0 -o 1 RC2 RC4 Figure C3 (a) Circuit showing a four-terminal resistance measurement. (b) Equivalent circuit for (a). 125 APPENDIX D Van der Pauw’s method The surface resistivity of a sample ,0 can be determined from four-terminal resistances in the vertical and horizontal directions, Rvertical and Rhofimm. For a flat sample of anisotropic material of arbitrary shape with four points, van der Pauw equation [241, 242] is given by exp “-fld Rverticall/2 +6Xp "7&1 Rhorizonltzlrlz =1 (1) (pxpy) (pxpy) where pJr and py are principal values of the resistivity in the plane of the sample, and d is 1/2 the thickness of a sample (or thin film). From equation (1), (,0x p y) is obtained. If the sample is rectangular in shape and the four contacts are placed at the comers of the sample, the second equation is derived as (px /py)1/2 : _—?—ln tanh MRhOl’llonijlz (2) a” 16(pxpy) where a and b are the sides of the rectangle [243]. 1/2 1/2 ) ) , the individual values of the principal resistivity From (,0pr and (px lpy can be obtained: 1/2 1/2 ) ) p. = (10.x)y py =(10..py Mex/py )1/2 +(Px/Py 1/2 ) In our transport measurements (Section 6.3), we assumed px = py for simplicity. Therefore the surface resistivity of the sample p is given by equation (1), which is p E (pxpyf’z. 126 APPENDIX E List of publications Nogami, J., B.Z. Liu, M.V. Katkov, and C. Ohbuchi, Self-Assembled rare earth silicide nanowires on Si( 001 ). Phys. Rev. B, 2001. 63: p. 233305. Ohbuchi, C. and J. Nogami, Holmium growth on Si( 001 ): surface reconstructions and nanowire formation. Phys. Rev. B, 2002. 66: p. 165323. Ohbuchi, C. and J. Nogami, Samarium induced surface reconstructions of Si(001) (published) Ohbuchi, C., W. Swiech, and J. Nogami, Topographic evolution of Dy silicide on Si( 001 ).' Shape transitions from nanowires to 30 islands (in preparation) 127 [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] BIBLIOGRAPHY C. Preinesberger, S. Vandre, T. Kalka, and M. Dahne-Prietch, Formation of dysprosium silicide wires on Si(001), J. Phys. D: Appl. Phys. 31, p.L43 (1998). http://www. nytimes. com/libra ry/tech/99/II/biztech/articles/OI nano. html. Y. Chen, D.A.A. Ohlberg, G Medeiros-Ribeiro, Y.A. Chang, and RS. Williams, Self-assembled growth of epitaxial erbium disilicide nanowires, Appl. Phys. Lett. 76(26), 114004-4006 (2000). M. Katkov and J. Nogami. Growth of Yttrium on Si(001). in Bulletin of the American Physical Society, March meeting. 2002. Y. Chen, D.A.A. Ohlberg, and RS. Williams, Nanowires of four epitaxial hexagonal silicides grown on Si( 001 ), J Appl. Phys. 91(5), p.3213-3218 (2002). R. Ragan, Y. Chen, D.A.A. Ohlberg, G Medeiros-Ribeiro, and RS. Williams, Ordered arrays of rare-earth silicide nanowires on Si(001), J. Crystal Growth 251, p.657 (2003). B.Z. Liu and J. Nogami, Growth of parallel rare earth silicide nanowire arrays on vicinal Si(001), Nanotechnology 14(8), p.873-877 (2003). D. Lee and S. Kin, Formation of hexagonal Gd disilicide nanowires on Si(100), Appl. Phys. Lett. 82(16), p.2619-2621 (2003). J. Nogami, B.Z. Liu, M.V. Katkov, C. Ohbuchi, and N .0. Birge, Self-Assembled rare earth silicide nanowires on Si(001), Phys. Rev. B 63, p.233305 (2001). B.Z. Liu and J. Nogami, An STM Study of the Si(001) (2x4)-Dy Surface, Surf. Sci. 488, p.399-405 (2001). C. Preinesberger, S.K. Becker, S. Vandre, T. Kalka, and M. Dahne, Structure of DySi2 nanowires on SI(001), J Applied Physics 91(3), p.1695-1697 (2002). B.Z. Liu and J. Nogami, An STM study of DySi2 nanowires growth on Si(001), J Applied Physics 93(1), p.593-599 (2003). J. Nogami, Growth and characterization of atomic and nanometer scale wires on the silicon surface, Surface Review and Letters 7(5-6), p.555-560 (2000). C. Ohbuchi and J. Nogami, Holmium growth on Si(001): surface reconstructions and nanowire formation, Phys. Rev. B 66, p. 165323 (2002). Y. Chen, D.A.A. Ohlberg, G. Medeiros-Ribeiro, Y.A. Chang, and RS. Williams, Growth and evolution of epitaxial erbium disilicide nanowires on Si(001), Appl. Phys. A 75, p.353-361 (2002). 128 [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] J .S. Yang, Q. Cai, X.D. Wang, and R. Koch, Morphological evolution of erbium disilicide nanowires on Si(001), Surf. Interface Anal. 36, p.104-108 (2004). M. Kuzmin, P. Laukkanen, R.E. Perala, R.-L. Vaara, and 1.]. Vayrynen, Formation of ytterbium silicide nanowires on Si(001), Appl. Surf. Sci. 222, p.394 (2004). Z. He, D.J. Smith, and RA. Bennett, unpublished. J.L. McChesney, A. Kirakosian, R. Bennewitz, J.N. Crain, J.-L. Lin, and F.J. Himpsel, Gd disilicide nanowires attached to Si(III) steps, Nanotechnology 13, p.545-547 (2002). Z. He, M. Stevens, D]. Smith, and RA. Bennett, Dysprosium silicide nanowires on Si(IIO), Appl. Phys. Lett. 83(25), [15292-5294 (2003). K.A. Gschneidner and LR. Eying, Handbook on the Physics and Chemistry of Rare Earths. Electronic structure of rare earth metals, ed. S.H. Liu. Vol. 1. 