ABSTRACT Title of Dissertation: IN SITU ENRICHMENT AND EPITAXIAL GROWTH OF 28Si FILMS VIA ION BEAM DEPOSITION Kevin Joseph Dwyer, Doctor of Philosophy, 2017 Directed By: Professor John Cumings, Department of Materials Science and Engineering Isotopically enriched 28Si is an ideal material for solid state quantum comput- ing because it interacts weakly with the spin states of embedded qubits (quantum bits) resulting in long coherence times. This is the result of eliminating the roughly 4.7 % 29Si isotopes present in natural abundance Si, which possesses nuclear spin I = 1/2 that is disruptive to qubit operation. However, high-quality 28Si is scarce and the degree to which it improves the performance of a qubit is not well under- stood. This leads to an important question in the Si-based quantum information field, which can be stated as “how good is good enough?” regarding the perfection of 28Si as a host medium for qubits. The focus of this thesis is to engineer a material that can address this question, specifically in terms of the enrichment. Secondary requirements for ideal 28Si films that are also pursued are crystalline perfection and high chemical purity. I report on the production and characterization of 28Si thin films that are the most highly enriched of any known 28Si material ever produced with a maximum 28Si enrichment of 99.9999819(35) % and a residual 29Si isotopic concentration of 1.27(29)× 10−7. A hyperthermal energy ion beamline is used to produce this ex- treme level of enrichment starting from a natural abundance silane gas (SiH4) source. The Si is enriched in situ by mass separating the ions in a magnetic field just before deposition onto Si(100) substrates. Initial proof of principle experiments enriching 22Ne and 12C were also conducted. In the course of achieving this 28Si enrichment, I also pursue the epitaxial deposition of 28Si thin films. Characterizations of the film morphology and crystallinity are presented showing that smooth, epitaxial 28Si films are achieved using deposition temperatures between 349 ◦C and 460 ◦C. Crystalline defects present in these films include {111} stacking faults. When using higher deposition temperatures, I find that trace impurity compounds such as SiC cause step pinning and faceting of the growth surface leading to severely rough films. As- sessments of the chemical purity of 28Si films are also presented, which show major impurities N, C, and O are present in the purest film at an atomic concentration of approximately 1× 1019 cm−3, resulting in a Si purity of 99.96(2) %. Additionally, I introduce a model that describes the residual 29Si and 30Si in 28Si films, i.e. the enrichment, as the result of adsorption of diffusive natural abun- dance SiH4 gas from the ion source into the 28Si films during deposition. This model correlates the measured enrichments of 28Si films with the SiH4 partial pressures dur- ing deposition. An incorporation fraction for SiH4 adsorption at room temperature of s = 6.8(3)× 10−4 is extracted. Finally, the temperature dependence of the sample enrichment is analyzed using a thermally activated incorporation model that gives an activation energy of Ec = 1.1(1) eV for the reactive sticking coefficient of SiH4. IN SITU ENRICHMENT AND EPITAXIAL GROWTH OF 28Si FILMS VIA ION BEAM DEPOSITION by Kevin Joseph Dwyer Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2017 Advisory Committee: Professor John Cumings, Chair/Advisor Dr. Joshua Pomeroy, Co-Advisor Professor Lourdes Salamanca-Riba Professor Ichiro Takeuchi Professor Neil Goldsman © Copyright by Kevin Joseph Dwyer 2017 Acknowledgments The path that led me to become a Ph.D. candidate studying materials science and engineering (MSE) at the University of Maryland, College Park began as I was an undergraduate physics major at Maryland and sought to join a lab to gain research experience. I contacted Professor Gottlieb Oehrlein, who invited me to work in his lab, for which I am grateful. His lab was a materials science lab researching plasma-surface interactions. Later, when applying to graduate programs, Professor Oehrlein’s graduate student, Robert Bruce, suggested that I apply to the Materials Science and Engineering Department at Maryland. I did apply, and, after being accepted, I decided to become a materials scientist. While taking first year classes in graduate school, my professor, John Cum- ings, put me in touch with a scientist at the National Institute of Standards and Technology (NIST) looking for a graduate student. That scientist was Dr. Joshua Pomeroy, and I soon joined his group at NIST in Gaithersburg after an impressive tour. Josh became my co-advisor and John my official MSE advisor for the duration of a fruitful six year collaboration that produced the work in this thesis. I thank John for not only introducing me to Josh and being the “guy that signs my forms”, but also for his practical advice for achieving academic milestones as well as the growth of my scientific career. Working with Josh at NIST has been pivotal in my development as a scientist, and I thank him for his role in guiding that development. The combination of practical engineering and analytical science knowledge imparted by Josh has made a strong impression on me and will no doubt serve me well moving ii forward. Additionally, Josh’s do-it-yourself attitude and adage that “data talks and everything else walks” will always stick with me. I am very proud of what we have accomplished and look forward to collaborating with him again. I want to thank Dr. Russell Lake, Josh’s first graduate student, for making me feel welcome after transitioning from Maryland to Josh’s group at NIST and for many insights and scientific discussions. I thank Josh’s current group members, Dr. Aruna Ramanayaka and Ke Tang for helpful discussions, and especially Hyun soo Kim, with whom I spent countless hours in the lab (breaking and fixing things) in the pursuit of data. I have had the fortune and pleasure of working with numerous other smart people at NIST. Thank you to former and current members of the Quantum Processes and Metrology group at NIST including Dr. Neil Zimmerman, Dr. Ted Thorbeck, Dr. Panu Koppinen, Dr. Justin Perron, Dr. Michael (Stew) Stewart, Dr. Roy Murray, Dr. Hamza Shakeel, and Zac Barcikowski for numerous helpful discussions and brainstorming sessions over the years. Interactions with these colleague have led to valuable scientific relationships and even more valuable friendships. Additionally, thank you to Dr. Garnett Bryant for paying my stipend during my tenure. I gratefully acknowledge the Laboratory for Physical Sciences for partially funding this work and look forward to joining them as a postdoctoral researcher. Other collaborators at NIST I want to thank include Terry Moore, Dr. Rick Silver, Dr. Kai Li, Dr. Pradeep Namboodiri, Xiqiao Wang, Dr. Kristen Steffens, Dr. June Lau, Dr. Vald Oleshko, Dr. Joshua Schumacher, and Dr. Alline Myers. I extend a special thank you to the most important collaborator I had the pleasure iii of working with, Dr. Dave Simons, whose expertise in isotope measurements and willingness to push the boundaries of his work played a critical role in this thesis. Thank you to the fellow students of my MSE Ph.D candidate class including Alex, Jen, Amy, Colin, Elliot, Ben, and Jeff for helping me survive my first year of graduate school. Thank you to the GRA NIST solfball team led by our fearless leader Jack for helping me blow off steam at the plate and on the pitching mound in between experiments. We may not have won any championships, but it was fun nonetheless! Thank you also to my roommates Dana, Somak, and Evan for putting up with my crazy hours while writing this thesis. Most crucially, I want to acknowledge and extend a very warm thank you to my parents Eddie and Linda and brother John as well as the rest of my family for their endless love and support over the years! Thank you to my parents for cultivating, encouraging, and enabling my scientific curiosity as a child. Lego building, physics and chemistry kits, model rockets, a telescope, trips to the Franklin Institute, and numerous other activities and discussions about science topics provided by my par- ents formed the foundation for the scientist that I have become. Further financial and emotional support from them throughout the years as I progressed through high school, college, and finally, graduate school is greatly appreciated. Ultimately, the work presented in this thesis would not have been possible without them. Finally, thank you to my committee members Josh, John, Professor Lourdes Salamanca-Riba, Professor Ichiro Takeuchi, and Professor Neil Goldsman for their careful consideration of this thesis. iv Table of Contents List of Tables ix List of Figures x 1 Introduction 1 1.1 28Si for Quantum Information . . . . . . . . . . . . . . . . . . . . . . 1 1.1.1 Si-Based Solid State Quantum Information . . . . . . . . . . . 1 1.1.2 Sources of 28Si . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 1.1.3 Single Spin Measurements in 28Si . . . . . . . . . . . . . . . . 15 1.2 Ion Beam Enrichment and Deposition . . . . . . . . . . . . . . . . . . 19 1.3 Objectives and Outline . . . . . . . . . . . . . . . . . . . . . . . . . . 23 1.3.1 Project Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 1.3.2 Strategy and Impact of Results . . . . . . . . . . . . . . . . . 25 1.3.3 Outline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 2 Experimental Apparatus and Methods 30 2.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 2.1.1 Ultra-High Vacuum Deposition . . . . . . . . . . . . . . . . . 30 2.1.2 Previous Operation . . . . . . . . . . . . . . . . . . . . . . . . 32 2.2 Hyperthermal Energy Ion Beamline . . . . . . . . . . . . . . . . . . . 33 2.2.1 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 2.2.2 Theory of Magnetic Mass Separation . . . . . . . . . . . . . . 40 2.2.3 Operating Parameters . . . . . . . . . . . . . . . . . . . . . . 50 2.2.4 Ion Beam Characterization . . . . . . . . . . . . . . . . . . . . 54 2.2.4.1 Ion Beam Mass Spectra . . . . . . . . . . . . . . . . 54 2.2.4.2 Ion Beam Energy, Ei . . . . . . . . . . . . . . . . . . 67 2.2.4.3 Ion Beam Spot Size . . . . . . . . . . . . . . . . . . 73 2.3 UHV Deposition and Analysis Chamber . . . . . . . . . . . . . . . . 75 2.3.1 Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 2.3.2 Vacuum Analysis . . . . . . . . . . . . . . . . . . . . . . . . . 77 2.3.3 Sample Manipulation . . . . . . . . . . . . . . . . . . . . . . . 79 2.3.4 In situ Sample Analysis . . . . . . . . . . . . . . . . . . . . . 87 v 3 Initial Experiments Enriching 22Ne and 12C 91 3.1 Context and Experimental Setup . . . . . . . . . . . . . . . . . . . . 91 3.2 22Ne Implantation and Characterization: Proof of Principle . . . . . . 94 3.2.1 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . . . 94 3.2.2 22Ne Implantation . . . . . . . . . . . . . . . . . . . . . . . . . 94 3.2.3 Enrichment Measurements via SIMS . . . . . . . . . . . . . . 99 3.3 12C Deposition and Characterization: First Enriched Thin Films . . . 102 3.3.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 3.3.2 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . . . 103 3.3.3 Deposition of 12C . . . . . . . . . . . . . . . . . . . . . . . . . 104 3.3.4 Enrichment Measurements via SIMS . . . . . . . . . . . . . . 109 3.4 Chapter 3 Summary: Outlook for 28Si . . . . . . . . . . . . . . . . . . 114 4 28Si Thin Film Deposition and Characterization Phase I: In Situ Enrichment116 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 4.1.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116 4.1.2 Experimental Configurations for 28Si Deposition . . . . . . . . 119 4.2 Si Deposition Proof of Principle: Ion Beam Chamber Samples . . . . 125 4.2.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . 125 4.2.2 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . . . 128 4.2.3 Deposition of 28Si . . . . . . . . . . . . . . . . . . . . . . . . . 128 4.2.4 Enrichment Measurements via SIMS for IC–1 Samples . . . . 134 4.2.5 Summary of Results for IC–1 Samples . . . . . . . . . . . . . 141 4.3 Achieving Highly Enriched 28Si: Lens Chamber Samples . . . . . . . 142 4.3.1 Experimental Setup . . . . . . . . . . . . . . . . . . . . . . . . 142 4.3.2 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . . . 145 4.3.3 Deposition of 28Si . . . . . . . . . . . . . . . . . . . . . . . . . 146 4.3.4 Enrichment Measurements via SIMS for LC–2 Samples . . . . 150 4.3.5 Crystallinity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157 4.3.6 Chemical Purity . . . . . . . . . . . . . . . . . . . . . . . . . . 159 4.3.6.1 SIMS . . . . . . . . . . . . . . . . . . . . . . . . . . 159 4.3.6.2 XPS . . . . . . . . . . . . . . . . . . . . . . . . . . . 164 4.4 Chapter 4 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 167 5 28Si Thin Film Deposition and Characterization Phase II: Crystallinity and Chemical Purity 169 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 5.2 Experimental Setup for Improving Crystallinity and Chemical Purity: Deposition Chamber Samples . . . . . . . . . . . . . . . . . . . . . . 175 5.3 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180 5.3.1 Ex Situ Cleaning . . . . . . . . . . . . . . . . . . . . . . . . . 180 5.3.2 In Situ Preparation . . . . . . . . . . . . . . . . . . . . . . . . 183 5.4 Deposition of 28Si . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191 5.5 Enrichment Measurements via SIMS for DC–3 Samples . . . . . . . . 201 5.5.1 Initial Tests at DC–3 . . . . . . . . . . . . . . . . . . . . . . . 201 vi 5.5.2 Enrichment Progression Timeline Samples . . . . . . . . . . . 205 5.5.3 Samples with Deposition T > 600 ◦C . . . . . . . . . . . . . . 214 5.5.4 High Pressure Mode Sample . . . . . . . . . . . . . . . . . . . 225 5.6 Epitaxial Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 5.6.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229 5.6.2 Morphology of Films with Deposition T > 600 ◦C . . . . . . . 238 5.6.2.1 RHEED . . . . . . . . . . . . . . . . . . . . . . . . . 238 5.6.2.2 STM . . . . . . . . . . . . . . . . . . . . . . . . . . . 243 5.6.2.3 SEM . . . . . . . . . . . . . . . . . . . . . . . . . . . 244 5.6.2.4 Step Pinning Induced Roughness . . . . . . . . . . . 250 5.6.3 Elimination Strategies for Step Pinning Sites . . . . . . . . . . 280 5.6.4 Morphology of Films with Deposition T < 600 ◦C . . . . . . . 292 5.6.4.1 RHEED . . . . . . . . . . . . . . . . . . . . . . . . . 292 5.6.4.2 STM . . . . . . . . . . . . . . . . . . . . . . . . . . . 297 5.7 Chemical Purity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306 5.7.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 306 5.7.2 XPS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309 5.7.3 SIMS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311 5.8 Crystallinity: Film Inspection via TEM . . . . . . . . . . . . . . . . . 334 5.9 Chapter 5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 347 6 Pressure and Temperature Dependent Adsorption of 29Si and 30Si During 28Si Deposition 352 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 352 6.2 Experimental Methods . . . . . . . . . . . . . . . . . . . . . . . . . . 357 6.2.1 28Si Samples and Enrichment Values . . . . . . . . . . . . . . 357 6.2.2 SiH4 Mass Spectrum and Mass Selectivity . . . . . . . . . . . 360 6.2.3 Determination of SiH4 Partial Pressures . . . . . . . . . . . . 365 6.2.4 Substrate Temperature Calibration . . . . . . . . . . . . . . . 369 6.3 Temperature Dependent Gas Incorporation Model . . . . . . . . . . . 370 6.4 Correlating Enrichment to SiH4 Partial Pressure . . . . . . . . . . . . 373 6.5 Temperature Dependence of 29Si and 30Si Adsorption . . . . . . . . . 379 6.6 Temperature Dependence of the Incorporation Fraction, s . . . . . . . 383 6.7 Determination of the Reactive Sticking Activation Energy, Ec . . . . 387 6.8 Chapter 6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . 390 7 Summary of Results and Future Experiments 392 7.1 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 392 7.2 Proposals for Future Experiments . . . . . . . . . . . . . . . . . . . . 400 7.2.1 Targeted Levels of Enrichment . . . . . . . . . . . . . . . . . . 400 7.2.2 Enriched Si/Ge Deposition . . . . . . . . . . . . . . . . . . . . 403 7.2.3 28Si sublimation . . . . . . . . . . . . . . . . . . . . . . . . . . 404 7.2.4 Al Dopant Devices with Hydrogen Lithography . . . . . . . . 405 7.2.5 Electrical Measurements and T2 in 28Si . . . . . . . . . . . . . 407 vii A Ion Source and Beamline: Additional Operating Parameters 410 B Experimental Apparatus Photographs 415 C Substrate Catalog 425 D Sample Catalogs 427 E SIMS Measurement Settings 441 F 28Si Deposition Fun Facts 448 Bibliography 449 viii List of Tables A.1 Ion source and beamline operating parameters for implanting 22Ne on 5/4/11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 413 A.2 Ion source and beamline operating parameters for depositing 12C on 2/7/12 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 413 A.3 Ion source and beamline operating parameters for depositing 28Si on 12/18/15 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 414 C.1 Substrate Catalog . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 426 D.1 22Ne Sample Catalog (3/1/11–5/4/11) . . . . . . . . . . . . . . . . . 428 D.2 12C Sample Catalog (1/25/12–2/7/12) . . . . . . . . . . . . . . . . . 429 D.3 28Si Sample Catalog: IC–1 (6/4/12–6/28/12) . . . . . . . . . . . . . . 430 D.4 28Si Sample Catalog: LC–2: I (2/4/13–3/4/13) . . . . . . . . . . . . . 431 D.5 28Si Sample Catalog: LC–2: II (3/7/13–9/24/13) . . . . . . . . . . . 432 D.6 28Si Sample Catalog: DC–3: I (2/7/14–6/26/14) . . . . . . . . . . . . 433 D.7 28Si Sample Catalog: DC–3: II (8/28/14–7/27/15) . . . . . . . . . . . 434 D.8 28Si Sample Catalog: DC–3: III (9/20/15–11/9/15) . . . . . . . . . . 435 D.9 28Si Sample Catalog: DC–3: IV (12/18/15–6/24/16) . . . . . . . . . . 436 D.10 Deposition parameters for the analysis of 28Si samples deposited at room temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 437 D.11 Enrichment measurements and analysis results for 28Si samples de- posited at room temperature . . . . . . . . . . . . . . . . . . . . . . . 438 D.12 Deposition parameters for the analysis of 28Si samples deposited at elevated temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . 439 D.13 Enrichment measurements and analysis results for 28Si samples de- posited at elevated temperature . . . . . . . . . . . . . . . . . . . . . 440 ix List of Figures 1.1 Si-based quantum dot and P-donor devices . . . . . . . . . . . . . . . 4 1.2 Schematic diagrams of the Bloch sphere for a qubit spin and the T2 coherence time due to spin dephasing . . . . . . . . . . . . . . . . . . 6 1.3 Cartoon depictions of a natural abundance Si crystal containing 29Si nuclear spins and a nuclear spin-free 28Si crystal . . . . . . . . . . . . 8 1.4 T2 coherence time measurements of the electron and nuclear spins of 31P atoms in 28Si . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 1.5 31P in Si energy levels and optical transitions, photoluminescence spectra of 31P in natural abundance Si and 28Si showing hyperfine splitting, and ESR frequency tuning of a 31P electron spin in 28Si . . 10 1.6 Process flow chart for production of a 28Si crystal from centrifugation of natural abundance SiF4 . . . . . . . . . . . . . . . . . . . . . . . . 12 1.7 Thermal conductivity measurements of natural abundance Si and 28Si 13 1.8 Schematic and photographs of the 28Si single-crystal boule and final sphere produced in the International Avogadro Coordination . . . . . 14 1.9 Predicted electron T∗2 coherence time dependance on 29Si concentra- tion for qubits in Si . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16 1.10 ESR measurement phase space diagram of 29Si concentration vs. number of measured 31P spins . . . . . . . . . . . . . . . . . . . . . . 18 1.11 Dual source ion beam deposition system schematic and an enrichment measurement of a deposited 28Si film from the Tsubouchi group . . . 22 1.12 Enrichment progression timeline of the best overall sample enrichments 26 2.1 Schematic of the ion beam deposition system . . . . . . . . . . . . . . 31 2.2 Ion beamline schematic with ion source inset and electrostatic lens potential landscape . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 2.3 RGA mass spectra of the ion beam chamber at two achieved base pressures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 2.4 Ion beamline schematic showing SiH4 diffusion and mass separation . 38 2.5 Calculated sector mass analyzer magnetic field relation to selected mass 43 2.6 Calculated mass dependance of the spatial separation, λ1, of ions with adjacent mass number in the ion beam . . . . . . . . . . . . . . . . . 45 x 2.7 Calculated accelerating voltage dependance of the spatial separation, λE, of ions accelerated with voltages differing by ∆V in the ion beam 48 2.8 Calculated mass dependance of the mass resolving power of the ion beamline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 2.9 Schematic diagrams of the gas-mode and solids-mode Penning ion sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 2.10 Ar, N2, and CH4 ion beam mass spectra . . . . . . . . . . . . . . . . 56 2.11 SiH4 ion beam mass spectrum . . . . . . . . . . . . . . . . . . . . . . 58 2.12 SiH4 ion beam mass spectrum showing shoulder peaks . . . . . . . . . 61 2.13 Ion beam mass spectra of chemical contaminants acquired when using SiH4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 2.14 SiH4 ion beam mass spectra showing the transition from the low pres- sure mode to the high pressure mode . . . . . . . . . . . . . . . . . . 65 2.15 Ion beam mass spectrum of Si ions generated from Si cathodes while using an Ar working gas . . . . . . . . . . . . . . . . . . . . . . . . . 66 2.16 Average ion energy measurement for Ar ions using a roll-off curve . . 68 2.17 Ion energy dependance on the anode and arc voltages . . . . . . . . . 70 2.18 Ion energy dependance of the relative energy spread, ∆E/Ei . . . . . 71 2.19 Calculated sputter yields for 28Si and 12C ions on a 28Si and a 12C surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72 2.20 Ion beam spot 2D current map . . . . . . . . . . . . . . . . . . . . . 74 2.21 Schematic cross section through the deposition chamber showing the sample location and relevant instruments . . . . . . . . . . . . . . . . 76 2.22 RGA mass spectrum of the deposition chamber at its base pressure . 78 2.23 Photograph and wiring diagram for heating a Si(100) substrate on the manipulator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 2.24 Au-Si eutectic phase diagram . . . . . . . . . . . . . . . . . . . . . . 85 2.25 DH power supply sample heating control curve . . . . . . . . . . . . . 86 2.26 RHEED diffraction pattern of a clean, flash annealed Si(100) substrate 89 3.1 Ne ion beam mass spectrum . . . . . . . . . . . . . . . . . . . . . . . 95 3.2 22Ne implantation depth for different ion energies based on TRIM . . 97 3.3 Optical micrograph of a 22Ne-implanted sample produced at room temperature at LC–2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 3.4 SIMS “depth” profile of a 22Ne-implanted sample produced at room temperature at LC–2 with 0.545 % 20Ne . . . . . . . . . . . . . . . . 100 3.5 CO2 ion beam mass spectrum . . . . . . . . . . . . . . . . . . . . . . 105 3.6 Optical micrograph of a 12C sample deposited at room temperature at LC–2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 3.7 SEM cross-sectional micrograph of a 12C sample deposited at room temperature at LC–2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 3.8 SIMS “depth” profile of a 12C sample deposited at room temperature at LC–2 with 39.2 ppm 13C . . . . . . . . . . . . . . . . . . . . . . . 110 3.9 SIMS “depth” profile of a 12C sample deposited at room temperature at LC–2 with 39.2 ppm 13C on a linear scale . . . . . . . . . . . . . . 112 xi 4.1 Schematic drawings of the ion beam chamber, lens chamber, and deposition chamber experimental setups used to deposit 28Si samples 120 4.2 Enrichment progression timeline of the best 28Si sample enrichments . 124 4.3 SiH4 ion beam mass spectrum for samples deposited at room temper- ature at IC–1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 4.4 SEM cross-sectional micrograph of a 28Si film deposited at room tem- perature at IC–1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 4.5 SIMS mass spectrum of 29Si and 28SiH showing well resolved peaks . 135 4.6 SIMS depth profile for a 28Si sample deposited at room temperature at IC–1 with 1130 ppm 29Si . . . . . . . . . . . . . . . . . . . . . . . 137 4.7 SIMS depth profile for the most highly enriched 28Si sample deposited at room temperature at IC–1 with 9.5 ppm 29Si . . . . . . . . . . . . 140 4.8 SiH4 ion beam mass spectrum for samples deposited at LC–2 . . . . . 148 4.9 Optical micrograph of a 28Si sample deposited at room temperature at LC–2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149 4.10 SIMS depth profile of a 28Si sample deposited at room temperature at LC–2 with 2.02 ppm 29Si . . . . . . . . . . . . . . . . . . . . . . . 151 4.11 SIMS depth profile of a 28Si sample deposited at room temperature at LC–2 with 0.993 ppm 29Si . . . . . . . . . . . . . . . . . . . . . . . 154 4.12 29Si isotope fractions vs. deposition rate for multiple SIMS measure- ments of a sample deposited at room temperature at LC–2 . . . . . . 156 4.13 SIMS depth profile of the most highly enriched 28Si sample deposited at room temperature at LC–2 with 0.691 ppm 29Si . . . . . . . . . . . 157 4.14 HR-TEM cross-sectional micrograph of an amorphous 28Si film de- posited at room temperature at LC–2 . . . . . . . . . . . . . . . . . . 158 4.15 SIMS depth profile measuring C in a 28Si sample deposited at room temperature at LC–2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 161 4.16 SIMS depth profile measuring C in a 28Si sample deposited at room temperature at LC–2 with varying deposition pressure . . . . . . . . 163 4.17 XPS spectra of a 28Si sample deposited at room temperature at LC–2 and a control Si sample . . . . . . . . . . . . . . . . . . . . . . . . . . 166 5.1 RHEED patterns of a Si(100) substrate before and after thermal des- orption of the native SiO2 . . . . . . . . . . . . . . . . . . . . . . . . 187 5.2 STM topography images of clean Si(100) (2×1) surfaces prepared in situ by flash annealing . . . . . . . . . . . . . . . . . . . . . . . . . . 189 5.3 RGA mass spectrum of the deposition chamber during deposition of a 28Si sample at DC–3 . . . . . . . . . . . . . . . . . . . . . . . . . . 193 5.4 SiH4 ion beam mass spectrum for samples deposited at DC–3 using the low pressure mode . . . . . . . . . . . . . . . . . . . . . . . . . . 194 5.5 SiH4 ion beam mass spectrum for samples deposited at DC–3 using the high pressure mode . . . . . . . . . . . . . . . . . . . . . . . . . . 196 5.6 Optical micrographs of three 28Si samples deposited at DC–3 . . . . . 198 5.7 SIMS depth profile of a 28Si sample deposited at room temperature at DC–3 with 0.58 ppm 29Si . . . . . . . . . . . . . . . . . . . . . . . 203 xii 5.8 SIMS depth profile of a 28Si sample deposited at 249 ◦C at DC–3 with 0.79 ppm 29Si . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 204 5.9 SIMS depth profile of a 28Si sample deposited at 610 ◦C at DC–3 with 300 ppb 29Si . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207 5.10 SIMS depth profile of a 28Si sample deposited at 712 ◦C at DC–3 with 132 ppb 29Si . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209 5.11 SIMS depth profile of the most highly enriched 28Si sample deposited at 502 ◦C at DC–3 with 127 ppb 29Si . . . . . . . . . . . . . . . . . . 212 5.12 SIMS depth profile of a 28Si sample deposited at 812 ◦C at DC–3 with 4.32 ppm 29Si . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216 5.13 29Si concentration calculations in 28Si from isotope diffusion during deposition at 700 ◦C to 900 ◦C . . . . . . . . . . . . . . . . . . . . . . 219 5.14 SEM cross-sectional micrograph of the surface roughness of a 28Si film deposited at 708 ◦C at DC–3 . . . . . . . . . . . . . . . . . . . . . . . 222 5.15 SIMS depth profile of a 28Si sample deposited at 808 ◦C at DC–3 with 3.97 ppm 29Si compared to a sample deposited at 812 ◦C . . . . . . . 224 5.16 SIMS depth profile of 28Si sample deposited at 421 ◦C at DC–3 using the high pressure mode with 0.303 ppm 29Si . . . . . . . . . . . . . . 227 5.17 TEM cross-sectional micrograph of a Si film showing the epitaxial- to-amorphous transition of low temperature epitaxy . . . . . . . . . . 231 5.18 Si ion beam deposition epitaxy phase diagrams of T vs. Ei . . . . . . 236 5.19 RHEED pattern of a rough crystalline surface of a 28Si sample de- posited at 708 ◦C at DC–3 . . . . . . . . . . . . . . . . . . . . . . . . 239 5.20 RHEED patterns showing diffraction due to microfacets on rough surfaces of 28Si samples deposited at 610 ◦C and 705 ◦C at DC–3 . . . 241 5.21 STM topography image of the rough surface of a 28Si sample de- posited at 708 ◦C at DC–3 . . . . . . . . . . . . . . . . . . . . . . . . 244 5.22 SEM tilted micrograph of the rough surface morphology of a 28Si film deposited at 708 ◦C at DC–3 . . . . . . . . . . . . . . . . . . . . . . . 245 5.23 SEM top-down micrographs of the surface morphology variation of 28Si films deposited between 610 ◦C and 1041 ◦C at DC–3 . . . . . . 247 5.24 STM topography images of pits formed at contaminants on a 28Si film deposited at 709 ◦C at DC–3 . . . . . . . . . . . . . . . . . . . . . . . 256 5.25 STM topography images of pits formed from contaminants in a nat- ural abundance Si film deposited at 713 ◦C at DC–3 . . . . . . . . . . 258 5.26 RHEED pattern of SiC contamination on a flash annealed Si(100) substrate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262 5.27 STM topography image of SiC clusters on a Si(100) substrate after an HF etch and 900 ◦C anneal . . . . . . . . . . . . . . . . . . . . . . 265 5.28 STM topography images of contaminant clusters on Si(100) sub- strates after an HF etch and flash annealing . . . . . . . . . . . . . . 268 5.29 STM topography images showing metal contamination on Si(100) substrates after flash annealing . . . . . . . . . . . . . . . . . . . . . 271 5.30 Mound size vs. deposition temperature of rough 28Si films deposited above 600 ◦C . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 276 xiii 5.31 Arrhenius plot of ln(A), the natural log of the mound area, vs. inverse deposition temperature for rough 28Si samples deposited above 600 ◦C 279 5.32 STM topography images of pits formed from contaminants on 28Si and natural abundance Si films deposited at 712 ◦C at DC–3 after CMOS cleaning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284 5.33 STM topography images of contaminant clusters on Si(100) sub- strates after CMOS cleaning and flash annealing . . . . . . . . . . . . 287 5.34 SEM top-down micrograph of the rough surface morphology of a 28Si film deposited after CMOS cleaning at 705 ◦C at DC–3 . . . . . . . . 289 5.35 RHEED pattern of a 28Si sample with a smooth crystalline surface with islands deposited at 357 ◦C at DC–3 . . . . . . . . . . . . . . . . 293 5.36 RHEED pattern variation of 28Si samples deposited between 249 ◦C and 920 ◦C at DC–3 . . . . . . . . . . . . . . . . . . . . . . . . . . . 295 5.37 STM topography images of islands on the smooth surface of a 28Si sample deposited at 357 ◦C at DC–3 . . . . . . . . . . . . . . . . . . 298 5.38 STM topography images of islands on the smooth surfaces of 28Si samples deposited at DC–3 at low temperature . . . . . . . . . . . . 301 5.39 STM topography images of the surface morphology variation of 28Si samples deposited between 249 ◦C and 804 ◦C at DC–3 . . . . . . . . 304 5.40 XPS spectra of a 28Si sample deposited at 812 ◦C at DC–3 . . . . . . 310 5.41 SIMS depth profile measuring 22 potential contaminants in a 28Si film deposited at 460 ◦C at DC–3 . . . . . . . . . . . . . . . . . . . . . . . 312 5.42 Phase diagrams for the N-Si, C-Si, and O-Si systems near the solid solubility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314 5.43 SIMS depth profile measuring N, C, and Cl in a natSi film deposited at DC–3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319 5.44 RGA mass spectrum of the deposition chamber during substrate de- gassing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321 5.45 SiH4 ion beam mass spectrum for N-contaminated 28Si samples de- posited at DC–3 using the low pressure mode . . . . . . . . . . . . . 323 5.46 SIMS depth profile measuring N, C, and O contaminants in a 28Si sample deposited at 712 ◦C at DC–3 . . . . . . . . . . . . . . . . . . 325 5.47 SIMS depth profile measuring contaminants in a 28Si film deposited at 460 ◦C at DC–3 after experimental improvements . . . . . . . . . . 329 5.48 N, C, and O atomic concentrations of 28Si samples vs. deposition ion current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333 5.49 TEM cross-sectional micrographs showing facets and stacking faults of a rough 28Si sample deposited at 708 ◦C at DC–3 . . . . . . . . . . 337 5.50 HR-TEM cross-sectional micrograph of a rough 28Si sample deposited at 708 ◦C at DC–3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339 5.51 TEM cross-sectional micrographs at two magnifications of a smooth 28Si sample deposited at 460 ◦C at DC–3 using the low pressure mode 341 5.52 HR-TEM cross-sectional micrograph of a smooth 28Si sample de- posited at 460 ◦C at DC–3 using the low pressure mode . . . . . . . . 343 xiv 5.53 HR-TEM cross-sectional micrograph of a smooth 28Si sample de- posited at 421 ◦C at DC–3 using the high pressure mode . . . . . . . 345 5.54 Isotope reduction timeline of the isotope reduction factors of the best 28Si sample enrichments . . . . . . . . . . . . . . . . . . . . . . . . . 348 6.1 SiH4 CVD surface reaction . . . . . . . . . . . . . . . . . . . . . . . . 355 6.2 SiH4 ion beam mass spectrum measuring the geometric selectivity . . 362 6.3 Ion beam mass spectrum of ThO+ showing a gas scattering peak tail 364 6.4 RGA mass spectra of the deposition chamber prior to and during operation of the ion beam with SiH4 . . . . . . . . . . . . . . . . . . 367 6.5 Gas sticking deposition model cartoon for 28Si deposition . . . . . . . 371 6.6 Room temperature correlation plot of isotope fraction vs. SiH4 flux ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374 6.7 Room temperature correlation plot of isotope fraction vs. SiH4 flux ratio including high pressure 28Si samples deposited at IC–1 . . . . . 377 6.8 29Si/30Si isotope ratios . . . . . . . . . . . . . . . . . . . . . . . . . . 378 6.9 Adjusted isotope fraction, cz(sT , k502), vs. temperature . . . . . . . . 381 6.10 High temperature correlation plot of converted isotope fraction vs. SiH4 flux ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 384 6.11 s vs. deposition temperature . . . . . . . . . . . . . . . . . . . . . . . 386 6.12 Arrhenius plot of ln(s) vs. inverse deposition temperature . . . . . . 389 7.1 Isotope reduction timeline of the isotope reduction factors of the best 22Ne, 12C, and 28Si sample enrichments . . . . . . . . . . . . . . . . . 394 7.2 Enrichment progression timeline of the best overall sample enrich- ments with references . . . . . . . . . . . . . . . . . . . . . . . . . . . 396 7.3 SIMS depth profile of a Si isotope heterostructure deposited at LC–2 401 7.4 Si sublimation rates for a 28Si chip and calculated values . . . . . . . 405 7.5 SIMS depth profile of an Al delta layer in Si . . . . . . . . . . . . . . 406 7.6 Cartoon schematic of a 28Si capacitor . . . . . . . . . . . . . . . . . . 408 A.1 Ion beamline lens element circuit diagram . . . . . . . . . . . . . . . 411 A.2 Ion source operating parameter scans . . . . . . . . . . . . . . . . . . 412 B.1 Photographs of the ion beam deposition system . . . . . . . . . . . . 416 B.2 Photograph of the gas manifold . . . . . . . . . . . . . . . . . . . . . 417 B.3 Photographs of the ion source elements . . . . . . . . . . . . . . . . . 417 B.4 Ion beamline electrostatic elements and magnetic sector mass analyzer418 B.5 Photograph of the mass-selecting aperture used for 22Ne and 12C sam- ples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419 B.6 Photographs of the mass-selecting aperture used for 28Si samples at IC–1 and LC–2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 419 B.7 Photograph of the mass-selecting aperture used for 28Si samples at DC–3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420 B.8 Photograph of a gas aperture on the inlet to the deceleration lenes in the ion beam . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 420 xv B.9 Photographs of the experimental setup and sample stage for 28Si sam- ples deposited at IC–1 . . . . . . . . . . . . . . . . . . . . . . . . . . 421 B.10 Photograph of the experimental setup of the lens chamber for samples produced at LC–2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 421 B.11 Photographs of the sample stage for producing sample at LC–2 . . . 422 B.12 Photographs of the manipulator at DC–3 . . . . . . . . . . . . . . . . 423 B.13 Photographs of interchangeable sample apertures at DC–3 . . . . . . 424 xvi Chapter 1 Introduction 1.1 28Si for Quantum Information 1.1.1 Si-Based Solid State Quantum Information Major technological advances are often driven by or require development of new or improved materials [1]. Development of visible light LEDs and lasers in the 1960s was possible due to improvements in GaAs crystals with engineered compo- sitions. Carbon-based materials have played an important role in both manufac- turing advances and basic physics research. Carbon fibers and carbon reinforced plastics were first developed in the 1960s and have found a wide variety of man- ufacturing applications as strong, lightweight materials. At the microscopic level, carbon nanotubes have spurred a tremendous amount of research into their unique and impressive physical and electrical properties with numerous potential applica- tions. Although they were initially discovered in 1952, interest in carbon nanotubes increased after further observations in the 1990s. The widespread use of small, high- energy density, rechargeable batteries for portable electronics was made possible by 1 the development and engineering of lithium ion-based electrodes such as LiCoO2 or later LiFeO2 starting in the 1980s. Another major materials-based technology that has made portable electronics and, more broadly, all modern computers possible is the transistor. Developed initially in the 1950s, the ubiquitous use of semicon- ductor transistors as the base computing component in integrated circuits in the microelectronics industry first required the engineering of extremely high purity, single-crystalline silicon material. Further advances in silicon processing for metal- on-semiconductor transistor technology would come to be guided on a large scale by the International Technology Roadmap for Semiconductors. The engineering of the properties of silicon has also been integral to the development of photovoltaic solar cells since the 1970s and silicon cells account for the majority of solar panels in use. High-quality, i.e. nearly perfect, silicon has perhaps been the most important material to the modern and increasingly computer oriented world. In solid state quantum information (QI), isotopically enriched 28Si is a critical material for the further development of Si-based quantum computing architectures. The abundance of high-quality (unenriched) Si and the established microelectronics infrastructure make Si an attractive medium for quantum computing, which holds the promise of significant increases in computing speed for certain tasks over clas- sical computers. For quantum coherent devices, both the classical aspects of device operation and the states of the qubits (quantum bits) utilize the electronic band structure of crystalline semiconductors like Si. A great deal of experimental QI re- search has leveraged these advantages with a high degree of success, as described by Zwanenburg et al. in a review article [2]. Si-based quantum computing architectures 2 include electron spin qubits in quantum dots defined electrostatically by gate elec- trodes. Quantum dots can be formed in a variety of structures including undoped Si surfaces, Si nanowires, and Si quantum wells in Si/SiGe heterostructures. Another popular quantum computing architecture is based on a proposal by Kane to use the nuclear spins of an array of single donor 31P atoms in Si as qubits [3]. In this design, metal “A-gates” on the surface manipulate the individual donor spins and they interact via electron-mediated coupling, which is controlled by “J-gates”. The proposal makes clear that for this scheme to be successfully implemented in a quantum computer, a host material for the donors needs to be free of nuclear spins (I = 0). This basic design principle of gate-controlled donors and dots has been adapted to produce architectures including implanted 31P atoms in Si transistor devices or in combination with quantum dots and 31P single atom transistors or quantum dots in Si produced by scanning tunneling microscopy H-lithography. A schematic of Kane’s architecture as well as examples of two physical implementations of these types of Si quantum computing devices are shown in Fig. 1.1. Panel (a) is the Kane schematic from Ref. [3] showing the relation between the 31P donors and the surface control gates. Panel (b) is a false colored top-down scanning electron microscope (SEM) micrograph of a Si quantum dot device from an experiment in Ref. [4] defined by electrostatic metal gates on a 28Si epilayer. The location of the quantum dot is shown by the representation of the cartoon electron spin (circle with arrow). A single electron transistor (SET) is formed by the gates at the bottom of the micrograph. A transmission line for sending microwave pulses to the dot to manipulate the qubit spin states is seen at the left. A schematic cartoon of the quantum dot system is 3 Figure 1.1: Si-based quantum dot and P-donor devices. (a) Schematic for a pro- posed quantum computing architecture for 31P donors in Si. 31P nuclear spin qubits are manipulated by metal “A-gates” and interact via their electrons, controlled by “J-gates” (from Ref. [3]). (b) False colored SEM micrograph of a Si quantum dot (cartoon arrow) defined electrostatically by metal gates on 28Si epilayer. Gates defin- ing a single electron transistor are seen below the quantum dot and a transmission line for ESR pulses is at the left. A cartoon schematic of the quantum dot is at the right (from Ref. [4]). (c) STM image of a H-terminated Si(100) surface with areas selectively depassivated for 31P dosing that will define electric gates and a quantum dot with single or multiple 31P donors (from Ref. [5]). shown at the right. Panel (c) is a scanning tunneling microscope (STM) image of a H-terminated Si(100) surface from Ref. [5] with electrostatic gates and a dot defined by selectively removing H in the bright and outlined areas of the image. These areas of bare Si substrate will be dosed with 31P atoms, producing conductive wires and a dot formed of one or several 31P atoms. Interest in 31P nuclear and electron spins as qubits (or memory) in Si has also 4 spurred research in electron spin resonance (ESR) and nuclear magnetic resonance (NMR) of 31P spin ensembles in Si crystals. One of the key performance metrics of spin qubits, which is measured in bulk ESR experiments, is their dephasing or coher- ence time, T2. This is the time that coherent spins comprising the qubit maintain their quantum phase before the information they encode is lost to the environment. For a continuous wave ESR measurement, T2 is inversely related to the ESR signal linewidth. For a viable quantum computer, the T2 time must be sufficiently long as to be approximately 1× 106 times longer than the average single gate operation time to account for dephasing errors. This is a general design rule which depends on a number of factors including use of error reduction codes. Two qubit gate op- erations with gate times of the order of 100 ns have been demonstrated [6]. If the spin interacts with local inhomogeneities in the magnetic field, enhanced dephasing will occur and the resulting coherence time is denoted as T∗2 . Certain manipula- tions of the spins using specific pulse sequences such as a Hahn echo or dynamical decoupling can reduce the effects of dephasing due to environmental noise and will result in a T2 echo signal measurement. The dephasing of spins in an ensemble is measured, for example, by projecting the spin states as spin-up (aligned to an ex- ternal magnetic field) or spin-down (anti-aligned to an external magnetic field) and the probability of the spin being in those projected states over time gives the decay characterised by T2. Other more complicated projections to states not aligned to the magnetic field are also possible. Figure 1.2 shows a cartoon schematic of the Bloch sphere construction for a spin qubit as well as a schematic of a T2 determi- nation from measurements of such a spin. Panel (a) shows the Bloch sphere for a 5 Figure 1.2: Schematic diagrams of the Bloch sphere for a qubit spin and the T2 coherence time due to spin dephasing (a) The Bloch sphere for two level quantum system. The state of an electron spin qubit in a semiconductor quantum dot exists as a vector on the sphere and is projected as spin-up or spin-down. A “σx” optical control pulse rotates the qubit spin around the x-axis to perform gate control opera- tions (from Ref. [7]). (b) Diagram of a T2 coherence time determination showing the decay of measurements of the spin-up probability after gate control operations. The T∗2 decay is caused by the dephasing of the spin ensemble (adapted from Ref. [8]). two level quantum system from Ref. [7]. The state of an electron spin qubit in a semiconductor quantum dot exists as a vector on the sphere and is projected by a measurement onto a basis state such as spin-up or spin-down. ESR or NMR control pulses rotate the qubit spin around an axis, e.g. the “σx” rotation shown in panel (a), to perform gate control operations. When the spin vector with a basis state aligned with the magnetic field in the positive z direction lies in the x-y plane, the 6 strongest dephasing can occur. Panel (b) shows how T∗2 is determined from many measurements of a spin where the spin-up population oscillates with varying spin evolution time, adapted from Ref. [8]. The spin evolution when the spin is dephas- ing is shown for three representations of the Bloch sphere at the top of the figure. The spin-up oscillations exhibit a certain decay related to T2 and environmental inhomogeneities. In Si, a significant source of inhomogeneous magnetic field noise is the Over- hauser field generated by nuclear and impurity spins in the crystal. Natural abun- dance Si is comprised of three stable isotopes, 28Si, 29Si, and 30Si, which have abun- dances of approximately 92.2 %, 4.7 %, and 3.1 %, respectively. The 29Si isotope has a nuclear spin I = 1/2, while 28Si and 30Si isotopes have no net nuclear spin. By eliminating 29Si nuclei, pure, isotopically enriched 28Si becomes an ideal spin-free environment in which to place the electron and nuclear spins of qubits. Without a randomly fluctuating global Overhauser field present, spins in 28Si interact with their environment far less than in unenriched material leading to a greatly enhanced T∗2 coherence time. Consequently, 28Si has been dubbed a “semiconductor vacuum” and is analogous to the isolation of trapped atoms in a vacuum chamber [9]. A cartoon model depicting the composition of 28Si and natural abundance Si (natSi) is shown in Fig. 1.3, adapted from Ref. [10]. Theoretical modeling and bulk ESR experiments predicted the enhancement in T∗2 to be proportional to the reduction in 29Si concentration [11,12], which further spurred interest in exploiting 28Si experimentally. Numerous research groups have shown through bulk ESR and NMR experiments of 31P spins in 28Si that nuclear 7 Figure 1.3: Cartoon depictions of natural abundance Si and enriched 28Si crystals. (a) Natural abundance Si contains 29Si atoms that possess nuclear spins as well as nuclear spin-free 28Si and 30Si atoms. (b) The 28Si is free of nuclear spins. (adapted from Ref. [10]) and electron spin coherence T2 times can exceed seconds [9, 10, 13, 14]. A recent measurement of the T2 time of a single 31P electron spin resulted in a value of 559 ms [10]. Two examples of T2 measurements of spins in 28Si for both bulk NMR and in the single spin regime are presented in Fig. 1.4. Panel (a) shows the measurement results from Ref. [10] mentioned above for a single 31P atom implanted into a 28Si quantum dot device. The dynamic decoupling pulse sequence used for this measurement is referred to as CPMG [15]. The 31P nuclear spin measured in this experiment was determined to have a T2 = 35.6 s. Those measurements were done at cryogenic temperatures. Panel (b) shows a T2 determination measurement from Ref. [14] of the nuclear spins of an ensemble of 31P atoms in 28Si in a bulk NMR experiment. When using a so-called XY-16 decoupling pulse sequence [16], the measurement resulted in a coherence time of T2 = 39 min, measured at room 8 Figure 1.4: T2 coherence time measurements of spins in 28Si. (a) An electron spin of a single 31P atom in 28Si has a T2 = 559 ms after dynamical decoupling (from Ref. [10]). (b) An ensemble of 31P atoms in 28Si have a nuclear spin T2 = 39 min at room temperature (from Ref. [14]). temperature. Quantum computing architectures that stand to benefit from or already have demonstrated benefits from using enriched 28Si include STM hydrogen lithography Si:P devices [17–20], single dopant qubits implanted near SETs [2,21–23], 28Si quan- tum wells in Si/Ge heterostructures [24,25], and fabrication of transistors (FinFETs) for QI [26]. A few of these groups have shown both long T2 times and coherent manipulation in 28Si for bulk donor spins [27] as well as single spins in quantum wells [24] and quantum dots [10,28]. In addition to long coherence times, using 28Si as a medium for 31P opens up the possibility of optical manipulation of the qubit system through the use of hyperfine transitions, which are unresolvable in natural Si. Typically in solid state QI systems, in the absence of optical addressability, electrostatic control gates are needed in close proximity to the dot to manipulate the qubit states and have the 9 Figure 1.5: Examples of experimental capabilities for qubit control in 28Si (a) 31P in Si energy level diagram for 12 optical transitions between the neutral donor (D0) and the donor bound exciton (D0X) (from Ref. [14]). (b) Photoluminescence spectra (bound exciton no-phonon) of 31P atoms in natural abundance Si and 28Si showing hyperfine splitting only resolvable in 28Si, making optical addressing possible (from Ref. [29]). (c) ESR frequency tuning of a 31P electron spin in 28Si using a control gate voltage to induce a Stark shift. This is due to the very narrow ESR linewidth of spins in 28Si (from Ref. [4]). possibility of introducing charge noise into the system [30]. Additionally, qubit ma- nipulation schemes, which have been proposed for arrays of quantum dot qubits, and which involve tuning the qubit ESR frequency through Stark or Zeeman shifts, have been demonstrated in single quantum dots in 28Si [4]. The ability to tune the qubit ESR frequency relies on qubit spins that have very narrow inhomogeneous ESR linewidths of a few kHz, which are only achievable in a material with homoge- neous mass such as highly enriched 28Si with exceptionally small strain fields [9,31]. 