1978: North-Holland Publishing Company. A. Chatterjee and AK. Singh, Pressure-Induced Electronic Collapse and Structural Changes in Rare-Earth Monochalcogenides, Phys. Rev. B 6(6), p.2285-2291 (1972). A. Stenborg, O. Bjorneholm, A. Nilsson, N. Martensson, J.N. Andersen, and C. Wigren, Valence changes and core-level shifts of Sm adsorbed on Mo(110), Phys. Rev. B 40(9). P.59l6-5923 (1989). H. Norde, J .d.S. Pires, F. d'Heurle, F. Pesavento, S. Petersson, and RA. Tove, The Schottky-barrier height of the contacts between some rare-earth metals (and silicides) and p-type silicon, Appl. Phys. Lett. 38(11), p.865 (1981). K.N. Tu, R.D. Thompson, and B.Y. Tsaur, Low Schottky barrier of rare-earth silicide on n-Si, Appl. Phys. Lett. 38(8), p.626-628 (1981). RD. Thompson, B.Y. Tsaur, and K.N. Tu, Contact reaction between Si and rare earth metals, Appl. Phys. Lett. 38(7), p.535-537 (1981). A. Franciosi, J.H. Weaver, P. Perfetti, A.D. Katnani, and G Margaritondo, Samarium valence changes and reactive interdifi‘usion at the Si(111)-Sm interface, Solid State Communications 47(6), p.427 (1983). G Rossi, J. Nogami, I. Lindau, L. Braicovich, I. Abbati, U.d. Pennino, and S. Nannarone, First Spectroscopic Investigation of the Yb / Si Interface at Room Temperature, Journal of Vacuum Science and Technology A 1(2), p.781 (1983). L.D. Woolf, Transport and magnetic properties of layered rare earth disilicides, Solid State Communications 47(7), p.519-523 (1983). 129 [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] E. Houssay, A. Rouault, 0. Thomas, R. Madar, and J .P. Senateur, Metallurgical Reinvestigation of Rare Earth Silicides, Appl. Surf. Sci. 38, p.156-161 (1989). A. Iandelli and A. Palenzona, in: Handbook on the Physics and Cherrristry of Rare Earths (North-Holland Publishing Company, 1978) vol.2 Alloys and intermetallics, ch.l3. G. Rossi, d and f metal interface formation on silicon, Surf. Sci. Rep. 7, p.1 (1987). FR Netzer, Rare earth overlayers on silicon, J. Phys. Condens. Matter 7, p.991-1022 (1995). S. Auffret, J. Pierre, B. Lambert, J.L. Soubeyroux, and J.A. Chromoczek, Crystallographic and magnetic-structures of Er3Si5, Physica B 163(3), p.271-280 (1990). V.M. Koleshko, V.F. Belitsky, and A.A. Khodin, Thin-films of rare-earth-metal silicides, Thin Solid Films 141(2), p.277—285 (1986). J .A. Knapp and ST. Picraux, Epitaxial growth of rare-earth silicides on (111) Si, Appl. Phys. Lett. 48(7). [1466-468 (1986). A. Iandelli, A. Palenzona, and GL. Olcese, Valence Fluctuations of Ytterbium in Silicon-rich compounds, Journal of Less Common Metals 64, p.213-220 (1979). 0. Thomas, C.S. Petersson, and EM. Dheurle, The reaction of scandium thin-films with silicon — diffusion, nucleation, resistivities, Appl. Surf. Sci. 53, p.138 (1991). R. Baptist, S. Ferrer, G Grenet, and RC. Poon, Surface Crystallography of YSi2-x Films Epitaxially Grown on Si(III): An X-Ray Photoelectron Drfiraction Study, Phys. Rev. Lett 64 p.311-314 (1990). M.-A.Nicolet and 8.8. Lau, Formation and characterization of transition-metal silicides, VLSI electronics -Mictostructure Science Vol.6 Eds N.G Einspruch, GB. Larrabee (1983), and references therein. C.C. Hsu, Y.X. Wang, J. Hu, J. Ho, and J .J . Qian, The efiect of thermal treatment on the thin-film reaction of La, Ce, and Nd on silicon surfaces, J. Vac. Sci. Technol. A 7(5). p.3016-3022 (1989). K.A. Gschneidner and L.-R. Eyn'ng, Handbook on the Physics and Chemistry of Rare Earths; and references therein. vol.1 Metals (1978) ; vol.2 Alloy and Intermetallics (1979); and references therein: North-Holland Publishing Company. 130 [43] [44] [45] [46] [47] [48] [49] [50] [51] [52] [53] [54] [55] A. Vantomme, M.F. Wu, S. Hogg, U. Wahl, W. Deweerd, H. Pattun, G Langouche, S. J in, and H. Bender, Stabilization and phase transformation of hexagonal rare-earth silicides on Si(III), Nuclear Instruments and Methods B 17, p.261-266 (1999). ML. Huang, J .H. Yang, Y.A. Chang, R. Ragan, Y. Chen, D.A.A. Ohlberg, and RS. Williams, Phase stabilities of ternary rare earth metal disilicides, Appl. Phys. A 78(1). [3.1 (2004). S. Auffret, J. Pierre, B. Lambert-Andron, R. Madar, E. Houssay, D. Schmitt, and E. Siaud, Magnetic properties versus crystal structure in heavy rare-earth silicides RSi2-x, Physica B 173, p.265-276 (1991). N. Frangis, GV. Tendeloo, J .V. Landuyt, G Kaltsas, A. Travlos, and AG Nassiopoulos, New Erbium Silicide Superstructures: A Study by High Resolution Electron Microscopy, Phys. Stat. Sol. 158, p.107-116 (1996). J .E. Baglin, FM. d'Heurle, and CS. Petersson, Appl. Phys. Lett. 36(7), p.594-596 (1980). I. Gerocs, G Molnar, E. Jaroli, E. Zsoldos, G Peto, J. Gyulai, and E. Bugiel, Epitaxy of orthorhombic gadolinium disilicide on <100> silicon, Appl. Phys. Lett. 52(25). p.2144-2145 (1987). J .Y. Duboz, P.A. Badoz, F.A. Davitaya, and J .A. Chroboczek, Electronic transport-properties of epitaxial erbium silicide silicon heterostructures, Appl. Phys. Lett. 55(1). P.84-86 (1989). L. Stauffer, A. Mharchi, C. Pirri, P. Wetzel, D. Bolmont, and G Gewinner, Electronic structure and interfacial geometry of epitaxial two-dimensional Er silicide on Si(III), Phys. Rev. B 47(16), P.10555-10562 (1993). M. Lohmeier, W.J. Huisman, Gt. Horst, P.M. Zagwijn, and E. Vlieg, Atomic structure and thermal stability of two-dimensional Er silicide on Si(lll), Phys. Rev. B 54(3), [32004-2009 (1996). C. Wigren, J.N. Andersen, and R. Nyholm, Sm- and Yb-induced reconstructions of the Si(III) surface, Phys. Rev. B 48, p.11014-11019 (1993). R. Hofmann, W.A. Henle, F.P. Netzer, and M. Neuber, Electronic-structure of epitaxial Yb silicide, Phys. Rev. B 46(7), p.3857-3864 (1992). C. Wigren, J.N. Andersen, R. N yholm, U.O. Karlsson, J. Nogami, A.A. Baski, and CF. Quate, Adsorption-site determination of ordered Yb on Si(III) Surfaces, Phys. Rev. B 47(15). [19663-9668 (1993). W.A. Henle, M.G Ramsey, F.P. Netzer, and L. Horn, Formation of divalent Eu silicides at the Eu-Si(111) interface, Surface Science 254(1-3), p.182-190 (1991). 131 r." .. - [56] [57] [58] [59] [60] [611 [62] [63] [64] [65] [66] [67] [68] S. Vandre, T. Kalka, C. Preinesberger, and M. Dahne-Prietch, Epitaxial growth and electronic structure of lanthanide silicides on n-type Si(III), J. Vac. Sci. Technol. B 17(4), p.1682-l690 (1999). T.V. Krachino, M.V. Kuz'min, M.V. Loginov, and M.A. Mittsev, Yb and Sm Adsorption and Silicide Formation on Si(III) Surface, Phys. Low-Dim. Struct. 9/10, p.95-106 (1999). M.V. Kuz'min, N .V. Mikhailov, and M.A. Mittsev, Formation and properties of a binary adsorbed layer in a two-component adsorption system (Sm+Yb)-Si(111), Phys. Sol. Sta. 45(3), P.579-585 (2003). T.V. Krachino, M.V. Kuz'min, M.V. Loginov, and M.A. Mittsev, Growth of an Eu-Si(111) thin film structure: The stage of silicide formation, Phys. Sol. Sta. 46(3). [1563-568 (2004). S. Vandre, T. Kalka, C. Preinesberger, I. Manke, H. Eisele, M. Dahne-Prietsch, R. Meier, E. Weschke, and G Kaindl, Growth and electronic structure of Dy silicide on Si(III), Appl. Surf. Sci. 123, p.100-103 (1998). CH. Luo, GH. Shen, and L]. Chen, Vacancy ordering structures in epitaxial RESig.Jr thin films on (III)Si and (001 )Si, Appl. Surf. Sci. 113/114, p.457 (1997). A. Travlos, N. Salamouras, and N. Boukos, Growth, structure and electrical properties of epitaxial thulium silicide thin films on silicon, J. Appl. Phys. 81(3), p.1217-1221 (1997). J.C. Chen, GH. Shen, and L]. Chen, Interfacial reactions of Gd thin films on ( III ) and (001 )Si, Appl. Surf. Sci. 142, p.291-294 (1999). Y.K. Lee, N. Fujimura, and T. Ito, Epitaxial-growth of Yttrium silicide YSi2-x on(100) Si, J. Alloys Comp. 193(1-2), [1289-291 (1993). S. Kinnou, J.Y. Veuillen, and T. Tan, Formation and electronic-properties of erbium silicide on Si(100), Surf. Sci. 307(Part A), p.258-263 (1994). G Kaltsas, A. Travlos, N. Salamouras, A.G Nassiopoulos, P. Revva, and A. Traserse, Erbium silicide films on (100) silicon, grown in high vacuum. Fabrication and properties, Thin Solid Films 275(1-2), p.87-90 (1996). N. Frangis, J .V. Landuyt, G Kaltsas, A. Travlos, and AG Nassiopoulos, Growth of erbium-silicide on (100) silicon as characterized by electron microscopy and diffraction, J. Crystal Growth 172, p.175-182 (1997). A. Travlos, N. Salamouras, and N. Boukos, Epitaxial dysprosium silicide films on silicon: growth, structure and electrical properties, Thin Solid Films 397, p.138-142 (2001). 132 EMEVJ ... ... l [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [31] [82] A. Travlos, P. Aloupogiannis, E. Rokofyllou, and C. Papastaikoudis, Epitaxial lutetium silicide: Growth, characterization and electrical properties, J Applied Physics 72(3). [1948-952 (1992). G Peto, GL. Molnar, Z.E. Horvath, E. Zsoldos, N.Q. Khanh, J. Gyulai, and J. Kanski, Formation of epitaxial HoSiz layer on Si( 100), Thin Solid Films 318, p.168-171 (1998). M.V. Katkov and J. Nogami, Yb and Nd growth on Si(001), Surf. Sci. 524(1-3), p.129-136 (2003). B.Z. Liu and J. Nogami, An STM study of the Si(001)(2x7)-Gd, Dy surface, Surf. Sci. 540(1), P.136-144 (2003). RJ. Godowski, J. Onsgaard, F. Orskov, and M. Christiansen, structural aspects of the Sm/Si(100) interface, J. Mater. Sci. Lett. 9, p.989 (1990). J.S. Yang, Q. Cai, X.D. Wang, and R. Koch, Initial stages of erbium disilicide formation on Si(001), Surf. Sci. 526, p.291-296 (2003). M. Kuzmin, R.E. Perala, P. Laukkanen, R.-L. Vaara, M.A. Mittsev, and 1.]. Vayrynen, Initial stages of Yb/Si( I 00) interface growth: 2x3 and 2x6 reconstructions, Appl. Surf. Sci. 214(1-4), p.196-207 (2003). L. Fitting, M.C. Zeman, W.C. Yang, and R.J. Nemanich, Influence of strain, surface diffusion and 0stwald ripening on the evolution of nanostructures for erbium on Si(001), J Appl. Phys. 93(7), 114180-4184 (2003). EM. Dheurle, M.O. Aboelfotoh, F. Pesavento, and CS. Petersson, Schottky barriers of scandium and scandium monosilicide on silicon, Appl. Surf. Sci. 53, p.237-239 (1991). EU. Hillebrecht, Interaction of Ce with Si(100), Appl. Phys. Lett. 55(3), p.277-279 (1989). M.H. Unewisse and J .W.V. Storey, Electrical and infrared investigation of erbium silicide, J Appl. Phys. 72(6), 132367-2371 (1992). A. T ravlos and N. Salamouras, Superconductivity of lanthanum silicide thin films, Vacuum 48(1), p.13-14 (1997). SM. Hogg and A. Vantomme, Growth and electrical characterization of GeSi1.7 epilayers formed by channeled ion beam systhesis, J Appl. Phys. 91(6), p.3664-3668 (2002). A. Travlos, N. Salamouras, and E. Flouda, Epitaxial erbium silicide films on (100) silicon: growth, structure and electrical properties, Appl. Surf. Sci. 120, p.355-364 (1997). 