10 Examples of the 31P hyperfine splitting and the tuning of a 31P ESR frequency are shown in Fig. 1.5. Panel (a) shows an energy level diagram for 12 hyperfine-split optical transitions of a 31P in Si from Ref. [14]. The transitions are between the neutral donor (D0) levels and the donor bound exciton (D0X) levels and can be excited with optical control pulses. Panel (b) shows the photoluminescence spectra (bound exciton no-phonon) of 31P atoms in natural abundance Si and 28Si from Ref. [29]. The 12 transitions of the hyperfine splitting are clearly resolvable only in the 28Si sample, which makes it possible to address the spins with the optical pulse in panel (a). Panel (c) shows experimental data from Ref. [4] displaying the change in the ESR frequency of a 31P electron spin in 28Si due to a Stark shift induced by a gate voltage. Multiple qubits can be addressed by one pulse signal using such a technique by bringing only one at a time into resonance with the control pulse. 1.1.2 Sources of 28Si Despite the demonstrated advantages of 28Si for solid state quantum informa- tion, only a very limited amount of highly enriched 28Si is available within the solid state quantum computing community for use in QI experiments. 28Si is primar- ily produced at great cost and effort through international collaborations requiring large centrifuge facilities. This lack of readily available 28Si is one of the prime mo- tivations for using ion beam enrichment and deposition to produce 28Si films in this work. The majority of the 28Si bulk crystals and epitaxial films that have been pro- duced are grown from chemical vapor deposition (CVD) of enriched 28SiH4. The 11 Figure 1.6: Process flow chart for the production of a 28Si bulk crystal at IKZ by 28SiH4 CVD, generated from centrifugation of natural abundance SiF4 at Centrotech. (from Ref. [32]) production of this 28SiH4 has predominately started at the Centrotech facility in St. Petersburg, Russia. There, natural abundance SiF4 is enriched using industrial gas centrifuges. The enriched 28SiF4 is then converted chemically into the silane ( 28SiH4) used to grow 28Si crystals at the Leibniz Institute (IKZ) in Berlin, Germany [32]. A process flow chart for this production of 28Si is shown in Fig. 1.6. Production of 28Si early on in the 1990s in these efforts was spurred on by interest in the thermal conductivity of 28Si. Evidence existed that 28Si had a higher thermal conductivity than natural abundance Si. A higher thermal conductivity was advantageous because heat dissipation was a major problem in the microelec- tronics industry at the time. This early 28Si had a 28Si enrichment of about 99.9 %. Measurements of the thermal conductivity of 28Si revealed that it significantly ex- ceeded that of natural abundance Si, but only at cryogenic temperatures [33]. This meant that 28Si was not a viable option to solve the heat dissipation crisis and thus 12 Figure 1.7: Thermal conductivity measurements of natural abundance Si and 28Si as a function of temperature showing a large enhancement in thermal conductivity for 28Si compared to natural abundance Si at low temperatures. (adapted from Ref. [33]) demand for 28Si plummeted. The thermal conductivity measurements of 28Si and natural abundance Si are shown in Fig. 1.7, adapted from Ref. [33]. Multiple 28Si bulk crystals were grown by CVD and zone refinement from enriched 28SiH4 at IKZ. The largest of these crystals were grown as part of the International Avogadro Coordination (IAC), which seeks to use single crystal 28Si spheres as a standard to measure the Avogadro constant, NA, by counting the num- ber of atoms in a kg of 28Si [32,34,35]. This effort is also related to the redefinition of the kg unit using the kg 28Si sphere. The accuracy of the measurements of NA using these spheres relies on them being very nearly perfect 1 kg 28Si spheres. Measure- ments of their properties show that they are almost perfect single crystals with no detectable dislocations. They have chemical impurity concentrations, including C and O, of approximately 5 ×1014 cm−3, and they have a 28Si enrichment of approx- 13 Figure 1.8: 28Si bulk crystals produced by the International Avogadro Coordination. (a) Schematic of the cutting plan to produce two 28Si spheres from a single-crystal boule and (b) a photograph of the 28Si boule (from Ref. [32]). (c) Photograph of one the final 28Si spheres on a weighing apparatus (from Ref. [34]). imately 99.995 % with a residual 29Si isotopic concentration, i.e. the concentration among the three Si isotopes, of approximately 50 ppm (parts per million, equal to the isotopic concentration times 106) [32]. A schematic cutting plan for producing two 28Si spheres from the 28Si boule produced for the IAC is shown in Fig. 1.8 (a) from Ref. [32]. Panel (b) shows a photograph from Ref. [32] of the IAC single-crystal 28Si boule and panel (c) shows a photograph from Ref. [34] of one of the final spheres sitting on a weighing apparatus. An effect of producing this much 28Si is that the pieces leftover after forming the spheres from the boule as well as from other bulk crystals produced at IKZ were then able to be used for research in the QI community. This of course is a 14 limited supply of 28Si and producing more in this manner using centrifuges is both extremely expensive and time consuming. Some enriched 28SiH4 was also acquired by the Isonics Corporation, USA as well as Dr. Kohei Itoh in the early 2000s who collaborated to grow CVD epilayer 28Si films on natural abundance Si substrates. Itoh has also grown bulk 28Si crys- tals from this source. The 28Si enrichment of these materials was measured to be approximately 99.927 % with a residual 29Si isotope concentration of approximately 730 ppm [36, 37]. The epilayers and other material from Itoh have been used by research groups for QI experiments as well [10], although this material is also lim- ited in supply. Some other less abundant sources of 28Si have also been used in QI research although the exact details of those sources are difficult to verify. 1.1.3 Single Spin Measurements in 28Si In addition to there being a general need for 28Si in semiconductor quantum computing research, a specific need exists for material with targeted levels of enrich- ment to map the dependence of T2 on 29Si concentration in the few-spin or single-spin regime. Recent ESR measurements of T∗2 for single 31P spins in 28Si [10, 28] have disagreed with both the theoretical predictions for the same systems made by Witzel et al. [12], as well as each other. These theoretical predictions are shown in Fig. 1.9. The T2 and T ∗ 2 coherence times for quantum dots and 31P electron spin qubits in Si are shown vs. the concentration of 29Si in the system. Solid and dashed lines represent T2 times for 31P-donor and quantum dots, respectfully, i.e. the two archi- tectures shown in Fig. 1.1. The theory predicts that for every order of magnitude 15 Figure 1.9: Predicted electron T2 (solid and dashed lines) and T ∗ 2 (solid and dashed line and symbols) coherence time dependance on 29Si concentration for qubits in Si including quantum dots and 31P-donor atoms. Experiments measuring T∗2 for single 31P electron spins in 28Si represented by the triangle (from Ref. [10]) and diamond (from Ref. [28]) do not agree with the prediction value (open circle). (figure adapted from Ref. [12]) decrease in 29Si concentration, there is roughly an order of magnitude increase in the T2 time. Bulk electron paramagnetic resonance (EPR) and ESR experiments of 31P spins in Si with various enrichments including Si enriched in 28Si to Si enriched in 29Si [11] have agreed with this theoretical work over a large portion of the predicted curve, as discussed by Witzel et al. in Ref. [12]. The dashed line with symbols shows the T∗2 times for a quantum dot, and the solid line with symbols shows the T∗2 times for 31P electron spins in Si. Also shown are the two results from the previ- ously mentioned ESR experiments (triangle and diamond) measuring a 31P electron spin T∗2 in 28Si with residual 29Si isotopic concentrations of approximately 800 ppm. 16 Both of these experiments involve a single 31P atom implanted in a 28Si SET device, which is used for readout, and manipulated with a nearby ESR line. One result by Muhonen et al. [10] gave a value of T∗2 = 268 µs, and the result by Tracy et al. [28] gave a value of T∗2 = 18 µs, which was believed to be limited by experimental non- idealities, e.g. magnetic field noise. The value predicted by the theory of T∗2 ≈ 2 µs is highlighted by the open circle, which lies below both experimental results. Although the experimental results give longer and thus more desirable coherence times than the theory, this outcome shows the fundamental mechanisms limiting coherence at the single-spin level require further study. In order for the field of solid state quantum computing, especially utilizing 31P or other donor spin qubits, to continue to progress, the effects of nuclear spins such as 29Si isotopes near a single qubit spin needs to be better understood. This requires further measurements of coherence times in the few or single-spin regime with varying concentrations of 29Si in the host material around the qubit atoms. The goal of such a measurement would be to recreate the 31P T∗2 curve from Fig. 1.9 with additional experimental data and compare it to the existing theoretical curve. 28Si material with 29Si concentrations as low as 1 ppm would make the measure- ment more robust and complete. The ability to measure single spins and predict coherence times may ultimately be required for a viable quantum computer. This concept is highlighted by a schematic ESR measurement phase space diagram of 29Si concentration vs. number of measured spins, Ns, in Fig. 1.10. Ovals represent ESR measurements that have been demonstrated already in QI research. These include bulk EPR experiments on a large number of donor spins (e.g. > 1010) for a range 17 Figure 1.10: ESR measurement phase space diagram of 29Si concentration vs. num- ber of measured 31P spins, Ns, in Si. Ovals represent ESR measurements that have been made in the QI field including bulk ESR (i.e. > 1010 spins) for 31P in Si with a range of enrichments (29Si concentrations), a large number of ESR measurements of hundreds or thousands of 31P in natural abundance Si, and single spin ESR mea- surements in natural abundance Si. Only two single spin measurements are known to have been made of 31P in 28Si, represented by the star. These two measurements are the two shown in Fig. 1.9. For a viable quantum computer, it is probable that single spin measurements in highly enriched 28Si is necessary, represented by the shaded region of the phase space. of 29Si concentrations down to approximately 800 ppm [11, 27, 38], represented by the large oval on the right. Bulk ESR experiments have also been done on 28Si with an approximately 50 ppm 29Si concentration [39]. Experiments on smaller numbers of donor spins ranging from 100s to 1000s have also been done but only in natural abundance Si [40,41], which is represented by the oval at the top. In the single spin regime, a large number of QI experiments have measured 31P spin coherence times in natural abundance Si, but only the two previously mentioned experiments have used enriched 28Si with a residual 29Si concentrations of approximately 800 ppm, 18 represented by the star. These two measurements are the two shown in Fig. 1.9 from Ref. [28] and [12]. The far bottom left corner of the phase space is where it is suspected that measurements required for a viable quantum computer will reside when considering scale-up, represented by the shaded region. It is this region of single 31P spins in 28Si with 1 ppm 29Si concentrations where further research is required. Enabling such measurements through production of highly enriched 28Si epitaxial films with targeted enrichments is another aspect of the goals of this work. 1.2 Ion Beam Enrichment and Deposition Si thin film epitaxial deposition can proceed by several techniques including CVD using SiH4, molecular beam epitaxy (MBE) by thermal evaporation, ion as- sisted deposition (IAD) which uses a separate source of ions to enhance deposition, and direct ion beam deposition or ion beam epitaxy (IBE). IBE has two advantages over the other methods. First, the energy of the ions can be used to enhance the deposition process leading to higher quality epitaxy [42, 43]. Typically, hyperther- mal energies below 200 eV are used, and this will be discussed further in Chapter 5. Second, ions can be mass filtered in a magnetic field to select a single isotope of an atom for deposition. This means that ion beam deposition can be used to isotopically enrich a material during the deposition process itself starting from a natural abundance source. This is referred to here as in situ enrichment because the enrichment occurs along the flight path of the ions from the source before being deposited on a substrate to grow a film of enriched material, e.g. 28Si. Mass sepa- 19 rated ion beam deposition and epitaxy is the technique used in this work to produce highly enriched 28Si films. One of the earliest and most well known uses of ion beam enrichment was in the calutron mass spectrometer developed by Ernest Lawrence for the the Manhat- tan Project in the United States in the 1940s [44]. The calutron was used to generate enriched quantities of the isotope 235U for use as fissile material in the development of nuclear weapons. This was accomplished by ionizing natural abundance U con- taining over 99 % 238U, accelerating the ions using electric fields, and then deflecting the ion trajectory using a magnetic field as in a mass spectrometer. The magnetic field separates the ions by mass, which generates an enriched ion beam of 235U that is collected at a target. In order for this process to produce significant quantities of enriched 235U, large, industrial scale apparatus were required, which were produced at great cost and effort during the Manhattan Project. Laboratory scale ion beam deposition systems have been studied and developed since the early 1970s. Fair developed an ion beam deposition system and demon- strated deposition of thin In films with energies between 100 eV and 500 eV [45]. Around the same time, Aisenberg and Chabot demonstrated deposition of diamond- like thin films at room temperature using a beam of C atoms with 40 eV of energy, which was transferred into the film to enhance the deposition [46]. Neither of these early experiments involved mass selecting the ions. Later, in the 1980s and 1990s, a number of other groups began experimenting with mass separated ion beam deposi- tion and IBE. Shimizu et al. developed an ion beam deposition capable of producing mass separated ion beams with mA level current and ion energies down to 10 eV. 20 Ar ion beams and Ca deposition was demonstrated [47]. Various groups have also demonstrated enrichment and ion beam deposition of materials significant to quantum computing research including semiconductors such as Si and Ge, for production of enriched Si/SiGe heterostructures and quan- tum wells, and C, for nitrogen-vacancy centers in enriched diamond. Herbots et al. demonstrated deposition of both 30Si and 74Ge ions with energies of 40 eV at deposition temperatures including 400 ◦C [48]. Zalm and Beckers deposited mass separated 28Si ions on both Si and Ge substrates [49], and Yagi et al. likewise de- posited 28Si as well as 74Ge films with energies of 100 eV at deposition temperatures of 300 ◦C [50]. 28Si IBE was extensively studied by several groups including by Tsubouchi et al. [43] and Rabalais et al. [51] who both developed dual ion beam deposition systems for single and compound enriched materials [52, 53]. 28Si epitaxial deposition was achieved by both groups using ions with energies of typically 20 eV at very low deposition temperatures of 100 ◦C to 400 ◦C. These parameters produced epitaxial thin films of 28Si with low defect densities. The highest 28Si enrichment reported by Rabalais et al. was a film enriched to approximately 99.99 %. Tsubouchi et al. reported a 28Si sample with a higher enrichment of approximately 99.9982 % with a residual 29Si isotopic concentration of approximately 16 ppm, which is more highly enriched than the material produced for the IAC. An isotope measurement of this enriched 28Si sample as well as a schematic of the ion beam system used for deposition are shown in Fig. 1.11. Panel (a) shows a dual source ion beam deposition system schematic drawing from Ref. [53]. Two ion sources are seen connected to 21 Figure 1.11: Ion beam enrichment by Tsubouchi et al. used for depositing 28Si. Panel (a) Schematic drawing of a dual source ion beam deposition system for en- riching and depositing single or multi-component materials (from Ref. [53]). Panel (b) Isotope depth profile of a 28Si film deposited from the ion beam. The isotope concentrations of 28Si, 29Si, and 30Si are shown vs. the depth from the surface into the film and Si substrate. The 28Si film is clearly enriched with the 29Si and 30Si concentrations reduced to below 1018 cm−3, or approximately 16 ppm for 29Si (from Ref. [43]). two 90◦ magnetic isotope separators that both feed into a deposition chamber. This system can be used to enrich and deposit single or multi-component materials. Panel (b) shows a depth profile measurement of the enrichment of a 28Si film deposited with the system in (a) from Ref. [43]. The concentrations of isotopes 28Si, 29Si, and 30Si are shown vs. the depth from the surface into the film and Si substrate. 29Si and 30Si concentrations in the 28Si film are reduced to below 1018 cm−3, resulting in the aforementioned enrichment values. Rabalais et al. have also used their ion beam system to implant 74Ge ions in SiO2 to create enriched Ge quantum dots [54], and deposited 28Si16O2 using the dual beam setup with 28Si and 16O [55]. Tsubouchi et al. have used the dual beam 22 system to deposit enriched compounds including 28Si12C and 12C14N [56,57]. Finally, the work presented here relies heavily on the previous work done using this same ion beam deposition system by the lead researcher of this effort, Dr. Joshua Pomeroy. This system was used to deposit thin films of Cu using mass separated ions and observe the effects of the ion energy on the epitaxial quality of the film using STM [58,59]. 1.3 Objectives and Outline 1.3.1 Project Goals Enrichment and thin film deposition of 28Si is pursued here with the objective of producing high-quality enriched material for solid state quantum computing. 28Si of sufficiently high quality (i.e. high enrichment, crystallinity, and purity) provides an ideal solid state environment to host qubit spins, as discussed previously. Un- wanted deviations from ideal 28Si material can be classified as three types of defects: isotopic defects, structural defects, and chemical defects. Controlling and limiting these defects is critical for successful integration of 28Si into quantum computing architectures. The 28Si materials goals of this work are stated as follows: (1) high enrichment in 28Si with a residual 29Si isotopic concentrations less than 50 ppm, (2) single-crystalline and smooth epitaxial structure with a low dislocation density below 1× 106 cm−3, and (3) high chemical purity including C and O with atomic concentrations below 2× 1015 cm−3. 23 These are believed (but not known) to be the criteria needed for the 28Si to be comparable to single-crystalline electronic grade (EGS) natural abundance Si as well as the enriched Si currently available in the QI research community from the IAC. Electronic grade Si that is purified and crystallized into single-crystalline boules of Si can have dislocation densities below roughly 1× 106 cm−3 [60]. Float zone refinement also produces Si with atomic concentrations of most residual impurities below 5× 1013 cm−3 and atomic concentrations of some impurities such as O below 1× 1018 cm−3 [61]. As mentioned previously, Si produced for the IAC has C and O concentrations below roughly 5× 1014 cm−3. The crystallinity of this material is nearly perfect with no detectable dislocations and a vacancy related defect density of roughly 3× 1014 cm−3 [35]. Additionally, the IAC 28Si has a residual 29Si isotopic concentration as low as 50 ppm. Producing 28Si with 29Si isotopic concentrations as low as 1 ppm is necessary to enable a robust and systematic study measuring electron coherence times vs. 29Si concentration in the single spin regime and compare it to the theoretical predictions discussed previously. Ultimately, the goal of this work is to produce 28Si material that can answer the question “how good is good enough?” for quantum information. This means determining the levels of enrichment, purity, and crystallinity that are necessary to satisfy the materials needs of QI and solid state quantum computing devices. These goals will be achieved using processing methods that are both common (e.g. vacuum deposition, sample heating) and fairly unique (e.g. mass selected ion beam deposition) to engineer the properties, such as enrichment in 28Si and chemical purity, and structure (crystallinity) of Si thin films. 24 1.3.2 Strategy and Impact of Results The experimental strategy used to achieve the materials goals described above for this work is to use an ultra-high vacuum (UHV) ion beam deposition system to prepare and deposit 28Si thin films several hundred nm in thickness. A hyperther- mal energy ion beam is used to achieve in situ enrichment to very high levels from a natural abundance silane gas (SiH4) source and deposit it on Si(100) substrates. Clean substrates are prepared in situ in a UHV environment for minimal incorpo- ration of chemical impurities and heated during deposition to facilitate epitaxial deposition. Proof of principle experiments enriching Ne and C are used to establish experimental techniques. Characterization methods used to support this effort include in situ analy- sis of the surface and crystallinity of 28Si films by reflection high energy electron diffraction (RHEED) and STM. These are used for quick feedback to fine tune the deposition parameters. Ex situ characterization used to assess the quality of the films includes secondary ion mass spectrometry (SIMS) to analyze the enrichment as well as the chemical purity, x-ray photoelectron spectroscopy (XPS) to determine chemical purity, SEM to inspect the surface morphology, and transmission electron microscopy (TEM) to inspect the crystallinity of the films. These characterization methods are used for feedback on the deposition process to make informed decisions on experimental changes leading to higher quality 28Si films. The main impact on the QI field of this work is that 28Si was produced with extremely high levels of enrichment. The most highly enriched sample produced 25 Figure 1.12: Enrichment progression timeline. A timeline of the progression of samples with the lowest residual isotope fractions of 20Ne (diamonds), 13C (circle), 29Si (squares), and 30Si (triangles), as measured by SIMS. These were achieved for 22Ne, 12C, and 28Si samples produced over approximately five years. by this work had an overall enrichment in 28Si of 99.9999819(35) % with a residual 29Si isotopic concentration of 127(29) ppb (parts per billion, equal to the isotopic concentration times 106). This level of enrichment exceeds that of all other known sources of 28Si by a factor of approximately 125. The enrichment of this sample and other highly enriched samples is seen in the enrichment progression timeline in Fig. 1.12 showing the best isotope fractions, which are a measure of the sample enrichment, measured by SIMS for the minor isotopes 29Si (squares) and 30Si (tri- angles) vs. the deposition date for selected samples produced throughout this work. Isotope fractions of a particular isotope are defined in a SIMS measurement as the 26 detected counts of that isotope divided by the total counts of the measurement. Also shown in this figure are measurements of the isotope fractions of 20Ne (diamonds) and 13C (circles). These measurements were part of initial proof of principle exper- iments producing enriched 22Ne and 12C samples done in preparation for enriching and depositing 28Si. The isotope fractions here are written generally as zX/Xtot., where z refers to the mass number of a particular isotope, e.g. 28 for 28Si, X refers to a particular element, and therefore zX is the counts of a particular isotope in the measurement. Xtot. is the sum of the counts of all isotopes being measured. Uncertainties in the isotope fractions are shown for all samples and are derived from isotope measurements described in later chapters. The samples depicted here in this enrichment timeline are those that had the best enrichments of any sample produced up to that point. In other words, Fig. 1.12 is a timeline of the record enrichments achieved in this work. 1.3.3 Outline ˆ Chapter 2: The experimental apparatus and methods used to produce 28Si thin films are presented. Descriptions of the hyperthermal energy ion beam- line used for in situ isotopic enrichment, as well as descriptions of the UHV deposition and analysis chamber used for substrate preparation and in situ sample analysis are included. ˆ Chapter 3: Initial proof of principle experiments enriching 22Ne and 12C are discussed. 22Ne is implanted into Si while 12C thin films are deposited 27 on Si substrates. Three 22Ne and three 12C samples were produced. SIMS enrichment measurements are presented showing a maximum achieved 22Ne isotope fraction of 99.455(36) % and a maximum achieved 12C isotope fraction of 99.9961(4) %. ˆ Chapter 4: Phase I of 28Si deposition is discussed involving two experimen- tal configurations. First, 28Si thin films are deposited in a proof of principle experiment for Si enrichment producing five 28Si samples out of 61 total 28Si samples produced in this work. Adjustments of the deposition parameters are discussed for improved depositions producing 16 28Si films. SIMS measure- ments of the enrichments and chemical purity are presented. A maximum 28Si isotope fraction of 99.999888(10) % was achieved with a residual 29Si isotope fraction of 0.691(74) ppm. ˆ Chapter 5: Phase II of 28Si deposition is discussed involving an experimental configuration that leverages the capabilities of the full system while depositing 40 28Si samples. SIMS enrichment measurements are presented showing a maximum achieved 28Si isotope fraction of 99.9999819(35) % with a residual 29Si isotope fraction of 127(29) ppb for the most highly enriched sample in this work. New substrate preparation procedures are discussed as is substrate heating to enable epitaxial deposition. RHEED, STM, and SEM analysis of 28Si film morphology and TEM analysis of film crystallinity is presented. Finally, chemical purity analysis of 28Si films by SIMS is presented. Smooth, epitaxial 28Si films were achieved for deposition temperatures between about 28 349 ◦C and 460 ◦C. These samples contain atomic concentrations of N, C, and O slightly below 1 ×1019 cm−3. ˆ Chapter 6: A model describing the adsorption of natural abundance SiH4 into 28Si films during deposition is presented and discussed. SIMS enrichment values and SiH4 partial pressures are correlated using the model to extract a room temperature incorporation fraction, s = 6.8(3)×10−4. The temperature dependance of the sample enrichment is explored and an activation energy for reactive SiH4 adsorption is determined to be Ec = 1.1(1) eV. ˆ Chapter 7: A summary of the main scientific results is presented. Then, experimental proposals enabled by this work are discussed including 28Si sam- ples with targeted levels of enrichment for measuring T2, deposition of enriched 28Si/28Si74Ge quantum well heterostructures, 28Si re-deposition, Al dopant de- vices, and several electrical measurements. 29 Chapter 2 Experimental Apparatus and Methods 2.1 Context 2.1.1 Ultra-High Vacuum Deposition These experiments involving the deposition of 28Si from a mass selected ion beam are conducted primarily in an ultra-high vacuum chamber. This provides the cleanest possible environment to prepare clean, flat surfaces on substrates before deposition. While the ion beam itself is not UHV, a gate valve separates it from the UHV deposition chamber and the bulk of the higher pressures gases generated by the ion source are differentially pumped before reaching the sample position. The surfaces of 28Si samples are also inspected in UHV after deposition, which is critical to prevent contaminants from adsorbing that would obscure the measurements. A top-down schematic of the UHV and ion beamline deposition system is shown in Fig. 2.1. The deposition system consists of four connected but isolated vacuum chambers. The hyperthermal energy ion beamline is pictured at the left. The UHV deposition and analysis chamber is at the right with a load lock branching 30 Figure 2.1: Top-down schematic of the ion beam deposition system including the hyperthermal energy ion beamline pictured at the left, the UHV deposition and analysis chamber at the upper right, the load lock used for sample loading into the deposition chamber at the right, and the scanning tunneling microscope pictured at the bottom. These four sections roughly separate four vacuum environments. A smaller scale schematic showing the full length of the magnetic transfer arm used to move samples from the deposition chamber into the STM is in the lower left. 31 off further to the right for sample loading. Separated from this chamber is the scanning tunneling microscope pictured at the bottom, which is separated from the deposition chamber by a gate valve. Photographs of the system are shown in Fig. B.1 in Appendix B. 2.1.2 Previous Operation This work relies heavily on the previous work done using this same ion beam deposition system by the lead researcher of the broader enriched Si project, of which this work is a part, Dr. Joshua Pomeroy. This system was used by Pomeroy at Cornell University to deposit thin films of Cu using mass separated ions and observe the effects of the ion energy on the epitaxial quality of the film using STM [58,59], as previously mentioned. Additionally, Baumann and Bethge have extensively tested and utilized the same type of Penning ion source used in these experiments [62–64], and one study in particular served as a reference for the operation of the source in this work under various settings [65]. These tests showed the dependence of the ion current on the gas flow, ion source magnetic field, and anode and cathode voltages. Versions of similar ion source parameter tests generated using the ion source in this work are shown in Fig. A.2 in Appendix A. Other references of previous operation of the ion source used here include studies by Nouri et al. [66] and the Handbook of Ion Sources by Wolf [67]. 32 2.2 Hyperthermal Energy Ion Beamline 2.2.1 Apparatus Much of the hyperthermal energy ion beamline used in this work was obtained from Physicon Corporation (MA, USA) including two ion sources, the first section of the beamline and housing chamber, and the magnetic sector mass analyzer. The first element of the ion beamline is the Penning-type ion source, also called a Penning Ionization Gauge (PIG) ion source. The principle of gas discharge, which is the basis of this type of ion source, was first investigate by Phillips [68]. The Penning source generates a plasma from a working gas in an electric field generated by a high voltage between the anode and the cathode that is contained radially by a magnetic field. This plasma ionizes and cracks gas molecules that are injected into the ion source. Mostly singly charged ions are then extracted by a high voltage applied to the extractor electrode, VExt, into the beamline. An einzel lens with the focus electrode, VF , focuses and transports the ions at the transport voltage, VT , which is typically about -4 kV. The transport voltage is applied to the entire ion beam vacuum chamber, which is isolated from the chamber frame and ground. This high transport voltage is needed in order to minimize the effect of space charge repulsion of the positively charged beam, which would cause the beam to expand and become unfocused. A schematic of the ion beamline is shown in Fig. 2.2 (from Ref. [59]). This diagram shows the electrostatic elements of the beamline. Inset is a schematic of the gas-mode Penning ion source showing the anode, cathodes, electromagnet, 33 Figure 2.2: Ion beamline schematic showing the electrostatic lens elements used to focus and control an ion beam. Inset is a schematic of the gas-mode Penning ion source showing the anode, cathodes, electromagnet, and extraction cusp (extractor). The gas plasma forms in the anode between the cathodes. The potential energy landscape seen by ions in the beamline is represented at the bottom of the figure with labels A through K referencing the electrostatic elements in the beamline. Ions are created in the source at close to the anode potential (A) and extracted with a large negative potential by the extractor (B). Ions are then focused by an einzel lens (D) and transported at a large negative potential (C). Deceleration lenses (C to K) lower the ion kinetic energy to their starting energy for deposition at a grounded target sample. (from Ref. [59]) and extraction cusp (extractor). At the bottom of the figure is a representation of the potential energy landscape seen by ions in the beamline with labels A through K corresponding to lens elements. A is the anode potential and the potential at which the ions are created, B is the extractor, C is the transport potential, D is the 34 focus, E is an electron suppression (rejection) electrode that has a potential fixed at approximately 100 V more negative than VT . Before the electron suppressor, there are X-Y deflectors for steering the ion beam and a skimmer element with monitoring current, Isk. F through K are the individual elements of the deceleration lenses that focus and slow the ions to deliver them onto the sample, which is at ground potential. Six lens elements are independently tunable and are referred to as lenses A2, A3, B2, B3, B4, and X. A typical voltage applied to some of these lenses is roughly -1 kV. Voltages for the deceleration lenes that were used when depositing a 28Si sample are given in Table A.3 in Appendix A. Additionally, a schematic circuit diagram of the power supplies and wiring of the various lens elements within the beamline is presented there in Fig. A.1. In order to isolate the transport voltage of the ion beam chamber and the cathode voltage applied to the housing of the ion source, an insulating flange is used to connect the two. While most components and flanges comprising the ion beam chamber use UHV seals, the insulating flange uses o-rings to form the vacuum seals. The components of the ion source itself are also sealed using o-rings. This means that these components of the vacuum chamber containing the beamline are only rated to high vacuum, i.e. a minimum achievable pressure of roughly 1.3× 10−7 Pa (1.0× 10−9 Torr). The ion beam chamber is pumped using a 350 L/s turbo pump (Pfeiffer Vacuum). The base pressure of the ion beam chamber ranged from ap- proximately 2.7× 10−6 Pa to 1.3× 10−5 Pa (2.0× 10−8 Torr to 1.0× 10−7 Torr) throughout the work presented here. While initially at the higher end of this range, the base pressure was reduced at one point by removing a gate valve that was also 35 only rated to high vacuum. The components of the vacuum for the first and sec- ond achieved base pressures can be observed using the residual gas analyzer (RGA) in the ion beam chamber. Residual gas mass spectra representative of these base pressures are shown in Fig. 2.3. Base pressure 1 (line) corresponds to the earlier base pressure before removing the gate valve. Base pressure 2 (diagonal line fill) corresponds to the lower achieved base pressure after removing the gate valve. Com- mon vacuum components are observed in both spectra as peaks vs. their mass un u (unified atomic mass unit) including H2, C, N, H2O, CO and N2 at 28 u, O2, and CO2. Ar is also observed due to it being used to vent the chamber previously. When a lower base pressure was achieved, the partial pressures of N at 14 u, N2 and CO at 28 u, and O2 at 32 u were all reduced within the chamber. After the ions are extracted and focused into a beam, they are transported into a re-focusing magnetic sector mass analyzer, which bends the ion trajectories in a 90◦ arc. The sector mass analyzer is a large electromagnet that is used to separate the ions according to their mass-to-charge ratio, m/q. Ions of different mass-to- charge ratios have different resulting trajectories. When the sector mass analyzer is tuned to transmit ions with a particular value of m/q, that value is referred to in discussions and in graphical representations of data only by the mass number, e.g. the mass analyzer is said to be tuned to a mass of 28 u for 28Si ions with m/q = 28 u/e. This is because ions generated from the ion beam and discussed in this work are assumed to be singly charged, unless otherwise noted. This mass analyzer is used to select a particular mass, u, and propagate those ions down the beamline. At the exit of the mass analyzer is the mass-selecting aperture. Nominally, 36 Figure 2.3: Residual gas mass spectra collected from the RGA in the ion beam chamber for two achieved base pressures. Partial pressure peaks of typical compo- nents of the vacuum are seen including H2, C, N, H2O, CO and N2 at 28 u, O2, and CO2. Ar is also observed due to it being used to vent the chamber previously. The initial base pressure 1 (line) corresponds to the use of a gate value only rated to high vacuum, while a later base pressure 2 (diagonal line fill) was achieved after removing that gate valve. The partial pressures of N, N2 and CO, and O2 were reduced with base pressure 2. ions of a single mass pass through the aperture, while those with other masses are rejected. Three mass-selecting apertures were used in the work presented here. Ini- tially, an aperture that consisted of a circular hole approximately 5 mm in diameter in a stainless steel spacer that was 16 mm thick was used to produce 22Ne and 12C samples. Then, a much thinner Cu gasket aperture in the shape of a slit approx- imately 1 mm in width, i.e. the same direction that different mass ion beams are spatially separated, was used for initial 28Si depositions. The aperture was also ap- proximately 15.25 mm tall and 2 mm thick. Finally, a second Cu slit approximately 2 mm in width and 12 mm tall with a beveled slit opening to reduce ion scattering 37 Figure 2.4: Ion beamline schematic showing the diffusion path of natural abundance SiH4 (“clouds”), including 29SiH4, from the ion source inlet (lower left), through the magnetic sector mass analyzer (top left), past the mass-selecting aperture and into the deposition chamber where the sample is located (upper right). The inset cartoon shows a magnification of the aperture area with mass separated Si ions. 29Si ions are blocked by the aperture while the 28Si ions as well as 29SiH4 (and other) gas molecules pass into the depositions chamber. off of the aperture was used to deposit the remaining 28Si samples. Photos of these three mass-selecting apertures are shown in Fig. B.5 to B.7 in Appendix B. Simultaneous with ions passing through the aperture, gas from the ion source diffuses along the beam path and through the aperture as well. This source gas is natural abundance and results in a partial pressure of unwanted isotopes at the sample, which may be incorporated into the sample during deposition. This phe- nomenon is discussed in great detail in Chapter 6. Figure 2.4 shows a schematic 38 of the ion beamline with a cartoon representation of the mass selection and gas diffusion for the case of 28Si deposition. Natural abundance silane gas (SiH4) is used as the source for generating Si ions in this work. This gas has a purity of 99.999 % according to the gas vendor (Matheson Tri-Gas). Gas is injected into the ion source using a UHV leak valve from a gas manifold used to regulate different gases used for operation of the ion source. A photograph of the gas manifold is shown in Fig. B.2 in Appendix B. When depositing 28Si, the natural abundance SiH4 diffuses from the ion source down the beamline and to the sample location in the deposition chamber, as represented by the “clouds” in the figure. The inset shows a magnification of the mass selection process occurring at the mass-selecting aperture. 29Si (and 30Si) ions are blocked by the aperture while the 28Si ions and SiH4 gas molecules pass into the deposition chamber. Beyond the mass-selecting aperture are the deceleration lenses. As mentioned, the purpose of these einzel lenses is to maintain the focus or de-focus of the ions while decelerating them from the transport voltage mentioned above to ground potential at the sample. The final kinetic energy of the ions depends on the voltage at which they were created in the ion source (plasma potential), which is typically similar to the positive voltage applied to the anode. This is discussed in more detail later in this chapter. Photographs of the electrostatic and magnetic elements comprising the ion beamline are shown in Fig. B.4 in Appendix B. 39 2.2.2 Theory of Magnetic Mass Separation When charged ions with different masses but otherwise equal properties tra- verse a region of magnetic field, they will follow circular motion. Each ion of a particular mass will follow different circular trajectories, resulting in a physical sep- aration of the ions by mass. This principle of magnetic mass separation of ions used in mass spectrometers relies primarily on the magnetic Lorentz force, FB, which is the force exerted on a charged particle in a magnetic field. The magnetic Lorentz force is given by FB = qv ×B, (2.1) where q is the charge state of the particle, here an ion, v is the ion velocity, and B is the magnetic field experienced by the ion. The cross product results in a force, FB, acting on the ion in a direction perpendicular to the direction of motion. This force thus results in circular motion of the ion. In order to determine how ions of different masses are separated as a function of the magnetic field in terms of the radii of curvature of their trajectories, a general equation for circular motion is used. Circular motion of any object can be described as being the result of a centripetal force, Fc, which acts on the object in the direction of the center of the circular path. Fc is given by Fc = mv2 r , (2.2) where m is the mass of the object, v is the velocity, and r is the radius of curvature of the circle defining the object’s trajectory. The centripetal force and the magnetic 40 Lorentz force can then be equated, FB = Fc, because they both describe circular motion. This expression is solved to get the radius of curvature of the ion trajectory, r, in terms of the mass, velocity, charge state, and magnetic field, yielding r = mv qB . (2.3) This means that for ions with different masses but equal charge states and velocities, the radius of curvature of the ion’s motion is proportional to its mass. However, for ions with different masses generated in a beamline and accelerated by electric fields, their velocities will not be equal. An expression for the ion velocity in a beamline can be determined from con- sidering the ion kinetic energy, EK , due to acceleration by an electric field generated from an applied voltage. For this scenario, the potential energy, EP , of the ion in the electric field is transformed into its kinetic energy. Thus, setting EK = EP gives the expression mv2 2 = qV, (2.4) with the kinetic energy term on the left side of the equation and the potential energy on the right. V is the accelerating voltage used to generate the ion beam. Then, solving for v yields v = √ 2qV m . (2.5) This expression for v is substituted into Eq. (2.3) yielding a general expression for the mass dependance of the radius of curvature of an ion trajectory given by ri = 1 B √ 2miV q . (2.6) 41 Here, ri is the radius for an ion with index i corresponding to a mass of mi. In other words, Eq. (2.6) shows that for ions with equal charge states of q and accelerated with voltage V that pass through a magnetic field of B, different masses, mi, will follow trajectories with different radii of curvature, ri. It is from this equation that the mass-to-charge ratio, m/q, is seen to be the important parameter for mass separation in an ion mass spectrum. In a mass spectrometer and in the ion beamline used in this work, the radius of curvature of the sector mass analyzer is fixed, giving a single value of the radius for selected (transmitting) ions, r0. Rearranging Eq. (2.6) to get an expression for B and substituting q = 1 e and r0 = 97.9 mm gives B = (1.47× 10−3) √ mV , (2.7) which is the magnetic field in units of T required for selecting ions with a given mass, m, in units of u using a given acceleration voltage, V , in units of V for the beamline used in this work. The quoted value of r0 is an approximation determined from analysis similar to Eq. (2.6) using experimental values. A plot of B vs. m calculated from Eq. (2.7) is shown in Fig. 2.5 (a). This calculation uses a value for the accelerating voltage into the sector mass analyzer of V = 4040 V, which is a typical value for the operation of the beamline in this work. B ranges from 0.09 T at 1 u to 0.84 T at 80 u. For 28Si at 28 u, B = 0.49 T in this calculation. Panel (b) of Fig. 2.5 shows the relation between the current applied to the sector mass analyzer magnet and the resulting mass of the ions being selected for accelerating voltage values of 4040 V (line) and 4740 V (dashed line). An accel- 42 Figure 2.5: Calculated relation of the sector mass analyzer magnetic field and applied current to the selected mass in the ion beam. These calculations are for singly charged ions and use a value for the sector mass analyzer radius of curvature of r0 = 97.9 mm. (a) The magnetic field required to select a certain mass (line) is calculated from Eq. (2.7) as a function of mass. This calculation uses an accelerating voltage V = 4040 V, typical of the operation of the beamline. (b) The selected mass is shown as a function of the current applied to the mass analyzer magnet for accelerating voltages of V = 4040 V (line) and V = 4740 V (dashed line). These curves are calculated from Eq. (2.9), which is an experimentally derived conversion from the calculated field to the current. erating voltage of V ≈ 4740 V is also used for operation of the ion beamline in addition to 4040 V. These curves are determined from the curves in panel (a), by converting the calculated values of B into the applied current based on experimen- tal measurements. B and the applied magnet current, Imag, are linearly related, as given by B = (9.69× 10−3)Imag + (3.43× 10−3). (2.8) The slope and intercept of this linear equation were experimentally determined. Here, Imag is in units of A and B is in units of T. Substituting Eq. (2.8) for B in Eq. (2.7) and solving for m yields an expression relating the selected mass and the 43 applied current given by m = ((6.59)Imag + 2.33) 2 V , (2.9) where, Imag is in units of A, V is in units of V, and m is in units of u. In Fig. 2.5 (b), values of Imag between approximately 10 A and 90 A are needed to select ions with masses from 1 u up to about 80 u. Inverting Eq. (2.9) yields an expression for Imag as a function of m and V given by Imag = (0.15) √ mV − 0.35. (2.10) Again, m is in units of u, V is in units of V, and Imag is in units of A. This equation gives the applied current needed for the mass analyzer in this work to select ions with a particular mass for a given accelerating voltage. The mass analyzer current needed for 28Si at 28 u for V = 4040 V is Imag ≈ 50.1 A. For a mass analyzer with a 90◦ bend, the physical separation, λ1, at a mass- selecting aperture of the trajectories of ions with different masses can be calculated using geometry and trigonometric relations. The two parameters that are required for the calculation are the radius of curvature of the analyzer, r0, and the distance, Y , from the exit of the analyzer to the aperture where mass selection occurs. An expression for λ1 as a function of r0, Y , and two masses is given by λ1 = r0 ({ m1 m0 } 1 2 cos { sin−1 [ 1− ( m0 m1 ) 1 2 ]} − 1 ) + Y tan { sin−1 [ 1− ( m0 m1 ) 1 2 ]} , (2.11) where m0 is the mass of the ions being selected and passing through the aperture and m1 is the mass of another ion not being selected. For this equation, m1 > 44 Figure 2.6: Calculated mass dependance of the spatial separation, λ1, of ions with adjacent mass number in a mass separated ion beam at the mass-selecting aperture. These calculations (line) use values for the physical parameters of the ion beamline including the sector mass analyzer radius of curvature of r0 = 97.9 mm and the distance from the analyzer outlet to the mass-selecting aperture of Y = 215 mm. The spatial separation between ions of a given mass and ions of one mass unit, u, higher decreases with increasing mass number from about 100 mm between masses 1 u and 2 u to about 2 mm between masses 80 u and 81 u. This calculated curve is derived from Eq. (2.11). The datum (triangle) is the result of measuring the distance between the aperture center and a deposited mark on the aperture due to the 29 u ion beam. m0 in general, and in this case, m1 is adjacent to m0, i.e. m1 = m0 + 1 u. λ1, the distance between m0 and m1 ion trajectories, depends only on the physical parameters (geometry) of the system and the two masses, and it is independent of the energy of the ions or the magnetic field used. The dependance calculated from Eq. (2.11) of λ1 as a function of mass, m0, is shown in Fig. 2.6. A measured value for the distance to the aperture of Y ≈ 215 mm was used for this calculation along with a value for the analyzer radius of 45 curvature of r0 = 97.9 mm. The calculated λ1 values (line) are highest at lower masses, decreasing with increasing mass. This shows that the physical distance between trajectories at the aperture of ions with adjacent mass numbers ranges from approximately 100 mm between 1 u and 2 u to approximately 2 mm between 80 u and 81 u. For 28Si at 28 u, the calculated distance to the 29Si beam at 29 u is λ1 ≈ 5.5 mm. A 2 mm wide mass-selecting aperture then allows a mass-to-charge range of approximately ± 0.18 u/e when selecting 28 u/e ions. This distance is compared to a measurement of the distance between these two beams (triangle), which has a similar value of approximately 6.2 mm. This measurement was made from the center of the aperture to the center of a visible deposition spot on the aperture created from the 29 u ion beam after depositing 28Si samples. A photograph of this measured aperture is shown in Fig. B.6 in Appendix B. The significance of λ1 being larger at lower masses is that ions at those masses have a better isolation with ions at adjacent masses leading to larger geometric selectivites for a selected mass. Ions with the same mass but different kinetic energies due to different accel- erating voltages will also have different radii of curvature in a mass analyzer. The physical separation of these ion trajectories at the aperture can be calculated in a similar manner as Eq. (2.11). For the case of a small kinetic energy spread amongst otherwise identical ions due to slight variations in the accelerating voltage, ∆V , a slight spreading of the ion trajectories would result. The physical size of this spread- ing of trajectories at the mass-selecting aperture, λE, from a selected ion to an ion 46 of the same mass but with a slightly higher kinetic energy is given by λE = r0 ({ 1 + ∆V V } 1 2 cos { sin−1 [ 1− ( 1 + ∆V V )− 1 2 ]} − 1 ) + Y tan { sin−1 [ 1− ( 1 + ∆V V )− 1 2 ]} . (2.12) This expression shows that the spreading of ion trajectories due to an energy spread, λE, for ions of a given mass depends only on the physical parameters of the system and the accelerating voltage and is independent of mass or magnetic field. A plot of λE calculated from Eq. (2.12) as a function of the accelerating voltage, V , for different values of the variation in accelerating voltage, ∆V , is shown in Fig. 2.7. Values in this calculation for the distance to the aperture (Y = 215 mm) and the analyzer radius of curvature (r0 = 97.9 mm) were the same as previously used. λE is calculated for ∆V = 5 V (line), ∆V = 7.5 V (dotted line), and ∆V = 7.5 V (dashed line). For ions accelerated with a nominal voltage of 4040 V with a typical ∆V = 6 V, λE ≈ 0.23 mm. It is observed from these calculations that λE decreases with increasing V , and although the effect is relatively small, it is therefore advantageous to transport ions in the beamline at higher voltages. For V increased to 4740 V in the previous calculation, λE is reduced to approximately 0.20 mm. For λE = 0.23 mm, the total width of the trajectories due to this spread is 0.46 mm, which is equivalent to a mass range of 0.08 u at a mass of 28 u. From the calculated values of λ1, the mass resolving power, m ∆m , of the ion beam system and sector mass analyzer can be calculated as m ∆m = m0λ1 (m1 −m0)∆λ , (2.13) 47 Figure 2.7: Calculated accelerating voltage dependance of the spatial separation, λE, of ions accelerated with voltages differing by ∆V in the ion beam at the mass- selecting aperture. These calculations use values for the physical parameters of the ion beamline including the sector mass analyzer radius of curvature of r0 = 97.9 mm and the distance from the analyzer outlet to the mass-selecting aperture of Y = 215 mm. λE is calculated for ∆V values of 5 V (line), 7.5 V (dotted line), and 10 V (dashed line). λE decreases with increasing accelerating voltage by about 35 % from 3500 V to 5000 V. These calculated curves are derived from Eq. (2.12). where ∆λ is the width of the selected ion beam at mass m0. m1 is the larger mass used in the calculation of λ1 in Eq. (2.11) and here, m1 −m0 = 1 u. For the mass resolving power of a mass spectrometer, the numerator represents the selected mass, here m0, and the denominator, ∆m, is a variant of the width of the selected ion beam. Calculated mass resolving power values from Eq. (2.11) and (2.13) are shown as a function of mass in Fig. 2.8. Values in this calculation for the distance to the aperture (Y = 215 mm) and the analyzer radius of curvature (r0 = 97.9 mm) were the same as previously used. Additionally, a value of the beam width of ∆λ = 2 mm was used, which is equal to the aperture width used for much of this work. 48 Figure 2.8: Calculated mass dependance of the mass resolving power, m ∆m , of the ion beamline. These calculations (line) use values for the physical parameters of the ion beamline including the sector mass analyzer radius of curvature of r0 = 97 mm, and the distance from the analyzer outlet to the mass-selecting aperture of Y = 215 mm. Additionally, they use an ion beam width, ∆λ = 2 mm. This calculated curve is derived from Eq. (2.13). The datum (triangle) is the highest measured mass resolution for a 28Si ion current peak. The calculated values of m ∆m (line) increase with increasing mass, which is counter intuitive because the ion beams of lower masses are better separated than those at higher masses. m ∆m increases sharply from a value of approximately 50.2 at a mass of 1 u to approximately 73.4 at 10 u and then increases much slower at higher masses up to approximately 77.6 at 80 u. A single experimental value of m ∆m (triangle) for the best mass resolving power measured in this work is shown for comparison. Although other experimental data exists, it can be problematic comparing them if the total ion beam current at various masses is very different because the beam widths can depend on the total current, which is often the case. 49 2.2.3 Operating Parameters Two versions of the Penning ion source were used in this work. One with a thin disc-like anode that is efficient at producing ions by sputtering the cathode material, and another with a longer cylindrical anode that is more efficient at ionizing atoms of the working gas. Schematic diagrams of the two ion sources are shown in Fig. 2.9. These schematics are viewed as a cross-section through the middle of the ion sources in a plane parallel to the axis of the ion beamline. Panel (a) shows a schematic obtained from Physicon Corp. of the gas-mode ion source with the gas inlet at the right and the ion beam exit facing the left with an outlet cone. This source uses an anode that is long and cylindrical that is surrounded on either side by the inlet and outlet cathodes. These cathodes have holes through their middles to allow gas to enter the source at the inlet side and ions to exit at the outlet. The anode and cathodes are isolated from each other by a ring shaped insulator seen inside the back plate. As mentioned, the voltage between the anode and cathodes, Varc, was typically about 3 kV, and it was used to generate the gas plasma between the cathodes and inside the anode (shaded oval). Surrounding the anode and cathode region is the source electromagnet solenoid. The anode is in the middle of the solenoid in this source. The source magnet provides a magnetic field (dashed curves) of typically 60 mT that radially confines the ions in the plasma within the anode. Ions are extracted from the plasma by an electric field generated by the high voltage applied to the extractor electrode, Vext. Photographs of the anodes and cathodes of the two ion sources are shown in Fig. B.3 in Appendix B. 50 Figure 2.9: Schematic diagrams of the gas-mode and solids-mode Penning ion sources. These sources are distinguished primarily by the size and configuration of the anode and cathodes. (a) The gas-mode ion source is viewed in cross-section through the middle with the gas inlet at the right and the ion beam exit at the left (from Physicon Corp.). This source uses a long cylindrical anode surrounded by inlet and outlet cathodes with holes through each. The voltage between the anode and cathodes, Varc, was typically 3 kV and was used to generate the plasma between the cathodes (shaded oval). The source electromagnet solenoid surrounds the anode and cathode section, providing a magnetic field (dashed curves) of typically 60 mT to radially confine the ions in the plasma. (b) The solids-mode ion source is viewed in cross-section in a similar orientation to the schematic in (a) but with only the middle area visible (from Ref. [62]). This source uses a short anode in the shape of a disk with a hole in the center surrounded by an outlet cathode with a hole on the left and a solid inlet cathode at the right. The plasma (shaded oval) is confined between the cathodes in a smaller volume than in the source in (a). Although the magnetic field is not shown, the electromagnet is used in a similar manner as in (a). Panel (b) of Fig. 2.9 shows a schematic cross-section from Ref. [62] of a portion of the solids-mode ion source in the same orientation as the source in (a). This source uses an anode that is in the shape of a thin disk with a hole in the center. As with the gas-mode source, the anode is surrounded by the inlet and outlet cathodes. The 51 outlet cathode on the right of the diagram is the same used for the outlet of the gas-mode source with a hole through it for ions to escape. The inlet cathode is larger and is solid without a hole in order to provide a surface for sputtering. Gas is injected into this source through holes at the base of the inset cathode support. Although not shown, the magnetic field used is similar to that of panel (b) but with anode and cathodes offset from the center of the solenoid. The plasma region is again between the cathodes and within the anode (shaded oval), however, the volume is much smaller in this source than for the gas-mode source. The solids-mode ion source was used to produce both 22Ne and 12C samples in this work. Working pressures of the source gas injected into the ion source ranged from 1.1× 10−3 Pa to ≈ 1.0× 10−2 Pa (8.0× 10−6 Torr to 7.8× 10−5 Torr). The anode voltage, VA, was typically about 500 V or more and the cathode was typically about -500 V. The source electromagnet typically current ranged between 1.0 A and 1.5 A. Arc currents, Iarc, between 15 mA and 30 mA were observed. The gas-mode ion source was used to produce 28Si in this work. There are two operating regimes of the gas-mode ion source. These are the typical “low pressure” plasma mode operating condition used to deposit the majority of 28Si samples in this work, as well as a “high pressure” mode that was used to deposit several samples. The typical working pressure used when injecting gas into the source for the low pressure mode was between 1.0× 10−4 Pa and 3.3× 10−4 Pa (7.5× 10−7 Torr to 2.5× 10−6 Torr). The operating pressure of the high pressure plasma mode was typically 1.3× 10−3 Pa (1.0× 10−5 Torr). Anode voltages, VA, between +30 V and +700 V were used for various depositions throughout this work to generate ions with 52 a similar range of final ion kinetic energies, Ei. The voltage of to the cathode, VC , was between -1 kV and -4 kV. This voltage was floating on top of the anode voltage, and the applied difference between the two was referred to as the “arc” voltage, Varc. The source electromagnet typically requires 1.5 A to 2 A to enable ignition of the plasma, which corresponds to magnetic fields of approximately 40 mT to 80 mT. A typical arc current, Iarc, of 0.5 mA was observed for this source. Additional operating parameters for both ion sources including those used to deposit the most highly enriched 28Si sample produced in this work are shown in Tables A.1 to A.3 in Appendix A. The operation and tuning procedures used with the ion beamline to produce a stable and focused ion beam are described in the following steps. 1. Inject the source gas into the ion source using the leak valve at an appropriate pressure, depending on the source and plasma mode being used, 2. set the anode voltage, VA, to a small positive voltage and the arc voltage, Varc, negative with reference to VA to define the cathode voltage, VC , 3. set the transport voltage, VT , to typically -4 kV, 4. set the extractor voltage, VExt, which floats on and is more negative than VT , and 5. set the focus voltage, VF , which also floats on and is more negative than VT and VExt. 6. With these voltages set, turn on the source electromagnet and as the current, ISM , is increased, watch the current reading on the transport power supply, IT . This will ignite a plasma in the source and then set ISM for when IT is observed to increase and stabilize. 7. Then, set the magnetic sector mass analyzer to a current corresponding to the desired ion mass and power on the deceleration lenses, monitoring the ion current at the target. 53 8. Tune the ion beam by adjusting each control element starting from the ion source and moving down the beamline to the deceleration lenses, including the source pressure, ISM , Varc, VExt, VF , and the X-Y deflectors, all while maximising the total detected ion current. 9. Finally, tune the deceleration lenses by adjusting the voltages of lenses A2, A3, B2, B3, B4, and X in order while maximising the ion beam current through the sample aperture. 2.2.4 Ion Beam Characterization 2.2.4.1 Ion Beam Mass Spectra Several techniques were used to characterize each ion beam in preparation for deposition of enriched materials. The components of the ion beam can be observed by sweeping the current, and thus the magnetic field, of the sector mass analyzer magnet, which corresponds to sweeping the selected mass, and recording the mea- sured ion current at the sample location. This type of measurement generates a mass spectrum of the components of the ion beam. Ion currents are measured using a picoammeter with a typical noise floor of 10 pA for the range setting typically used. Several working gases are used in the ion source throughout this work includ- ing Ne, carbon dioxide (CO2), and SiH4 for 28Si deposition. Mass spectra for Ne and CO2 are discussed in Chapter 3, and mass spectra for SiH4 are discussed later in this chapter and throughout this work. Collecting and analyzing a mass spectrum is an important aspect to preparing a mass separated ion beam for deposition. The mass spectrum is used for tuning the mass analyzer to the desired ion, assessing the isolation of that ion beam from adjacent beams, and detecting contaminants within 54 the vacuum system, among other things. The mass spectra collected while using several other working gases of Ar, N2, and methane (CH4) in the ion source are shown in Fig. 2.10. The measured ion currents (circles and line) are shown as a function of the ion mass-to-charge ratio, m/q. As discussed previously, in this work, mostly singly charged ions are produced, so the mass-to-charge ratio is simply referred to and represented as the ion mass when discussing mass spectra. The current applied to the sector mass analyzer used for the magnetic field sweep is shown on the top axes. The spectra in panels (a) and (b) were acquired using the gas-mode ion source, while the spectrum in panel (c) was acquired using solids-mode source. Panel (a) is a portion of a mass spectrum for Ar showing current peaks corresponding to 36Ar (36 u), 38Ar (38 u), and 40Ar (40 u). The relative sizes of these peaks are similar to the natural abundance of Ar, which is comprised of approximately 0.33 % 36Ar, 0.07 % 38Ar, and 99.6 % 40Ar. Panel (b) of Fig. 2.10 is a portion of a mass spectrum for N2 showing current peaks corresponding to 13C (13 u), 14N (14 u), 15N (15 u), and 16O (16 u). The presence of C and O in the beam is due to partial pressures of molecular species containing those elements within the vacuum chamber or possibly contamination in the N2 gas itself. The peak at 14.5 u likely corresponds to a doubly-charged ion of mass 29 u. A large shoulder peak is seen on the lower mass side of the 14N peak and is likely a result of poor ion beam focusing causing ion scattering from a lens element. The relative sizes of the 14 u and 15 u peaks are similar to the natural abundance of N, which is comprised of approximately 99.6 % 14N and 0.4 % 15N. Panel (c) of Fig. 2.10 is a portion of a mass spectrum for CH4 showing current 55 Figure 2.10: Ion beam mass spectra for different working gases used in the ion source. Ion currents (circles and line) are recorded while sweeping the mass analyzer current, and thus the magnetic field (top axes) (a) Ar mass spectrum showing current peaks corresponding to 36Ar (36 u), 38Ar (38 u), and 40Ar (40 u). The relative sizes of these peaks are similar to the natural abundance of Ar. (b) N2 mass spectrum showing current peaks corresponding to 13C (13 u), 14N (14 u), 15N (15 u), and 16O (16 u). The peak at 14.5 u likely corresponds to a doubly-charged ion of mass 29 u. A large shoulder peak is seen on the lower mass side of the 14N peak due to poor ion beam focusing causing scattering from a lens element. The relative sizes of the 14 u and 15 u peaks are similar to the natural abundance of N. (c) CH4 mass spectrum showing current peaks corresponding to 12C (12 u) and various C hydrides that are cracked in the ion source plasma. The peak at 13 u is mostly 12CH with a small amount of 13C. Likewise, the peaks at 14 u, 15 u, 16 u are mostly 12CH2, 12CH3, 12CH4, respectively. 56 peaks corresponding to 12C (12 u) as well as various C hydrides that are the result of different numbers of H atoms being removed when CH4 is cracked in the ion source plasma. The ion current peak at 13 u is mostly 12CH with a small amount of 13C. Similarly, the peak at 14 u is mostly 12CH2 and possibly 14N, the peak at 15 u is mostly 12CH3, and the peak at 16 u is mostly 12CH4 and possibly 16O. As mentioned, natural abundance SiH4 gas is used for 28Si deposition and several spectra representing various deposition conditions for different samples will be presented in chapters 4, 5, and 6. Figure 2.11 shows a portion of a mass spectrum of the ion beam collected while using SiH4 as the working gas. The corresponding magnetic sector mass analyzer current used for the field sweep is shown on the top axis. This spectrum was generated while using the gas-mode ion source. Ion current peaks (circles) corresponding to 28Si, and various Si hydrides are observed from 28 u to 33 u. The ion current peak at 28 u is 28Si, and the ion current peak at 29 u is mostly 28SiH containing ≈ 5 % 29Si. This estimated relative 29Si is based on the peak heights of 28 u and 29 u being similar and the expected natural abundance of 29Si relative to the 28Si at 28 u. Similar to the peak at 29 u, the peak at 30 u is mostly 28SiH2, the peak at 31 u is mostly 28SiH3, and the peak at 32 u is mostly 28SiH4. The peak at 33 u is likely a combination of 30SiH3 and 29SiH4. These molecular species are formed by SiH4 being cracked in the ion source and losing different numbers of H atoms. The overall efficiency of generating 28Si is fairly low as it makes up roughly 10 % of the total SiH4 ion current. The 28Si ion current observed in this spectrum is Ii ≈ = 620 nA, which is typical for 28Si depositions. The 28 u peak and the 29 u peaks in Fig. 2.11 show a high degree of separa- 57 Figure 2.11: SiH4 mass spectrum representative of the ion beam settings for deposi- tion of many 28Si samples. The ion current (circles) is recorded while sweeping the mass analyzer current, and thus the magnetic field (top axis). Gaussian fits (line, Eq. (2.14)) to the 28 u, 29 u, 30 u, and 31 u peaks are superimposed on the data. The peak at 28 u is 28Si and the peak at 29 u peak is 28SiH and ≈ 5 % 29Si. Sev- eral higher order hydrides are also observed corresponding to mostly 28SiH2 (30 u), 28SiH3 (31 u), and 28SiH4 (32 u). These molecular species are cracked in the ion source plasma. The centers of the 28 u and 29 u fits are separated by ≈ 10 σ. tion with no detectable ion current signal occurring between the peaks. Secondary electrons generated by the ion beam cause the current between the peaks to be ≈ -0.5 nA. Also shown is a sum of Gaussian fits to the data (line) of the form Ii = I 0 i + B σ √ 2pi exp ( −1 2 ( m−mc σ )2) , (2.14) where Ii is the measured ion current, I 0 i is an offset to the current due to the noise floor of the measurement, B is the area of the Gaussian, σ is the standard deviation, m is the mass, and mc is the center mass of the peak. The value of 58 I0i used for fits to ion current peaks is typically 10 pA, which is the noise of the picoammeter mentioned previously. One can see that the data matches the form of a Gaussian very well. This indicates a symmetric and optimally tuned beam shape with minimal perturbations such as scattering. The ion current peaks generated in this system are expected to be approximately Gaussian due to several factors. A Gaussian shaped current peak can result only if the width of the ion beam is greater than the aperture width, which is the case for this system. For a beam with a width smaller than the aperture width, the current peak would have a flat top as the full beam traversed the aperture. The approximately Gaussian shaped profile represents a spacial distribution of beam fluxes that likely originates at the ion source where a higher flux of ions is extracted from the center of the source exit and a smaller distribution coming from the edges. For this mass spectrum in Fig. 2.11, the centers of the 28 u and 29 u fits are separated by ≈ 10 σ (standard deviation). The mass resolving power of the system in this configuration can also be derived from this mass spectrum as m ∆m ≈ 58 at 28 u. ∆m in this calculation is determined by taking the full width of the fit to the 28 u peak at 10 % of the peak height, which is approximately 0.48 u. This width is equivalent to a physical width at the mass-selecting aperture of 2.6 mm. As mentioned previously, for a typical ion energy spread of ± 6 eV due to a spread of accelerating voltages in the beam of ∆V = 6 V, the width in mass due to the energy spread at mass 28 u is approximately 0.08 u. This width due to the energy spread is therefore approximately 17 % of the total width of the 28Si beam as measured in this mass spectrum. The more significant contribution to the full width is likely related 59 to the width of the ion beam exiting the ion source and may be an intrinsic width for the system, although there may be different focusing conditions which could reduce this width. The hole in the exit cathode of the ion source is approximately 2.54 mm, which is perhaps not coincidentally very similar to the measured 28Si beam width. Ions with a small spread of angles entering the sector mass analyzer could contribute to this width as well. If the total width was only due to the mentioned energy spread, then the resulting mass resolving power would be m ∆m ≈ 350. An example of a SiH4 mass spectrum where the ion current peaks are not symmetric and cannot be described by single Gaussian fits is presented in Fig. 2.12. This spectrum shows four ion current peaks (circles) including 28Si at 28 u and a combination of 29Si and 28SiH at 29 u. The corresponding magnetic sector mass analyzer current used for the field sweep is shown on the top axis. Unlike the peaks in the mass spectrum in Fig. 2.11, the peaks here are not symmetric. Shoulders appear on the current peaks to the lower mass side. This is likely due to imperfect tuning and focusing of the ion beam causing a small amount of ion scattering from a lens element. This type of asymmetry decreases the isolation of adjacent peaks, which could possibly affect the realized selectivity and enrichment of a sample. Gaussian fits (line, Eq. (2.14)) to the 28 u and 29 u peaks are superimposed on the data. An additional Gaussian fit of a shoulder peak (dotted line) of the 29 u main peak is also shown, which is needed to accurately determine the peak separation. In this case, the separation from the 29 u main peak to 28 u is ≈ 13 σ, however, when considering the shoulder peak, that separation drops to ≈ 10 σ. This value is still very large, but it illustrates the effects of non-ideal beam tuning. Much more 60 Figure 2.12: SiH4 ion beam mass spectrum showing shoulder peaks on the main ion current peaks to lower mass side. The ion current (circles) is recorded while sweeping the mass analyzer current, and thus the magnetic field (top axis). The peak at 28 u is 28Si and the peak at 29 u peak is 28SiH and ≈ 5 % 29Si. Gaussian fits (line, Eq. (2.14)) to the 28 u and 29 u peaks are superimposed on the data. A Gaussian fit of the shoulder peak (dotted line) of the 29 u main peak is also shown. significant shoulder peaks are sometimes observed as well. Other elements and contaminants in the ion beam generated from the chamber or the gas being used can be observed in the mass spectra as well. Typically, when analyzing the mass spectrum of SiH4, several other mass regions of the full spectrum that contain common contaminants are inspected and recorded. Four of these ion beam contaminant mass spectra acquired when using SiH4 as the working gas are shown in Fig. 2.13. In each of these spectra, the ion current (circles and lines) is recorded while sweeping the mass analyzer current, and thus the magnetic field, 61 Figure 2.13: Ion beam mass spectra of several chemical contaminants in the ion beam. These spectra were acquired while using SiH4 as the working gas. In each of these spectra, the ion current (circles and line) is recorded while sweeping the mass analyzer current, and thus the magnetic field (top axes). (a) H (1 u) and H2 (2 u) current peaks are typically observed in the ion beam. (b) Most of the contaminants have masses between 11 u and 20 u, as seen here. These current peaks correspond to elements including 12C (12 u), 14N (14 u), 16O (16 u), H2O at mass 18 u, and F (19 u). Doubly charged Si-hydride peaks 28SiH2+ and 28SiH2+2 are also present at masses 14.5 u and 15 u, respectively. Other unknown peaks appear as well. (c) A CO2 current peak is also typically observed at 44 u, although usually at much lower current levels than other peaks in (a) and (b). The species corresponding to the other three peaks observed within this range are not known. (d) 63Cu (63 u) and 65Cu (65 u) ion current peaks are observed due to ions generated through sputtering of the Cu cathodes in the ion source. The relative peak heights are similar to the natural abundance of Cu. 62 which is shown on the top axes. The spectra in panels (a), (b), and (c) were acquired while using the gas-mode ion source and the spectrum in panel (c) was acquired while using the solids-mode source. Panel (a) shows H (1 u) and H2 (2 u) current peaks, which are typically observed in the ion beam. These can originate from vacuum chamber, which always contains H2, or from the ion source which is cracking SiH4 and releasing H2. Panel (b) shows a spectrum with a mass range between 11 u and 20 u containing the most common and abundant contaminants observed. These current peaks correspond to elements including 12C (12 u), 14N (14 u), 16O (16 u), H2O at mass 18 u, and F (19 u). Doubly charged Si-hydride peaks 28SiH2+ and 28SiH2+2 are also present at masses 14.5 u and 15 u, respectively. If a large 14N or 12C is present, it could indicate the presence of N2 or CO, which both would appear at approximately 28 u along with 28Si. Panel (c) of Fig. 2.13 shows a mass range beyond the SiH4 peaks where several peaks are present. The current peak at 44 u is likely CO2, which is also typically observed, although with lower currents than the peaks in panels (a) and (b). It is not obvious what ionic species correspond to the other three peaks in this spectrum. Panel (d) shows 63Cu (63 u) and 65Cu (65 u) ion current peaks, which are due to ions generated through sputtering of the Cu cathodes in the ion source. The relative peak heights are similar to the natural abundance of Cu, which is comprised of approximately 69.2 % 63Cu and 30.8 % 65Cu. Note the linear scale of panel (d). As previously mentioned, a low pressure plasma mode and an high pressure plasma mode of the ion source were used to deposit 28Si samples in this work. These working modes result in qualitatively and quantitatively different SiH4 ion beam 63 mass spectra. To demonstrate the transition between these two pressure modes, SiH4 mass spectra were acquired for several different ion source pressures. These spectra are presented in Fig. 2.14. Note that these spectra are shown on a linear scale to emphasize the differences. Four current peaks are shown comprised of 28Si at 28 u, 28SiH and 29Si at 29 u, primarily 28SiH2 at 30 u, and primarily 28SiH3 at 31 u. Initially, a typical working pressure for the low pressure mode of 1.2× 10−4 Pa (9.0× 10−7 Torr) is used. This pressure results in a standard low pressure mode spectrum (open circles and line), with the 28 u peak appearing smaller than the 30 u and 31 u peaks. This is qualitatively similar to the spectrum shown in Fig. 2.11 on a semi-log scale. When the pressure of the ion source is increased to 6.7× 10−4 Pa (5.0× 10−6 Torr), the relative peak heights of the ion peaks in the spectrum shift (open squares and lines). The 28 u peak increases as the 29 u, 30 u, and 31 u peaks decrease. This is because the cracking efficiency of the plasma mode increases with higher pressures and additional H atoms are being cracked from the hydrides, producing more 28Si. This trend continues for the spectrum corresponding to a pressure of 1.1× 10−3 Pa (8.0× 10−6 Torr), which shows the 28 u peak even larger (dotted line) but still lower than the 30 u and 31 u peaks. Finally, when the working pressure of the source is increased to 1.3× 10−3 Pa (9.5× 10−6 Torr), the mass spectrum (line) appears as a nominal spectrum for the high pressure mode. Here the 28 u peak is the largest of the four peaks resulting from a further increase in the cracking efficiency of hydrides into 28Si. In these spectra, the total efficiency for producing 28Si increased from approximately 8 % to 33 %. The highest efficiency for producing 28Si observed for a high pressure mode spectrum was approximately 64 Figure 2.14: SiH4 ion beam mass spectra for several working pressures showing the transition of the relative peak heights from the low pressure mode of the ion source to the high pressure mode. Note the linear scale. The four current peaks are comprised of 28Si at 28 u, mostly 28SiH and ≈ 5 % 29Si at 29 u, primarily 28SiH2 at 30 u, and primarily 28SiH3 at 31 u. For a typical pressure used for the low pressure mode of 1.2× 10−4 Pa, a standard low pressure mode spectrum is observed (open circles and line) with the 28 u peak appearing smaller than the 30 u and 31 u peaks. As the pressure of SiH4 in the source is increased, leading to a higher cracking efficiency, the 29 u, 30 u, and 31 u peaks decrease in height while the 28 u peak increases. For a pressure of 1.3× 10−3 Pa, typical of the high pressure mode, the spectrum has fully transitioned to a high pressure mode spectrum (line) with the 28 u peak larger than the other peaks. 47 %. Note also that the total 28Si ion current increased roughly a factor of 2.7 up to approximately 1.5 µA. The high pressure mode is thus able to produce the highest 28Si ion currents, which leads to the highest growth rates and thickest films. While all 28Si samples produced in this work were deposited using SiH4 as the Si source, the viability of using solid Si cathodes in the ion source that are sputtered to generate ions was explored. Tests with Si cathodes were conducted using the gas- 65 Figure 2.15: Ion beam mass spectrum of Si ions generated by sputtering natural abundance Si cathodes while using Ar as the working gas in the gas-mode ion source. The ion current (circles and line) is recorded while sweeping the mass analyzer current, and thus the magnetic field (top axis). Ion current peaks corresponding to 28Si (28 u), 29Si (29 u), and 30Si (30 u) are observed. The relative heights of the three peaks are similar to the natural abundance of Si. mode ion source instead of the solids-mode source because the gas-mode source was deemed more reliable at the time. To sputter the Si cathodes, Ar was used as the working gas. A Si mass spectrum acquired when using Si cathodes is shown in Fig. 2.15. The ion current (circles and line) is recorded while sweeping the mass analyzer current, and thus the magnetic field, which is shown on the top axes. Ion current peaks corresponding to 28Si (28 u), 29Si (29 u), and 30Si (30 u) are seen. There also appears to be a small peak at 29.5 u, but the origin of this peak is not known. The relative heights of the three peaks approximately match the natural abundance of Si. With only approximately 35 nA of 28Si ion beam current, which 66 is at least a factor of ten lower than when using SiH4 as a source, this approach was abandoned in favor of continuing to use SiH4. 2.2.4.2 Ion Beam Energy, Ei The average ion energy, Ei, of ions in the beam at the target can be determined in a measurement referred to as a “roll-off” curve. In this measurement, a positive bias voltage, Vbias, is applied to the target to repel the incoming ions and is increased in steps while the ion current, Ii, is recorded. The ion current will drop significantly at a voltage equal to the average ion energy. An example of a roll-off curve produced in this work is shown in Fig. 2.16 (a). This roll-off curve is for Ar ions while using an anode voltage VA = 30 V. The Ar ion current (circles and line) initially has a value over 1.5 µA and initially decreases slowly as Vbias is increased. The current then drops much more sharply from approximately 1.25 µA to near 0 µA between 20 V and 30 V. Finally, beyond 30 V, the ion current levels off and again degreases slowly with further increasing Vbias. Note that the ion current becomes negative for voltages above 30 V. This is because electrons are attracted to the positively biased target. Panel (b) is the numerical derivative of the current signal in (a) with respect to the bias voltage, |dIi/dVbias|. The absolute value of the derivative is taken for clarity in the figure. The derivative (circles) shows the peak change in ion current at the average ion energy. |dIi/dVbias| is then fit to a Gaussian (line) of the same form as Eq. (2.14) but with different variables given by ∣∣∣∣ dIidVbias ∣∣∣∣ = C0 + Bσ√2pi exp ( −1 2 ( Vbias − Ei/q σ )2) , (2.15) 67 Figure 2.16: Average ion energy measurement for Ar ions using a roll-off curve. (a) Roll-off curve for an anode voltage, VA = 30 V. As the positive bias voltage on the sample, Vbias, is increased, positively charged Ar ions are repelled and the ion current, Ii, decreases (circles and line). At a bias voltage corresponding to the average ion energy, Ei, Ii decreases significantly. The measured current becomes negative as electrons are attracted into the target. (b) The absolute value of the derivative, |dIi/dVbias|, of the data in (a) (circles) shows a peak in the derivative at the average ion energy. The data is fit to a Gaussian (line, Eq. (2.15)) giving a peak center Ei ≈ 26 eV. where C0 is the vertical offset from zero of the derivative, B is again the area of the Gaussian, and the ion charge state, q = 1, is needed as a conversion of the voltage value of the peak center to ion energy. From this fit, the average ion energy, i.e. the energy corresponding to the peak center, is determined to be Ei ≈ 26 eV, which is close to the 30 V applied to the anode, as expected for ions created at the plasma potential defined by the anode in the source. Ei can be mapped as a function of anode voltage, VA, to determine their re- lationship. The plasma potential is also affected by the cathode voltage, however, and so the relation between Ei and VA will change depending on the cathode volt- age, or rather the arc voltage between the anode and cathode, Varc. Experimental 68 measurements of the dependance of Ei on VA and Varc is shown in Fig. 2.17. The values of Ei are determined in roll-off curve measurements such as the one shown in Fig. 2.16. These measurements are repeated for several values of VA to get their relationship while keeping the value of Varc constant. This measurement was re- peated for different values of Varc including 1.74 kV (circles), 2 kV (squares), 3 kV (triangles), and 4 kV (diamonds). The resulting values of Ei are always smaller than VA. Additionally, for a given value of VA, Ei increases with increasing values of Varc. The data for each Varc value is fit to a line (solid, dashed, and dotted lines) given by Ei = αVA + β, (2.16) with slope α and offset β. These fits give values of α ≈ 1 e, and in general α should ideally be the ion charge state, q. This means that the offset fit parameter, β, approximately describes the difference between the applied VA and the resulting Ei of ions, written as β ≈ Ei − qVA. β is plotted as a function of Varc in the inset of Fig. 2.17, which shows that the magnitude of β becomes smaller, i.e. less negative, with increasing Varc. The energy spread of the ions can also be determined from the fits to the derivatives of the roll-off curves. The energy spread, ∆E, is determined as half the peak width at half the height, with the full range of expected energies being twice this value. For the measurements of Ei in Fig. 2.17, ∆E varies between approximately 17 eV for the highest values of VA and 4 eV for the lowest ones. The relative energy spread of the ion is then calculated as ∆E/Ei, which is plotted vs. 69 Figure 2.17: Dependance of the average ion energy, Ei, on the anode voltage, VA, and arc voltage, Varc. Measured values of Ei are plotted vs. VA for a four values of Varc: 1.74 kV (circles), 2 kV (squares), 3 kV (triangles), and 4 kV (diamonds). The data for each arc voltage value are fit to a line (solid, dashed, and dotted lines, Eq. (2.16)). The inset shows the dependance of the intercept fit parameter of the linear fits, β, on Varc. With slopes of α ≈ 1 e for the linear fits, β ≈ Ei − qVA, the magnitude of which becomes smaller, i.e. less negative, with increasing Varc. Ei in Fig. 2.18. The measured values of ∆E/Ei (circles) increase with decreasing ion energy from roughly 0.1 at 160 eV to 0.35 at 15 eV. This means that nominally 15 eV ions will have an approximate range of energies between 4.5 eV and 25.5 eV. The uncertainty on the values of ∆E/Ei come from the standard error of the fit values. This data is fit using a power low plus a constant given by ∆E Ei = (4.5)E−1i + 0.04, (2.17) where the prefactor and the constant are the fit values, Ei is in units of eV, and 70 Figure 2.18: Dependance of the relative energy spread ∆E/Ei of ions on the average ion energy, Ei. ∆E/Ei (circles) increases from below 0.1 at about 150 eV to above 0.3 at lower ion energies below about 20 eV. A fit to the data of the form of a power law plus a constant (line, Eq. (2.17)) is also shown. ∆E/Ei is unitless. This fit equation is derived by first fitting ∆E vs. Ei to a line. These quantities are expected to be proportional with a constant offset from an intrinsic energy spread possibly due to other potentials near the source. The resulting linear equation is then divided by Ei to get a fit for ∆E/Ei. For deposition of 28Si, ions with average energies of Ei ≈ 35 eV were typically used, and these ions would have a ∆E of approximately 5.9 eV and range in energy from approximately 29 eV to 41 eV. Additionally, at 28 u, a ∆E of 5.9 eV would represent a physical width of of the beam at the mass-selecting aperture of approximately 0.23 mm and a width in mass of approximately 0.08 u, as mentioned previously. The ion energy is an important parameter in hyperthermal energy ion beam de- position because ions with these energies will sputter the film as they are deposited. 