133 [83] [84] [85] [86] [87] [88] [89] [90] [91] [92] [93] [94] [95] [96] [97] J. Pierre, S. Auffret, J.A. Chroboczek, and T.T.A. Nguyen, Magnetotransport in the rare earth silicides RSiZ-x, J. Phys.: Condens. Matter 6, p.79-92 (1994). C. Boragno and MR. Bonansinga, Electrical and magnetic properties of ErSi2 and GdSiZ alloy thin films, Solid State Communications 92(6), p.515-518 (1994). A. Travlos, P. Aloupogiannis, E. Pokofyllou, and C. Papastaikoudis, Growth, characterization and electrical properties of gadolinium silicide thin layers, phil. Mag. B 67(4), p.485-495 (1993). G Guizzetti, E. Mazzega, M. Michelini, and F. Nava, Electrical and optical characterization of GdSiz and ErSiz alloy thin films, J Applied Physics 67(7), p.3393-3399 (1989). EH. Kaatz, J .V.d. Spiegal, and W.R. Graham, Anomalous magnetotransport in epitaxial TbSi2.x, J. Vac. Sci. Technol 9(3), p.426-429 (1991). SM. Hogg, A. Vantomme, M.F. Wu, and G Langouche, Electrical properties of rare earth silicides produced by channeled ion beam synthesis, Microelectron Eng. 50, p.211-215 (2000). J. Chroboczek, Transport and magnetic properties of rare earth compounds epitaxially grown on semiconductors, Acta Physica Polonica A 80(2), p.179-192 (1991). BR Wohlfarth, Ferromagnetic materials: a handbook on the properties of magnetically ordered substances ( North-Holland, Amsterdam, 1980). J .M. Ziman, Electrons and phonons: the theory of transport phenomena in solids, Oxford, 2001, chapter 9. A. Franciosi, P. Perfetti, A.D. Katnani, and J .H. Weaver, Samarium chemisorption on group-IV semiconductors, Phys. Rev. B 29(10), p.561] (1984). R.G Long, M.C. Bost, and J .E. Mahan, Metallic behavior of lanthanum disilicide, Appl. Phys. Lett. 53(14), p.1272-1273 (1988). F. Canepa, S. Cirafici, F. Merlo, and A. Palenzona, Electrical resistivity in the R5Si3 systems (R=La, Ce, Pr; Nd, Sm), J. Alloys Comp. 203, p.Lll-L13 (1994). WE. Henry, Bulletin of the American Physical Society 7, p.474 (1962). X. Zhenjia, Electrical transport in RE silicides, Properties of metal silicides: Eds K. Maex, M. Rossum (1995) IMEC. Leuven, Belgium and references therein, (1994). D. Tsamakis, M. Vlachos, A. Travlos, and N. Salamouras, Electrical properties of crystalline Er and Dy silicide layers, Thin Solid Films 418, p.211-214 (2002). 134 [98] [99] [100] [101] [102] [103] [104] [105] [106] [107] [108] [109] [110] F. Canepa, S. Cirafici, F. Merlo, and A. Palenzona, Electrical resistivity measurements on some R55i3 phases: R=Gd, Tb, Lu and Y, J. Magn. Magn. Mater. 118, p.182-186 (1993). V.N. Eremenko, V.E. Listovnichii, S.P. Luzan, Y.I. Buyanov, and RS. Martsenyuk, Phase diagram of the holmium-silicon binary system and physical properties of holmium silicides up to 1050C, J. Alloys Comp. 219, p.181-184 (1995). G Binnig, H. Rohrer, C. Gerber, and E. Weibel, Tunneling through a controllable vacuum gap, Appl. Phys. Lett. 40(2), p.178-180 (1982). G Binnig, H. Rohrer, C. Gerber, and E. Weibel, Surface Studies by Scanning Tunneling Microscopy, Phys. Rev. Lett. 49(1), p.57-60 (1982). G Binnig, H. Rohrer, C. Gerber, and E. Weibel, 7x7 reconstruction on Si( 111 ) resolved in real space, Phys. Rev. Lett. 50, p.120 (1983). D.M. Eigler and BK. Schweizer, Positioningsingle atoms with a scanning tunneling microscope, Nature 344, p.524-526 (1990). L.J. Whitman, J.A. Stroscio, R.A. Dragoset, and R.J. Celotta, Manipulation of adsorbed atoms and creation of new structures on room-temperature surfaces with a scanning tunneling microscope, Science 251, p.1206-1210 (1991). H. Uchida, D. Huang, P. Grey, and M. Aono, Site-Specific Measurement of Adatom Binding Energy Differences by Atom Extraction with the STM, Phys. Rev. Lett. 70(13), 112040-2043 (1993). TC. Shen, C. Wang, GC. Abeln, J.R. Tucker, J .W. Lyding, P. Avouris, and RE. Walkup, Atomic scale desorption through electronic and vibrational excitation mechanisms, Science 268(1590), ( 1995). P. Avouris, Manipulation of matter at the atomic and molecular levels, Acc. Chem. Res. 28, p.95-102 (1995). S. Hosaka, S. Hosoki, T. Hasegawa, H. Koyanagi, T. Shintani, and M. Miyamoto, Fabrication of nanostructures using scanning probe microscopes, J. Vac. Sci. Techno] Bl3, p.2813 (1995). TA. Jung, R.R. Schlittler, J.K. Gimzewski, H. Tang, and C. Joachim, Controlled Room-Temperature Positioning of Individual Molecules: Molecular Flexure and Motion, Science 271(5246), p.181-184 (1996). G Dujardin, A. Mayne, 0. Robert, F. Rose, C. Joachim, and H. Tang, Veneical Manipulation of Individual Atoms by a Direct STM Tip-Surface Contact on Ge(III), Phys. Rev. Lett 80(14), 113085-3088 (1998). 135 [111] [112] [113] [114] [115] [116] [117] [118] [119] [120] [121] [122] [123] [124] [125] [126] U.J. Quaade, K. Stokbro, R. Lin, and F. Grey, Single-atom reversible recording at room temperature, Nanotechnology 12, p.265-272 (2001). Y. Okawa and M. Aono, Nanoscale control of chain polymerization, Nature 409(683), (2001). R. Bennewitz, J.N. Crain, A. Kirakosian, J.-L. Lin, J.L. McChesney, D.Y. Petrovykh, and F.J. Himpsel, Atomic scale memory at a silicon surface, Nanotechnology 13(499-502), (2002). SW. Hla and K.H. Rieder, Engineering single molecules with a scanning tunneling microscope tip, Superlarrices and microstructures 31(1), p.63-72 (2002). K. Sattler, Nanolithography using the scanning tunneling microscope, J ap. J. Appl. Phys. Part] 42(7B), P.4825-4829 (2003). O. Dudko, A.E. Filippov, J. Klafter, and M. Urbakh, Manipulatioons of individual molecules by scanning probes, Nano Letters 3(6), p.795-798 (2003). SD. Barrett, http://www. imagesxrn. org. ukl. J.A. Kubby and J.J. Boland, Scanning tunneling microscopy of semiconductor surfaces, Surface Science Reports 26, p.61-204 (1996). J .A. Stroscio and W.J. Kaiser, Scanning Tunneling Microscopy, Academic Press Inc. (1993) ch.l and ch.4 and references therein. W. Telieps and E. Bauer, An analytical reflection and emission UHV surface electron microscope, Ultramicroscopy 17, p.57-66 (1985). E. Bauer, in: Proc. 5th Intern. Congr. on Electron Microscopy (Academic Press, New York, 1962), p.D-11. R.M. Tromp and MC. Reuter, Design of a new photo-emission/low-energy electron microscope for surface studies, Ultrarnicroscopy 36, p.99-106 (1991). R.M. Tromp, Low-energy electron microscopy, IBM J. Res. Develop. 44(4), p.503-516 (2000). 13, Matter E. Bauer, Photoelectron Microscopy, J. Phys.: Condens. p.11391-11404 (2001). E. Bauer, Methods of surface studies depending on inelastic-scattering of electrons, Vacuum 22(11), p.539-552 (1972). R.M. Tromp and MC. Reuter, Imaging with a low-energy electron microscope, Ultramicroscopy 50, p.171-178 ( 1993). 136 [127] [128] [129] [130] [131] [132] [133] [134] [135] [136] [137] [138] [139] [140] [141] http://www. ulvac-ph i. co. jp/PDF/taguch i0] .pdf. M. Aono, Y. Hou, C. Oshima, and Y. Ishizawa, Low-energy ion scattering from the Si( 001 ) surface, Phys. Rev. B 49(8), p.567 (1982). W.S. Yang and F. Jona, Atomic structure of Si(001) 2x1, Phys. Rev. B 28(4), p.2049 (1983). H. Over, J. Wasserfall, W. Ranke, C. Ambiatello, R. Sawitzki, D. Walf, and W. Moritz, Surface atomic geometry of Si(001)-(2x1): A low-energy electron-diffraction structure analysis, Phys. Rev. B 55(7), p.473] (1997). J .A. Appelbaum, GA. Baraff, and DR. Hamann, Si(001) surface reconstruction: spectroscopic selection of a structural model, Phys. Rev. Lett. 35, p.729 (1975). DJ. Chadi, Si(100) surfaces: Atomic and electronic structures, J. Vac. Sci. Technol 16, p.1290-1296 (1979). A. Redondo and W.A. Goddard, Electronic correlation and the Si(100) surface: Buckling versus nonbuckling, J. Vac. Sci. Technol. 21(2), p.344-350 (1982). Z. Jing and J .L. Whitten, Ab initio studies of H chemisorption on Si(100), Phys. Rev. B 46(15), [3.9544 (1992). Y. Jung, Y. Shao, M.S. Gordon, D]. Doren, and M. Head-Gordon, Are both symmetric and buckled dimers on Si( 100) minima? Density fimctional and multireference perturbation theory calculations, J. Chem. Phys. 119(20), p.10917 (2003). S. Gasiorowicz, Quantum Physics - second edition: Wiley, p.334 (1996). J.A. Appelbaum, GA. Baraff, and DR. Hamann, The Si(100) surface. 111. Surface reconstruction, Phys. Rev. B 14(2), p.588-601 (1976). CL. Alerhand, D. Vanderbilt, R.D. Meade, and J .D. Joannopoulos, Spontaneous formation of stress domains on crystal surfaces, Appl. Phys. Lett. 61(17), p.1973-1976 (1988). http://www. research. ibm. com/leem/. R.M. Tromp, R.J. Hamers, and J .E. Demuth, Si(001) dimer structure observed with scanning tunneling microscopy, Phys. Rev. Lett. 55, p.1303-1306 (1985). T. Tabata, T. Aruga, and Y. Murata, Order-disorder transition on Si( 001 ) -C(4X2)T0(2X1), Surface Science 179, p.L63 (1987). 137 [142] [143] [144] [145] [146] [147] [148] [149] [150] [151] [152] [153] [154] Y. Enta, S. Suzuki, and S. Kono, Angle-resolved-photoemission study of the electronic structure of the Si(001) c(4x2) surface, Phys. Rev. Lett 65, p.2704 (1990). RA. Wolkow, Direct observation of an increase in buckled dimer on Si(001) at low-temperature, Phys. Rev. Lett 68, p.2636 (1992). A. Ramstad, G Brooks, and R]. Kelly, Theoretical study of the Si(001) surface reconstruction, Phys. Rev. B 51(20), p.14504-14523 (1995). G LeLay, A. Criecenti, C. Ottaviani, P. Perfetti, T. Tanikawa, I. Matsuda, and S. Hasegawa, Evidence of asymmetric dimers down to 40 K at the clean Si(100) surface, Phys. Rev. B 66, p.153317 (2002). K. Hata, S. Yoshida, and H. Shigekawa, p(2x2) Phase of Buckled Dimers of Si(100) Observed on n-Type Substrates below 40 K by Scanning Tunneling Microscopy, Phys. Rev. Lett 89, p.286104 (2002). M. Matsumoto, K. Fukutani, and T. Okano, Low-Energy Electron Diffraction Study of the Phase Transition of Si( 001 ) Surface below 40 K, Phys. Rev. Lett 90, p.106103 (2003). R. Stalder, H. Sirringhaus, N. Onda, and H.V. Kane], Observation of misfit dislocations in epitaxial CoSi;/Si( 111 ) layers by scanning tunneling microscopy, Appl. Phys. Lett. 59(16), 111960-1962 (1991). K. Ojima, M. Yochimura, and K. Ueda, STM observation of the 2x3 and c(2x6) structures on Ba/Si(001), Surf. Sci. 491(1-2), p.169 (2001). DE. Jones, J.P. Pelz, Y. Hong, E. Bauer, and I.S.T. Tsong, Striped Phase and Temperature Dependent Step Shape Transition on Highly B-Doped Si(001)-(2x1) Surfaces, Phys. Rev. Lett. 77(2), p.330 (1996). J .-F. Nielsen, H.-J. Im, and J .P. Pelz, Scanning tunneling microscope studies of boron-doped Si( 001 ), J. Vac. Sci. Tec. A 17(4), p.1670 (1999). DE. Jones, J .P. Pelz, Y.H. Xie, P.J. Silverman, and GH. Gilmer, Enhanced Step Waviness on SiGe( 001 H 2x1 ) Surface under Tensile Strain, Phys. Rev. Lett. 75(8), p.1570 (1995). K.M. Chen, D.E. Jesson, SJ. Pennycook, M. Mostoller, and T. Kaplan, Step Instabilities: A New Kinetic Route to 3D Growth, Phys. Rev. Lett. 75(8), p.1582 (1995). G Ye, J. Nogami, and M.A. Crimp (submitted). 