71 Figure 2.19: Calculated sputter yields for 28Si ions bombarding a crystalline 28Si surface (circles and line) and an amorphous 28Si surface (triangles and line) are shown vs. the ion energy. Sputter yields for 12C ions bombarding an amorphous 12C surface (squares and line) are also shown. Sputter yield values are based on TRIM calculations [69]. The 28Si sputter yield decreases from about 0.6 sputtered atoms per ion down to about 0.05 when decreasing the ion energy from 550 eV to 50 eV. 12C sputter yields are lower and only reach about 0.25 at an ion energy of 550 eV. If the sputter rate is too large, then the efficiency of the deposition decreases. Sput- ter yields for 28Si and 12C deposition have been calculated using TRIM [69] and are shown in Fig. 2.19. Sputter yields for the average number of atoms sputtered for every incident ion were calculated for 28Si ions bombarding an amorphous 28Si film (triangles and line), 28Si ions bombarding a crystalline 28Si film (circles and line), and 12C ions bombarding an amorphous 12C film (squares and line). These calculated sputter yields are shown as a function of the ion energy, Ei. As one would expect, the sputter yields decrease with decreasing ion energy. The 28Si sputter yield decreases from approximately 0.6 sputtered atoms per ion down to approximately 0.05 when decreasing the ion energy from 550 eV to 50 eV. 12C sputter yields are 72 lower and only reach approximately 0.25 at an ion energy of 550 eV, showing that the issue of sputtering is more significant for Si deposition. For 28Si ions with Ei ≈ 35 eV, the sputter yield is approximately 0.03, which means that for every 100 28Si ions bombarding the surface, roughly 3 28Si atoms on the surface are sputtered off. This results in a very high deposition efficiency. 2.2.4.3 Ion Beam Spot Size The beam spot shape and ion distribution can be mapped using a small detec- tor at the sample position that is scanned in two dimensions. The detector consists of a collector plate behind an aperture referred to as the sample aperture. At the detector, the ion beam passes through the roughly 3 mm diameter aperture and is collected. The measured ion current through aperture is typically used for assessing the beam flux to estimate the expected deposition rate when producing a sample. A 2D current current map of an Ar ion beam spot is shown in Fig. 2.20. The max- imum current measured in the map is approximately 550 nA. The size of the spot is approximately 5 mm by 7 mm with a resulting area of approximately 35 mm2. This spot size is convolved with the size of an aperture at the detector used for this measurement. The aperture is a roughly 3 mm diameter circular hole, which is large compared to the beam spot size. A simple estimate of the true beam spot size can be made by subtracting half of the aperture diameter from each size of the spot. This give dimensions for the ion beam spot size of roughly 2 mm by 4 mm with a resulting area of roughly 8 mm2. 2D maps such as these can help to tune the beam 73 Figure 2.20: Ar ion beam spot 2D current map. The ion current is recorded while moving the detector in two directions, y and z, perpendicular to the axis of the ion beam to produce a 2D grid of data. The maximum current measured in the map is about 550 nA. The area of the spot in this measurement is about 35 mm2, although the spot size is convolved with the 3 mm diameter of the detector. better and produce a more circular shape for deposition so the spot location on the sample is better known. While mass analyzed ion beam currents are typically of the order of 1 µA, the ion source actually generate as much as 1.5 mA of ion current. This total current can be measured on the power supply for the transport voltage, because all ions exiting the ion source must be neutralized in the ion beam chamber except those ions that pass through the mass-selecting aperture. This loss of ion current (fluence) from the total output to the analyzed current at the target was investigated by measuring ion currents on various electrostatic lens elements along the beamline. These investigation found that for a total of 1.3 mA, 0.56 mA was captured on the 74 extractor, 0.11 mA was captured on the focus, 0.2 mA was captured on the skimmer, and 0.02 mA was captured on the electron suppressor. A total of 0.28 mA was transmitted into the sector mass analyzer. As previously mentioned, approximately 10 % of the total current of SiH4 ions is 28Si when using the low pressure mode, so one would expect about 28 µA of 28Si current to be transmitted, which is much larger than typical. This may be because the re-focusing properties of the analyzer only affect ion trajectories in the plane of the bend of the analyzer and so ions are perhaps spreading out in a vertical direction perpendicular to the bend of the analyzer. Some modeling using SIMION ion trajectory software was used to model the electrostatic lenes of the ion beamline to try to find better settings for transmitting more ions, but a suitable solution was not found. 2.3 UHV Deposition and Analysis Chamber 2.3.1 Apparatus The deposition and analysis chamber pictured in the right part of Fig. 2.1 consists of a UHV system connected to the ion beamline on one side and the STM chamber on another. Samples can be loaded into the deposition chamber through the load lock, which can accommodate up to four samples. The load lock is pumped by a 67 L/s turbo pump (Agilent Technologies), which results in a base pressure of approximately 1.3× 10−7 Pa (1.0× 10−9 Torr). Typically, the load lock must be pumped out for one day before samples can be transferred into the deposition chamber. At that time, the pressure is about 1.3× 10−6 Pa (1.0× 10−8 Torr). The 75 Figure 2.21: Schematic cross section through the deposition chamber in the plane of the sample normal to the ion beam. The sample location on the manipulator and instruments are shown including the ion beam lenses, the RHEED gun and screen, and the natural abundance Si evaporator. Samples facing the ion beam during deposition can simultaneously be monitored with RHEED. (from Ref. [59]) deposition chamber also contains a 5-axis sample manipulator, where the samples sit during deposition. Several other analytic instruments are also present in the chamber including UHV ion gauges, a residual gas analyzer (RGA), reflection high energy electron diffraction (RHEED), a natural abundance Si electron beam thermal evaporator, and an Auger electron spectrometer (Physical Electronics), which was not used significantly in this work. Figure 2.21 shows a schematic cross section through the deposition chamber at the sample location normal to the ion beam, i.e. from the perspective of the ion beam (from Ref. [59]). The sample manipulator is pictured at the right of the schematic with the sample at the center. Normal to the sample (on the manipulator) are the deceleration lenses of ion beam optics, and aligned in the plane of the sample are the RHEED gun and screen and the Si 76 evaporator. Samples facing the ion beam during deposition can simultaneously be monitored with RHEED. The Si thermal evaporator (Omicron) is a model EFM 3 and is used for natural abundance Si (natSi) deposition. The EFM is capable of producing deposition rates of typically 0.1 nm/min. The Si source used was either standard Si wafer pieces in a crucible or a high purity Si rod. Throughout the experiments discussed here, the deposition chamber has also contained other instruments including a sputter gun for Ar sputter cleaning, an Al- deposition source, a STM tip preparation tool, a H-passivation cracking filament, and a quartz crystal microbalance (QCM). These components were also not signifi- cantly used in the work reported here. Two external infrared pyrometers, one from Omega and the other from Process Sensors, were used to measure the temperature of samples through a window. The Process Sensors pyrometer has a measurement range between 300 ◦C and 1300 ◦C, while the Omega pyrometer can only measure temperatures above 600 ◦C. When comparing the reading from the two pyrometers, a correction is applied to the Omega pyrometer, which can differ from the Process Sensors readings by ≈ 25 ◦C. A long magnetic transfer rod is used to transport sam- ples between the manipulator, the load lock, and the STM chamber. Additionally, wobble sticks are used to manipulate samples in vacuum at sample stations. 2.3.2 Vacuum Analysis The deposition chamber is pumped using two turbo pumps and two ion pumps. The turbo pumps are a 300 L/s pump (Edwards Vacuum) located at the decel- eration lens section, and a 685 L/s pump (Pfeiffer Vacuum) located at the main 77 Figure 2.22: Residual gas mass spectrum collected from the RGA in the deposition chamber for a base pressure of 8.3× 10−9 Pa obtained after baking out the chamber to UHV conditions. Partial pressure peaks of typical components of the vacuum are seen including H2, C, a very small H2O peak, CO and N2, and CO2. F is also seen in this system. deposition chamber below the manipulator. Typically, the base pressure of the deposition chamber for the experiments discussed in this work has been between 6.7× 10−9 Pa (5.0× 10−11 Torr) and about 1.3× 10−8 Pa (1.0× 10−10 Torr). This UHV environment is achieved by heating the chamber to around 150 ◦C for sev- eral days, commonly referred to as baking out. This process removes water and other impurities from inside the chamber, which otherwise get pumped very slowly. The RGA (SRS Vacuum Instruments) can detect the components of the vacuum as partial pressures, giving insight into possible contaminants within the chamber. A residual gas mass spectrum collected from the RGA of the deposition chamber 78 at a base pressure of approximately 8.3× 10−9 Pa (6.2× 10−11 Torr) is shown in Fig. 2.22. Typical components of the vacuum are seen including H2, C, a very small H2O peak, CO and N2, and CO2. F is also seen in this system. This indicates that some component in the chamber is outgassing F, which may act as a contaminant in the deposition of thin films. The impurities and vacuum pressures do increase from these base values when the deposition chamber is exposed to the ion beam cham- ber and ion beam. Another RGA residual gas mass spectrum from the deposition chamber collected while the ion beam was being operated with SiH4 will be shown and discussed in Fig. 6.4 in Chapter 6. 2.3.3 Sample Manipulation As mentioned previously, a 5-axis manipulator is used to position samples within the deposition chamber for various purposes. The manipulator itself was produced by VG Scienta, but a sample stage from Omicron was attached to the end. It has motion in x, y, and z directions as well as rotation about two axes. One rotation axis is aligned with the insertion axis of the manipulator and is used to face a sample towards or away from the ion beam as well as other instruments. The other rotation axis is an azimuthal rotation about the normal to the sample surface. This rotation is used to position the sample so that particular crystallo- graphic directions of the substrate can be aligned with the RHEED electron beam. The sample manipulator also provides heating capabilities to samples, which are held by sample holders that are loaded onto the manipulator. Sample holders are made of Mo and clamp the Si chips between clips, one of which is electrically iso- 79 lated from the base. Two methods of sample heating include radiative heating from a tungsten wire back heater, which is referred to as “RH”, and direct current re- sistive heating where current is passed through the Si substrate resistively heating it, which is referred to as “DH”. Heating of the substrates and samples are used for several purposes. Deposition at elevated temperatures is critical for facilitating epitaxial deposition of the films. Additionally, substrates can be prepared in situ to have clean surfaces by flash annealing them using the DH method for heating. Flash annealing involves rapidly increasing the temperature of the substrate typi- cally from approximately 600 ◦C to as high as 1250 ◦C in a few seconds. This high temperature anneal step removes oxide from the substrate surface and reconstructs the surface to form the well known Si(100) (2×1) dimer rows. The oxide typically present on substrates introduced into the vacuum chamber in this work was either a native SiO2 or a deliberately grown oxide. After remaining at a high temperature for typically 10 s, the substrate is cooled quickly back to 600 ◦C. This flash annealing method produces flat, nominally clean surfaces on the Si(100) substrates. A wiring diagram on a photograph of the sample stage on the manipulator is shown in Fig. 2.23 indicating the current path for the two heating methods. A Si(100) substrate with dimensions of approximately 10 mm wide by 4 mm long is seen mounted in a sample holder that is being held on the manipulator. The sample is mounted in the holder ex situ using two Mo foil clips that clamp either end of the substrate chip, as mentioned. The Si(100) substrate is glowing due to it being at a temperature of roughly 1000 ◦C. This demonstrates the DH heating method where current is being driven through the sample to resistively heat it. The DH 80 Figure 2.23: Photograph and wiring diagram for heating a Si(100) substrate on the sample manipulator. The Si(100) substrate shown is heated to about 1000 ◦C using the DH power supply to drive current through the substrate. The DH supply connects to “finger” contact on the left of the sample holder and the grounded base of the sample stage. The current and voltage of the DH supply used for sample flashing to 1200 ◦C are IDH ≈ 9 A, and VDH ≈ 5 V, respectively. The RH power supply connects to terminals of a tungsten wire heater beneath the sample, one of which is grounded to the sample stage base. The RH supply uses a current and voltage of IRH ≈ 8 A, and VRH ≈ 12 V, respectively, for sample degassing at about 600 C. power supply is represented on the left side of the diagram with the positive lead being connected to the “finger” contact on the left side of the sample stage. The finger is isolated from the rest of the sample stage and contacts the left sample holder mounting clip, which is also isolated from the rest of the sample holder. Therefore, when a voltage, VDH is applied to the finger, a current, IDH flows through the substrate. The other (right) side of the substrate is grounded through the sample holder to the sample stage so that the current flows to the negative terminal pictured on the right. The current and voltage of the DH supply used for sample flashing to 1200 ◦C are IDH ≈ 9 A, and VDH ≈ 5 V, respectively. The RH tungsten back heater shares the negative terminal on the right side of the sample stage in Fig. 2.23. The positive lead of the RH power supply represented 81 on the right connects to the positive terminal above the negative terminal at the right. The positive terminal of the RH heater is isolated from the sample stage, unlike the negative one. The tungsten wire is connected to these two terminals and runs beneath the sample location in a snaking pattern. The RH supply uses a current and voltage of IRH ≈ 8 A, and VRH ≈ 12 V, respectively, for degassing samples at approximately 600 C. The RH heater is also typically required to pre-heat the Si substrate before the DH method can be used. This is because the resistance of the substrate is often too large at room temperature to begin flowing current. As the sample is heated by the RH back heater, the resistance of the substrate drops and the Si can then be heated using DH. Photographs of the manipulator and sample stage are shown in Fig. B.12 in Appendix B. When depositing a 28Si film while simultaneously heating the substrate, a different wiring setup is used in order to be able to monitor the ion beam current on the sample. In this setup, the RH leads are disconnected and the DH power supply is isolated from ground with its leads remaining connected to the sample stage as shown in Fig. 2.23. The input of a picoammeter, which is itself grounded, is then connected to the negative terminal of the sample stage. This setup allows for the ion beam current to flow from the sample through the picoammeter to be measured. There is some leakage current of typically about -10 µA registered using this setup, which produces an offset in the ion current measurement. In order to achieve the sample heating procedures discussed in this section including rapidly changing the sample temperature as well as using precise depo- sition temperatures, accurate temperature measurements are important. As was 82 mentioned previously, an infrared pyrometer is used in this work to measure the temperature of samples. The pyrometers are external to the deposition chamber and view the sample surface through a window. The temperature reading of the pyrometer depends on the value of the emissivity, e, to which it is referenced. There- fore, the emissivity of the Si substrates, which is a temperature dependant quantity, needs to be determined for the temperature range being measured. An indication of the emissivity values for Si can be taken from literature, however, differences in experimental setups such as the type of window that the pyrometer views the sample through require calibration to the system being used. The pyrometer from Process Sensors, which was used for the most samples, was calibrated for the work reported here, while the Omega pyrometer was not. Experimental data and calcu- lations show that for Si at temperatures above approximately 800 ◦C, e ≈ 0.68 and below 100 ◦C, e can be as low as 0.1 [70,71]. To calibrate the pyrometer and determine e, two Si eutectics were used. These Si compounds have melting temperatures that are much lower than the Si melting temperature of approximately 1414 ◦C. By forming a Si eutectic in a small region on a Si substrate, the substrate temperature can be brought to the melting point of the eutectic while the Si substrate remains solid. To calibrate the pyrometer, it is used to monitor the temperature of the substrate while the eutectic is heated until it melts. With the temperature adjusted so that the eutectic is held at its melting point, the emissivity of the pyrometer is adjusted until the temperature that it reads matches the known melting temperature of the eutectic. This gives a calibration for the pyrometer at a specific temperature for the experimental setup being used. 83 The two Si eutectic systems that were used in this work to calibrate the Process Sensors pyrometer are Au-Si, which has a melting temperature of approximately 363 ◦C and Al-Si, which has a melting temperature of 577 ◦C. The Au-Si eutectic phase diagram is shown in Fig. 2.24 taken from Ref. [72]. The phase boundary for the liquid phase is shown as a function of Si atomic concentration as a percentage. The eutectic forms at a Si concentration of approximately 18.6 % where the melting temperature is at a minimum of approximately 363 ◦C. The Au-Si eutectic was formed here by pressing a small Au wire onto the edge of a Si substrate. The substrate was then heated in the vacuum chamber using the DH method. The substrate temperature was increased until part of the Au wire was seen to melt. This signified a Au-Si eutectic being formed due to a small amount of Si diffusing into the Au where it contacting the substrate. Upon repeated cycling of the temperature above and below the the point that the (mostly) Au wire melts, the temperature at the melting point is observed to decrease as the Si concentration equilibrates to the optimal eutectic Si concentration of 18.6 %. When this occurs, the temperature reading at the melting and freezing of the wire remains constant. This temperature is then known to be approximately the 363 ◦C melting temperature of Au-Si and the pyrometer emissivity was adjusted to match this temperature, yielding a value of e = 0.25. A similar procedure was followed for the Al-Si eutectic although instead of a Al wire, an approximately 2 mm by 2 mm square of Al approximately 500 nm thick was deposited on the Si substrate. This method was found to be less reliable than using a larger wire. For the Al- Si calibration at 577 ◦C, an emissivity value of e = 0.42 was determined. These 84 Figure 2.24: Au-Si eutectic phase diagram showing the phase boundaries as a func- tion of Si concentration. The melting temperature of a Au-Si system is lowest for a Si atomic concentration of 18.6 %. This corresponds to the eutectic melting tem- perature of 363 ◦C. (from Ref. [72]) measured values were then combined with the literate value for temperatures above approximately 800 ◦C of e ≈ 0.68 to get an approximate function of e vs. T used for experiments here. In this function, e = 0.25 is used for Si substrates temperatures between room temperature and 470 ◦C, e = 0.42 is used between temperatures of 470 ◦C and 625 ◦C, and e = 0.68 is used for temperatures above 625 ◦C. These crossover points were chosen based on both the measured e values and literature values of e for a range of temperatures [70, 71]. The calibrated pyrometer was then used to determine a calibration curve for the current applied for DH sample heating and flashing. This control curve of substrate temperature as a function of the substrate current using the DH, IDH, is shown in Fig. 2.25. To generate this control curve, the temperature of the substrate was recorded for different values of IDH using the three previously discussed values of e in their appropriate temperature ranges. These data are seen appearing in three distinct curves (circles) marked with each value of e. Note that the current is shown 85 Figure 2.25: Experimental control curve for the DH power supply used for sam- ple heating. Temperature measurements (circles) are shown vs. DH current, IDH, applied through the substrate. Note that the current axis is shown as a log scale. Temperatures are measured using an infrared pyrometer for three emissivity val- ues, 0.25, 0.42, and 0.68. These three values correspond to measurements in three temperature ranges with crossover points between the ranges at 470 ◦C and 625 ◦C. To transition smoothly between the three regions, the data is fit to a sum of three exponentials (line, Eq. (2.18)), which is then used for sample heating. This data is for a Si(100), phosphorous-doped substrate 300 µm thick with a resistivity between 7 Ω · cm and 20 Ω · cm. on a log axis. This data is for a Si(100), phosphorous-doped substrate 300 µm thick with a resistivity between 7 Ω · cm and 20 Ω · cm acquired from Virginia Semiconductor. Substrates with different specifications, especially the thickness, will have different control curves. In order to smoothly transition between the three ranges, the data is fit to a sum of three exponentials (line) given by T = 1360− (290) exp (−IDH 0.026 ) − (240) exp (−IDH 0.024 ) − (810) exp (−IDH 4.7 ) , (2.18) which shows the fit parameters. IDH is in units of A and T is in units of ◦C. 86 This fit allows for the substrate temperature to be controlled continuously using the applied DH current for both sample degassing and flash annealing. Including the uncertainties from the pyrometer calibration, the temperature readings of the substrate are estimated to have a 5 % relative uncertainty due to fluctuations in the current used for sample heating as well as temperature gradients across the sample. 2.3.4 In situ Sample Analysis After flash annealing a Si(100) substrate to clean and prepare it in situ, the surface can be inspected by RHEED. This and other capabilities including STM are used to produce 28Si thin films and analyze them in situ. The RHEED system (STAIB Instruments) used here typically uses an electron energy of approximately 15 keV, and an electron gun filament current of approximately 1.55 mA. As men- tioned previously, Si(100) substrates are positioned using the manipulator to be able to do RHEED during deposition. The sample can be rotated about its azimuth normal to the surface to align the RHEED electron beam along different crystal- lographic directions, but typically the beam was aligned with the 〈110〉 direction. The other axis of rotation of the manipulator can be used to adjust the angle of incidence of the RHEED electron beam, and typically, an angle of approximately 2◦ was used for capturing diffraction patterns although angles as high as 5◦ were sometimes used. The RHEED diffraction patterns are captured using a camera in a dark enclosure that images the patterns produced by diffracted electrons on a phosphorous screen. An example RHEED diffraction pattern of a flashed Si(100) surface showing 87 diffraction corresponding to the (2×1) surface reconstruction is shown in Fig. 2.26, which was acquired with the surface at approximately 600 ◦C. The observed semi- circle of spots is the zeroth Laue zone, which is typically imaged. The outer diffrac- tion spots are due to the bulk Si diffraction and the inner two spots are due to the reconstruction. The reflected specular spot appears in the middle. As this pattern is a representation of reciprocal space, the distance from the central (00) spot to the bulk Si (11) spots are inversely related to the atomic spacing of atoms on the Si(100) surface. Knowing the physical spacing of these spots on the screen and the working distance to the sample, the Si lattice constant can be extracted. The spacing of Si(100) (2×1) dimer rows on the surface is double the atomic spacing, and so the (2×1) spots appear at a distance from the central spot that is half of that to the bulk spots. These diffraction spots and a lack of a diffuse background indicate a crystalline surface that is fairly clean, i.e. free of oxide, although RHEED is not sensitive to trace impurities. The observed diffraction patterns in RHEED are the result of the combined diffraction from a macroscopic area on the sample (≈ 0.25 mm2) due to an elongated electron beam spot on the surface of several mm resulting from the low incidence angle. Typically, diffraction from a flat surface produces long vertical streaks in the pattern, which can be seen in Fig. 2.26. Stronger intensity spots in the middle of the streaks are the result of the finite and relatively large size of atomic terraces (e.g. 100 nm) on the Si(100) surface with the streak length being inversely proportional to the terrace size. Kikuchi lines are also seen running diagonally through the pattern, which are due to surface resonances. 88 Figure 2.26: RHEED diffraction pattern of a clean, flash annealed Si(100) substrate showing bulk Si diffraction spots and (2×1) reconstruction spots. A lack of a bright diffuse background and the presence of the (2×1) spots indicates that the native oxide has been removed. This pattern was acquired with the substrate at about 600 ◦C and the electron beam in the 〈110〉 direction. RHEED is used during deposition of 28Si films to monitor the growth structure, i.e. smooth crystalline vs. rough crystalline vs. amorphous, as a function of film thickness. Additionally, by inspecting different areas of the substrate, the location of the deposition spot, which is typically smaller than the substrate chip, can be roughly located. This is because the diffraction pattern of a deposited film is never exactly the same as the diffraction pattern of the bare Si(100) substrate. Thus the spot can be located based on the changing pattern and by comparing those patterns to the patterns of the substrate before deposition. After depositing 28Si films, samples are typically moved into the STM chamber for inspection of the film surface. As previously mentioned, the STM resides in a UHV chamber separated from the deposition chamber by a gate valve. A magnetic transfer arm was used to transfer samples between the two. The STM chamber has a typical base pressure of approximately 6.7× 10−9 Pa (5.0× 10−11 Torr) and 89 was pumped by a 500 L/s ion pump (Varian) and a Ti sublimation pump (TSP). A sample holding area in this chamber can store up to 12 samples at a time. The STM used here is an Omicron variable temperature model. STM images were typically acquired using a tip bias of -2 V for Si substrates and -1.8 V for 28Si samples. Tunneling currents used were typically 150 pA for Si substrates and 100 pA for 28Si samples. STM was used to assess the cleanliness and surface quality of Si(100) substrates prepared in situ by flashing before deposition. Scanning was typical done in an area on the surface of 100 nm by 100 nm where contaminants, residual surface oxide, adsorbates, or other surface defects are clearly visible. The quality of the visible Si(100) (2×1) dimer rows can also give an indication of the presence of contaminants. After depositing 28Si films, STM was also used to image the film surface morphology to gain qualitative information about island formation or other features as well as determine some qualitative parameters such as the total local height variation of a film surface. STM images presented in this work were acquired in collaboration with Hyun soo Kim. 90 Chapter 3 Initial Experiments Enriching 22Ne and 12C 3.1 Context and Experimental Setup Before depositing 28Si films, proof of principle experiments were conducted involving the enrichment of 22Ne, which was implanted into Si substrates, and 12C, which was deposited as thin films onto Si. These experiments helped establish the experimental practices needed for later adapting the system to 28Si deposition. Ad- ditionally, they showed that a high enough level of enrichment was possible to war- rant investment in 28Si enrichment and deposition, as can be seen on the enrichment progression timeline in Fig. 1.12 in Chapter 1. The experiments discussed in this chapter were conducted with an experimen- tal setup where the ion beamline was disconnected from the deposition chamber. A schematic of the ion beam and lens chambers in this setup is shown in the upper right section of Fig. 4.1 and is discussed further in Chapter 4. The mass-selecting aper- ture used in these experiments was a circular hole that was approximately 5 mm 91 in diameter and 16 mm thick. In this configuration, samples were located after the deceleration lenses, and they were mounted on the end of an electrical vacuum feedthrough which acted as a sample stage. A sketch of the sample stage feedthrough and sample location is shown in the blowup of the schematic of the “LC–2” setup in Fig. 4.1. Samples were mounted using a strip of conductive carbon tape on the back side of the chip. The purpose of the feedthrough was to isolate the sample elec- trically from the chamber so that the ion beam current could be monitored during beam tuning and deposition. Additional electrical feedthroughs on the sample stage were used to mount a masking element above the sample location consisting of a metal shim for collecting current. This was positioned between the sample and the path of the ion beam and had a small circular aperture directly above the sample. This mask and aperture was fixed in place for the duration of a deposition. The fixed sample aperture was approximately 3 mm in diameter and provided a mech- anism to monitor the focusing of the ion beam on the sample by maximizing the ion current detected on the sample while minimizing the current detected on the sample mask. Additionally, the mask and sample aperture allowed for precise location of the ion beam spot on the sample substrate. The feedthrough (and sample) were posi- tioned on axis with the ion beamline optics, and the sample aperture allowed for the ion beam to be tuned and steered onto the exact sample location under the aperture. A photograph of a sample mounted on the vacuum feedthrough on this intermediate sample stage with the sample mask is shown in Fig. B.11 in Appendix B. A lack of sample motion in this setup means that the ion beam was constrained to be on 92 axis, which may not have been the optimal tuning position. There was no method for sample heating in this setup. One difference between the 12C and 22Ne experiments is that, as a gas at room temperature, Ne needs to be implanted into a substrate as a means of capturing it so that the enrichment can be measured later. Ne implantation required much higher ion kinetic energies than the carbon deposition. To accomplish this, a bias voltage as high as approximately -4 kV was applied to the substrates during deposition to attract the positively charged 22Ne+ ions into the substrate. The final energy of the 22Ne ions would then be the energy due to the bias potential plus the starting energy of the ions, which was approximately 500 eV for Ne ions, for a total energy of roughly 4.5 keV. As will be discussed below, this high energy was needed to achieve an implantation depth of at least 10 nm for a reliable SIMS measurement of the enrichment. A sample bias of -4 kV was the highest bias achievable for the experimental setup used in Ne implantation experiments. Section 3.2 of this chapter will discuss the enrichment and implantation of 22Ne. Section 3.3 will discuss the enrichment and deposition of 12C. Finally, Section 3.4 will summarize those results. 93 3.2 22Ne Implantation and Characterization: Proof of Principle 3.2.1 Sample Preparation For these 22Ne implantation tests, there was very little ex situ sample prepara- tion. Substrates consisted of “lightly doped” natural abundance commercial Si(100) wafers that were cleaved by hand into approximately 1.5 cm by 1.5 cm chips. The chips were handled with clean teflon tweezers and mounted on the end of the feedthrough using a strip of carbon tape on the back of the wafer, and they were loaded into the vacuum chamber with a native oxide. No further sample prepara- tion occurred in situ. The role of these substrates is to simply be a “catcher foil” to collect the 22Ne ions. Typically, after being loaded, samples sat several day in the vacuum chamber before deposition. 3.2.2 22Ne Implantation For Ne implantation experiments, the solids-mode Penning ion source was used (see discussion and Fig. 2.9 in Chapter 2). While the gas-mode ion source should be more efficient for generating Ne ions, it was not yet purchased at the time of this experiment. Natural abundance Ne gas was used to generate mostly singly charged Ne ions with a working pressure in the ion source of ≈ 1.0× 10−2 Pa (7.8× 10−5 Torr). Sweeping the magnetic field of the mass analyzer and monitoring the ion current onto the vacuum feedthrough and sample through the fixed sample 94 Figure 3.1: A mass spectrum obtained using Ne as a working gas shows ion current peaks (circles) corresponding to atomic species of 20Ne (20 u), 21Ne (21 u), and 22Ne (22 u). The ion current is recorded while sweeping the mass analyzer current, and thus the magnetic field (top axis). Gaussian fits (line, Eq. (2.14)) to the Ne peaks are superimposed on the data. The centers of the 20 u and 22 u fits are separated by ≈ 14 σ. aperture at the end of the ion beamline, a mass spectrum of the individual molecular and atomic species is generated. Figure 3.1 shows a portion of the mass spectrum obtained using Ne as a source gas with ion current peaks (circles) at masses of 20 u, 21 u, and 22 u corresponding to atomic species of 20Ne, 21Ne, and 22Ne, respectively. The top axis shows the applied current used to sweep the magnetic field of the analyzer. The three stable Ne isotopes have a natural abundance of approximately 90.5 % 20Ne, 0.3 % 21Ne, and 9.2 % 22Ne, and the relative peak heights of the three peaks in Fig. 3.1 fairly accurately reflect this abundance. A sum of Gaussian fits (line, Eq. (2.14)) to the Ne peaks are superimposed on the data and show 95 fairly good agreement with some divergence to the low mass side in the tails of the peaks, indicating possible ballistic scattering of a small portion of the ion beam. As discussed in Chapter 2, the spacial distribution of ions in the beam is expected to be roughly Gaussian for this system. The minor isotope, 22Ne, was chosen for enrichment in these experiments because enriching in the minor isotope and rejecting the major isotope is a stronger proof of principle demonstration of the enriching power of the ion beam system than enriching the more abundant major isotope. The Gaussian fits to the mass spectrum peaks can be used to indicate the separation of the 22 u peak from the 20 u peak. From the fits, the 20 u peak has a standard deviation of σ ≈ 0.14, resulting in a separation to the major isotope 20Ne at 20 u of approximately 14 σ. The mass resolving power at the 22Ne peak of the ion beam in this configuration is derived from the mass spectrum to be m ∆m ≈ 26 (measured at 10 % of the peak height). This fairly low mass resolution and the presence of an ion current of roughly 1 nA between the peaks indicates that the beam may not be optimally focused at the mass-selecting aperture. Tuning the mass analyzer to a mass of 22 u (unified atomic mass units) results in the 22Ne ion beam passing through the mass-selecting aperture, through the focusing lenses, and to the sample for implantation. The 22Ne sample discussed below in the analysis of the sample enrichment here was implanted with the substrate at room temperature with an unknown back- ground pressure. This pressure was possibly around 6.5× 10−5 Pa (5.0× 10−7 Torr) based on later experiments. This partial pressure during growth was mostly Ne leak- age from the ion source as the base pressure of the deceleration lens chamber before 96 Figure 3.2: 22Ne stopping range (implantation depth) vs. ion energy (circles and line) based on TRIM calculations [69]. In the calculation, the ions hit a crystalline Si target with a 2 nm SiO2 surface layer. implanting this sample was ≈ 3.1× 10−6 Pa (2.8 × 10−8 Torr). The anode voltage used for generating Ne ion beams in these experiments was approximately 500 V resulting in an initial energy of the ions in the ion source of 500 eV. As mentioned previously, the sample was biased with a negative high voltage to achieve implanta- tion. The starting energy of the ions is added to the energy gained from the sample bias voltage of approximately -4 kV. The total average ion energy at the target was then approximately 4.5 keV, as discussed previously. This high energy was required to achieve an implant depth of at least 10 nm, which was needed for the SIMS mea- surements of the enrichment to separate the Ne signals of the implant from that of surface contaminants. Calculations of the stopping range of energetic ions based on TRIM [69] predict that 4.5 keV 22Ne ions will implant into Si with a native oxide at a peak implantation 97 Figure 3.3: Optical micrograph of an implanted 22Ne sample produced at room temperature at LC–2. The Si(100) substrate is seen with scribe marks used to align the chip to the expected ion beam spot location. The 22Ne implanted area is not clear but is likely to the right of the crossing of the scribe marks in an expected area of about 2 mm2. Other marks visible on the chip are due to sputtering from the SIMS measurement. depth of approximately 12.5 nm. The stopping range, i.e. the peak implantation depth, of 22Ne ions hitting a Si target (circles and line) is shown for a range of ion energies in Fig. 3.2. The Si target used in these calculations is crystalline and has a 2 nm SiO2 surface layer representing the native oxide on the substrates used in these experiments. The implantation depth varies between approximately 4 nm for an ion energy of 1 keV to approximately 14.5 nm for an ion energy of 5.5 keV. The ion current of 22Ne achieved for implantation was approximately 140 nA over an implantation area, estimated from other samples, of approximately 2 mm2. The associated ion dose for these parameters is approximately 7.7× 1016 cm2, re- sulting in an atomic concentration of 22Ne in the Si substrate of approximately 4.9× 1022 cm−3, based on TRIM calculations. This means the implantation volume in the Si substrate was roughly 50 % 22Ne. An optical micrograph of the 22Ne sample analyzed for enrichment here is 98 shown in Fig. 3.3. The Si(100) substrate is seen in the micrograph with scribe marks used to align the chip to the expected ion beam spot location just under the fixed sample aperture. It is not obvious where the 22Ne-implanted area is, although it is believed to be just to the right of the crossing of the scribe marks. Other marks visible on the chip are due to sputtering from the SIMS measurement. A total of three 22Ne samples were produced in this work, and two were analyzed for enrichment, the second of which is discussed below. 3.2.3 Enrichment Measurements via SIMS SIMS was used to determine the enrichment of two 22Ne-implanted samples, both of which are represented on the enrichment progression timeline in Fig. 1.12. For the SIMS measurement of these samples, the primary ion sputter beam was composed of Cs+ ions with a current of 40 nA and an energy of 10 keV. The sputter beam was raster-scanned over a nominal 280 µm by 430 µm area on the sample, although the actual area sputtered was considerably larger because of the finite size of the ion beam. Measurements of the counts of the major isotope, 20Ne, and the minor isotope chosen for enrichment, 22Ne, were made after each sputter cycle, producing a “depth” profile of the isotope ratios (assuming a constant sputter rate). Isotope ratios are then used to determine the isotope fractions, which indicate the enrichment level. As mentioned in Chapter 1, isotope fractions of a particular isotope are defined as the counts of that isotope divided by the total counts of the measurement. The isotope fractions of Ne can be written as zNe/Netot., where z is the mass number denoted as 20 for 20Ne counts and 22 for 22Ne counts and Netot. is 99 Figure 3.4: SIMS “depth” profile of a 22Ne-implanted sample produced at room temperature at LC–2. 22Ne was implanted into Si(100) with an energy of about 4.5 kV. Isotope ratios of 22Ne/20Ne (open diamonds and line) are shown vs. the sputter time. The second point at 63 s is the largest ratio value (183), representing the maximum enrichment in 22Ne. The inset shows the raw count rates vs. sputter time for 22Ne (triangles and line) and 20Ne (diamonds and line). the sum of 20Ne and 22Ne counts. The first 22Ne sample produced, which was also the first enriched sample of any kind produced in this work, had a lower level of enrichment than the second one. This first sample had a 22Ne isotope fraction of 84.4(10) %, seen on the enrichment progression timeline as the highest residual isotope fraction of those samples. The second 22Ne sample was also analyzed by SIMS, and the resulting “depth” profile is shown in Fig. 3.4. The isotope ratios of 22Ne/20Ne (open diamonds and line) are shown vs. the sputter depth into the sample. At a sputter time of 63 s, which likely corresponds to the peak implantation depth of the 22Ne (expected to be ap- 100 proximately 12.5 nm), the isotope ratio has a maximum value of 183. This value represents a 22Ne isotope fraction (assumes no 21Ne is present) of 99.455(36) % and a residual 20Ne isotope fraction of 0.545(36) %, as seen on the enrichment progres- sion timeline. The inset of Fig. 3.4 shows the raw count rates vs. sputter time of 22Ne (triangles and line) and 20Ne (diamonds and line) for reference. Another useful parameter that describes the measured enrichment of a sample is the isotope reduction factor, which gives the amount by which an excluded isotope is reduced from the natural abundance. This reduction factor is determined by dividing the natural abundance of an isotope of an element by the measured isotope fraction described previously. The natural abundance of an isotope is represented by az, where z again is the mass number of the isotope. For the reduction factor of the excluded isotope 20Ne, az is a20, i.e. the natural abundance of 20Ne. The reduction factor for 20Ne is thus written as a20/( 20Ne/Netot.). The measured isotope fractions discussed here give an isotope reduction factor for 20Ne of 166(1), i.e. the 20Ne in the sample is approximately 166 times lower than in natural abundance Ne. These measurements indicate a realized mass selectivity of 1785:1 for the ion beam system in this configuration. This value gives a sense of the performance of the system independent of the natural abundance of Ne. 101 3.3 12C Deposition and Characterization: First Enriched Thin Films 3.3.1 Context Solid state QI using spin states of atoms or quantum dots as qubits is lim- ited by the need for host materials that are as minimally-interacting as possible, e.g., 28Si. Another potential host material for QI devices is diamond. Natural abundance diamond has isotopes possessing a non-zero nuclear spin, and so isotopic enrichment can increase the coherence time of qubits in diamond by reducing the density of randomly fluctuating spins in their local environment. Qubits in 12C enriched diamond have received a lot of attention in this context [73]. Isotopic en- richment eliminates the 13C nuclear spins I = 1/2 present in natural abundance C (natural stable isotope abundances: 98.9 % 12C, 1.1 % 13C). Nitrogen-vacancy (NV) centers in enriched, highly pure diamond can have electron spin T2 coherence times of milliseconds [74,75]. Additionally, qubits with long nuclear spin coherence times approaching a second have been demonstrated for isolated 13C atoms coupled to NV centers in 12C diamond [73]. This chapter describes the enrichment and deposition of 12C films as a potential host for qubits, but it also serves as a demonstration for the production of 28Si films. Significant portions of the data and analysis presented in this section was previously published in Ref. [76]. The 12C reported here has an enrichment in the solid state comparable to that of commercially available methane (up to 99.999 % 12CH4) used as a source 102 gas to grow enriched diamond by CVD [73, 77, 78], where enrichment and purity may be diminished in processing. CVD diamonds grown from enriched 12CH4 can suffer from incorporation of excess N (14N has nuclear spin I = 1) that can also limit coherence times. Several groups growing enriched 12C diamond by CVD have made efforts to reduce the atomic concentration of paramagnetic impurities like N below 1× 10−9 cm−3 [74, 75, 78, 79]. The mass selecting method used here isolates the 12C to avoid contaminants such as N and O. N may still be physisorbed from the background vacuum but vacuum improvements and heated substrates could mitigate this liability. As previously mentioned, some work has been done depositing enriched C materials including 12C, 28Si12C and 12C14N [56, 57, 80]. The numerous research efforts utilizing enriched diamond and other C allotropes [74,81,82] can also benefit from the availability of 12C films. 