138 [155] [156] [157] [158] [159] [160] [161] [162] [163] [164] [165] [166] [167] OF Balkashin, A.GM. Jansen, O. Laborde, U. Gottlieb, GL. Sukhodub, P. Wyder, and I.K. Yanson, High-Frequency Point-Contact Spectroscopy of IiSiz, TaSiz, and VSi2, J. Low Temp. Phys. 129(314), [1105-116 (2002). CC. Lin, W.S. Chen, K.Y.J. Hsu, H.K. Liou, and K.N. Tu, Reliability study of sub-micron titanium silicide contacts, Appl. Surf. Sci. 92, p.660-664 (1996). N.M. Ravindra, L. Jin, D. Ivanov, V.R. Mehta, L.M. Dieng, G Popov, O.H. Gokce, J. Grow, and AT Fiory, Electrical and Compositional Properties of TaSiz Films, J. of Elec. Matls. 31(10), 111074-1079 (2002). G Medeiros-Ribeiro, D.A.A. Ohlberg, D.R. Bowler, R.E. Tanner, GAD. Briggs, and RS. Williams, Titanium disilicide nanostructures: two phases and their surfaces, Surf. Sci. 431, p.116-127 (1999). KL. Saenger, J. C. Cabral, LA. Clevenger, R.A. Roy, and S. Wind, A kinetic study of the C49 to C54 IiSiZ conversion using electrical resistivity measurements on single narrow lines, J. Appl. Physics 78(12), p.7040-7044 (1995). P. Gallais, J .J . Hantzpergue, and LC. Remy, Sputter deposition of thin tantalum layers and low temperature interactions between tantalum and SiOz and tantalum and silicon, Thin Solid Films 165, p.227-236 (1988). M.T. Huang, T.L. Martin, V. Malhotra, and IE. Mahan, Electronic transport properties of tantalum disilicide thin films, J. Vac. Sci. Technol 33(3), p.836-845 (1985). H. Oppolzer, F. Neppl, K. Hieber, and V. Huber, Influence of slight deviations from TaSiZ stoichiometry on the high-temperature stability of tantalum silicide/silicon contacts, J. Vac. Sci. Technol B2(4), p.630-635 (1984). SB. Hemer and KS. Jones, Investigation of mechanisms of vacancy generation in silicon in the presence of a TaSiz film, J. Appl. Phys. 82(2), p.583-588 (1997). CV Hul'ko, R. Boukherroub, and GP. Lopinski, Chemical and thermal stability of titanium disilicide contacts on silicon, J Appl. Phys. 90(3), p.1655-1659 (2001). N. Lundqvist, J. Aberg, S. Nygren, C.-A. Bjormander, and S.-L. Zhang, Effects of substrate bias and temperature during titanium sputter-deposition on the phase formation in IiSiz, Microelectron Eng. 60, p.211-220 (2002). A.W. Dunn, B.N. Cotier, A. Nogaret, P. Moriarty, and RH. Beton, Molecular scale alignment strategies: An investigation of Ag adsorption on patterned fidlerene layers, Appl. Phys. Lett. 71(20). P.2937-2939 (1997). RA. Bennett, B. Ashcroft, Z. He, and R.M. Tromp, Growth dynamics of titanium silicide nanowires observed with low-energy electron microscopy, J. Vac. Sci. Tec. B 20(6), p.2500-2504 (2002). 139 in!!! [168] [169] [170] [171] [172] [173] [174] [175] [176] [177] [178] [179] [180] [181] [182] W.-C. Yang, H. Ade, and R.J. Nemanich, Shape stability of TiSiz islands on Si( 111 ), J. Appl. Physics 95(3), P.1572-1576 (2004). E. Abramof, A.F.d. Silva, B.E. Semelius, J.P.d. Souza, and H. Boudinov, Metal-nonmetal transition and resistivity of silicon implanted with bismuth, J. Mater. Res. 12(3), [1641-645 (1997). http://www. tydex. ru/mate rial s/material s2/S i. html . P. Gambardella, A. Dallmeyer, K. Maiti, M.C. Malagoli, W. Eberhardt, K. Kern, and C. Carbone, Ferromagntism in one-dimensional monatomic metal chanis, Nature 416(6878), P301 (2002). P. Gambardella, Magnetism in monatomic metal wires, J. Phys. Condens. Matter 15, p.82533-S2546 (2003). J. Pierre, E. Siaud, and D. Frachon, Magnetic Properties of Rare Earth Disilicides RSiZ, J. Less-Common Met. 139, p.321-329 (1988). K. Sekizawa and K. Yasukochi, Antiferromagnetism of Disilicides of Heavy Rare Earth Metals, J. Phys. Soc. Jpn. 21, p.274 (1966). C. Kittel, Introduction to Solid State Physics. 6 ed. 1986. S. Labroo, X. Zhang, P. Hill, and N. Ali, Electrical and magnetic properties of antiferromagnetic rare earth disilicides, J. Less-Common Met. 149, p.337 (1989). H. Yashima, T. Satoh, H. Mori, D. Watanabe, and T. Ohtsuka, Thermal and magnetic properties and crystal structures of CeGez and CeSiz, Solid State Communications 41(1), p.1-4 (1982). J. Pierre, S. Auffret, E. Siaud, R. Madar, E. Houssay, A. Rouault, and J .P. Senateur, Magnetic properties of rare earth silicide crystals RSiZ-x (R = Pt; Nd, Gd), J. Magn. Magn. Mater. 89, p.86-96 (1990). ET Matthias, E. Corenzwit, and W.H. Zachriasen, Phys. Rev. 112, p.89 (1958). K.S.V.L. Narasimhan and H. Steinfink, Magnetic Investigations on AlBZ Type Structures, J. Solid State Chem. 10, p.137-141 (1973). C. Pescher, J. Pierre, A. Ermolieff, and C. Vannuffel, Magnetic properties of gadolinium silicide thin films produced by different fabrication processes, Thin Solid Films 278, p.140-143 (1996). J. Pierre, B. Lambert-Andron, and J .L. Soubeyroux, Magnetic-structures of rare earth silicides NdSi2.x, HoSi2.x, DySi2.x, J. Magn. Magn. Mater. 81, p.39 (1988). 140 [183] [184] [185] [186] [187] [188] [189] [190] [191] [192] [193] [194] [195] MC. Hersam, N.P. Guisinger, and J.W. Lyding, Atomic-level study of the robustness of the Si(100)-2x1:H surface following exposure ro ambient conditions, Appl. Phys. Lett. 78(7), p.886-888 (2001). T. Sakurai and H.D. Hagstrum, Interplay of the monohydride phase and a newly discovered dihydride phase in chemisorption of H on Si(100)2x1, Phys. Rev. B 14, p.1593-1596 (1976). Y]. Chabal and K. Raghavachari, New Ordered Structure for the H-Saturated Si( 100) Sun‘ace: The (3x1) Phase, Phys. Rev. Lett 54(10), 111055-1058 (1985). M. Niwa, H. Iwasaki, and S. Hasegawa, Hydrogen terminated Si( 100) surfaces studied by scanning tunneling microscopy, X-ray photon spectroscopy, and auger-electron spectroscopy, J. Vac. Sci. Tec. A 8(1), p.266-269 (1990). J.J. Boland, Structure of the H-saturated Si( 100) surface, Phys. Rev. Lett. 65, p.3325-3328 (1990). J .J. Boland, Role of bond-strain in the chemistry of hydrogen on the Si( 100) surface, Surface Science 261(1-3), p.17-28 (1992). Y. Okada, H. Shimomura, and M. Kawabe, Atomic image observation of hydrogen-saturated Si(100) prepared by atomic-hydrogen irradiation, J ap. J. Appl. Phys. Part2-Letters 31(8A). P-L1121-L1123 (1992). K. Kitahara and O. Ueda, Observation of atomic-structure by scanning-tunneling -microscopy of vicinal Si(100) surface annealed in hydrogen gas, Jap. J. Appl. Phys. Part2-Letters 33(llB). P.L157l-L1573 (1994). M. Yoshimura and K. Ueda, Evolution of surface structures on Si(001) during hydrogen desorption, Appl. Surf. Sci. 121, p.179-182 (1997). T. Komeda and Y. Kumagai, Si( 001 ) surface variation with annealing in ambient H2, Phys. Rev. B 58(3). 111385-1391 (1998). K. Arima, K. Endo, T. Kataoka, Y. Oshikane, H. Inoue, and Y. Mori, Scanning tunneling microscopy study of hydrogen-terminated Si( 001 ) surfaces after wet cleaning, Surface Science 446(1-2), p.128-136 (2000). . T. Hashizume, H. Kajiyama, Y. Suwa, S. Heike, S. Matsuura, J. Nara, and T. Ohno, Adsorption of Si atom on H-terminated Si(001)-2x1 surface, Appl. Surf. Sci. 216(1-4), P.15-18 (2003). Y. Suwa, S. Matsuura, M. Fujimori, S. Heike, T. Onogi, H. Kajiyama, T. Hitosugi, K. Kitazawa, T. Uda, and T. Hashizume, Dopant-pair structures segregated on a hydrogen-terminated Si(100) surface, Phys. Rev. Lett 90(15), p.156101 (2003). 141 [196] [197] [198] [199] [200] [201] [202] [203] [204] [205] [206] [207] T. Hirosugi, T. Hashizume, S. Heike, Y. Wada, S. Watanabe, T. Hasegawa, and K. Kitazawa, Scanning tunneling spectroscopy of dangling-bond wires fabricated on the Si(100)-2x1 -H surface, Appl. Phys. A 66, p.8695-S699 (1998). J .A. Schaefer, Electronic and structural-properties of hydrogen on semiconductor surfaces, Physica B ”004), p.45-68 (1991). E.J. Buehler and J.J. Boland, Dimer preparation that mimics the transition state for the adsorption of H-2 on the Si(100)-2x1 surface, Science 290(5491), p.506-509 (2000). K. Bobrov, G Comtet, G Dujardin, and L. Hellner, Electronic structure of partially hydrogenated Si(100)-(2 x 1) surfaces prepared by thermal and nontherrnal desorption, Phys. Rev. Lett 86(12), p.2633-2636 (2001). D.X. Chen and J .J . Boland, Chemisorption-induced disruption of surface electronic structure: Hydrogen adsorption on the Si(100)-2x1 surface, Phys. Rev. B 65(16). 13.165336 (2002). J .J . Boland, Evidence of pairing and its role in the recombinative desorption of hydrogen from the Si(100)-2x1 surface, Phys. Rev. Lett 67(12), p.1539 (1991). J .J . Boland, Scanning tunneling microscopy study of the adsorption and recombinative desorption of hydrogen from the Si(100)-2x1 surface, J. Vac. Sci. Tec. A 10(4), p.2458-2464 (1992). X.D. Wang, H. Lu, T. Hashizume, H.W. Pickering, and T. Sakurai, Atomic-hydrogen chemisorption on Si(100)(2x1) studied by FI-STM, Appl. Surf. Sci. 57(1-4), P.266-274 (1993). GC. Abeln, T.C. Shen, J .R. Tucker, and J .W. Lyding, Nanoscale STM-patteming and chemical modification of the Si( 100) surface, Microelectron Eng. 27(1-4), p.23-26 (1995). P. Avouris, R.E. Walkup, A.R. Rossi, T.C. Shen, GC. Abeln, J .R. Tucker, and J .W. Lyding, S TM-induced H atom desorption from Si(100): Isotope effects an site selectivity, Chem. Phys. Lett. 257(1-2), p.148-154 (1996). J.W. Lyding, T.C. Shen, GC. Abeln, C. Wang, and JR. 'Ducker, Nanoscale patterning and selective chemistry of silicon surfaces by ultrahigh-vacuum scanning tunneling microscopy, Nanotechnology 7(2), p.128-133 (1996). J .W. Lyding, T.C. Shen, GC. Abeln, C. Wang, P.A. Scott, J .R. Tucker, P. Avouris, and RE. Walkup, Ultrahigh vacuum scanning tunneling microscope-based nanolithography and selective chamistry on silicon surfaces, Israel J. Chem. 36(1), p.3-10 (1996). 142 [208] [209] [210] [211] [212] [213] [214] [215] [216] [217] [218] [219] D.H. Huang and Y. Yamarnoto, Manipulating atoms one by one with a scanning tunneling microscope, Surf. Rev. Lett. 3(3), p.1463-1472 (1996). T. Hitosugi, T. Hashizume, S. Heike, S. Watanabe, Y. Wada, T. Hasegawa, and K. Kitazawa, Scanning tunneling spectroscopy of dangling-bond wires fabricated on the Si(100)-2x1 -H surface, Jap. J. Appl. Phys. Part2-Letters 36(3B), p.L361-L364 (1997). M. Sakurai, C. Thirstrup, T. Nakaya, and M. Aono, Local modification of hydrogen-terminated silicon surfaces by clean and hydrogen-covered STM tips, Surf. Sci. 386(1-3), P.154-160 (1997). C. Thirstrup, M. Sakurai, T. Nakayama, and M. Aono, Atomic scale modifications of hydrogen-terminated silicon 2x1 and 3x1 surfaces by scanning tunneling microscope, Surf. Sci. 411, p.203-214 (1998). C. Syrykh, J.P. Nys, B. Legrand, and D. Stievenard, Nanoscale desorption of H—passivated Si(100)-2x1 surfaces using an ultrahigh vacuum scanning tunneling microscope, J Appl. Phys. 85(7), p.3887-3892 (1999). Q. Shi, D.H. Huang, and GS. Zhu, Vibrational-energy redistribution in single-atom manipulation by scanning tunneling microscope, Jap. J. Appl. Phys. Part] 38(6B), p.3856-3659 (1999). K. Stokbro, U.J. Quaade, R. Lin, C. Thirstrup, and G Grey, Electronic mechanism of STM-induced difi‘usion of hydrogen on Si(100), Faraday Discussions 117, p.231-240 (2000). M. Durr, Z.H. Hu, A. Biedermann, U. Hofer, and TR Heinz, Real-space study of the pathway for dissociative adsorption of H—2 on Si( 001 ), Phys. Rev. Lett 88(4), p.046104 (2002). MC. Hersam, N.P. Guisinger, J. Lee, K.G Cheng, and J.W. Lyding, Variable temperature study of the passivation of dangling bonds at Si(100)-2x1 reconstructed surfaces with H and D, Appl. Phys. Lett. 80(2), p.201-203 (2002). Z.H. Hu, Quantitative study of adsorbate-adsorbate interactions of hydrogen on the Si(100) surface, Phys. Rev B 68(15), p.155418 (2003). H. Tsurumaki, K. Iwamura, T. Karato, S. Inanaga, and A. Namiki, Adsorption and desorption of deuterium on partially oxidized Si(100) surfaces, Phys. Rev. B 67(15). 