3.3.2 Sample Preparation Similar to the Ne samples, for these 12C deposition tests, there was very little ex situ sample preparation. Substrates consisted of “lightly doped” natural abundance commercial Si(100) wafers that were cleaved by hand into approximately 1.5 cm by 1.5 cm chips. Clean teflon tweezers were used to mount these substrates onto the end of the same electrical vacuum feedthrough used in Ne implantation experiments using a strip of carbon tape on the back of the wafer. Substrates were loaded into the vacuum chamber with a native oxide. No further sample preparation occurred in situ before the deposition of 12C films. Similar to the substrates used in the Ne experiments, these substrates serve as a simple “catcher foil” to collect the 12C ions 103 as a film in these experiments. Typically, after being loaded, substrates sat several days in the vacuum chamber before deposition occurred. 3.3.3 Deposition of 12C In this experiment, natural abundance CO2 is used as the carbon source to gen- erate 12C ions and grow 12C films. CH4 gas was also tested but was found to produce a lower ion beam flux than CO2 (see Fig. 2.10 in Chapter 2). The solids-mode Pen- ning ion source was used in these experiments (see Fig. 2.9 and discussion in Chapter 2). The working pressure of the ion source used in these experiments was between approximately 1.1× 10−3 Pa (8.0× 10−6 Torr) and 1.3× 10−3 Pa (1.0× 10−5 Torr). CO2 molecules are cracked and ionized by the plasma in the ion source to gener- ate the ion beam. By scanning the magnetic sector mass analyzer and monitoring the ion current through the mass-selecting aperture onto the sample stage, a mass spectrum of the individual molecular and atomic species is generated. Figure 3.5 shows a portion of the mass spectrum acquired using CO2 as a source gas that is representative of the ion beam conditions used to deposit samples discussed in this section. Ion current peaks (circles) are visible at masses of 12 u, 13 u, 14 u, and 16 u corresponding to atomic species of 12C, 13C, 14N, and 16O, respectively. The 14N peak is due to a small amount of residual nitrogen background pressure present in the ion source chamber. The top axis shows the applied current used to sweep the magnetic field of the analyzer. The relative peak heights of the 12C and 13C peaks fairly accurately reflect the natural abundance of stable C isotopes. The 12 u and 13 u mass peaks are fitted to Gaussian functions (line, Eq. (2.14)), which are super- 104 Figure 3.5: A mass spectrum obtained using CO2 as a working gas shows current peaks (circles) corresponding to atomic species of 12C (12 u), 13C (13 u), 14N (14 u), and 16O (16 u). The ion current is recorded while sweeping the mass analyzer current, and thus the magnetic field (top axis). Gaussian fits (line, Eq. (2.14)) to the 12C and 13C peaks with 95 % confidence bands are superimposed on the data. The centers of the fits are separated by ≈ 4 σ. imposed on the data along with 95 % confidence bands (shaded area). The peaks are fairly symmetric and the data can be seen to match the form of a Gaussian, indicating minimal scattering of ions by lens elements. The mass resolving power of the ion beam at the 12C peak is derived from the fits to the mass spectrum to be m ∆m ≈ 13.5 (measured at 10 % of the peak height). This mass resolution is fairly low and the presence of a measurable ion current of roughly 0.4 nA between the 12 u and 13 u peaks indicates that the beam may not be optimally focused at the mass-selecting aperture. The separation of the 105 peaks can also be estimated from the Gaussian fits, which show that the center of the 13C peak is ≈ 4 σ away from the center of the 12C peak. The overlap of the 13 u peak on the 12 u peak due to this high level of separation can be determined using the parameters G12(m) and G13(m). These are the values (calculated at mass m in units of u) of the Gaussian fits to the current peaks at 12 u and 13 u. The geometric selectivity of the system for 12C relative to 13C for the conditions used when producing a 12C beam is then determined from the ratio of the Gaussian fit functions at 12 u, given by G12(12) G12(12) +G13(12) = 99.9999(3) % 12C, (3.1) where m = 12 for the above parameters signifying that the values of the Gaussian fits are calculated at a mass of 12 u. The uncertainty is found in a similar manner using the 95 % confidence bands of the fits. By further tuning the ion beam and adjusting the parameters of the electrostatic lenses, better isolation of the masses can be achieved in this setup at the expense of beam flux. This analysis sets an upper bound for a sample enrichment grown under these conditions. By tuning the magnetic field to the center of the peak at 12 u, 12C ions are transmitted with a high degree of selectivity to the target substrate for deposition. Thus, the enrichment should not be limited by the geometric mass selectivity of the beamline but by other sources of contamination, e.g. neutral 13C from un-ionized source gas diffusing to the sample during deposition or gas scattering of ions. 12C samples were grown with typical beam currents ranging from 600 nA to 900 nA over a ≈ 1 mm2 beam spot, resulting in deposition rates between approxi- 106 Figure 3.6: Optical micrograph of a12C sample deposited at room temperature at LC–2. The Si(100) substrate is seen with the 12C deposition spot near the upper center appearing darker. The spot is roughly 1.3 mm in diameter. Lighter dust is also visible on the substrate surface. Inset is a lower magnification image with the edges of the chip visible. mately 0.5 nm/min and 0.75 nm/min. An optical micrograph of a 12C sample after deposition is shown in Fig. 3.6. The Si(100) substrate is seen with the 12C deposition spot visible near the upper center appearing darker than the substrate. The spot is roughly 1.3 mm in diameter. Lighter dust is also visible on the substrate surface from handling the chip. Inset is a lower magnification image of the sample with the edges of the chip visible giving an indication of the chip size. The sample discussed in the analysis of this section was deposited with an average ion energy of ≈ 560 eV in a background pressure of ≈ 2.1× 10−5 Pa (1.6× 10−7 Torr). This partial pres- sure during growth was mostly CO2 leakage from the ion source and the typical base pressure for the deposition chamber is ≈ 1.3× 10−7 Pa (1.0× 10−9 Torr). A total of three 12C samples were produced in this work, and two samples were analyzed for enrichment, one of which is the focus of the analysis below. SEM was used to inspect one of the deposited 12C films in cross-section and 107 Figure 3.7: SEM cross-sectional micrograph of a 12C sample deposited at room temperature at LC–2. The top layer is ≈ 27 nm of Au-Pt deposited for imaging. The middle layer is the deposited 12C film which is ≈ 66 nm thick. Below that is the crystalline Si(100) substrate. measure the thickness. Figure 3.7 is a cross-sectional SEM micrograph of a cleaved edge of the analyzed 12C sample. The micrograph shows the crystalline silicon substrate on the bottom that has diagonal lines going from upper left to lower right, which are a consequence of the cleaving. On top of the silicon is the deposited 12C film, which also shows lines running vertically through the entire thickness of the film. These lines suggest continuity and uniformity throughout the material that would not be expected if the sample is more polycrystalline consisting of multiple grains from the bottom to the top of the film. In the area of the sample shown in the micrograph, the film is ≈ 66 nm thick. Above the 12C film is ≈ 27 nm of Au-Pt deposited for imaging. The upper and lower edges of the carbon film are nearly parallel, indicating smooth growth. While there are no obvious signs that the film 108 is amorphous, it is not expected that a crystalline film can be deposited at room temperature on a substrate with a surface oxide layer. 3.3.4 Enrichment Measurements via SIMS The enrichment in 12C and the residual 13C in the carbon films were deter- mined using SIMS. One of these analyzed samples is represented on the enrichment progression timeline in Fig. 1.12 in Chapter 1. For the SIMS measurements, a Cs+ primary ion beam with an impact energy of 20 keV was raster-scanned to sputter a 150 µm by 150 µm area of the sample. Secondary negative ions of 12C and 13C sput- tered from the surface were alternately directed by magnetic peak-switching into a secondary electron multiplier where the counts were recorded. Under the measure- ment conditions, the instrument has a mass resolving power m ∆m ≈ 5000, allowing it to distinguish between 13C and 12CH peaks, which are separated by approximately 0.0045 u. A set of measurements were made by pre-sputtering the target area with an 8 nA primary ion current and then recording counts of 40 isotopic ratio pairs with a sputter beam current of 23 pA. Isotope fractions of 12C and 13C were determined from the averages of the 40 ratios by calculating zC/Ctot., where z is the mass number denoted as 12 for the 12C average counts and 13 for the 13C average counts. Ctot. is the sum of the 12C and 13C average counts. The isotope fractions of 12C and 13C from eight sets of SIMS isotopic ratio measurements after successive pre-sputter cycles are presented in Fig. 3.8. The 12C (squares and line) and 13C (circles and line) isotope fractions are shown in a semi-log plot of the “depth” profile vs. the cumulative pre-sputter 109 Figure 3.8: SIMS “depth” profile of a 12C sample deposited at room temperature at LC–2. Isotope fractions of 12C (squares and line) and 13C (circles and line) representing eight successive data runs as the primary SIMS beam sputtered through the film are shown vs. the pre-sputter time. The inset shows 40 measurements of 13C/12C ratios (open circle and line) vs. sputter time, which are averaged (line) to determine the values of the isotope fractions of the sixth points at a pre-sputter time of 540 s. The middle six points between 180 s and 900 s were averaged for the final enrichment and 13C isotope fraction of 39.2(13) ppm. Also shown is the natural abundance value for 13C (dashed line) as a reference. time, i.e. different depths into the film. The inset shows one set of 40 isotopic ratio measurements of 13C/12C (open circles and line) vs. sputter time into the film for the sixth data points of the isotope fractions at a pre-sputter time of 540 s. The average (line) of the 40 measurements is used to determine the 13C and 13C isotope fractions. The natural abundance of 13C (dashed line) is also shown for reference highlighting the reduction of 13C. To determine the average isotope fractions of the film, only the middle six data between pre-sputter times of 180 s and 900 s were 110 averaged, excluding the first and last data points. The first datum at 90 s shows a lower enrichment that is likely due to natural abundance carbon contamination on the film surface. The last datum at 900 s has lower enrichment but also corresponds to a dramatic drop in absolute 12C counts indicating that the interface between the carbon film and the Si(100) substrate had likely been reached. This depth was later estimated to be roughly 100 nm. For the remaining six points, the average measured 13C/12C ratio is 3.79(5)× 10−5. After accounting for an instrumental mass fractionation for carbon, the average measured 12C isotope fraction in the film was determined to be 99.99608(13) %. The average residual isotope fraction of 13C in the film was measured to be 39.2(13)× 10−6 or 39.2(13) ppm, as seen in the enrichment progression timeline. The isotope reduction factor for 13C based on the isotope fraction is written as a13/( 13C/Ctot.). The isotope fraction of the measurement here represents an isotope reduction factor for 13C of 270(20), i.e. the 13C in the film is approximately 270 times lower than in natural abundance C. The data from Fig. 3.8 is also shown on a linear scale and over a smaller isotope fraction range in Fig. 3.9 to highlight the structure of the “depth” profile. Here, the 12C isotope fractions (squares and line) corresponding to the left axis and 13C isotope fractions (circles and line) in ppm corresponding to the right axis are shown again vs. the pre-sputter time. Small changes in the isotope fractions throughout the film are clearly visible with the sixth points at a pre-sputter time of 540 s representing the highest local enrichment within the film. As mentioned, the first datum at a pre- sputter time of 90 s and the last datum at 1260 s (crosses) were discarded when the averages were determined. The middle six points between 180 s and 900 s represent 111 Figure 3.9: SIMS “depth” profile of a 12C sample deposited at room temperature at LC–2. The data from Fig. 3.8 is shown on a linear scale. Isotope fractions of 12C (squares and line, left axis) and 13C (circles and line, right axis) representing eight successive data runs as the primary SIMS beam sputtered through the film are shown vs. the pre-sputter time. The middle six points between 180 s and 900 s were averaged for the final enrichment. The outside points of the 12C isotope fractions (crosses) at 90 s and 1260 s were discarded due to surface contamination and a loss of signal moving into the Si(100) substrate, respectively, represented by vertical dashed lines. The average 13C isotope fraction of the middle six points is 39.2(13) ppm. the bulk of the deposited film, represented by the vertical dashed lines, with surface contamination and the Si(100) substrate on either side. Each error bar is derived from the standard deviation of the mean of the 40 individual measurements at each data point. The enrichment of this sample is comparable to that of CVD grown 12C diamonds used for NV center QI experiments [74,78]. A second 12C sample was also analyzed by SIMS as a check on the reproducibility, and it was found to have a nearly identical 12C enrichment. These 12C enrichment measurements represent a 112 realized mass selectivity of 276:1 for 12C in this system. This value gives a sense of the performance of the system independent of the natural abundance of C. The 12C realized mass selectivity is less than that achieved for 22Ne, but that may be due to the estimated spatial peak separation between 12C and 13C at the mass-selecting aperture (12.4 mm) being smaller than that between 20Ne and 22Ne (14.7 mm). As mentioned above, it is believed that the limiting factor for the enrichment of these samples is not the ultimate mass selectivity of the ion beam system, but rather other sources of unwanted isotopes such as background CO2 gas diffusing from the ion source. Therefore, the contribution to the 13C concentration from this CO2 partial pressure is considered. Using the deposition conditions of a CO2 partial pressure of 2.1× 10−5 Pa (1.6× 10−7 Torr) and a natural 13C abundance in the gas of 1.1 %, an incorporation fraction, s, of about 0.037 is needed to account for the measured isotope fraction of 13C in the film of ≈ 39 ppm. Representative sticking coefficients found in the literature range from values that are similar in magnitude to several times lower [83]. The former case provides a reasonable ex- planation for the 13C observed while the latter may indicate an additional source of contamination such as scattered ions in the beam path. These sources of 13C can be mitigated by reducing the background pressure during deposition. For 13C due to adsorption from the vacuum, every order of magnitude decrease in background pressure would correspond to an order of magnitude increase in the enrichment of the deposited material. Reducing the background gas in the above estimate to 2× 10−6 Pa (1.5× 10−8 Torr) would lower the isotope fraction of 13C to ≈ 4× 10−6 (4 ppm) under the same deposition conditions. An improved vacuum would also 113 reduce chemical impurities incorporated during deposition, which could be param- agnetic. Increasing the beam flux and therefore deposition rate will further reduce the effect of unwanted background gases adsorbing into the films by reducing the relative 13C gas flux compared to the deposition rate. Time-of-Flight SIMS (TOF-SIMS) was also used to look at chemical contami- nants present in a sample similar to the one analyzed for enrichment. All signals of organic and inorganic species being monitored for analysis were near the noise floor of the measurement and at least 100 times lower than the 12C signal. The derived concentrations of chemical impurities in the film are unreliable because the sample was not being efficiently sputtered and ionized for detection. Any signal from O adsorbed during deposition from a CO2 background partial pressure is below the detection of this measurement. Further assessment needs to be done to determine the concentrations of spin and chemical impurities present in these films. 3.4 Chapter 3 Summary: Outlook for 28Si This chapter demonstrated successful proof of principle enrichment by im- planting the minor isotope 22Ne enriched with an isotope fraction of 99.455(36) % into Si as well as depositing thin films of 12C enriched to an isotope fraction of 99.9961(4) % onto Si using the mass selected ion beam system. A total of three 22Ne and three 12C were produced. The isotope fractions of 20Ne and 13C from these measurements can be seen in the enrichment progression timeline in Fig. 1.12. Re- alized mass selectivities of 1785:1 for 22Ne (selection between masses separated by 114 2 u) and 276:1 for 12C (selection between masses separated by 1 u) were achieved. While these selectivities are not directly translatable to 28Si because the peaks of Si are spatially closer than for Ne or C, these experiments do show that very high levels of in situ enrichment are possible in thin film deposition using this system. Similar levels of 28Si enrichment are likely possible, especially if background gas adsorption has a more significant effect on the realized sample enrichment than does the mass selectivity. 115 Chapter 4 28Si Thin Film Deposition and Characterization Phase I: In Situ Enrichment 4.1 Introduction 4.1.1 Context In this chapter as well as Chapter 5, the ion beam deposition of 28Si films enriched in situ is discussed as well as the characterization of their properties, par- ticularly their enrichment level. Chapter 3 demonstrated successful proof of principle enrichment experimenubots by implanting 22Ne enriched to 99.455(36) % into Si as well as depositing thin films of 12C enriched to 99.9961(4) % using the mass selected ion beam system. Realized mass selectivities of 1785:1 for 22Ne (selection between masses separated by 2 u) and 276:1 for 12C (selection between masses separated by 1 u) were achieved. Because the 28 u and 29 u mass peaks are spatially closer together at the mass-selecting aperture than the Ne or C peaks, the selectivity for 116 Si deposition is lower than these values. Nevertheless, these experiments showed that high levels of isotopic enrichment can be achieved with this method, laying the groundwork for adapting the system for 28Si deposition. As stated in Chapter 1, enrichment and thin film deposition of 28Si is pursued here with the objective of producing high-quality enriched material for Si based solid state quantum computing. 28Si of sufficiently high quality (i.e. high enrichment, crystallinity, and purity) provides an ideal solid state environment to host qubit spins. The minimal interactions between 28Si and nuclear and electron spins of qubits result in a level of isolation akin to a trapped atom in a vacuum chamber. This leads to extremely long coherence times which have earned 28Si the moniker of “semiconductor vacuum” [9]. Unwanted deviations from ideal 28Si material can be classified as three types of defects: isotopic defects, structural defects, and chemical defects. Controlling and limiting these defects is critical for successful integration of 28Si into quantum computing architectures. The 28Si materials goals of this work are discussed in Chapter 1 and can be restated as follows: (1) high enrichment in 28Si with a residual 29Si isotopic concentration less than 50 ppm, (2) single-crystalline and smooth epitaxial structure with a low dislocation density below 1× 106 cm−3, and (3) high chemical purity including C and O with atomic concentrations below 2× 1015 cm−3. These are believed to be the criteria needed for the 28Si to be comparable to elec- tronics grade natural abundance Si as well as the enriched Si currently available 117 in the QI research community. The bulk of that 28Si is produced by the Interna- tional Avogadro Coordination [32], which has a residual 29Si isotopic concentration as low as 50 ppm. Producing 28Si with 29Si isotopic concentrations as low as 1 ppm is necessary to enable a robust and systematic study measuring electron coherence times vs. 29Si concentration in the single spin regime and compare it to theoretical predictions (see Fig. 1.9), as discussed in Chapter 1 [12]. The experiments producing 28Si discussed in this chapter and Chapter 5 rely not only on prior work using this ion beam deposition system, such as the Ne and C experiments of Chapter 3 and previous thin metal film deposition experi- ments [59], but also on previous work by other groups that have deposited Si via an ion beam [43, 48–51]. The results from these groups were reviewed in Chap- ter 1 showing that they demonstrated both enrichment in 28Si to approximately 99.9982 % in one experiment [43] as well as epitaxial deposition using hyperthermal energy ions and a range of substrate temperatures [51]. The experiments described in this chapter and Chapter 5 seek to use process- ing methods that are both common (e.g. vacuum deposition, sample heating) and fairly unique (e.g. mass selected ion beam deposition) to engineer the properties, such as enrichment in 28Si and chemical purity, and structure (crystallinity) of Si thin films. Characterization methods including SIMS for assessing sample enrich- ment and chemical purity, STM, RHEED, and SEM for inspecting the film surface and crystallinity, TEM to inspect the bulk crystallinity of films, and XPS for de- tecting chemical impurities are used to assess the 28Si films in terms of the materials goals and guide the experimental adjustments needed to improve their quality. The 118 experiments discussed in this chapter focus on achieving very high levels of 28Si enrichment. The experiments of Chapter 5 will focus on maintaining a high en- richment while assessing and improving the chemical purity and crystallinity of 28Si samples. 4.1.2 Experimental Configurations for 28Si Deposition Three distinct experimental configurations were used for deposition of 28Si films in this work. Experiments involving the first two will be discussed in this chapter, and experiments involving the third one will be discussed in Chapter 5. These experimental configurations are defined partly by the location of samples during deposition at three positions in the vacuum system. Additionally, they are defined as a chronology of initial, intermediate, and final experimental configura- tions with materials characterizations and subsequent experimental improvements occurring between segments of 28Si deposition at each one. The final configuration is the final (last) setup used in this work, although not necessarily the final setup used in the larger enriched Si project of which this work is a part. These three setups are illustrated in Fig. 4.1, which shows top down schematics of the ion beam chamber, deceleration lens chamber, and deposition and analysis chamber in the three experimental configurations used to deposit 28Si samples. Schematic drawings of these chambers were previously shown in Fig. 2.1 in Chapter 2 and described in detail there. Highlighted here are the three sample locations used in each of the setups, which are discussed further below. The initial experimental configuration was one in which samples were located 119 Figure 4.1: Schematic top down drawings of the ion beam chamber, lens chamber, and deposition chamber experimental configurations used to deposit 28Si samples. The top left shows the setup for Sample Location 1 in the ion beam chamber (IC–1). This setup consists of the ion beam with the sample located on a feedthrough at the end of the beamline, shown in the blowup. The top right shows the setup for Sample Location 2 in the lens chamber (LC–2). This setup consists of the ion beam chamber connected to the lens chamber with the sample located on a feedthrough at the end of the deceleration lenses, shown in the blowup. The bottom shows the setup for Sample Location 3 in the deposition chamber (DC–3). This setup consists of the ion beam and lens chambers connected to the deposition chamber with the sample located on the manipulator, shown in the blowup. 120 just after the mass analyzer and mass-selecting aperture at the end of the ion beam chamber. This is the location corresponding to the schematic setup shown in the upper left section of Fig. 4.1 marked as “Sample Location 1: Ion Beam Chamber”, which will be referred to as IC–1. The sample was mounted on an electrical vac- uum feedthrough for deposition, which is shown in the transparent blowup of this schematic. It consists of an isolated metal rod connecting two sides of a vacuum flange. For the samples deposited at IC–1, the ion beamline was separated from the other vacuum chambers just after the mass-selecting aperture. Initially this setup was designed to maximize the total ion current onto the sample by position- ing the sample before the electrostatic deceleration lenses. Additionally, this simple configuration facilitated changing the mass-selecting aperture relatively quickly to test apertures with different dimensions in an effort to optimize ion beam fluence through it. This setup also had a higher background pressure during deposition due to the beamline being disconnected from the rest of the vacuum system, which nor- mally differentially pumps the sample location. Only a single turbo pump, marked in Fig. 4.1, pumps this setup. Next, the intermediate experimental configuration was used to deposit 28Si samples in the deceleration lens chamber. This is the sample location correspond- ing to the schematic setup shown in the upper right section of Fig. 4.1 marked as “Sample Location 2: Lens Chamber”, which will be referred to as LC–2. In this configuration, the ion beam chamber was reconnected to the deceleration lens cham- ber (separated by a gate valve), and the samples were placed on a new sample stage just after the lenses. This sample stage consisted, in part, of an electric feedthrough 121 to mount the sample, which is shown in the transparent blowup in this schematic. This setup was used for almost all the samples deposited at room temperature. It enabled better control of the ion beam by providing a means of focusing the beam using the deceleration lenses. The lens chamber also has additional vacuum pump- ing in the form of a second turbo pump, marked in Fig. 4.1 which differentially pumped the sample location compared to the ion beam chamber. This resulted in depositions at lower background pressures. This setup is also the same one used in the Ne and C deposition experiments from Chapter 3. Depositing samples at LC–2 was advantageous because it provided relatively easy access to the sample stage. This enabled moderately quick sample exchanges as well as the ability to perform quick modifications to the sample stage in efforts to optimize the deposition process. Photographs of the experimental setups for IC–1 and LC–2 are shown in Fig. B.9 and B.10 in Appendix B. Lastly, the final experimental configuration used for depositing 28Si in this work is one where the entire ion beamline, including the ion beam chamber and the deceleration lens chamber, is connected to the deposition and analysis chamber. This is the sample location corresponding to the schematic setup in the lower sec- tion of Fig. 4.1 marked as “Sample Location 3: Deposition Chamber”, which will be referred to as DC–3. In this setup, samples were placed on the 5-axis manip- ulator for deposition of 28Si, shown in the transparent blowup of this schematic. This configuration enabled lower background pressures, sample heating, and use of analytic instruments, and it is discussed further in Chapter 5. This lower schematic in Fig. 4.1 also marks the approximate locations of samples on the full combined 122 system for each of the three experimental configurations for depositing 28Si. In Chapter 1, the timeline progression of the best sample enrichment values achieved throughout this work using the ion beam deposition system was shown in the enrichment progression timeline in Fig. 1.12. This timeline will be discussed throughout this chapter as well as Chapter 5 in terms of the experimental changes that led to improvements in the 28Si enrichment. A version of this timeline showing the isotope fractions measured by SIMS of 29Si (squares) and 30Si (triangles) vs. de- position date for just the 28Si samples is presented in Fig. 4.2 along with indicators of the sample location during deposition. As mentioned in Chapter 1, isotope fractions of a particular isotope are defined in a SIMS measurement as the average detected counts of that isotope divided by the total average counts of the measurement. The isotope fractions of Si are written as zSi/Sitot., where z is the mass number denoted as 29 for 29Si average counts and similarly for 30Si and 28Si and Sitot. is the sum of 28Si, 29Si, and 30Si average counts. Uncertainties in the isotope fractions are derived from uncertainties in the SIMS measurements of those samples. Nine 28Si samples out of a total of 61 produced in this work are represented on this timeline. Each of the nine samples were the most highly enriched of any samples produced up to that point on the timeline, that is, they represent new record enrichments for 28Si samples achieved for this work. One can see that overall, the 29Si isotope fraction was reduced from 2822(18) ppm in the initial sample, down to a minimum of 127(29) ppb with an overall enrichment in 28Si of 99.9999819(35) % for the most highly enriched sample produced in this work. This level of enrichment exceeds that of all other known sources of 28Si and will be discussed in Chapter 5. This chapter 123 Figure 4.2: Enrichment progression timeline. A timeline of the progression of the lowest residual isotope fractions of 29Si (squares) and 30Si (triangles), as measured by SIMS. These were achieved for 28Si samples deposited over approximately three and one half years. Groups of samples are labeled based on their deposition loca- tions, IC–1 in the ion beam chamber, LC–2 in the lens chamber, and DC–3 in the deposition chamber. will focus on the enrichment of the samples in the first two sections of this timeline. In total, 21 28Si samples were deposited at IC–1 and LC–2 combined. Section 4.2 of this chapter will discuss samples deposited at IC–1 in the initial experimental configuration and SIMS measurements of their enrichment. Section 4.3 will discuss samples deposited at LC–2 in the intermediate configuration, SIMS measurements of their enrichment, and initial characterizations of their crystallinity by TEM and chemical purity by SIMS and XPS. Finally, Section 4.4 will give a brief summary of this chapter and the achieved enrichment values in the context 124 of the enrichment progression timeline. Some data presented in this chapter was previously published in Ref. [84]. 4.2 Si Deposition Proof of Principle: Ion Beam Chamber Samples 4.2.1 Experimental Setup This section discusses the initial deposition of 28Si samples in an experimental configuration with the sample located in the ion beam chamber at IC–1 during de- position. A schematic of the ion beam chamber in this setup is shown in the upper left section of Fig. 4.1. Depositing 28Si at IC–1 was a proof of principle experiment for the adaptation from depositing enriched C films to depositing enriched Si. One experimental change implemented for the transition from C deposition to Si deposi- tion was the initial use of the gas-mode Penning ion source described in Chapter 2. This source was designed to more efficiently crack and ionize gas to generate an ion beam as opposed to ions being generated by sputtering a solid target. Using this ion source, an increase in the mass analyzed ion beam fluence was achieved from an average of 0.55 µA for 12C depositions to an average of 0.92 µA for the initial 28Si samples described in this section. Additionally, the gas-mode source allows a reduction in the source working pressure of nearly an order of magnitude from ≈ 1.3× 10−3 Pa (1.0× 10−5 Torr) used for the 12C deposition to ≈ 1.7× 10−4 Pa (1.3× 10−6 Torr) for better integration with UHV deposition environments. Samples at IC–1 were located immediately after the mass-selecting aperture, 125 and they were mounted on the end of an electrical feedthrough using a strip of conductive carbon tape on the back of the chip. This feedthrough and the sample location of IC–1 are shown in Fig. 4.1 in the blowup of the upper left section of the figure. A ceramic electric break was used to isolate the mounting flange of the electric vacuum feedthrough from the high voltage ion beam chamber. The feedthrough and sample are both on axis with the ion beamline optics, but there was no sample motion available in this configuration. Therefore, the ion beam must be precisely tuned to be on axis as well. The purpose of this feedthrough was to isolate the sample electrically from the chamber so that the ion beam current could be monitored during beam tuning and deposition. A photograph of a sample mounted on the vacuum feedthrough after a 28Si deposition is shown in Fig. B.9 in Appendix B. Using a small feedthrough to mount the sample in this way is advantageous in that its simplicity allows for quick and easy sample loading and unloading. However, it is disadvantageous in that there is no way to shield the sample substrate before deposition, which means that it is exposed to and accumulates material from the ion beam during initial tuning procedures and sweeps of the ion beam mass spectrum. Additionally, no sample heating capabilities are available when using this setup, which precludes in situ sample annealing and makes crystalline growth unlikely. The mass-selecting aperture consisted of a slit approximately 1 mm in width, i.e. the same direction that different mass ion beams are spatially separated. The aperture was also approximately 15.25 mm tall and 2 mm thick. This aperture was expected to provide an improvement in mass resolving power over the aperture used for the Ne 126 and C samples described in Chapter 3, which was an approximately 5 mm diameter circular hole that was 16 mm thick. As mentioned, the sample location was not differentially pumped in this con- figuration, and so the base pressure and deposition pressures were that of the ion beam chamber itself. The base pressure of the ion beam chamber for these sam- ples ranged from approximately 6.5× 10−6 Pa to 1.3× 10−5 Pa (4.9× 10−8 Torr to 1.0× 10−7 Torr) before deposition. A residual gas mass spectra of the base pressure of this chamber was shown in Fig. 2.22 in Chapter 2. This high base pressure is due to ambient air leakage through the o-rings used to seal the ion source as well as a gate valve, which are only rated for high vacuum, i.e. a minimum pressure of roughly 1.3× 10−7 Pa (1.0× 10−9 Torr). As previously mentioned, SiH4 was used as the source gas for 28Si deposition. SiH4 was injected into the ion source via a UHV leak valve to generate a plasma that cracks and ionizes the SiH4, producing the Si ion beam. The SiH4 gas used for these samples and all 28Si samples in this work had a natural abundance of isotopes and a purity of 99.999 % according to the gas vendor (Matheson Tri-Gas). During operation of the ion source, a working pressure for the low pressure plasma mode typically around 1.7× 10−4 Pa (1.3× 10−6 Torr) was chosen, as measured by the ion gauge. It should be noted that because there is no differential pumping at the sample location and thus no outlet for gas diffu- sion, there may be a slightly higher local pressure at the sample location. This is because the ion beam itself carries a lot of hydrogen in the form of a H+2 beam as well as H from SiH4 fragments. This fluence of particles is delivered close to the sample before being blocked at the aperture where it can increase the local partial 127 pressures of H2 and SiH4. It is estimated that this effect may only contribute a 15 % higher total pressure at the sample, but it is nevertheless undesirable to have the sample in an environment with unknown partial pressures which may lead to increased adsorbates in the deposited film. SiH4 adsorption is of particular concern in this work because, as described in Chapter 2, naturally abundant SiH4 (including 29SiH4) adsorbed onto the deposition surface becomes incorporated into the film resulting in higher concentrations of 29Si and 30Si in the sample. This subject is explored in great detail in Chapter 6. 4.2.2 Sample Preparation For the initial 28Si deposition tests, very little ex situ sample preparation occurred. Substrates consisted of “lightly doped” natural abundance commercial Si(100) wafers that were cleaved by hand into approximately 1 cm by 1 cm chips. The chips were handled with clean teflon tweezers and mounted on the end of the feedthrough using a strip of carbon tape on the back of the wafer, and they were loaded into the vacuum chamber with a native oxide. No further sample preparation occurred in situ. The role of these substrates is to simply be a “catcher foil” to collect the 28Si ions, and one substrate used here actually was a Ag foil. Typically, after being loaded, samples sat several days in the vacuum chamber before deposition. 4.2.3 Deposition of 28Si The background deposition pressure for these samples was roughly the operat- ing pressure of the ion source, and so during deposition, the pressure in the chamber 128 rose to between approximately 1.6× 10−4 Pa and 2.3× 10−4 Pa (1.2× 10−6 Torr to 1.7× 10−6 Torr), as measured by an ion gauge in a different section of the chamber. 28Si ions were deposited onto the Si(100) substrates at room temperature. This tem- perature was not measured but assumed to be similar to the ambient temperature outside of the vacuum chamber, which was typically measured to be 21 ◦C ± 2 ◦C. Initially, an average ion energy, Ei, at the sample of approximately 455 eV was used. This value of Ei was chosen to be similar to the ion energy previously used for 12C deposition. The energy was lowered to around 64 eV for the remainder of these samples to reduce the sputter yield during deposition. Sputter yield values for Si are shown in Fig. 2.19 in Chapter 2, as determined from TRIM calculations [69]. For Si ions striking a Si target, the sputter yield is around 53 % at 455 eV. This means that on average, every incident ion sputters 0.53 atoms from the surface leading to a deposition rate that is effectively reduced by half. The sputter yield for Si at an energy of 64 eV is reduced to only about 8 %. Sputtering is more significant for Si deposition than for C deposition, which has sputter yields that are about half the value of those of Si for a given ion energy. The other significant aspect of the incident ion energies used here is that they are in the hyperthermal energy regime, as discussed in Chapter 2. The energy pos- sessed by the ions can be transferred into the depositing film promoting epitaxial deposition. However, only amorphous deposition is expected for the samples dis- cussed in this section due to the low substrate temperature and the presence of a surface oxide on the substrates. For these samples, 28Si ion beam currents, Ii, ranged from 770 nA to 1.1 µA. A mass spectrum from the measured ion current for 129 28Si and other SiH4 ions, which is representative of the configuration and ion beam settings for samples deposited at sample location IC–1, is shown in Fig. 4.3. The corresponding magnetic sector analyzer current used for the field sweep is shown on the top axis of this figure. Ion current peaks on this semi-log plot (circles) are observed between 28 u and 32 u. As discussed in Chapter 2, the 28 u current peak is 28Si, and the 29 u current peak is 28SiH and approximately 6 % 29Si based on the peak heights and the expected natural abundance. Ions of other Si hydrides up to 28SiH4 are also generated in the ion source. A sum of Gaussian fits to the peaks (line, Eq. (2.14)) are also shown superimposed on the data, which they match fairly well. As discussed in Chapter 2, the spacial distribution of ions in the beam is expected to be roughly Gaussian for this system. The mass resolving power of the ion beam in this configuration derived from this mass spectrum is m ∆m ≈ 38 (measured at 10 % of the peak height). The high ion current level between peaks of more than one tenth of the 28 u maximum (630 nA vs. 75 nA) possibly indicates a poorly focused beam leading to a large overlap between peaks. This could limit the separation of the 28Si ion beam from the 29Si ion beam and reduce the enrichment that is achieved with the settings used for these samples. The separation of the peaks can also be determined from the Gaussian fits, which give a standard deviation of the 29 u peak of σ ≈ 0.14. This results in a peak separation between the 28 u and 29 u peaks of approximately 7 σ. The 28Si ion beam in this setup had a fairly large average spot size leading to a deposition spot on the substrate of about 20 mm2. The 28Si spot is easily visible on the substrates due to the difference in color from the underlying native oxide. This 130 Figure 4.3: SiH4 mass spectrum representative of the ion beam settings for samples deposited at room temperature at IC–1. The ion current (circles) is recorded while sweeping the mass analyzer current, and thus the magnetic field (top axis). The 28 u peak is 28Si and the 29 u peak is both 28SiH and ≈ 6 % 29Si. Other higher order hydrides are also observed. Gaussian fits (line, Eq. (2.14)) to the peaks are shown superimposed on the data. The centers of the 28 u and 29 u fits are separated by ≈ 7 σ. large spot size is a result of not being able to use the focusing deceleration lenses in this configuration. Despite relatively high ion currents achieved in this configuration, the large spot size reduces the achievable growth rate for a given ion fluence, as compared to a more compact spot. The resulting thickness, d, of one deposited film was about 85 nm with a deposition rate, R, of 0.41 nm/min derived from dividing the thickness by the deposition time. The thickness was measured by SEM cross- sectional microscopy. A SEM micrograph of this 28Si sample is shown in Fig. 4.4. A photograph of the sample as deposited is inset showing the Si(100) substrate with 131 Figure 4.4: SEM cross-sectional micrograph of an amorphous 28Si thin film deposited at room temperature at IC–1. The Si(100) substrate is seen in the lower portion of the image and the 28Si film is above it. A dashed line marks the interface between the film and substrate. The thickness of the 28Si film here is approximately 86 nm. Au-Pt is deposited on the top surface of the sample to to protect the film during sample cleaving. Inset is a photograph of the sample after deposition showing the roughly 6 mm wide 28Si spot on the substrate. the 28Si deposition spot, which is roughly 6 mm wide. This SEM micrograph was acquired in collaboration with Dr. Michael Stewart (NIST). The Si(100) substrate is seen in the lower portion of the micrograph with the 28Si film appearing as a slightly lighter region above it. A dashed line marks the interface between the film and substrate. The thickness of the 28Si film seen here is approximately 86 nm. The 28Si appears to have a different texture and possibly structure than the substrate, possibly indicating grains about 10 nm in size, although it is more likely that the film is amorphous. Amorphous deposition is expected for these samples because the presence of a native oxide prevents crystalline registration of the atoms in the 132 depositing film which could otherwise lead to epitaxy. Additionally, depositing at room temperature means that the likelihood of forming polycrystalline grains is low [85]. The thickness of two other samples was inferred from the calibration of the SIMS depth profiles (discussed below). The thickness of one sample was estimated to be only about 15 nm with a corresponding deposition rate of 0.07 nm/min, and the other sample thickness ranged from approximately 75 nm to 92 nm across the deposition spot with deposition rates between 0.33 nm/min and 0.40 nm/min. To summarize, the typical deposition procedure for samples deposited at IC–1 was as follows: 1. a substrate is cleaved before being mounted onto the electrical feedthrough and loaded into the ion beam chamber, which is pumped out for > 12 h, 2. SiH4 is then introduced into the chamber and the ion source is turned on. The beam is tuned including characterization of the mass spectrum by collecting the ion current on the substrate, 3. the ion beam is then tuned to the 28 u peak to commence deposition of 28Si while monitoring the ion current onto the sample as a measure of the deposition rate, 4. after typically three to four hours of deposition, the ion source is turned off to end deposition, and the SiH4 leak valve is closed to reduce the ion beam chamber pressure back to its base, and 5. the sample is then removed from the vacuum chamber for ex situ analyses by venting the ion beam chamber. A total of five 28Si samples were deposited at IC–1 and all used this procedure. 133 4.2.4 Enrichment Measurements via SIMS for IC–1 Samples The enrichment of these samples and all samples discussed in this work were measured ex situ by SIMS using a CAMECA IMS-1270E7 large geometry spec- trometer as mentioned in Chapter 3. 28Si, 29Si, and 30Si isotopes were measured in collaboration with Dr. David Simons (NIST) and Dr. Shinichiro Muramoto (NIST). Dr. Simons collected the depth profile data for all but one of the samples discussed in this thesis. The raw data was then mostly analyzed by myself. The basic SIMS process for measuring Ne and C isotope ratios and isotope fractions was introduced in Chapter 3, and the measurement is similar for Si isotope fractions. 28Si samples are bombarded with a primary ion beam of O+2 which sputters the sample at a constant rate and is rastered across an area typically 50 µm across. Ejected Si ions are collected into a secondary ion beam for analysis in a mass spec- trometer. 