13.155316 (2003). L. Soukiassian, A.J. Mayne, M. Carbone, and G Dujardin, Atomic-scale desorption of H atoms from the Si(100)-2x1 : H sudace: Inelastic electron interactions, Phys. Rev. B 68(3), p.035303 (2003). 143 [220] [221] [222] [223] [224] [225] [226] [227] [228] [229] [230] [231] T. Hashizume, S. Heike, M.I. Lutwyche, S. Watanabe, and Y. Wada, Atom structures on the Si(100) surface, Surface Science 386(1-3), p.161-165 (1997). T. Hitosugi, T. Hashizume, S. Heike, H. Kajiyama, Y. Wada, S. Watanabe, T. Hasegawa, and K. Kitazawa, Electronic structures of dangling-bond structures fabricated on hydrogen-terminated Si(100)-2x1 surfaces, Deffect and Diffusion Forum/Journal 162, p.43-57 (1998). T. Hitosugi, S. Heike, T. Onogi, T. Hashizume, S. Watanabe, Z.Q. Li, K. Ohno, Y. Kawazoe, T. Hasegawa, and K. Kitazawa, Jahn-Teller distortion in dangling-bond linear chains fabricated on a hydrogen-terminated Si(100)-2 x 1 surface, Phys. Rev. Lett 82(20), P.4034-4037 (1999). R.P. Chen and D.S. Lin, Distribution of dangling bond pairs on partially hydrogen-terminated Si(100) surface observed by scanning tunneling microscopy, Surf. Sci. 454, p.196-200 (2000). H. Jungblut, D.J. Muller, M. Aggour, and HI. Lewerenz, Scanning-tunneling-microscopy observation of atomic structures on silicon (100) surface in air, Electrochimica Acta 38(10), p.1367-137l (1993). N. Kramer, M.R. vandenBerg, and C. Schonenberger, Scanning tunneling microscope-induced oxidation of hydrogen passivated silicon surfaces, Thin Solid Films 282(1-2), [1637-639 (1996). H. Kijiyama, S. Heike, Y. Wada, and T. Hashizume, Initial stage oxidation at an unpaired dangling bond site on a Si(100)-2 X I-H surface, Thin Solid Films 344(Sp. Iss. SI), p.550-553 (1999). Z.H. Mai, Y.F. Lu, W.D. Song, and WK. Chim, Nano-modification on hydrogen-passivated Si surfaces by a laser-assisted scanning tunneling microscope operating in air, Appl. Surf. Sci. 154, p.360-364 (2000). DA. MacLaren, NJ. Curson, P. Atkinson, B. Holst, D.J. Johnson, and W. Allison, Simple design for the transportation of ex situ prepared hydrogen passivated silicon, J. Vac. Sci. Tec. A 20(1), p.285-287 (2002). S. Maruno, H. Iwasaki, K. Horioka, S.T. Li, and S. Nakamura, Electronic structures of the monohydride (2x1):H and dihydride (1x1 )::2H Si(001) surfaces studied by angle—resolved electron-energy-loss spectroscopy, Phys. Rev. B 27(7), p.4110—4116(1983). D.T. Jiang, GW. Anderson, K. Griffiths, T.K. Sham, and RR. Norton, Adsorption of atomic hydrogen on Si(199)-2x1 at 400 K, Phys. Rev. B 48(7), p.4952-4955 (1993). J .Y. Veuillen and T.A.N. Tan, Hydrogen adsorption on ErSi1,7(0001), Phys. Rev. B 52(15), p.10796-10799 (1995). ' 144 [232] [233] [234] [235] [236] [237] [238] [239] [240] [241] [242] [243] S. Saintenoy, P. Wetzel, C. Pirri, J.C. Peruchetti, D. Bolmont, and G Gewinner, Hydrogen adsorption on erbium silicide surface, Solid State Communications 94(9), [3719-723 (1995). S. Saintenoy, P. Wetzel, C. Pirri, D. Bolmont, and G Gewinner, Interaction of H with epitaxial Er silicide layers on Si(III) adsorption versus absorption, Surface Science 349, p. 145-154 (1996). P. Wetzel, C. Pirri, and G Gewinner, Buckling reversal of the Si(III) bilayer termination of 2-dimensional ErSiz upon H dosing, Europhysics Letters 38(5), p.359-364 (1997). P. Wetzel, T. Angot, C. Pirri, and G Gewinner, Origin of the semimetal-to-semiconductor transition observed in two-dimensional Er silicide upon H exposure: evidence at two chemisorption sites, Surf. Sci. 383, p.340-349 (1997). P. Louis, T. Angot, D. Bolmont, and G Gewinner, Pairing mechanism in interaction of atomic hydrogen with epitaxial erbium silicide, Surf. Sci. 422, p.65-76 (1999). M.-H. Thilier, C. Pirri, D. Berling, D. Bolmont, G Gewinner, and P. Wetzel, Structure of clean and H-saturated epitaxial two-dimensional Er silicide on Si(III) studied by SEXAFS, Surf. Sci. 555(1-3), 1394-100 (2004). http://www. icknowl edge. com/glossa ry/s. html . http://www. ece. gatech. edu/resea rch/labs/vc/theory/sheetRes. html . L. Maisse] and R. Glang, Handbook of Thin Film Technology. 1983: McGraw-Hill. J. Homstra and L.J.v.d. Pauw, J. Electron. Control 7, p. 169 (1959). W.L.V. Price, Extension of van der Pauw's theorem for measuring specific resistivity in disks of arbitrary shape to anisotropic media, J. Phys. D: Appl. Phys. 5(6), p.112? (1972). W.L.V. Price, Electric potential and current distribution in a rectangular sample of anisotropic material with application to the measurement of the principal resistivities by an extension of van der Pauw's method, Solid State Electronics 16, p.753-762 (1973). 145