28Si, 29Si, and 30Si ions are collected separately using an electron multiplier and the counts are recorded for a given time interval, or SIMS cycle, to get a count rate. With a mass resolving power m ∆m = 6000 (measured at 10 % of the peak height) for this instrument, 28SiH is easily distinguished from 29Si in these measurements. An example SIMS mass spectrum (line) of these two separated peaks is shown in Fig. 4.5. The ion masses are separated by only approximately 0.008 u, but are well resolved in the SIMS instrument. This resolution is necessary for an accurate measurement of the ratio of 29Si to 28Si. The count rates are used to determine the isotope ratios 29Si/28Si and 30Si/28Si. To reduce discrete counting noise, isotope 134 Figure 4.5: SIMS mass spectrum of 29Si and 28SiH ion currents (line) showing that the two peaks, which are separated by about 0.008 u, are well resolved in the SIMS instrument. This resolution is necessary for an accurate measurement of the ratio of 29Si to 28Si. ratios for each cycle are then averaged together from the highly enriched portion of the 28Si films after confirming no systematic trend to calculate the total average isotope ratios for that measurement. Isotope count ratios are converted into isotope fractions of the form (zSi/Sitot.), which was previously described. The uncertainty of the isotope fractions was determined from the standard deviation of the mean of the measurements. The depth profiles are calibrated by measuring the depth of the crater formed from sputtering the sample during the measurement. Crater depths were measured using a stylus profilometer, and then this value was used to determine the depth at each cycle by assuming a constant sputter rate. Because the measurement area is much smaller than the typical deposition area, multiple SIMS measurements are 135 sometimes made on the same sample. Measurements are typically made in the thickest portion of the 28Si film because better measurement statistics result from analyzing more material within a single measurement. The results corresponding to the highest individual enrichment measured in each sample is presented in this work instead of an average of the multiple measurements. This is because the average of multiple measurements depends on factors such as the number of measurements and their location across the deposition spot, which are not consistent between different samples. Different locations across the deposition spot have different enrichments because the deposition rate and thus relative SiH4 adsorption is not constant across the sample, which will be discussed further later. The multi-spot averages are there- fore not a reliable metric for comparing the overall enrichments of samples, although they can give an idea of the variation in enrichment across a single sample. Addi- tionally, the best measured enrichment for a sample is a preferred metric because it gives a lower bound on the best possible enrichment achievable by the deposition system. Further details of the specific SIMS measurements used for the different sets of samples discussed in this chapter can be found in Appendix E. Of the five samples deposited at IC–1, three are represented on the enrichment progression timeline in Fig. 4.2 and will be discussed here. A SIMS depth profile for a 28Si sample deposited at IC–1 is shown in Fig. 4.6. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth into the sample. This measurement as done using a TOF-SIMS instrument. At very shallow depths below the sample surface (0 nm to 10 nm), the isotope ratios are inflated to higher values due primarily to surface contamination from the sample being exposed to the ambi- 136 Figure 4.6: SIMS depth profile for a 28Si sample deposited at room temperature at IC–1. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth into the sample. The average 29Si isotope fraction is 1130(14) ppm. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. The 30Si data fluctuates between two values due to either zero (not shown), one, or two counts being detected in each measurement cycle. At 92 nm, the isotope fractions return to their natural abundance values indicating the interface with the substrate (shaded region). This measurement was done using a TOF-SIMS instrument. ent environment before the SIMS measurements. This initial “surface tail” artifact is typical of SIMS and is seen in all the measurements discussed here. Isotope ratios from each measurement cycle were averaged together from about 20 nm to 80 nm to determine total isotope fractions for this sample. At a depth of around 92 nm, the isotope fractions return to their natural abundance values (dotted and dashed lines) indicating the interface with the natural abundance Si substrate (shaded region) and giving an estimate of the 28Si film thickness. This value for the depth of the substrate interface is determined as the point at which the 29Si and 30Si isotope fractions re- 137 turn to half of their natural abundance values. It is difficult to infer any quantitative information about the width of the interface between the film and substrate from the depth profile because SIMS measurements tend to exaggerate interface widths. This can be due to interface roughness, which can be intrinsic to the sample or caused by the SIMS sputtering process itself [86]. For the sample measured in Fig. 4.6, the average measured 28Si isotope fraction in the 28Si film is 99.8850(14) %. The average 29Si isotope fraction is 1.130(14)× 10−3 (1130(14) ppm), and the average 30Si isotope fraction is 2.03(14)× 10−5 (20.3(14) ppm). The 30Si signal appears in two bands because of discrete counting fluctuations between one and two counts in each SIMS data cycle and most 30Si data being zero counts in some cycles. This measurement was the more highly enriched of two SIMS measurements performed on different spots on this sample. The average 29Si isotope fraction of the two spots for this sample is 1332(13) ppm. As previously discussed in Chapter 3, the isotope reduction factor is another useful parameter that describes the measured enrichment of a sample and, more specifically, gives the amount by which an excluded isotope is reduced from the natural abundance of that element. The reduction factor is determined by dividing the natural abundance of an isotope of an element by the measured isotope fraction. az represents the natural abundance of an isotope and, again, z is the mass number of the isotope denoted as 29 for 29Si and 30 for 30Si. The isotope reduction factors of the minor Si isotopes in the film are thus written as az/( zSi/Sitot.). In the SIMS measurement discussed here, the isotope reduction factor for 29Si is determined to be 41.5(5) (i.e. approximately 41 times lower than the natural abundance of 29Si). 138 The reduction factor of 30Si in this measurement is higher than that of 29Si, having a value of 1.5(1)× 103. This is likely due to the 30 u peak being farther away spatially from the 28 u peak at the mass-selecting aperture. These measurements indicate a realized mass selectivity for this configuration of approximately 45:1 for 29Si, which is lower than what was achieved for 12C. This sample is less enriched than the 12C samples by a factor of about 30 despite the mass resolving power determined from the mass spectrum being similar. This is perhaps not surprising considering that the 28 u and 29 u peaks have a spatial separation (5.5 mm) at the mass-selecting aperture approximately 44 % that of the 12 u and 13 u peaks (12.4 mm). Additionally, the natural abundance of 29Si relative to 28Si is higher than the natural abundance of 13C relative to 12C, so a 28Si sample deposited with a given ion peak separation would be less enriched than a 12C sample deposited with the same peak separation. A second 28Si sample deposited at IC–1 was measured by SIMS as well, but it showed a 29Si isotope fraction of 2822(18) ppm, which is approximately twice as large as the previous sample. This is potentially due to slightly different beam tuning parameters and a slightly lower mass selectivity achieved at the time of deposition. After depositing these two samples and reviewing the SIMS data, a third 28Si sample was deposited in an effort to better tune the ion beam parameters and increase the mass spectrum geometric selectivity, i.e. reduce the overlap of the 29 u peak onto the 28 u peak. This effort resulted in a lowering of the residual 29Si isotope fraction by nearly a factor of 200. A SIMS depth profile of the most highly enriched spot measured for this sample is shown in Fig. 4.7. 28Si (circles), 29Si (squares), 139 Figure 4.7: SIMS depth profile for the most highly enriched 28Si sample deposited at room temperature at IC–1. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction is 9.5(10) ppm. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. At a depth of 75 nm, the isotope fractions return to their natural abundance values indicating the interface between the film and the substrate (shaded region). and 30Si (triangles) isotope fractions are shown vs. sputter depth. At a depth of around 70 nm, the isotope fractions return to their natural abundance values (dotted and dashed lines) giving an estimate of the 28Si film thickness. The substrate is marked by the shaded region. The average measured 28Si isotope fraction in the 28Si film is 99.99846(19) %. The average 29Si isotope fraction is 9.5(10)× 10−6 (9.5(10) ppm), and the average 30Si isotope fraction is 5.9(16)× 10−6 (5.9(16) ppm). Isotope fractions were determined by averaging the data from depths between 20 nm to 65 nm. The isotope fraction of 29Si in this measurement is a nearly a factor of 120 lower than that of the previous sample. These measurements indicate a much 140 larger average isotope reduction factor of approximately 5.1(8)× 103 for 29Si and 30Si. The large improvement in 29Si reduction factor but modest improvement in 30Si reduction factor shows that the ion beam was indeed tuned better for 29Si rejection during deposition of this sample. For this sample, a total of four areas across the deposition spot were measured by SIMS with similar results to the most highly enriched one mentioned above. The average 29Si isotope fraction of the four measurements from this sample is 11.28(54) ppm. This result shows that a very high level of enrichment in 28Si is achievable, which surpasses the enrichment goal stated at the beginning of this chapter. It is difficult to determine the exact cause of the large reduction in the 29Si isotope fraction, as seen in the enrichment progression timeline in Fig. 4.2. Pre- sumably, the peak separation between the 28 u and 29 u ion current peaks in the mass spectrum was significantly larger for this final sample than for earlier samples resulting in a realized selectivity, although that data was not recorded. 4.2.5 Summary of Results for IC–1 Samples The 28Si samples deposited at IC–1 in the initial experimental configuration showed that the mass selected ion beam deposition method, demonstrated first in in Chapter 3, could be adapted to produce enriched 28Si thin films. It established the experimental parameters and methods needed for depositing 28Si from a nat- ural abundance SiH4 source gas and subsequently characterizing the enrichment. The most highly enriched 28Si sample produced in this configuration has an average measured 28Si isotope fraction of 99.99846(19) %, an average residual 29Si isotope 141 fraction of 9.5(10) ppm, and an average 30Si isotope fraction of 5.9(16) ppm, similar to that of the previous 12C samples. This level of enrichment not only matches the enrichments of 28Si produced by other sources currently available in the QI community, such as the IAC [32], but it surpasses them and meets the enrichment goal stated at the beginning of this chapter. The 28Si enrichment of this sample also surpasses the best reported enrichment of previous 28Si ion beam deposition by Tsubouchi et al. [43]. However, the enrichment of some samples was significantly worse, probably due to a lower geometric mass selectivity achieved for 28Si in these experiments. It is obvious from these initial depositions that the geometric selec- tivity of the ion beam needs to be consistently larger in order to, at a minimum, consistently and predictably match the residual 29Si isotope fractions of 28Si mate- rial produced by other sources including the IAC. Large variations and uncertainty in the mass selectivity of the ion beam system is a clear drawback of this experi- mental configuration. A cross section of a 28Si film was also analyzed using SEM as a secondary method for measuring the film thickness, and it showed that the film likely has a different structure than the substrate. 4.3 Achieving Highly Enriched 28Si: Lens Chamber Samples 4.3.1 Experimental Setup This section discusses the deposition of 28Si samples in an experimental con- figuration with the sample located in the lens chamber at LC–2 during deposition. 142 A schematic of the ion beam and lens chambers in this setup is shown in the upper right section of Fig. 4.1. These experiments expand upon the work of the previous section by implementing a number of experimental improvements including recon- necting the deceleration lens chamber to the ion beam chamber. The mass-selecting aperture used between the ion beam and the lens chamber was the same 1 mm wide slit discussed in the previous section. In this configuration, samples were located after the deceleration lenses, and they were mounted on the end of an electrical feedthrough which acted as a sam- ple stage for this intermediate experimental setup. A sketch of the sample stage feedthrough and sample location is shown in the blowup of the schematic of the LC–2 setup in Fig. 4.1. Samples were mounted using a strip of conductive carbon tape on the back side of the chip. As was the case with the previous setup, the pur- pose of the feedthrough was to isolate the sample electrically from the chamber so that the ion beam current could be monitored during beam tuning and deposition. Unlike the previous setup, additional electrical vacuum feedthroughs on the sample stage were used to mount a masking element above the sample location consisting of a metal shim for collecting current. This was positioned between the sample and the path of the ion beam and had a small circular aperture directly above the sample. This mask and aperture was fixed in place for the duration of a deposition. The fixed sample aperture was approximately 3 mm in diameter and provided a mechanism to monitor the focusing of the ion beam on the sample by maximizing the ion current detected on the sample while minimizing the current detected on the sample mask. Additionally, the mask and sample aperture allowed for precise loca- 143 tion of the ion beam spot on the sample substrate. The feedthrough (and sample) were positioned on axis with the ion beamline optics, and unlike the previous setup, IC–1, the sample aperture allowed for the ion beam to be tuned and steered onto the exact sample location under the aperture. However, a lack of sample motion in this setup means that the ion beam was constrained to be on axis, which may not have been the optimal tuning position. A photograph of a sample mounted on the vacuum feedthrough on this intermediate sample stage with the sample mask is shown in Fig. B.11 in Appendix B. As was the case with the previous setup, no method for sample heating exists on this sample stage here. This precludes in situ sample preparation and limits the available experimental phase space for achieving epitaxial deposition. For these experiments, the ion source was operated in the low pressure mode with a working pressure of SiH4 similar to that which was used for the previous setup of around 2.7× 10−4 Pa (2.0× 10−6 Torr). An additional turbo pump in the lens chamber provides differential pumping at the sample location, which results in lower partial pressures during deposition. Unlike the ion beam chamber, the lens chamber is rated to UHV, although it was never baked prior to these experiments. The base pressure of the lens chamber for these samples ranged from approximately 3.6× 10−8 Pa to 1.3× 10−6 Pa (2.7× 10−10 Torr to 1.0× 10−8 Torr) before deposi- tion, although for the majority of samples, the base pressure was typically around 1.3× 10−7 Pa (1.0× 10−9 Torr). The decelerating lenses themselves are described in detail in Chapter 2. As was mentioned previously, the benefit of the lenses is that they provide focusing of the 144 28Si ion beam as it exits the mass-selecting aperture. The lenses help maintain a tight beam spot as the ions are smoothly decelerated from the transport voltage (-4 kV) to ground potential at the sample. A more focused beam spot is advantageous because it corresponds to a higher ion flux, Fi, on the substrate and a higher deposition rate. This results in a lower relative rate of adsorption of gaseous species from the vacuum into the 28Si film for a given background pressure. Another significant experimental change implemented for the experiments of this section is the replacement and reconfiguration of several power supplies that controlled the voltages of various electrostatic lens elements in the ion beamline. These include the “arc” voltage, which defines the potential between the anode and cathode in the source, the “extractor” element voltage, and the “focus” element volt- age. The additional control and degrees of freedom provided by this reconfiguration allowed for better ion beam tuning to maximize ion fluence as well as produce a more confined beam before mass separation. This improved ion beam tuning yield- ing a more consistent geometric selectivity similar to and exceeding that which was achieved with the final sample deposited at IC–1. The circuit diagram of the power supplies controlling the ion beam lens elements up to the sector mass analyzer is shown in Fig. A.1 in Appendix A and discussed in Chapter 2. 4.3.2 Sample Preparation Substrates used for depositing 28Si in this section consisted primarily of “lightly doped” natural abundance commercial Si(100) wafers that were cleaved by hand into approximately 1 cm by 1 cm chips. Additionally, several 28Si samples were 145 deposited onto silicon-on-insulator (SOI) chips which had a 40 nm Si “device” layer on top of 400 nm of buried thermal oxide. The motivation for using these SOI wafers was to address a potential difficulty in using electron microscopy to measure the properties of a Si film deposited onto Si. If perfect epitaxy is achieved, then the 28Si film would become indistinguishable from the substrate except by means of isotope measurements such as SIMS. By introducing the buried oxide below an ultra-thin Si surface layer of known thickness, the 28Si film-substrate interface, and thus the film, would always be easily located in microscopy studies. Before being loaded into the vacuum chamber, substrates were pre-treated with hydrofluoric acid (HF) to remove the native oxide and provide a clean surface for deposition. This treatment was necessary to enable the possibility of epitaxial deposition because in situ substrate heating was not available to thermally desorb the surface oxide before deposition. The substrates were only handled with clean teflon tweezers and were mounted on the end of the sample feedthrough using a strip of carbon tape on the back of the wafer. No further sample preparation occurred in situ. Typically, after being loaded into the chamber, samples sat several days in the vacuum chamber before deposition. 4.3.3 Deposition of 28Si After tuning and focusing the ion beam according to the procedures described in Chapter 2, 28Si ions were deposited onto the Si(100) substrates at room tem- perature (≈ 21 ◦C, see Section 4.2.3). For the deposition of one 28Si sample, a different sample stage was used that had a tungsten wire back heater for sample 146 heating. This was used to to deposit the 28Si film with a substrate temperature of approximately 550 ◦C. This experiment, however, did not have a significant impact on the work presented here and will not be discussed further. During deposition of the 28Si films at room temperature, the pressure in the chamber rose to between 3.3× 10−6 Pa and 7.2× 10−6 Pa (2.5× 10−8 Torr to 5.4× 10−8 Torr) due to gas diffusion from the operation of the ion source. These pressures are more than a factor of 20 improvement over the previous setup. Ions were deposited with a range of average ion energies at the sample between 50 eV and 170 eV. A lower value of Ei generally produced a larger geometric selectivity as observed in mass spectra during the tuning procedures prior to each deposition. For these samples, 28Si ion beam currents ranged from 200 nA to 800 nA. A mass spectrum for an ion beam with an average ion energy energy Ei ≈ 122 eV collected through the fixed sample aperture prior to deposition of a 28Si sample is shown in Fig. 4.8. The corresponding magnetic sector mass analyzer current used for the field sweep is shown on the top axis. Ion current peaks (circles) corresponding to 28Si and Si hydrides are observed from 28 u to 33 u. The 28Si ion peak at 28 u and the ion peak at 29 u, containing ≈ 5 % 29Si, show a high degree of separation on this semi-log plot with no detectable ion current signal occurring between the peaks. Secondary electrons generated by the ion beam cause the current between the peaks to be ≈ -0.5 nA. Gaussian fits to the data (line, Eq. (2.14)) are also shown for the 28 u and 29 u current peaks, and one can see that the data matches the form of a Gaussian very well. This indicates a symmetric and optimally tuned beam shape with minimal perturbations such as scattering off of lens elements. The Gaussian fits give a separation of the 28 u peak 147 2 7 2 8 2 9 3 0 3 1 3 2 3 3 3 4 1 0 - 4 1 0 - 3 1 0 - 2 1 0 - 1 1 0 0 D a t a G a u s s i a n F i t s 2 9 S i + 2 8 S i H Ion Cu rren t (µ A) M a s s ( u ) 2 8 S i 1 3 0 2 1 5 - 2 8 S i - S i 4 9 5 0 5 1 5 2 5 3 5 4 5 5M a g n e t C u r r e n t ( A ) Figure 4.8: SiH4 mass spectrum representative of the ion beam settings for samples deposited at LC–2. The ion current (circles) is recorded while sweeping the the analyzer current, and thus the magnetic field (top axis). The peak at 28 u is 28Si and the peak at 29 u peak is 28SiH and ≈ 5 % 29Si. Several higher order hydrides are also observed. Gaussian fits (line, Eq. (2.14)) to the 28 u and 29 u peaks are superimposed on the data. The centers of the 28 u and 29 u fits are separated by ≈ 11 σ. from the 29 u peak of approximately 11 σ. The mass resolving power of the ion beam in this configuration derived from this mass spectrum is m ∆m ≈ 78 (measured at 10 % of the peak height), which is significantly better than for the mass spectrum shown in the previous section. Focusing the 28Si ion beam with the deceleration lenses resulted in a more compact beam spot size and an average deposition area of about 6 mm2. The beam spot focusing procedure was not yet optimized for the initial sample deposited here resulting in a larger deposition area of 16 mm2. The 28Si spots on these samples are 148 Figure 4.9: Optical micrograph of a 28Si sample deposited at room temperature at LC–2. The Si(100) substrate consisting of a roughly 8.5 mm wide chip is seen with the 28Si spot measuring about 3.8 mm long. still visible to the naked eye although they appear different from the samples in the previous section likely due to the native oxide being stripped prior to deposition. A visible deposition spot indicates that the 28Si film is structurally different from the substrate and probably not epitaxial. A representative optical micrograph of a 28Si sample deposited at LC–2 is shown in Fig. 4.9. The Si(100) substrate is approximately 8.5 mm wide and the 28Si spot is approximately 3.8 mm long. The resulting thicknesses of these 28Si films ranged from approximately 50 nm to 350 nm as inferred from SIMS depth profiles with corresponding deposition rates between 0.51 nm/min and 1.49 nm/min. To summarize, the typical deposition procedure for samples deposited at LC–2 was as follows: 1. a substrate is cleaved and dipped in HF immediately before being mounted onto the sample stage feedthrough and loaded into the deceleration lens cham- ber which is pumped out for > 12 h, 2. after turning on the ion source, the gate valve to the ion beam chamber is 149 opened and the ion beam is tuned and characterized on the substrate using the fixed sample aperture on the feedthrough. The mass spectrum, ion beam energy, and beam spot focusing are analyzed and recorded, 3. the ion beam is then tuned to the 28 u peak to commence deposition of 28Si while monitoring the ion current onto the sample as a measure of the deposition rate, 4. after typically three to five hours of deposition, the gate valve to the ion beam chamber is closed to end deposition and reduce the lens chamber pressure back to its base, and 5. the sample is then removed from the vacuum chamber for ex situ analyses by venting the lens chamber. In total, 16 28Si samples were deposited at LC–2 under these conditions. 4.3.4 Enrichment Measurements via SIMS for LC–2 Samples SIMS was used to assess the enrichment of several samples deposited at LC– 2 to determine if the experimental improvements including the higher degree of control over the ion beam, which were discussed at the beginning of this section, yielded more consistent and lower overall 29Si and 30Si isotope fractions. Of the 16 samples produced at LC–2, three are represented on the enrichment progression timeline in Fig. 4.2, and will be discussed in this section. A SIMS depth profile of the highest enrichment measured for the first 28Si sample deposited at LC–2 is shown in Fig. 4.10. Measurements of the 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. the sputter depth into the sample. At a depth of around 80 nm, the isotope fractions begin to increase to their natural abundance 150 Figure 4.10: SIMS depth profile of a 28Si sample deposited at room temperature at LC–2. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction is 2.02(32) ppm, and the average 30Si isotope fraction is 1.41(20) ppm. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. At a depth of 80 nm, the isotope fractions increase to their natural abundance values in the transition into the Si(100) substrate (shaded region). The film-substrate interface is estimated to be at a depth of 105 nm (not shown). values (dotted and dashed lines) in the transition into the Si(100) substrate, marked by the shaded region. This relatively gradual increase in isotope fraction over 25 nm is partially an artifact of the SIMS measurement, which was described in the previous section. It is also probably due to beam tuning, including mass spectrum sweeps, which would deposit all three Si isotopes on the substrate before 28Si deposition. The film thickness is determined from the location of the interface between the film and substrate, which is estimated to be at a depth of 105 nm (not shown). The average measured 28Si isotope fraction in the 28Si film is 99.999657(38) %. 151 The average residual 29Si isotope fraction is 2.02(32)× 10−6 (2.02(32) ppm), and the average 30Si isotope fraction is 1.41(20)× 10−6 (1.41(20) ppm). These mea- surements indicate an average isotope reduction factor for 29Si and 30Si of nearly 2.3(2)× 104. A second area on this sample was also measured by SIMS giving an average 29Si isotope fraction of these two measurements of 2.16(21) ppm. This level of 29Si isotope fraction is nearly a factor of five lower than for sample with the highest enrichment achieved at IC–1, as seen in the jump from IC–1 to LC–2 in the enrichment progression timeline in Fig. 4.2. The similar reduction factors for 29Si and 30Si indicate that the upgraded ion beam tuning control introduced in this sec- tion enables more consistent geometric selectivites similar to and surpassing those achieved for the final sample deposited at IC–1 in the last section. This sample was deposited with a background pressure 20 times lower than previous samples, and the resulting enrichment appears to support the conjecture discussed previously that the enrichment is partially a consequence of the adsorption of SiH4 from the background gas. Despite this improvement in enrichment, this initial sample had a relatively large deposition spot size as was mentioned previously. This resulted in the lowest deposition rate of any of the samples discussed in this section, which may limit the enrichment that was achieved. This is because a lower 28Si deposition rate will result in a higher relative adsorption rate of gaseous species for a given background pressure during deposition. As was mentioned previously in Chapter 2 and in this chapter, part of this background pressure is natural abundance SiH4, which, if adsorbed into the film, will result in a lower 28Si enrichment in the sample. After depositing this initial sample, another 28Si sample was deposited under 152 similar conditions except with a more focused beam spot resulting in a smaller deposition spot size of about 7 mm2. The SIMS measurement of the most highly enriched region of this second sample is shown as a depth profile in Fig. 4.11. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. At a depth of around 340 nm, the isotope fractions begin to increase due to ion beam tuning and mass spectrum sweeps prior to deposition of 28Si. Beyond this, the isotope fractions continue to rise to their natural abundance values (dotted and dashed lines) in the transition into the substrate, marked by the shaded region. The film-substrate interface is estimated to be at a depth of 384 nm, giving a value for the 28Si film thickness. The data appears to reside in two main bands because of discrete counting fluctuations in the measurement, as mentioned for previous SIMS measurements. The average measured 28Si isotope fraction in the 28Si film for this sample is 99.9998308(82) %. The average 29Si isotope fraction is about a factor of two lower than the previous sample at a value of 0.993(64)× 10−6 (0.993(64) ppm). The average 30Si isotope fraction is 0.699(51)× 10−6 (0.699(51) ppm). These measure- ments indicate an average isotope reduction factor for 29Si and 30Si of 4.6(2)× 104. Achieving this very high enrichment potentially enables a robust measurement of the dependance of electron coherence times on 29Si concentration in the single spin regime, as discussed at the beginning of this chapter. Four other areas of the depo- sition spot of this sample were also measured by SIMS giving an average 29Si isotope fraction for the five measurements of this sample of 1.388(38) ppm. This result again indicates that the measured enrichment could be due to the 153 Figure 4.11: SIMS depth profile of a 28Si sample deposited at room temperature at LC–2. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction in the film is 0.993 ppm, and the average 30Si isotope fraction is 0.699 ppm. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. At 340 nm, the isotope fractions begin to increase to their natural abundance values in the transition to the substrate (shaded region). The film-substrate interface is estimated to be at a depth of 384 nm. presence of a SiH4 background gas at the sample. The smaller spot size of this sample led to the highest deposition rate for any of the samples discussed in this section, and one that is almost three times higher than the previously discussed sample. The higher deposition rate of this second sample results in a shorter time for gaseous species to adsorb into the film for a given volume of material being deposited and thus lower 29Si and 30Si residual isotope fractions. This effect is evident not just when comparing the previous sample to this one, but also within the five SIMS measurements of this sample. Due to the 28Si 154 ion beam flux being nonuniform throughout the beam spot with a roughly Gaussian spatial distribution, the five areas on the deposition spot were measured to have different thicknesses and thus different corresponding deposition rates. The SIMS measurements of 29Si from this sample are shown vs. the deposition rate in Fig. 4.12. As expected, the residual 29Si isotope fraction (squares) varies inversely with the deposition rate for these five measurements. These deposition rates are derived from the film thickness at each spot (top axis), which are inferred from the SIMS depth profiles. The uncertainty of the deposition rates are determined from uncertainty in the SIMS depth scales and deposition times. This shows that improvements to either the background pressure during deposition or the deposition rate can improve the enrichment of these 28Si samples. Having establishing the experimental procedures and processes to produce this last 28Si film with a residual 29Si isotope fraction of 0.993 ppm, several other 28Si samples were deposited with isotope fractions of 29Si and 30Si consistently at or below 1 ppm. One of these samples was deposited on a SOI substrate and was measured by SIMS to have an average 28Si isotope fraction of 99.999863(16) %. The average residual 29Si isotope fraction is 0.77(11)× 10−6 (0.77(11) ppm), and the average 30Si isotope fraction is 0.60(11)× 10−6 (0.60(11) ppm). The most highly enriched of these samples deposited at LC–2 has an average measured 28Si isotope fraction of 99.999888(10) %. The average residual 29Si isotope fraction in this sam- ple is 0.691(74)× 10−6 (0.691(74) ppm), and the average 30Si isotope fraction is 0.432(67)× 10−6 (0.432(67) ppm). The SIMS depth profile of this sample is shown in Fig. 4.13. The isotope fraction values were determined by averaging the data be- 155 Figure 4.12: 29Si isotope fractions vs. the deposition rate for multiple SIMS mea- surements made on a sample deposited at room temperature at LC–2. Five different areas across the deposition spot experienced different deposition rates due to an in- homogeneous ion beam flux. These rates were derived from the thickness at each spot (top axis) inferred from the SIMS depth profiles. The 29Si isotope fractions (squares) vary inversely with the deposition rate across the sample. tween 25 nm and 100 nm. The data from the rest of the 28Si film was not included because during the first half of the deposition, the pressure at LC–2 was being varied as part of a separate experiment resulting in slightly increased 29Si and 30Si isotope fractions due to SiH4 adsorption. The average isotope reduction factor for 29Si and 30Si in this sample is 7.0(7)× 104. At a depth of around 225 nm, the isotope frac- tions begin to increase to the natural abundance values in the substrate (dotted and dashed lines). The interface between the film and the substrate is estimated to be at a depth of 249 nm and is marked by the shaded region. Overall, SIMS measurements of samples deposited at LC–2 show that a reduc- tion of nearly a factor of 14 in the 29Si isotope fraction of the most highly enriched 156 Figure 4.13: SIMS depth profile of the most highly enriched 28Si sample deposited at room temperature at LC–2. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction in the film is 0.691 ppm, and the average 30Si isotope fraction is 0.432 ppm. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. At a depth of 225 nm, the isotope fractions begin to increase to their natural abundance values in the substrate (shaded region). The interface between the film and the substrate is estimated to be at a depth of 249 nm. sample was achieved compared to that of the previous most highly enriched sam- ple deposited at IC–1, which can be seen in the enrichment progression timeline in Fig. 4.2. Additionally, the 28Si enrichment values of these samples are consistently high unlike those of the samples deposited at IC–1, which varied a great deal. 4.3.5 Crystallinity Substrates were stripped of their native oxide ex situ in the experiments de- scribed in this section to facilitate the possibility of polycrystalline or epitaxial 157 Figure 4.14: HR-TEM cross-sectional micrograph of a 28Si film deposited at room temperature at LC–2. The Si substrate is seen in the lower right area of the image in (a), a layer of glue (light) and Pd cap (dark) are seen in the upper left, and the deposited 28Si film resides between them. (b) is a FFT of the 28Si region from box (b) and shows that it is amorphous. By comparison, (c) is a FFT of the Si(100) substrate from box (c) and it clearly shows a crystalline pattern. This image was taken on the 〈001〉 zone axis. deposition. However, depositing at room temperature severely limits Si epitaxy to very thin layers if it occurs at all [55, 87], and Si solid phase epitaxy (SPE) is expected to be negligible [88] making the formation of crystalline grains unlikely. The suspected amorphous nature of these 28Si films is verified using TEM. A cross- 158 sectional TEM micrograph of a 28Si sample deposited at LC–2 is shown in Fig. 4.14. This microscopy was done in collaboration with Dr. June Lau (NIST), and the TEM specimen was prepared by mechanical polishing. Panel (a) is a high resolu- tion (HR-TEM) image showing the deposited 28Si film between the substrate and a Pd capping layer. The Si(100) substrate is the region in the lower right of the micrograph and to the left of that is the 28Si film. In this area of the film, the 28Si layer thickness is about 20 nm and the average thickness over all the areas surveyed was 37 nm. Just to the left of the 28Si layer is a thin dark Pd capping layer deposited to protect the film during the TEM specimen thinning process by Ar milling. In the upper left of the micrograph is glue, also from specimen preparation. The sample was tilted to the 〈001〉 silicon zone axis in this micrograph. One can see that the 28Si film is amorphous because the lattice rows of the substrate do not continue into the film region. This is made more clear by the fast fourier transform (FFT) analysis of the film and substrate. Panel(b) in Fig. 4.14 is a FFT of the 28Si region (box (b) in panel (a)) and corresponds to an amorphous pattern. In contrast, panel (c), which is the FFT of the Si(100) substrate (box (c) in panel (a)), corresponds to a crystalline pattern. 4.3.6 Chemical Purity 4.3.6.1 SIMS To begin to assess the chemical purity of the 28Si samples in the context of the materials goals laid out at the beginning of this chapter, SIMS was used to detect 159 C contamination in two samples deposited at LC–2. 12C and 13C were monitored in the 28Si samples both as a marker for general chemical contaminants in the films, but also because 13C possesses nuclear spin (I = 1/2), which will cause decoherence of qubit spins in a quantum computing device in the same way that 29Si does. A SIMS depth profile of the C atomic concentrations in a 28Si sample is shown in Fig. 4.15. 12C (solid line) and 13C (circles and line) are plotted vs. sputter depth into the sample. A profile of 30Si (triangles and line) measured at the same time is shown as a reference to the boundary between the deposited 28Si film and the natural Si(100) substrate (shaded region). The film-substrate interface is estimated to be at a depth of around 425 nm. It is difficult to precisely determine the atomic concentration of carbon in the silicon film because the carbon background due to the SIMS instrument is not precisely known under the measurement conditions, however it must be a factor of 10 lower than in the film because the C concentration drops by that much in the substrate. For simplicity, C was monitored using the same measurement conditions as for Si detection, and so the measurement was not optimized for accurate C measurements. The slow roll off observed for the carbon signals moving into the substrate is probably a measurement artifact. After apply- ing a value obtained from literature of 0.007 for the relative sensitivity factor of carbon to silicon for SIMS under similar analytical conditions [89], the average mea- sured atomic fraction of 12C (measuring only Si and C isotopes) in the film is found to be approximately 3.39(8) %. Considering the nominal atomic concentration for amorphous Si (≈ 4.9× 1022) [90], the measured atomic fraction of 12C represents an atomic concentration of 1.66(4)× 1021 cm−3. The average measured 13C atomic 160 Figure 4.15: SIMS depth profile showing the atomic concentration of 12C (solid line), 13C (circles and line), and 30Si (triangles and line) in a 28Si sample deposited at room temperature at LC–2. The sharp rise in 30Si concentration marks the boundary between the deposited 28Si film and the natural Si(100) substrate (shaded region). The slow roll off of the carbon profiles into the substrate is an artifact of the SIMS measurement. fraction is 0.031(1) % (310(10) ppm), which is about 110 times smaller than the 12C fraction and close to the natural abundance fraction of about 1.1 %. For refer- ence, the 30Si atomic concentration in this particular measurement is approximately 1.6 ppm in the film. The concentration of spins in the 28Si film due to 13C is over 300 times larger than the residual 29Si spin concentration in this sample. The source of this C con- tamination is probably predominately gaseous carbon containing compounds in the vacuum such as CO and CO2, which can adsorb into the 28Si film during deposition. A buildup of gaseous C compounds is also likely the source of the small spike in 161 C concentration which can be seen at the interface with the substrate because the sample sat approximately one day in the chamber before deposition. In this ex- perimental configuration there is no residual gas analyzer available to measure the relative abundance of such compounds. It is also possible that some C adsorbs into the amorphous film from the ambient environment after the sample is removed from vacuum. To examine the dependance of the C contamination in the 28Si film on back- ground pressure, a sample was deposited using the nominal procedure described above except that during deposition, the pressure in the chamber was raised for about one third of the deposition time. This was achieved by closing the gate valve to the turbo pump in the lens chamber. This caused the pressure in the chamber to rise almost two orders of magnitude, which was partially due to SiH4 gas from the ion source, but also probably due to an increase in partial pressures of typical residual gasses found in high vacuum systems such as H2, CO, N, and CO2. The C contamination of this sample was measured by SIMS in the same man- ner as the last sample. A SIMS depth profile of the C concentration in this sample is shown in Fig. 4.16. 12C (solid line) and 13C (circles and line) are plotted vs. sputter depth into the sample. A profile for 30Si (triangles and line), which was measured separately, is shown as a reference to the boundary between the deposited 28Si film and the natural Si(100) substrate. The film-substrate interface, marked by the shaded region, is estimated to be at a depth of 240 nm. Also plotted is the ratio of 28SiH/28Si (dotted line), which is typically monitored during SIMS mea- surements and corresponds to the right axis in the figure. The 28SiH signal gives 162 Figure 4.16: SIMS depth profile showing the atomic concentration of 12C (solid line), 13C (circles and line), and 30Si (triangles and line) in a 28Si sample deposited at room temperature at LC–2. The sharp rise in 30Si concentration marks the boundary between the deposited 28Si film and the natural Si(100) substrate (shaded region). Also shown is the measured SIMS ratio of 28SiH/28Si (dotted line) corresponding to the right hand axis. The chamber pressure was increased almost two orders of magnitude during the portion of the deposition bounded by the vertical dashed lines between approximately 80 nm and 190 nm. The 12C, 13C, and 28SiH concentrations increase in this region. a qualitative idea of relative amounts of H in the different layers of the film. The region of the film corresponding to the higher pressure during deposition is approx- imately between 80 nm and 190 nm, bounded by the vertical dashed lines. In this portion of the film, one can see that the C concentrations increase more than an order of magnitude and the 28SiH signal increases slightly less than that. A spike in C concentration is observed at the interface between the film and the substrate due to carbon containing adsorbates accumulating when the sample sat for about three days in vacuum before deposition. 163 In the low pressure region of the film, i.e. the top layer from 0 nm to 80 nm, the average 12C atomic concentration is determined to be 2.67(3)× 1020 cm3, which is 0.545(6) %, and the average 13C atomic concentration is 42(3) ppm. These val- ues are around six times lower for this sample than for the previous one, prob- ably due in part to the deposition pressure being slightly lower for this second sample. In the high pressure region of the film, the 12C atomic concentration in- creases to 3.2(2)× 1021 cm3 (6.6(4) %). The 13C atomic concentration increases to 610(40) ppm. For reference, the 30Si atomic concentration throughout the enriched film is approximately 0.9 ppm. A lower bound on the total achievable Si chemical purity for samples deposited at LC–2 can be determined from the low pressure re- gion of this second sample to be approximately 99.451(6) %. This sample shows that adsorption and incorporation of chemical contaminants from the vacuum back- ground pressure into the depositing 28Si film is a significant issue in this experimental configuration. 4.3.6.2 XPS In addition to the chemical purity analysis provided by SIMS for the C con- centrations in 28Si films, XPS was used to search for a broader range of chemical contaminants. XPS spectra were acquired and analyzed in collaboration with Dr. Kristen Steffens (NIST). The 28Si sample used for XPS analysis was deposited at LC–2 under similar conditions as the previous samples described in this chapter. Additionally, a control chip accompanied the sample through the deposition pro- 164 cess but was not irradiated by the ion beam. XPS spectra were collected after sputter-cleaning the samples with Ar to remove surface contamination from the en- vironment. Both low resolution survey scans (pass energy 160 eV, step size 0.5 eV) and high resolution region scans (20 eV pass energy, step size 0.1 eV) for C 1s, N 1s, Si 2p, and O 1s were performed on a Kratos Axis-Ultra DLD Photoelectron Spec- trometer with a monochromated Al Kα x-ray source (1486.6 eV). Multiple spots were measured on each sample to ensure consistency. Peak positions were calibrated to the Si 2p1/2 peak at 99.3 eV. Initially, prior to sputter cleaning, the 28Si film showed C, N and O peaks in the XPS spectrum. However, after sputter cleaning, the N and the adventitious C disappeared, but a C 1s peak due to SiC persisted. This 28Si XPS spectrum is shown in Fig. 4.17 (upper spectrum) as a plot of count rate vs. electron binding energy. The data for the 28Si sample was shifted up for clarity. The C 1s peak indicates a relatively constant atomic fraction of approximately 3 %, consistent with the first SIMS result. After sputtering, two small O 1s peaks also remain at 531.4 eV and 532.5 eV, corresponding to an atomic fraction of approximately 4 %. Also shown are references for relevant elemental orbital level positions. In the Si 2p region, a SiO2 peak from the native oxide was no longer present after sputtering, however, in addition to the elemental Si peaks, a small shoulder attributed to SiC is seen in finer scans. The control Si sample (lower spectrum) shows reduced C and O peaks corresponding to atomic fractions of approximately 1 % to 2 % each. These values are taken as upper limits which can give an indication of the instrumental background because the actual C and O content of the wafer is expected to be 165 Figure 4.17: XPS spectra of a 28Si sample deposited at room temperature at LC–2 and a control Si sample. Count rates vs. electron binding energy for survey scans are shown for the 28Si sample after Ar sputter cleaning (upper spectrum) and the control chip (lower spectrum). The data for the 28Si sample was shifted up for clarity. References for elemental orbital level positions are included above relevant peaks. O 1s and C 1s peaks are visible with larger amplitudes in the 28Si film than the control sample. much lower. Both the 28Si and control scans show Ar peaks due to the Ar sputter cleaning process. This XPS analysis agrees with the SIMS result that at least some samples deposited at LC–2 can have as much as 3 % C throughout the 28Si film, and in addition they can also contain an approximately equal amount of O. The Si chemical purity for this sample measured by XPS is roughly 95 %. Although the previous SIMS measurement showed that a sample can have lower level of C, it is not known if the O contamination is similarly reduced in that sample, and so it is difficult to place a bound on the total Si purity that is achievable for these samples. 166 4.4 Chapter 4 Summary The 28Si samples deposited at IC–1 served as a successful proof of principle for the adaptation of the in situ enrichment and ion beam deposition method from 12C deposition to 28Si. Experimental procedures for deposition using a natural abundance SiH4 source gas and the characterization of the enrichment via SIMS were established. The samples deposited at LC–2 showed that 28Si films could be deposited with residual 29Si and 30Si isotope fractions consistently below 1 ppm. A total of five 28Si samples were produced at IC–1 and 16 were produced at LC–2. The overall improvement in 28Si enrichment achieved from the initial samples deposited at IC–1 to the samples deposited at LC–2 can be seen in the enrichment progression timeline in Fig. 4.2. The residual 29Si isotope fractions of these samples was reduced from 2822(18) ppm to 0.691(74) ppm. This most highly enriched sample deposited at LC–2 had a 28Si isotope fraction of 99.999888(10) %. The achieved reduction in 29Si and 30Si isotope fractions in samples deposited at LC–2 was likely due to several factors. First, better control of the ion beam tuning lead to consistently higher geometric selectivities. Additionally, the lens chamber has significantly lower background pressures resulting in less SiH4 adsorption during deposition. Finally, the deceleration lenses enabled focusing of the ion beam spot resulting in smaller deposition spots, higher growth rates, and less SiH4 adsorption. The enrichment values of these samples meet the enrichment materials goal laid out at the beginning of this chapter to achieve 28Si enrichments that surpass those of any other known sources of 28Si including the IAC, and they do it consistently. Additionally, they 167 attain enrichments sufficient to enable a robust measurement of the dependance of electron coherence time on 29Si concentration in the single spin regime and compare it to theoretical predictions (see Fig. 1.9), as proposed in Chapter 1. However, initial observations of 28Si film crystallinity and chemical purity show that these samples deposited at IC–1 and LC–2 do not meet the second and third materials goals stated at the beginning of this chapter of being crystalline and highly chemically pure. SEM and TEM cross-sectional micrographs of two samples show that the films are unsurprisingly amorphous, probably due to deposition occurring at room temperature. Chemical analysis of several samples deposited at LC–2 by both SIMS and XPS show that both C and O are present in these films at rela- tively high atomic concentrations up to approximately 3 %. Achieving both single crystal epitaxial deposition and eliminating chemical contaminants require further experimental improvements, and will be the subject of Chapter 5. 168 Chapter 5 28Si Thin Film Deposition and Characterization Phase II: Crystallinity and Chemical Purity 5.1 Introduction Chapter 4 demonstrated that extremely high levels of 28Si enrichment (< 1 ppm 29Si) were achievable for thin film samples produced via ion beam deposition at LC–2. The most highly enriched sample deposited at LC–2 had a 28Si isotope fraction of 99.999888(10) % and a residual 29Si isotope fraction of 0.691(74) ppm. However, these samples were found to be amorphous, likely due to them being deposited at room temperature, and they were measured to contain C and O in atomic concentrations up to approximately 3 %. In order for the 28Si to be viable for use in Si based solid state quantum computing, it must be of very high quality and meet the three materials goals of this work discussed in Chapters 1 and 4. The experiments in Chapter 4 produced samples which surpassed the first materials goal of achieving enrichments with residual 29Si isotopic concentrations below 50 ppm 169 to match other sources of 28Si including the International Avogadro Coordination. The residual 29Si isotope fractions of these samples was also potentially low enough to facilitate a robust measurement of the dependence of electron coherence times on 29Si concentration in the single spin regime, also discussed in Chapters 1 and 4. Assessments of samples produced in the experiments discussed in Chapter 4 were done in the context of the second and third materials goals of single-crystalline, epitaxial thin films with dislocation densities below 1 × 106 cm−3 and chemical impurity concentrations below 2× 1015 cm−3, however these goals were not pursued further in those experiments. Discussion of the ion beam deposition of 28Si films enriched in situ and the characterization of their properties including enrichment, crystallinity, and chem- ical purity is continued in this chapter. The experiments discussed here seek to maintain and improve upon the already high level of achieved 28Si enrichment while depositing samples in the final experimental configuration for this system used in this work, which enables epitaxial deposition. This continued improvement in en- richment and reduction in residual 29Si and 30Si isotope fractions is illustrated in Fig. 4.2, the enrichment progression timeline. As previously described in Chapter 1, the measured isotope fractions of Si are written as zSi/Sitot., representing the average detected counts of an isotope divided by the total average counts of the SIMS mea- surement. It shows the 29Si isotope fraction reduction at LC–2 from 0.691(74) ppm down to a minimum at DC–3 of 127(29) ppb with an overall enrichment in 28Si of 99.9999819(35) % for the most highly enriched sample produced in this work. SIMS measurements of the enrichment of 28Si presented in this chapter will be discussed in 170 terms of the progression of samples at DC–3 in Fig. 4.2 and the experimental factors that resulted in those higher enrichments. The experiments discussed here also seek to leverage the improved capabilities of this new experimental setup to facilitate high-quality epitaxial deposition as well as study chemical contaminants to improve the purity of the 28Si films. Chemical impurities present in 28Si films can act as or induce scattering sites and charge traps [91, 92]. This reduces electron mobility and other electronic properties that are important for the successful operation of QI devices. Additionally, chemical impurities can possess nuclear spin, which will cause decoherence of qubit spins in a quantum computing device in a similar manner as the nuclear spin of 29Si in natural abundance Si. A total of 40 28Si samples were produced at DC–3. The experimental configuration used for depositing samples in this chapter is the third setup discussed in Chapter 4 where samples are located at DC–3 in the deposition chamber, as shown in the deposition chamber schematic in Fig. 4.1. This setup had several advantages over the previous two including a further reduction in the background pressure during deposition due to additional differential pumping. In addition to a third turbo pump and an ion pump present in the deposition chamber, an ion pump was added to the lens chamber for this setup, which are all marked in Fig. 4.1. Lower partial pressures of gaseous species containing C and O should result in higher chemical purities of the film. Depositing samples at DC–3 crucially enabled heating of the Si substrate for in situ preparation and heating of the sample during 28Si deposition to facilitate epitaxial growth. This setup also enables the use of the analytic instruments in the deposition chamber in conjunction 171 with 28Si deposition including RHEED, RGA, and STM, as discussed in Chapter 2. These features will be discussed further in the next section. Epitaxial deposition of 28Si thin films with a high degree of crystallinity, i.e. a low defect density similar to electronics grade Si, is a key achievement towards producing material suitable for use with solid state quantum computing devices. The crystallinity of a film and the roughness of its surface depend on the characteristics of the deposition and growth. Si MBE is ideally categorized as either Volmer-Weber growth or Frank-van der Merwe growth [93]. Volmer-Weber growth is characterized by the formation of 3D islands consisting of multiple layers growing at once as islands coalesce into a continuous film. This type of growth is the result of the ratio of the depositing atomic flux to the surface diffusivity being large such that it is more likely that adatoms will cluster with each other and form islands before they can diffuse to and over step edges [94]. These islands result in the buildup of roughness on the surface. Conversely, Frank-van der Merwe growth is characterized by smooth layer-by-layer growth where single layer 2D islands merge to ideally form a continuous first monolayer before the second layer begins to form [95]. This type of growth is the result of the ratio of atomic flux to surface diffusivity being small such that adatoms are more likely to reach step edges as well as diffuse down to lower steps before clustering to form second layer islands. With the presence of atomic steps on the substrate, this growth mode can also lead to so-called step flow growth where no islands form and all adatoms can diffuse to step edges before any other interactions. In reality, most thin film deposition occurs between these two ideal cases. In general, deposition dominated by 3D island growth can lead to 172 rougher films with more crystalline defects than deposition dominated by smooth layer-by-layer growth. Crystalline defects such as vacancies, dislocations, and stacking faults can de- grade the performance of semiconductor electronic devices because they can act as charge traps and scattering sites which reduces electron mobility. Additionally, these types of defects have been theoretically predicted and experimentally shown using electric dipole spin resonance (EDSR), EPR, and deep-level transient spec- troscopy (DLTS) to introduce electric defect states in the band gap of Si [96–99]. Si divacancies can exist in multiple charge states and have multiple deep electronic levels in the band gap, and stacking faults in a Si crystal lead to a defect state in the band gap which is approximately 100 meV above the valence band [100]. Crystalline defects in 28Si are potentially more detrimental to quantum co- herent devices because they introduce scattering centers and local time varying electric and magnetic fields in the crystal that can contribute to the decoherence of a quantum system nearby. Crystalline defects also introduce local strain fields which are known to cause the appearance of unintentional quantum dots in Si wires due to strain induced conduction band modulation [101]. Local strain fields around a qubit, such as a 31P donor in Si, can also cause internal electric fields which have been shown to Stark shift the donor’s electron energy levels and make the qubit spin more sensitive to electric field noise [102, 103]. The nature and magnitude of the effect of local crystalline defects on the performance of quantum coherent devices is still an open area of research and a question that will need to be addressed as this field progresses. 173 In any device comprised of multiple layers, smooth interfaces can be important in addition to a low level of structural defects within the bulk of the layer. At a basic level, fabricating additional layers or electric gates on top of a device layer that has a rough surface can be challenging and is generally undesirable. Dangling bonds and other defects present at a rough surface may result in an increased density of interface charge traps typically seen at oxide interfaces. Surface or interface roughness has also been predicted to affect the valley states of electrons in Si quantum dots and Si/SiGe quantum wells by causing mixing of valley, spin, and orbital states as well as random fluctuations in the phase of the valley-orbit coupling [104–108]. These effects would vary across devices and be impossible to predict, making operation of quantum devices utilizing the valley degree of freedom more difficult. The overall smoothness or roughness of a deposited film can also be a general indicator of their epitaxial quality or crystallinity, and so it is presented here as an important aspect of the 28Si films discussed in this chapter. Section 5.2 of this chapter discusses the specifics of the experimental setup for samples produced at DC–3 in the final experimental configuration. Sections 5.3 and 5.4 go over the experimental methods for in situ and ex situ sample preparation as well as the deposition conditions used for 28Si samples at DC–3. Section 5.5 discusses the results of SIMS measurements of the enrichment of significant 28Si samples deposited at DC–3. Sections 5.6 explores experiments to deposit films epitaxially at elevated temperatures and characterize their morphology by RHEED, STM and SEM. Section 5.7 discusses measurements of the chemical purity of these samples as measured by SIMS and XPS. Section 5.8 discusses the crystallinity of 174 deposited 28Si films observed by TEM. Finally, Section 5.9 summarizes these results. 5.2 Experimental Setup for Improving Crystallinity and Chemical Purity: Deposition Chamber Samples This chapter discusses the deposition of 28Si samples in the final experimental configuration for this ion beam deposition system as it was designed as a whole and used in previous work [59]. In this experimental setup, samples are located at DC–3 as shown in Fig. 4.1. A number of experimental improvements and use of new analytic capabilities are enabled by connecting the ion beamline (including the deceleration lens chamber) to the deposition chamber. In this configuration, samples were located after the deceleration lenses, but in the deposition chamber. As described in Chapter 2, the lenses protrude into the deposition chamber and stop approximately 1 cm before the sample location. As a consequence, the lens chamber and the deposition chamber are actually a single connected vacuum environment. The ion beam chamber is separated from these chambers just before the deceleration lenses by a gate valve. The mass-selecting aperture used between the ion beam and the lens chamber here had a slightly differ- ent geometry than the one used for the depositions in Chapter 4. The mass-selecting aperture slit width was increased to 2 mm to allow a larger 28Si ion fluence to pass into the deceleration lens section. Previously, a thinner slit width was chosen to limit the range of mass values that can pass through from the beam thus increasing 175 the mass resolving power. However, the experiments at LC–2 in Chapter 4 demon- strated better beam tuning and very high geometric selectivity as represented by both SIMS measurements and mass spectra. Higher geometric selectivities allow for the aperture slit width to be increased without impacting the realized selectivity. The height of the slit was decreased from 15.25 mm to 12 mm. This was done to decrease the conductance of gas from the higher pressure ion beam chamber into the deposition chamber while maintaining the fluence of the ion beam, which has a typical spot size at the aperture of < 10 mm in the slit height direction. The thick- ness of the aperture was reduced to a knife edge at the opening to reduce potential scattering of ions as well as sputtering of the aperture as they pass through. A secondary aperture with dimensions 12.7 mm by 6.4 mm was also installed at the beginning of the deceleration lenses for this configuration. This functioned as a gas aperture to block the gas diffusing past the mass-selecting aperture from entering the deceleration lens column. The only outlet for this gas would be at the sample, but with the gas aperture in place, it is instead diverted around the lens column where it may be better pumped away in the lens chamber. A photograph of this secondary gas aperture is shown in Fig. B.8 in Appendix B. Samples were mounted onto sample holders which were then introduced into the vacuum chamber via the load lock and placed onto the 5-axis manipulator, all of which are described in more detail in Chapter 2. The manipulator and sample are positioned to face the deceleration lenses prior to deposition of 28Si. Similar to the sample mounting setup described of Chapter 4 using an electrical vacuum feedthrough at LC–2, samples discussed in this section are electrically isolated from 176 the chamber as they sit in the sample holder on the manipulator. This isolation allows for the ion beam current to be monitored during deposition while controlling the sample potential. In the previous experimental setup with samples at LC–2, the fixed sample aperture that sat over the sample allowed for current collection and focusing of the ion beam through the sample aperture (see Section 4.3.1). Here, an interchangeable sample aperture, which can be inserted into the sample position and then removed prior to deposition, provides similar capabilities. One advantage of the interchangeable sample aperture is that the sample can be completely removed from the path of the ion beam while beam tuning, focusing, and mass spectrum sweeps are performed on the aperture, thus keeping the sample pristine prior to deposition. A sample aperture which was 2.2 mm in diameter was used for most samples, although other diameters including 3 mm, 2.5 mm, and 1 mm were also used. Additionally, the 5-axis manipulator allows for sample motion in three spacial dimensions and precise positioning in the vacuum chamber relative to the ion beam optics. The advantage of this is that, unlike in the setups of Chapter 4, the ion beam can be optimally tuned and the sample repositioned so that the beam spot is located at any desired location on the sample, typically at the center. A photograph of a sample mounted in a sample holder and on the manipulator at DC–3 was shown in Fig. 2.23 in Chapter 2. Photographs of the interchangeable sample apertures can be found in Fig. B.13 in Appendix B. A significant experimental improvement enabled by this experimental configu- ration with samples located at DC–3 is sample heating. As described in Chapter 2, the 5-axis manipulator provides two methods of sample heating. First is a tungsten 177 wire back heater that sits behind the sample holder and can radiatively heat it up to around 900 ◦C, although with a relatively slow response time of several minutes. This is the RH method of sample heating referred to in Chapter 2. The second method uses electrical contacts on the manipulator that interface with the sample holder and provide the ability to pass current directly through the Si substrate, heating it resistively. This is the DH method of sample heating referred to in Chap- ter 2, and it can be used to heat the Si chip as high as its melting point with a very fast response time of less than a second. Sample heating crucially enables in situ substrate preparation and cleaning, deposition at elevated temperatures, and post-deposition sample annealing. These abilities allow for control over a critical experimental degree of freedom for achieving high-quality epitaxial deposition of 28Si [51] and will be discussed later in this chapter. Connecting the ion beam and lens chambers to the deposition chamber pro- vides additional differential pumping at the sample location from a turbo pump and an ion pump, which result in a lower chamber base pressure and a lower back- ground pressure during deposition. Additionally, for this configuration, an ion pump was added to the lens chamber. The typical base pressure of the deposition cham- ber for the experiments described in this chapter was approximately 6.7× 10−9 Pa (5.0× 10−11 Torr). Heating the sample before deposition usually causes some de- gassing that results in slightly elevated pressures. Typically, the pressure immedi- ately before starting deposition was approximately 2.6× 10−8 Pa (2.0× 10−10 Torr), although it varied almost an order of magnitude from approximately 7.1× 10−9 Pa to 5.7× 10−8 Pa (5.3× 10−11 Torr to 4.3× 10−10 Torr). Additionally, the depo- 178 sition chamber and lens chamber, being of all UHV construction, were baked to approximately 150 ◦C prior to these experiments to reduce the partial pressures of water as well as carbon and oxygen containing compounds. The remaining domi- nant partial pressure in these chambers was H2. The RGA in the deposition chamber gives insight into the specific atomic and molecular species that make up the base background and deposition pressures in this chamber. An example of the residual gas mass spectrum corresponding to the base pressure of the deposition chamber was shown in Fig. 2.22 in Chapter 2. For the experiments described in this chapter, the ion source was operated with a working pressure of SiH4 between 1.0× 10−4 Pa and 3.3× 10−4 Pa (7.5× 10−7 Torr to 2.5× 10−6 Torr), similar to that of samples deposited at LC–2 in Chapter 4. The most common working pressure used was ap- proximately 2.0× 10−4 Pa (1.5× 10−6 Torr). In addition to these typical low pres- sure operating conditions, the high pressure mode of the ion source plasma was also explored in which the typical operating pressure was approximately 1.3× 10−3 Pa (1.0× 10−5 Torr). In this experimental setup, another RGA was also installed in the ion beam chamber to diagnose the base pressure gas components and contaminants in the SiH4 gas, both of which may diffuse into the deposition chamber. A residual gas mass spectrum of the base pressure from the ion beam chamber was shown in Fig. 2.3 in Chapter 2. Another significant experimental advantage to depositing 28Si samples at DC– 3 vs. at IC–1 or LC–2 is having access to the various analytical and other features of the deposition and analysis chamber, which are described in Chapter 2. The load lock allows for quicker loading of multiple samples while maintaining the very high 179 vacuum level of the deposition chamber. Sample heating is critical for epitaxial deposition and feedback mechanisms in the form of analysis tools in the deposi- tion chamber are necessary for developing the correct experimental procedures to achieve epitaxy. RHEED and the STM provide information on the state of the pre-deposition substrate, the growth mode during deposition, and surface charac- teristics of the deposited sample. The RGA and AES can provide ambient and surface chemical information for samples. Finally, deposition parameters for 28Si can be modified to produce higher quality films by comparing them to the natural abundance Si films deposited from the electron beam evaporator (i.e. the EFM) in the same system. 5.3 Sample Preparation 5.3.1 Ex Situ Cleaning A variety of Si substrates were used for the 28Si samples discussed in this chapter, but they were all natural abundance, electronic grade, single crystalline commercial Si(100) wafers. A complete table of substrates used here is given in Ta- ble C.1 in Appendix C. Si(100) has been shown to facilitate higher quality epitaxial deposition at lower temperatures than other Si surface orientations such as Si(113), Si(111) or Si(110) [87,109,110]. N-type, p-type, and undoped (intrinsic) wafers were used. The p-type Si wafers were all boron-doped with resistivities between 1 Ω · cm and 20 Ω · cm and thicknesses between 300 µm and 380 µm. These wafers were used to deposit 22 samples at DC–3 and were obtained from both ITME and University 180 Wafer. Substrates from University Wafer have an unknown history because they are reclaimed wafers. The n-type wafers were all phosphorous-doped, mostly with resistivities of between 1 Ω · cm and 10 Ω · cm, and were 300 µm thick, although the substrate used for one sample originated from 600 µm thick stock. These wafers were used with eight samples and were obtained from University Wafer. Five samples were deposited on intrinsic Si wafers in these experiments, and they were obtained from University Wafer, were 380 µm thick, and had a resistivity of > 20 kΩ · cm. The 28Si samples that were deposited using the high pressure plasma deposition mode (a total of five) were deposited onto phosphorous-doped wafers with a resis- tivity between 7 Ω · cm and 20 Ω · cm. They were 300 µm thick with a minimal misalignment or miscut angle relative to the (100) plane of ± 0.05◦ and were ob- tained from Virginia Semiconductor. The wafers from Virginia Semiconductor were float-zone refined with an atomic concentration of O of < 9× 1017 cm−3 and an atomic concentration of C of < 5× 1013 cm−3. Initially, for about one quarter of the samples deposited at DC–3, there was no ex situ cleaning performed on the substrates, which were loaded into the vacuum chamber with a native oxide. This oxide was then thermally desorbed in situ in a process described below. As a comparison to those samples with no cleaning, two samples were treated with HF to strip off the oxide immediately prior to loading them into the vacuum chamber. These initial samples were cleaved by hand into approximately 5 mm by 12 mm chips, which is the approximate maximum sample size that can be accepted by the sample holders used in these experiments. The substrates used for the remainder of the samples deposited at DC–3 were cut using 181 a dicing saw into 4 mm by 10 mm chips after depositing a thin photoresist layer to protect the surface from the accumulation of Si dust during dicing. Individual chips were then cleaned using a more rigorous cleaning procedure designed for complemen- tary metal-oxide-semiconductor (CMOS) technology to remove metals and organics from Si surfaces. This clean consists of a piranha etch, HF strip, and “Standard Clean 2” (SC–2) [111], and it was adopted to improve the substrate surface clean- liness and quality after feedback from analysis of earlier samples. The full cleaning procedure is as follows: 1. photoresist remover (PG Remover) at 70 ◦C for 10 min, 2. fresh PG Remover at room temperature for 5 min, 3. isopropanol (IPA) rinse for 1 min, 4. deionized water rinse, and N2 blow dry, 5. 6:1 H2SO4:H2O2 for 12 min with no deliberate heating (piranha etch), 6. deionized water rinse, and N2 blow dry, 7. 50:1 H2O:HF for 10 s, 8. deionized water rinse and, and N2 blow dry, 9. 5:1:1 H2O:HCl:H2O2 at 80 ◦C for 12 min (SC–2), 10. deionized water rinse, and N2 blow dry. PG Remover is a standard solvent used to remove the protective photoresist from the substrates here. That is followed by a rinse in IPA and deionized water to remove residual photoresist and PG Remover residue and prepare the substrate surface for chemical cleaning. The first clean is a piranha etch designed to remove organics from the surface through oxidation. This was chosen over the “Standard Clean 1” 182 (SC–1) solution, which also cleans organics, because (SC–1) is known to sometimes roughen the wafer surface. A water rinse and N2 blow dry is used between each cleaning step to remove chemical residues. The substrates are next etched with HF to remove the surface oxide containing contaminants and also remove some ionic metal contaminants. Finally, SC–2 solution is used to remove remaining ionic metal contaminants from the surface before a final deionized water rinse. As a result of the SC–2 clean, the chips are left with a thin protective oxide before being mounted onto sample holders and loaded into the vacuum chamber via the load lock. This protective oxide has been shown to reduce carbon contamination on the Si(100) surface as compared to a substrate which was treated with HF to strip the oxide prior to loading [112], and that result was confirmed in this work and will be discussed later in this chapter. Substrates were only handled with clean teflon tweezers. Typically, after being loaded, samples sat about a week in the vacuum chamber while being prepared before deposition. 5.3.2 In Situ Preparation After substrates are introduced into the vacuum chamber via the load lock, they are prepared in situ using the degassing and flash annealing procedures first described in Chapter 2. This well known UHV high temperature Si “flashing” procedure is used to prepare an atomically clean Si(100) (2×1) reconstructed sur- face [113,114], enabling epitaxial deposition. Before flashing, samples are loaded into the deposition chamber after sitting in the load lock for approximately one day un- til the pressure drops to around 1.3× 10−6 Pa (1.0× 10−8 Torr). Sample substrates 183 and the sample holders next need to be degassed slowly and thoroughly before they can be flashed to high temperature in order to limit the pressure increase during the flashing procedure. The magnitude of the pressure spike that occurs during a high temperature flash is a critical parameter for forming a clean Si surface. A bare Si surface should not be exposed to a background pressure higher than roughly 1.3× 10−7 Pa (1.0× 10−9 Torr), or it will become contaminated and more defec- tive. This pressure is based on experience from this work and communications with other labs as well as the literature [114]. Substrates are degassed using both the RH back heater to heat the sample holder and DH power to heat the chip itself. Degassing occurs at temperature of approximately 600 ◦C for at least 12 hours, and the pressure in the deposition chamber is monitored during the temperature ramp up to ensure that the pressure stays below 1.3× 10−7 Pa. This pressure criteria is only precautionary for the degassing step if the substrate has a protective oxide. After degassing, the DH power is used to rapidly flash anneal the sample in a few seconds up to higher temperatures for a short period of time. Typically, a 1 min flash to 1050 ◦C is initially used to desorb the thin oxide layer (if one is present) that was either a native oxide or remained from the ex situ cleaning process [115]. This initial flash also degasses the sample holder more thoroughly before further flashes to higher temperatures. After each flash anneal, the sample temperature is dropped rapidly back to around 600 ◦C. The flash temperature is then sequentially increased during the next several (one to five) flashes depending on the level of outgassing during flashing. These initial higher temperature flashes have a duration of approximately 15 s. The pressure spike that occurs from the higher temperature 184 outgassing (mostly from the sample holder) determines the duration and number of times that each flash is repeated at a certain temperature. If the pressure during flashing reaches the critical 1.3× 10−7 Pa value, the flash is interrupted to reduce the pressure. This cycle is repeated until a final flash temperature of approximately 1150 ◦C to 1200 ◦C is reached. The sample sits at these higher temperatures for a max of 10 s each. The purpose of the higher temperature flashes is to desorb any trace amounts of oxide or other contaminants remaining on the surface and anneal surface defects such as missing or buckled dimers to produce a clean, well ordered Si(100) (2×1) surface. After the final flash, the sample is cooled slowly at a rate of approximately 1 ◦C/s, which additionally helps to recrystallize the surface, down to either the deposition temperature or near room temperature. Typically, as few as five total flashes are needed to prepare a substrate, although some samples/sample holders outgas more, requiring tens of flashes. A typical flash-anneal sequence for substrates used for many of the 28Si samples deposited at DC–3 is as follows: 1. degas initially at ≈ 400 ◦C for 1 h, 2. degas at ≈ 600 ◦C for > 12 h, 3. flash to ≈ 1050 ◦C for 1 min then return to 600 ◦C, 4. flash to ≈ 1150 ◦C for 15 s then return to 600 ◦C, 5. repeat step 3, 6. flash to ≈ 1200 ◦C for 10 s then return to 600 ◦C, 7. repeat step 5, 8. flash to ≈ 1150 ◦C for 10 s then return to 850 ◦C, 9. ramp down the temperature at 1 ◦C/s to the desired final temperature. 185 Again, each flashing step is repeated until the pressure in the chamber remains less than 1.3× 10−7 Pa for the maximum duration of the flash before the next flash step proceeds. In order to provide feedback on this high temperature in situ substrate prepa- ration process and assess its effectiveness for each sample before deposition, two analysis techniques are used: RHEED and STM. RHEED is used during the flash- ing process to identify oxide remaining on the surface and verify the (2×1) recon- struction of the Si surface. RHEED images are typically captured between flashes when the substrate is at ≈ 600 ◦C, but not typically after step 6 above so as to not introduce possible contamination due to the RHEED electron beam interacting with the surface. The number of flashes required to prepare a sample is dictated by the quality of the RHEED pattern in addition to the amount of outgassing. The highest temperature flash may be repeated until a sufficient RHEED pattern typical of a clean Si (2×1) surface is observed (discussed below). RHEED is also used to screen for samples contaminated with silicon carbide (SiC) or those with particu- larly rough surfaces. Figure 5.1 shows two RHEED images of a Si substrate before and after flash annealing at DC–3 to prepare a clean surface. Panel (a) shows the diffraction pattern from an intrinsic Si(100) substrate before flashing. Faint spots are seen in a (1×1) pattern originating from diffraction to rods in reciprocal space, but the image is dominated by the diffuse background caused by the amorphous native SiO2 layer. This image was acquired at a substrate temperature of ≈ 600 ◦C with the electron beam in the 〈110〉 direction. Panel (b) shows the sample substrate after flash annealing seven times to a maximum temperature of ≈ 1197 ◦C for 8 s. 186 Figure 5.1: RHEED diffraction patterns showing the transition from Si with a native oxide to a reconstructed Si(100) (2×1) surface during the substrate flash annealing procedure. (a) Before flashing the substrate, a diffraction pattern showing weak Si (1×1) spots on a diffuse background is seen which indicates the presence of the SiO2 surface layer. (b) After flashing the substrate to ≈ 1197 ◦C, the diffraction pattern shows strong Si(100) (2×1) spots indicating a clean, reconstructed surface. Both images were acquired at a substrate temperature of about 600 ◦C with the electron beam in the 〈110〉 direction. After flashing, strong Si (2×1) diffraction spots appear in the pattern between the outer (1×1) spots, and the diffuse background is gone. These features plus the ab- sence of additional spots due to chemical contaminants such as C indicates that the substrate has a reasonably clean, crystalline, and reconstructed Si(100) surface. RHEED serves as a quick feedback mechanism during the substrate flashing procedure by indicating the general state and quality of surface, and a high-quality 187 RHEED pattern corresponding to a clean Si surface is a necessary condition for producing an atomically clean Si(100) ideal for epitaxial deposition. A RHEED pattern indicating a clean surface is not, however, a sufficient condition because the RHEED pattern is an amalgam resulting from diffraction over a macroscopic area (≈ 0.25 mm2) on the substrate. RHEED is thus insensitive to trace surface defects and contaminants. To assess the surface quality of prepared substrates on an atomic scale, STM imaging is used. The STM used in this system was briefly described in Chapter 2. After samples are flashed and cooled at a deliberate rate to approximately 250 ◦C using the DH power, they cool radiatively for at least 10 min to > 50 ◦C before being transferred into the STM chamber. The typical base pressure in the STM chamber is approximately 6.7× 10−9 Pa (5.0× 10−11 Torr), and it is separated from the deposition chamber by a gate valve. The samples remain protected in the isolated UHV environment of the STM during ion beam tuning in the deposition chamber. The substrates are typically scanned both on relatively large areas (1 µm × 1 µm) and smaller areas (50 nm × 50 nm) to screen for both larger particulates or other features as well as atomic scale defects. Most STM images shown in this chapter were acquired in conjunction with Hyun soo Kim, who assisted with some aspects of these experiments. Several typical STM images of three different sub- strate types after flash annealing are shown in Fig. 5.2. These STM topography images all show clean Si(100) (2×1) reconstructed surfaces with atomic steps and atomically flat terraces of various widths. Panels (a) and (b) show images of boron- doped Si substrates with a relatively small average terrace width of approximately 188 Figure 5.2: STM topography filled state images of clean Si(100) (2×1) surfaces for three substrate types prepared in situ by flash annealing. Images were typically acquired with a tip bias ≈ -2 V and a tunneling current ≈ 100 pA. These images show atomic steps on the Si surface and flat terraces of varying widths. Si (2×1) dimer rows are also seen in (a), (b), (d), and (f) with a minimal density of dimer defects (dark spots). (a) and (b) are boron-doped Si substrates with an average terrace width of around 15 nm. (c) and (d) are intrinsic Si substrates with an average terrace width of around 55 nm. (e) and (f) are phosphorous-doped Si substrates with an average terrace width of about 100 nm. 189 15 nm. The terrace size is a result of a misalignment during manufacturing of the cut direction of the wafer and the (100) crystal plane. When this miscut angle is small, larger terraces result. Panels (c) and (d) show images of intrinsic Si substrates with an average terrace width of approximately 55 nm. Finally, panels (e) and (f) show images of the phosphorous-doped Si substrates with a small miscut angle of ± 0.05◦ resulting in a large average terrace width of approximately 100 nm. The increase in average terrace width is apparent when noting the difference in scale between panels (a), (c), and (e). Panel (c) shows an area roughly 25 times larger than (a), and panel (e) shows an area four times larger than (c). Si (2×1) dimer rows are also seen predominantly in panel (b) and are just visible in (a), (d), and (f). A minimal density of dimer row defects (dark spots) are seen, which confirms that the surface cleaning procedures were sufficient. These defects can simply be missing surface atoms or due to chemical contaminants [116]. After the substrate surface is inspected via STM, and the ion beam is tuned for deposition, the sample is moved back to the deposition chamber and flashed a final time to remove possible adsorbates which may have accumulated during STM scanning. The substrate temperature is then lowered to the deposition temperature before 28Si deposition commences. This cool down typically takes around 30 min as the substrate comes into equilibrium. 190 5.4 Deposition of 28Si Once substrates are prepared and ion beam tuning similar to that of Chapter 4 is complete, 28Si is deposited in the deposition chamber at DC–3 using a range of deposition conditions. As was mentioned previously, one of the critical degrees of freedom for achieving epitaxial deposition that is enabled in this final experimental setup is the substrate temperature. 28Si samples were deposited using a range of substrate temperatures, T , from room temperature (≈ 21 ◦C) up to a maximum of 1080 ◦C and many temperatures in between. Depositing at room temperature allows for comparison between these samples and those deposited in the previous setup at LC–2. The most common sample deposition temperatures used were either around 450 ◦C or 700 ◦C. The range of temperatures was used to determine the effect of temperature on both the epitaxial quality of the 28Si films and the adsorption of SiH4 gas into samples, which then increases the 29Si and 30Si isotopic concentrations. The first of these effects is described in a later section of this chapter, and the second is described in Chapter 6. For the samples deposited at DC–3, both the nominal low pressure working mode of the ion source as well as the high pressure working mode were used, which led to a range of deposition parameters. When using the low pressure mode, during deposition, the pressure in the deposition chamber rose to between 1.6× 10−7 Pa and 1.5× 10−6 Pa (1.2× 10−9 Torr to 1.1× 10−8 Torr) due to the SiH4 gas diffusion from the ion beam chamber. Most samples, however, were deposited in a background pressure of about 1.1× 10−6 Pa (8.0× 10−9 Torr). This pressure is a factor of three 191 lower than the low end of the range of background pressures measured during de- position of samples at LC–2 in Chapter 4. The RGA in the deposition chamber is routinely used to monitor the partial pressures of various chemical contaminants present in the deposition chamber during 28Si deposition. A typical RGA mass spectrum recorded while depositing a 28Si sample with a substrate temperature of 705 ◦C at DC–3 is shown in Fig. 5.3. H2 dominates the spectrum and is the result of both residual H2 in the deposition chamber and also cracking of SiH4 gas in the ion source releasing H2 which diffuses to the deposition chamber. SiH4 gas also diffuses into the chamber, however, cracking from the RGA itself reduces the SiH4 signal to mostly just the 28 u peak. This peak also contains N2 and CO. SiH4 2+ hydride peaks do appear between 14 u and 16 u at 0.5 u increments. Partial pressures of other chemicals of interest can be seen including C, H2O, F, and CO2. The peak appearing at 26 u is not known, but it may be related to an alcohol. When using the high pressure mode, the pressure in the deposition chamber rose to approximately 4.0× 10−6 Pa (3.0× 10−8 Torr), which is comparable to the samples deposited at LC–2. Ions were deposited with a range of average ion energies, Ei, at the sample between 20 eV and 40 eV. This energy range is smaller with lower overall values than the samples discussed in Chapter 4 in order to standardize the deposition process and minimize sample sputtering. The estimated sputter yield for 30 eV 28Si ions hitting a crystalline Si surface is approximately 2 % (see Fig. 2.19 in Chapter 2). The most common average ion anergy used here was approximately 35 eV. For the low pressure mode, the average ion beam current achieved was approximately 192 Figure 5.3: Residual gas mass spectrum collected from the RGA in the deposition chamber while operating the ion beam and depositing a 28Si sample at 705 ◦C at DC–3. The spectrum is dominated by H2 which diffuses from the ion beam chamber as SiH4 gas is cracked in the ion source. The SiH4 complex of peaks normally is reduced to only the 28 u peak due to cracking of the molecules by the RGA itself. N2 and CO also make up the 28 u peak. Doubly charged SiH4 hydrides do appear at around 15 u. Other potential chemical contaminants are present including C, F, H2O, and CO2. 0.55 µA. One major advantage of the high pressure plasma mode is that it yields a much larger ion current on average, and for these samples, an ion current of approximately 2.5 µA was achieved. A mass spectrum for a Si ion beam with an ion energy of 37 eV generated using the low pressure mode of the ion source is shown in Fig. 5.4. This spectrum was collected using the interchangeable sample aperture at DC–3 prior to depositing a 28Si sample. The corresponding magnet current used for the field sweep of the magnetic sector mass analyzer is shown on the top axis. Ion current peaks (circles) 193 Figure 5.4: SiH4 mass spectrum representative of the ion beam settings for samples deposited at DC–3 using the low pressure mode of the ion source. The ion current (circles) is recorded while sweeping the the analyzer current, and thus the magnetic field (top axis). The 28 u peak is 28Si and the 29 u peak is both 28SiH and ≈ 5 % 29Si. Several higher order hydrides are also shown. Gaussian fits (line, Eq. (2.14)) to the 28 u and 29 u peaks are shown superimposed on the data. The centers of the 28 u and 29 u fits are separated by ≈ 10 σ. corresponding to 28Si and Si hydrides are observed between 28 u and 33 u. The ion current peaks in this spectrum appear qualitatively similar to the mass spectrum acquired in the previous experimental configuration (Fig. 4.8). The ion peak at 28 u is 28Si and the ion peak at 29 u contains ≈ 5 % 29Si based on the similar peak height to that of 28 u and the expected natural abundance, as discussed in Chapter 2. These peaks show a high degree of separation with no detectable ion current signal occurring between the peaks. Gaussian fits to the peaks (line, Eq. (2.14)) are also shown superimposed on the data, and one can see only a slight asymmetry in 194 the beam profile and no shoulder peaks, which indicates a fairly optimally tuned beam with minimal scattering off of lens elements. The mass resolving power of the ion beam in this configuration derived from this mass spectrum is m ∆m ≈ 57 (measured at 10 % of the peak height). This value is similar to but slightly lower than the mass resolving power achieved in the previous experimental configuration perhaps due to the wider mass-selecting aperture used in this setup. However, no current is detected between the peaks, and the geometric selectivity is still expected to be sufficient to produce highly enriched 28Si films comparable to those produced at LC–2. The Gaussian fits give a separation of the 28 u peak from the 29 u peak of approximately 10 σ. A mass spectrum for a Si ion beam with an ion energy of 31 eV generated using the high pressure mode of the ion source is shown in Fig. 5.5. This spectrum was also collected using the interchangeable sample aperture prior to depositing a 28Si sample. Again, the corresponding current used for the field sweep for the magnetic sector mass analyzer is shown on the top axis. The ion current peaks (circles) observed between 28 u and 33 u show the peak shapes typical of the high pressure mode. This spectrum shows a dramatic increase in the 28Si ion current over the low pressure spectrum and a decrease in the 30 u and 31 u peaks. The 28Si ion current becomes larger while the hydride peaks become smaller. This is because the hydrides are cracked more efficiently in the high pressure ion source plasma mode, as discussed in Chapter 2. The 28Si ion current peak at 28 u and the ion current peak at 29 u show a fairly good separation but with a current level between the peaks of about 8 nA, which is higher than for the low pressure mode. Using the high pressure 195 Figure 5.5: SiH4 mass spectrum representative of the ion beam settings for samples deposited at DC–3 using the high pressure mode of the ion source. The ion current (circles) is recorded while sweeping the mass analyzer current, and thus the magnetic field (top axis). The 28 u peak is 28Si and the 29 u peak is both 28SiH and ≈ 24 % 29Si. Several higher order hydrides are also shown. Gaussian fits (line, Eq. (2.14)) to the 28 u and 29 u peaks are shown superimposed on the data. The centers of the 28 u and 29 u fits are separated by ≈ 8 σ. mode, 29Si makes up approximately 24 % of the 29 u peak based on the peak heights and the expected natural abundance, as discussed in Chapter 2. Gaussian fits to the peaks (line, Eq. (2.14)) are also shown superimposed on the data as a sum. An asymmetry in the lower portion of the peaks possibly indicates some beam scattering to the lower mass side. The mass resolving power derived from this mass spectrum is m ∆m ≈ 48 (measured at 10 % of the peak height), which is reduced further from the previous low pressure mode mass spectrum. The asymmetric peak and the lower mass resolving power may reduce the geometric selectivity and lower the achievable 196 enrichment level in 28Si samples deposited using this mode. However, the roughly factor of five increase in 28Si current compared to the 29 u peak containing 29Si may offset any deleterious effects of a lower observed peak separation. The Gaussian fits give a separation of the 28 u peak from the 29 u peak of approximately 8 σ. Tuning and focusing the ion beam onto the interchangeable sample aperture on the sample manipulator resulted in deposition spot sizes similar to those of samples deposited at LC–2. The average deposition area for samples deposited at DC–3 was approximately 6 mm2 and as small as 2 mm2 for a few samples. During deposition, RHEED was typically used to periodically monitor the structure of the depositing film and verify the location of the deposition spot on the chip. The appearance of the 28Si deposition spot on the substrate of samples deposited at DC–3 varied depending on the deposition temperature. Optical micrographs of three samples deposited at three different temperatures at DC–3 are shown in Fig. 5.6. These micrographs show the Si(100) substrates with the 28Si deposition spot appearing at the center as three visually distinct areas. Panel (a) shows 28Si deposited on an intrinsic Si chip with a substrate temperature of approximately 249 ◦C in an area about 4.7 mm long. The deposition spot is clearly visible as a discolored patch on the chip and is visually similar to the samples deposited at room temperature at LC–2 in Chapter 4. This may indicate poor epitaxial quality. Panel (b) shows 28Si deposited on a phosphorous-doped Si chip at a substrate temperature of about 421 ◦C in an area approximately 2.6 mm wide. The deposition spot is nearly indistinguishable from the substrate and only a faint outline is visible. This may indicate that the film is structurally similar to or even epitaxially matched with the substrate. The 197 Figure 5.6: Optical micrographs of three 28Si samples deposited at DC–3 at three different temperatures. The Si(100) substrates are seen in all three micrographs with the 28Si deposition spot at the center. (a) 28Si deposited on an intrinsic Si chip at 249 ◦C in an area 4.7 mm long is clearly visible. (b) 28Si deposited on a phosphorous- doped Si chip at 421 ◦C in an area 2.6 mm wide is nearly indistinguishable from the substrate. (c) 28Si deposited on a boron-doped Si chip at 705 ◦C in an area 3.7 mm wide is clearly visible. 198 brightness and contrast of this micrograph was altered to highlight the edges of the deposition spot, which is why the substrate appears lighter. Panel (c) shows 28Si deposited with a substrate temperature of approximately 705 ◦C on a boron-doped Si chip in an area approximately 3.7 mm wide. The deposition spot is clearly visible on the chip and appears as a diffuse whitish color. Light scattering from the surface of the deposited area is probably due to a higher degree of surface roughness, which will be discussed later in this chapter. The thicknesses of these 28Si films ranged from around 45 nm to 370 nm as measured by TEM and/or inferred from SIMS depth profiling. The corresponding deposition rates for samples deposited using the low pressure mode of the ion source were between 0.32 nm/min and 1.41 nm/min with the most common deposition rate around 0.80 nm/min. The high pressure mode of the ion source resulted in a higher range of rates between 2.2 nm/min and 4.6 nm/min. These higher rates are desirable because they enable production of thicker samples in a typical deposition time (assuming similar spot sizes), and they lessen the relative adsorption rate of gaseous species from the background pressure into the 28Si film during deposition. To summarize, the typical deposition procedure for samples deposited at DC–3 was as follows: 1. a substrate is diced from a wafer and chemically cleaned ex situ before being mounted onto a sample holder and loaded into the load lock which is pumped out for > 12 h, 2. the substrate is moved into the deposition chamber and degassed on the sample manipulator at 600 ◦C for > 12 h, 3. the substrate is flash annealed to ≈ 1200 ◦C to prepare a clean (2×1) surface 199 which is inspected by RHEED, 4. after a deliberate temperature ramp down to ≈ 250 ◦C following flash anneal- ing, the substrate cools radiatively for at least 10 min to ≈ 50 ◦C and is then transferred to the STM surface inspection, 5. with the sample in the STM, the ion beam is tuned and characterized using the interchangeable sample aperture on the manipulator. The mass spectrum, ion beam energy, beam spot focusing, and the deposition chamber RGA are all analyzed and recorded, 6. the substrate is transferred back to the deposition chamber after the gate valve to the ion beam is closed to reduce the background pressure where it is flashed once and set to the deposition temperature and position, 7. the gate valve to the ion beam is opened to commence deposition of 28Si while monitoring the ion current onto the sample as a measure of the deposition rate, 8. RHEED is periodically used to monitor the crystallinity of the deposition and locate the deposition spot, 9. after typically three to five hours of deposition, the gate valve to the ion beam is closed to end deposition and reduce the deposition chamber pressure back to its base, then the sample is cooled to room temperature, 10. the sample is then transferred to the STM to inspect the surface morphology of the deposited 28Si film, and finally, 11. the sample may be annealed at around 600 ◦C back on the manipulator and later inspected by STM again before ultimately being removed from the vac- uum chamber via the load lock for ex situ analyses. In total, 40 28Si samples were deposited at DC–3 including one room temper- ature sample, 32 samples deposited at elevated temperatures using the low pressure mode, and seven samples deposited at elevated temperatures using the high pressure mode. 200 5.5 Enrichment Measurements via SIMS for DC–3 Samples 5.5.1 Initial Tests at DC–3 SIMS was used to assess the enrichment of a large number of samples deposited at DC–3. The basic procedure for SIMS isotope measurements of 28Si samples was discussed in Chapter 4. Isotope ratio measurements by SIMS were used to calcu- late the isotope fractions (zSi/Sitot.). SIMS depth profiles presented in this chapter were acquired in collaboration with Dr. David Simons (NIST). Initial SIMS mea- surements were done to verify that the high levels of 28Si enrichment achieved for samples deposited at LC–2 in Chapter 4 were reproducible for samples deposited at DC–3 after the significant experimental reconfiguration of connecting the ion beam to the deposition chamber. Recall the mass-selecting aperture width was increased, which may decrease the realized mass selectivity. Additionally, lower partial pres- sures of SiH4 during deposition reduced the potential for adsorption of SiH4 into the 28Si films, which results in increased 29Si and 30Si isotope fractions. Along those same lines, it must be determined what role the substrate temperature plays in the enrichment of the samples. While the substrate deposition temperature is increased to facilitate epitaxial deposition, it may also change the kinetics and chemistry of any possible SiH4 adsorption into the samples. Higher temperatures may reduce gaseous adsorption or perhaps the added energy may enhance incorporation into the depositing films. Addressing this topic required many SIMS measurements of 201 samples deposited at a large range of temperatures (some of which are shown here) and is the subject of Chapter 6. Depositing a sample at room temperature should exclude any potential sub- strate temperature effects on the enrichment and provide a SIMS measurement more comparable to those of the samples deposited at LC–2 to verify the enrichment level of samples is no worse for this setup. A SIMS depth profile for a 28Si sample de- posited at room temperature (≈ 21 ◦C) using the low pressure mode of the ion source at DC–3 is shown in Fig. 5.7. Measurements of the 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. the sputter depth into the sample. This sample was relatively thin, and so the data density of the measurement is lower than for other samples. Also, a fairly short region (≈ 20 nm) exists where the iso- tope fractions reach a minimum and are averaged for the measurement. The count rate for a number of SIMS measurement cycles in this region was zero and thus that data does not appears on this semi-log plot. At a depth of around 40 nm, the isotope fractions begin to increase to their natural abundance values (dotted and dashed lines) in the transition into the Si(100) substrate, which is marked by the shaded region. The interface between the film and the substrate is estimated to be at a depth of about 53 nm, which is also the film thickness. The average measured 28Si isotope fraction in this sample is 99.999898(35) %. After some initial surface contamination, the average residual 29Si isotope fraction measured between 13 nm and 41 nm is 0.58(26)× 10−6 (0.58(26) ppm), and the average 30Si isotope fraction is 0.44(23)× 10−6 (0.44(23) ppm). Not only is the measured level of enrichment not diminished after the experimental transition to 202 Figure 5.7: SIMS depth profile of a 28Si sample deposited at room temperature using the low pressure mode at DC–3. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction in the film is 0.58 ppm, and the average 30Si isotope fraction is 0.44 ppm (dashed lines). Note that data are zero counts in some SIMS cycles. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. At a depth of 40 nm, the isotope fractions begin to increase to their natural abundance values in the Si(100) substrate. The interface between the film and the substrate (shaded region) is estimated to be at a depth of 53 nm. this final experimental configuration, it is slightly better than the most highly en- riched sample deposited at LC–2 (0.691(74) ppm 29Si). This may be the result of a combination of the background pressure during deposition of this sample being roughly a factor of three lower than that of the sample deposited at LC–2 and the deposition rate being nearly a factor of three higher. These counter acting condi- tions would in fact result in only a slightly better enrichment. Also, apparently the widening of the mass-selecting aperture did not measurably decrease the realized selectivity in the 28Si samples. 203 Figure 5.8: SIMS depth profile of a 28Si sample deposited at 249 ◦C using the low pressure mode at DC–3. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction in the film is 0.79 ppm, and the average 30Si isotope fraction is 0.229 ppm (dashed lines). Note that most data are zero counts in particular SIMS cycles for 30Si. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. At a depth of about 280 nm, the isotope fractions begin to increase to their natural abundance values in the substrate (shaded region). The interface between the film and the substrate is estimated to be at a depth of 305 nm. The room temperature sample can also be compared with samples deposited at elevated temperatures. Figure 5.8 shows a SIMS depth profile for a sample deposited at approximately 249 ◦C at DC–3 using the low pressure mode. The 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. the sputter depth into the sample. This much thicker sample shows a large flat region in the isotope fraction profile and an abrupt increase up to the natural abundance values in the substrate indicating a sharp interface. At a depth of about 280 nm, the isotope fractions begin to increase to their natural abundance values (dotted and dashed 204 lines) in the transition into the substrate, which is marked by the shaded region. The interface between the film and the substrate is estimated to be at a depth of approximately 305 nm, giving a value for the film thickness. For this sample, the average measured 28Si isotope fraction in the film is 99.999898(13) %. After sputtering through some initial surface contamination, the average residual 29Si isotope fraction is 0.79(12)× 10−6 (0.79(12) ppm), and the av- erage 30Si isotope fraction is 0.229(64)× 10−6 (0.229(64) ppm). For the 30Si data, most of the SIMS measurement cycles in this region contained zero counts and thus the data does not appears on this semi-log plot. These averages were calculated using the data between 32 nm and 285 nm. The level of enrichment of this sample is very similar to that of the room temperature sample. In fact, the 28Si isotope frac- tion is identical with slightly different proportions of 29Si and 30Si. When this result is combined with a SIMS measurement of a second area of this sample, the average 29Si isotope fraction of the two measurements is 0.90(10) ppm. These measurements shows that 28Si samples can successfully be deposited at elevated temperatures us- ing the low pressure mode of the ion source at DC–3 while maintaining a very high level of enrichment. 5.5.2 Enrichment Progression Timeline Samples Out of the 40 samples produced at DC–3, three are represented on the en- richment progression timeline (Fig. 4.2) in Chapter 4 as new record enrichments achieved in this work at the time of their deposition, and they will be discussed in this section. In fact, these samples achieved the highest enrichments of any samples 205 produced in this work. A SIMS depth profile for the first of these 28Si samples deposited using the low pressure mode of the ion source at DC–3 with a higher level of enrichment than the previous ones deposited at LC–2 is shown in Fig. 5.9. This sample was deposited at a substrate temperature of approximately 610 ◦C using the low pressure mode of the ion source. The 28Si (circles), 29Si (squares), and 30Si (tri- angles) isotope fractions are shown vs. the sputter depth into the sample. Beyond some initial surface contamination, there is only a relatively thin region between 20 nm and 70 nm over which the 29Si isotope fraction reaches a sustained minimum value. In this region of the film, the isotope fractions were averaged. At a depth below 70 nm, the 29Si isotope fraction increases more than two orders of magnitude. This increase is likely a result of extended ion beam tuning off of the 28 u peak after deposition had commenced. At a depth of around 125 nm, the 29Si and 30Si isotope fractions together begin to increase to their natural abundance values (dotted and dashed lines) in the transition into the substrate, which is marked by the shaded region. The interface between the film and the substrate is estimated to be at a depth of approximately 162 nm, giving a value for the total film thickness. The length over which the isotope fractions return to their natural abundance values is relatively wide in this sample compared to the previous sample, possibly indicating a larger degree of surface roughness. Within the highly enriched portion of this sample, the average measured 28Si isotope fraction is 99.9999570(70) %. The average residual 29Si isotope fraction in the film is 3.00(60)× 10−7 (300(60) ppb), and the average 30Si isotope fraction is 1.30(37)× 10−7 (130(37) ppb). Some of the SIMS measurement cycles in this region 206 Figure 5.9: SIMS depth profile of a 28Si sample deposited at 610 ◦C using the low pressure mode at DC–3. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The isotope fraction values were averaged between 20 nm and 70 nm. The average 29Si isotope fraction in this part of the film is 300 ppb, and the average 30Si isotope fraction is 130 ppb (dashed lines). Note that most data are zero counts in particular SIMS cycles. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. At a depth of about 70 nm, the 29Si isotope fraction increases two orders of magnitude likely as a result of ion beam tuning off of the 28 u peak. At a depth of 125 nm, the 29Si and 30Si isotope fractions begin to increase to their natural abundance values. The interface between the film and the substrate (shaded region) is estimated to be at a depth of 162 nm. for the 29Si data and most of the measurement cycles for the 30Si data contained zero counts and thus the data does not appears on this semi-log plot. These mea- surements indicate an isotope reduction factor (az/( zSi/Sitot.), discussed in Chapter 4) for 29Si of 1.6(3)× 105, i.e. the 29Si is approximately 1.6(3)× 105 times lower than in natural abundance Si. The isotope reduction factor for 30Si is slightly higher at 2.4(4)× 105, although it is similar to that of 29Si within the uncertainty of the 207 values. The 29Si isotope fraction of this sample is a little more than a factor of two less than that of the most highly enriched sample deposited at LC–2 in Chapter 4, as seen in the jump from LC–2 to DC–3 in Fig. 4.2. This decrease in 29Si and 30Si isotope fractions was probably due to the significantly lowered background deposi- tion pressure for these samples, which would result in a lower amount of adsorbed SiH4. Additionally, this result shows that very high enrichments continue to be achieved while the substrate deposition temperature is increased further compared to the previous sample discussed in this section. Despite the high level of enrichment achieved in a portion of the film, this sample is clearly not ideal due to the region of elevated 29Si. Additional SIMS measurements of samples deposited under similar conditions show the elevated 29Si to be an anomaly. The second sample represented on the enrichment progression timeline in Fig. 4.2, which was deposited at DC–3 and achieved an even higher level of en- richment over the previous sample is one that was deposited with a substrate tem- perature of approximately 712 ◦C using the low pressure mode. A SIMS depth profile of the area of highest enrichment of this sample is shown in Fig. 5.10. The 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. the sputter depth into the sample. Beyond the initial signal due to surface contamina- tion, the 29Si and 30Si isotope fractions show an extended minimum. At a depth of around 227 nm, the isotope fractions begin to increase to their natural abundance values (dotted and dashed lines) in the transition into the substrate, which is marked by the shaded region. The interface between the film and the substrate is estimated to be at a depth of approximately 256 nm, giving a value for the film thickness. 208 Figure 5.10: SIMS depth profile of a 28Si sample deposited at 712 ◦C using the low pressure mode at DC–3. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction in the film is 132 ppb, and the average 30Si isotope fraction is 70 ppb (dashed lines). Many data in this region are zero counts in the SIMS cycles. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. The isotope fraction values were averaged between 21 nm and 227 nm, where the 29Si and 30Si isotope fractions begin to increase to their natural abundance values. The interface between the film and the substrate (shaded region) is estimated to be at a depth of 256 nm. The average measured 28Si isotope fraction in the highly enriched portion of the film is 99.9999797(30) %. The average residual 29Si isotope fraction is mea- sured to be 1.32(27)× 10−7 (132(27) ppb), and the average 30Si isotope fraction is 7.0(12)× 10−8 (70(12) ppb) (dashed lines). Many of the SIMS measurement cycles in this region contained zero counts and thus the data does not appears on this semi- log plot. The averages were calculated from the data between 21 nm and 227 nm. These measurements indicate an isotope reduction factor for 29Si of 3.5(7)× 105. 209 The isotope reduction factor for 30Si is slightly higher at 4.4(8)× 105, although they are similar within the uncertainty of the values. A second area of this sam- ple was also measured by SIMS giving an average 29Si isotope fraction of the two measurements of 163(18) ppb. As mentioned in Chapter 4, this variation between measurements on the same sample is likely due to a difference in deposition rate at the two areas of the film. The enrichment of this sample represents yet another decrease of more than a factor of two in the isotope fraction of 29Si from the 610 ◦C sample, which can be seen in Fig. 4.2, the enrichment progression timeline. This reduction is possibly due to a lower rate of incorporation of SiH4 into the film from a slight decrease in deposition background pressure and also a nearly twofold increase in deposition rate for this sample compared to the previous one. The deposition was approximately 1.22 nm/min, which is fairly high compared to other samples deposited using the low pressure mode. This was a result of both a slightly higher average ion beam current of about 740 nA and better beam spot tuning that led to a more compact than average deposition spot of about 4.2 mm2. A complicating factor which may effect the enrichment of this sample is an anomalous amount of nitrogen observed in the ion beam mass spectrum before depositing this sample. This matter will be addressed in the later sections of this chapter. The third and final sample represented on the enrichment progression timeline in Fig. 4.2 that was deposited at DC–3 is a sample deposited at approximately 502 ◦C using the low pressure mode. This 28Si sample is the most highly enriched sample produced in this entire work. A SIMS depth profile of the area of highest 210 enrichment within this sample is shown in Fig. 5.11. The 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. the sputter depth into the sample. Beyond the initial signal due to surface contamination, the 29Si and 30Si isotope fractions show an extended minimum with very little variation in value. At a depth of around 290 nm, the isotope fractions begin to increase to their natural abundance values (dotted and dashed lines) in a fairly abruptly transition into the substrate, which is marked by the shaded region, indicating a sharp interface. The interface between the film and the substrate is estimated to be at a depth of approximately 321 nm, which also gives a value for the film thickness. The average measured 28Si isotope fraction in this highly enriched film is 99.9999819(35) %. This value is the highest enrichment of any sample measured in this work. Additionally, the average residual 29Si isotope fraction is the lowest for any sample with a value of 1.27(29)× 10−7 (127(29) ppb), and the average 30Si isotope fraction is 5.5(19)× 10−8 (55(19) ppb) (dashed lines). The averages were calculated from the data between 30 nm and 290 nm. They lie well below where the data appears because many of the SIMS measurement cycles in this region contained zero counts and thus the data is not represented on this semi-log plot. These en- richment measurements of 29Si and 30Si represent the best isotope reduction factors for this entire work. The isotope reduction factor for 29Si is 3.7(8)× 105, and the isotope reduction factor for 30Si is higher with a value of 5.6(19)× 105. These values are actually similar to each other to within their uncertainties due to a large relative uncertainty in the measurement of the isotope fraction of 30Si in this sample. The level of enrichment of this sample is similar to that of the 712 ◦C sample, and while 211 Figure 5.11: SIMS depth profile of the most highly enriched 28Si sample deposited at DC–3. This sample was deposited with a substrate temperature of 502 ◦C using the low pressure mode. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction in the film is 127 ppb, and the average 30Si isotope fraction is 55 ppb (dashed lines). Many data in this region are zero counts in the SIMS cycles. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. The isotope fraction values were averaged between 21 nm and 292 nm, where the 29Si and 30Si isotope fractions begin to increase to their natural abundance values. The interface between the film and the substrate (shaded region) is estimated to be at a depth of 321 nm. the center value of this measurement is lower than that of the previous sample, they agree within their uncertainties as seen in the enrichment progression timeline in Fig. 4.2. Like the 712 ◦C sample, the extremely low isotope fractions of 29Si and 30Si observed in this sample are likely due to a lower rate of incorporation of SiH4 into the film. This is probably due in part to a relatively high deposition rate of 1.41 nm/min, which is actually the highest deposition rate for samples deposited using the low pressure mode at DC–3. Again, the deposition temperature may also 212 be playing a part in the reduction of 29Si and 30Si although it is not clear from this data alone. A total of three SIMS measurements were made on different areas of this sample resulting in an average 29Si isotope fraction for the three areas of 233(39) ppb. The single area best value of 127(29) ppb 29Si residual isotope fraction is given more weight in this discussion than the multiple measurement average because the average depends on the distribution of the measurement areas across regions of the film with different deposition rates. The single area best value also gives a lower bound on the best possible enrichment achievable by the deposition system for the specific ion beam selectivity, deposition rate, background pressure, and substrate temperature used. The samples already described in this chapter, which were deposited with el- evated substrate temperatures, demonstrate enrichment levels similar to or better than those of the highly enriched room temperature samples deposited at LC–2, i.e. isotope fractions < 1 ppm of 29Si. However, some samples deposited at DC–3 with elevated substrate temperatures showed significantly higher isotope fractions of 29Si and 30Si. One sample deposited at approximately 357 ◦C using the low pressure mode was measured by SIMS to have a average 28Si isotope fraction of 99.999405(93) %. The average residual 29Si isotope fraction in the film is mea- sured to be 4.18(70)× 10−6 (4.18(70) ppm), and the average 30Si isotope fraction is 1.77(61)× 10−6 (1.77(61) ppm). These higher isotope fractions are likely due in part to a low deposition rate for this sample of 0.33 nm/min and a higher background pressure during deposition of approximately 1.5× 10−6 Pa (1.1× 10−8 Torr), which 213 equals the highest pressure for any sample deposited at DC–3 using the low pressure mode of the ion source. Another sample deposited with a substrate temperature of 421 ◦C using the low pressure mode was measured in one area to have an average 28Si isotope frac- tion of 99.999812(25) %. The average residual 29Si isotope fraction measured for this sample is 1.30(22)× 10−6 (1.30(22) ppm), and the average 30Si isotope fraction is 5.8(12)× 10−7 (0.58(12) ppm). Averaging the measurements of three areas on the deposition spot of this sample gives an overall average 29Si isotope fraction of 1.48(13) ppm. The average deposition rate corresponding to the three measured areas of this sample was also lower than average at approximately 0.46 nm/min. The background deposition pressure was similar to that of the 357 ◦C sample. 5.5.3 Samples with Deposition T > 600 ◦C The highest residual 29Si and 30Si isotope fractions were measured in sam- ples deposited with a substrate temperature above 600 ◦C, suggesting that higher substrate deposition temperatures effect the enrichment. This temperature depen- dence will be discussed in Chapter 6. SIMS measurements of a sample deposited at approximately 705 ◦C using the low pressure mode show a maximum average 28Si isotope fraction of 99.999488(48) %. The average residual 29Si isotope frac- tion is 3.30(25)× 10−6 (3.30(25) ppm), and the average 30Si isotope fraction is 1.82(42)× 10−6 (1.82(42) ppm). The 29Si isotope fraction averaged between two measurements on this sample is 4.06(37) ppm. Unlike the 357 ◦C and 421 ◦C sam- ples, this 705 ◦C sample had a higher estimated average deposition rate of approxi- 214 mately 0.74 nm/min, which is average for samples deposited at DC–3. This sample was also deposited in a slightly lower background deposition pressure than the pre- vious two sample. These results are counter intuitive when comparing them to past results discussed throughout Chapter 4 and this chapter, and thus they seem to provide supporting evidence for the substrate temperature having an affect the enrichment. Another similar sample with elevated 29Si and 30Si isotope fractions was de- posited with a substrate temperature of approximately 812 ◦C using the low pres- sure mode. The average 28Si isotope fraction measured by SIMS for this sam- ple is 99.99907(10) %. The average residual 29Si isotope fraction in the film is 4.32(46)× 10−6 (4.32(46) ppm), and the average 30Si isotope fraction is slightly higher at 4.96(93)× 10−6 (4.96(93) ppm). This sample was deposited in a slightly lower background pressure than the 705 ◦C sample and had a slightly higher es- timated deposition rate of 0.90 nm/min, which is again counter intuitive because higher deposition rates were seen to lower the residual isotope fractions previously. In addition to higher deposition temperatures, other differences exist between these samples and others discussed in this chapter that may affect the measurement of the isotope fractions. Unlike the SIMS measurements of the lower temperature samples, the measurements of some of the samples deposited above 600 ◦C result in only a very short range in the depth profile where the isotope fractions reach a minimum value. A SIMS depth profile of the 812 ◦C sample discussed above is shown in Fig. 5.12 and exhibits this effect. The 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. the sputter depth into the sample. 215 Figure 5.12: SIMS depth profile of a 28Si sample deposited at 812 ◦C using the low pressure mode at DC–3. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. After some surface contamination, the 29Si and 30Si isotope fractions reach a minimum at a depth of 25 nm and then gradually increase through the rest of the film to the natural values in the substrate. The natural abundance values for each isotope (dotted and dashed lines) are shown for reference. The 29Si and 30Si isotope fractions are averaged at this minimum between 14 nm and 32 nm. The average 29Si isotope fraction at the film minimum is 4.32 ppm, and the average 30Si isotope fraction is 4.96 ppm (dashed lines). The interface between the film and the substrate (shaded region) is estimated to be at a depth of 158 nm. This depth profile is qualitatively very different from previously shown SIMS depth profiles. After sputtering through the typical surface contamination, the 29Si and 30Si isotope fractions reach a minimum value at a depth of about 25 nm and then gradually and immediately increase through the remaining 133 nm of deposited film up to the natural values in the substrate (dotted and dashed lines). The interface between the film and the substrate is estimated to be at a depth of approximately 158 nm, marked by the shaded region. The 29Si and 30Si isotope fraction averages 216 (dashed lines) were calculated for a small region where the signals reach its minimum between 14 nm and 32 nm. SIMS depth profiles of other samples deposited at temperatures of approximately 708 ◦C, 759 ◦C, 804 ◦C, and 1041 ◦C also exhibit the same qualitative depth profile seen in Fig. 5.12. The apparent gradual increase in 29Si and 30Si isotope fractions throughout the enriched film of these samples may be explained by considering self-diffusion of 29Si and 30Si isotopes from the naturally abundant Si substrate into the deposited film. As the 28Si film is depositing, it is effectively being annealed by the elevated tem- perature of the substrate. At deposition temperatures of about 700 ◦C and above, the thermal activation of Si self-diffusion may be enough to produce the isotope fraction gradients seen in the SIMS depth profiles. To explore this hypothesis, Si diffusion profiles are calculated and compared to the measured SIMS depth profiles. Si self-diffusion is believed to be dominated by self-interstitials above 800 ◦C with an activation energy of approximately 4.75 eV [117]. The concentration of 29Si or 30Si in an enriched 28Si film, C(x), at a depth below the film surface, x, due to Si isotope diffusion is given by C(x) = Csub + Cfilm 2 + Csub − Cfilm 2 erf ( x− d 2(DSDSi t) 0.5 ) , (5.1) where Csub is the concentration of 29Si or 30Si in the substrate, Cfilm is the concen- tration of 29Si or 30Si in the enriched film without diffusion, d is the film thickness, t is the deposition time (annealing time), and the Si self-diffusion coefficient, DSDSi , is given by DSDSi = (530) exp ( − Ea kBT ) . (5.2) 217 Here, Ea is the activation energy, mentioned above. Equations (5.1) and (5.2) as well as the exponential prefactor of Eq. (5.2), 530 cm2 · s−1, were taken from Ref. [117]. These equations are used to calculate the expected concentration of 29Si from diffusion into enriched 28Si films and compare it to the measured SIMS depth profile shown in Fig. 5.12. The calculated concentration profiles for 29Si in a 28Si film resulting from several different deposition temperatures are shown in Fig. 5.13. In these calculations, 29Si diffuses from the natural abundance Si substrate (shaded region) into the 28Si film during deposition with elevated substrate temperatures which anneal the film. The film thickness and thus film/substrate interface is arbitrarily set at a depth of 100 nm below the film surface. Calculated 29Si concentration profiles are shown for deposition temperatures of 700 ◦C (solid line), 750 ◦C (dotted line), 800 ◦C (dash-dot line), 850 ◦C (dash-dot-dot line), and 900 ◦C (dashed line). These calculations also used a total deposition time of 4 h, which is an average deposition time for samples produced at DC–3, and a nominal 29Si isotope fraction in the 28Si film of 1 ppm. The 29Si profiles show that there is relatively little diffusion expected from the substrate into the 28Si film for these time and temperature combinations. Most of the samples deposited with substrate temperatures within the range of these calculation were deposited at around 700 ◦C or 800 ◦C, and the profiles for these temperatures show significant concentrations of 29Si only within approximately 1 nm of the interface. Even for the 900 ◦C profile, the 29Si concentration drops to less than double the nominal film concentration beyond a distance of 10 nm from the substrate interface. This small amount of diffusion would be not even be detected in the SIMS depth profiles because it is still smaller 218 Figure 5.13: Calculated 29Si concentration profiles in a 28Si film from isotope diffu- sion at elevated temperatures. 29Si diffuses from the natural abundance Si substrate (shaded region) at a depth of 100 nm into the 28Si film during deposition. Atomic concentration profiles for five deposition temperatures from 700 ◦C (solid line) to 900 ◦C (dashed line). These profiles were calculated from Eq. (5.1) for a deposition time of 4 h and a nominal 29Si isotope fraction in the 28Si film of 1 ppm. than the typical 25 nm region over which the isotope signals transition from low isotope fractions in the film to high isotope fractions in the substrate. Figure 5.13 shows that isotope self-diffusion from the substrate into the depositing 28Si films is not responsible for the SIMS depth profile shape in Fig. 5.12. Another possible explanation for the apparent gradual increase in 29Si and 30Si isotope fractions throughout the films of some samples deposited at higher temperatures is that a large amount of surface roughness is causing an artifact in the SIMS measurement. Producing an accurate SIMS depth profile relies on the assumption that all of the sputtered atoms contributing to the signal at a given time 219 in the measurement originated from the same plane in the sample relative to the film- substrate interface. The presence of surface roughness invalidates this assumption because it results in the SIMS beam sputtering atoms from different depths at the same time. Here, surface roughness is characterized by a surface width, ∆z, which is the total width of the region at the surface between the highest peak and lowest valley. This surface roughness will be transferred down through the film during the SIMS measurement, and when the sputter beam reaches the interface between the 28Si film and the substrate, it will begin sampling the substrate in some areas while still sampling the film in other areas. This effect artificially inflates the measured 29Si and 30Si isotope fractions near the substrate interface by mixing the higher 29Si and 30Si signals from the substrate with the lower signals from the 28Si film. The relationship between ∆z and the total film thickness, d, determine whether a reliable SIMS measurement is possible. The range over which this artifact will manifest is always similar to the size of ∆z, and it only manifests near the substrate interface. So, for ∆z  d, the effect does not impact the ability to make a good measurement of the film. However, for ∆z ∼ d, the signal from the measurement artifact would be comparable or dominant to the signal of the true measurement of the film. Surface roughness appears to be a reasonable explanation for the 28Si samples deposited at higher temperatures that exhibit a SIMS depth profile qualitatively similar to the one in Fig. 5.12, with a gradual 29Si and 30Si isotope fraction increase through the depth profile of the film. Many of these samples are observed to have much rougher surfaces as deposited than samples deposited at lower temperatures, and for several samples, ∆z was of the order of d, as determined by SEM and 220 TEM cross-sectional microscopy. Details of this observed roughness is discussed in a later section. A tilted SEM cross-sectional micrograph of a sample deposited with a substrate temperature of approximately 708 ◦C using the low pressure mode is shown in Fig. 5.14. This micrograph was acquired in collaboration with Dr. Joshua Schumacher (NIST). The Si(100) substrate and the 28Si film both appear dark in the lower half of this micrograph with the substrate at the bottom and the film above it. The 28Si film is indistinguishable from from the substrate in this image without obvious grains or other features. A dashed line represents the approximate location of the interface between the substrate and the 28Si film, determined from TEM cross-sectional microscopy. The light region above the film is Pt deposited to protect the sample, which was cross-sectioned using a focused ion beam (FIB). The top surface of the Pt is also visible because the sample is viewed at a tilt angle of 52◦. The surface of the 28Si appears very rough with a maximum ∆z of approximately 80 nm. The thickest areas of the film are approximately 120 nm. The inset shows a cartoon of a rough 28Si film on a substrate being sputtered by SIMS in an isotope measurement. SIMS sputter beams (vertical arrows) sample both a thick and thin region of the film. Sputtering in the thick region results in isotope signals from the 28Si film, while sputtering at the same time in the thin region results in isotope signals partially originating from the natural abundance substrate. This is the process that leads to the measurement artifact which inflates the 29Si and 30Si isotope fractions above the nominal film values as the measurement approaches the interface. The measured 28Si enrichment values for all samples that exhibit the roughness 221 Figure 5.14: SEM tilted cross-sectional micrograph of a 28Si film deposited at 708 ◦C at DC–3. The tilt angle is 52◦. The Si(100) substrate and 28Si film are indistin- guishable and appear dark in the lower half of this micrograph with a dashed line representing the film-substrate interface. The light region above the film is Pt de- posited to protect the sample. The top surface of the Pt is visible in this tilted image. The surface of the 28Si film appears very rough with a maximum surface width ∆z ≈ 80 nm. The thickest areas of the film are almost 120 nm. Sputtering this surface in a SIMS measurement introduces a measurement artifact. The inset shows a cartoon of a rough 28Si film being sputtered by SIMS beams (arrows) in a thick and thin region of the film. Sputtering in the thick region results in iso- tope signals from the 28Si film, while sputtering in the thin region results in isotope signals partially from the substrate. induced SIMS measurement artifact need to be considered as a lower bound on the true film enrichment. Conversely, the measured 29Si and 30Si isotope fraction for these samples are an upper bound. This means that the isotope fractions are not larger than the measured values, but they may be smaller due to the SIMS artifact inflating the results. In this work, however, only SIMS measurements which are believed to show an accurate minimum value for 29Si and 30Si isotope fractions are reported. 222 The accuracy of the reported enrichment values of the 812 ◦C sample shown above in Fig. 5.12 can be evaluated by comparing them to those of another sample similar sample which does not show evidence for a large amount of surface roughness. A sample deposited with a substrate temperature of approximately 808 ◦C using the low pressure mode of the ion source was measured by SIMS and showed a more typical depth profile without obvious effects from a measurement artifact. Also, large scale roughness was not observed in a top down SEM micrograph of this sample due to it being mostly amorphous from a higher than usual N content, which will be discussed later. The SIMS depth profile of area of highest enrichment for this sample is shown in Fig. 5.15. The 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. the sputter depth into the sample. Beyond the initial signal due to some surface contamination, the 29Si and 30Si isotope fractions do show an extended minimum between a depth of 22 nm and 81 nm. The average isotope fraction values were calculated by averaging the data in this region. Beyond a depth of about 81 nm, the isotope fractions begin to increase to their natural abundance values (dotted and dashed lines) in the transition into the substrate, which is marked by the shaded region. The interface between the film and the substrate is estimated to be at a depth of approximately 112 nm, which also gives a value for the film thickness. Also shown for comparison is the 29Si isotope fraction depth profile (squares and solid line) from the 812 ◦C sample shown above in Fig. 5.12. The depth scale of this profile was shifted by compressing it so that the locations of the substrate interfaces of this profile and the depth profiles of the 812 ◦C sample were aligned 223 Figure 5.15: SIMS depth profile of a 28Si sample deposited at 808 ◦C using the low pressure mode at DC–3. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. After some surface contamination, the 29Si and 30Si isotope fractions reach an extended minimum between a depth of 22 nm and 81 nm where they are averaged. The average 29Si isotope fraction in the film is 3.97 ppm, and the average 30Si isotope fraction is 2.23 ppm (dashed lines). At a depth of 81 nm the 29Si and 30Si isotope fractions increase to their natural abundance values in the substrate (shaded region). The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. The interface between the film and the substrate is estimated to be at a depth of 112 nm. For reference, the 29Si depth profile for the 812 ◦C sample (squares and solid line) is plotted showing the effect of surface roughening on the SIMS measurement. The depth scale of this profile was shifted to match the film thickness of the 808 ◦C sample. with each other. Clearly, the depth profiles of the two sample are qualitatively different. The 29Si and 30Si profiles of the 808 ◦C sample do not show signs of the surface roughness induced SIMS measurement artifact. Unlike the 812 ◦C sample, the 29Si and 30Si isotope fraction values of the 808 ◦C sample remain low through most of the film thickness. These depth profiles seem to support the conclusion that the depth profiles with elevated isotope fractions throughout the film, such 224 as those of the 812 ◦C sample, are not due to Si isotope self-diffusion at elevated temperatures. Despite the qualitative differences of the depth profiles of the two samples, the minimum in 29Si isotope fractions for the two samples appears to be similar. The average measured 28Si isotope fraction in this film is 99.999380(36) %. The average residual 29Si isotope fraction is 3.97(31)× 10−6 (3.97(31) ppm), and the average 30Si isotope fraction is 2.23(19)× 10−6 (2.23(19) ppm) (dashed lines). A second area of this sample was also measured giving an average 29Si isotope fraction for the two areas on the sample of 4.33(23) ppm. The average 29Si isotope fraction in the more highly enriched area of this sample is quite similar in value to the average measured value of the 812 ◦C samples, which is 4.32(46) ppm, and they agree within their uncertainties. This agreement is possibly partially coincidental because they were deposited with slightly different deposition parameters. However, this result does show that SIMS measurements of samples exhibiting the roughness induced measurement artifact can give accurate enrichment values, with the caveat that they still may be bounds on the true film enrichment values. 5.5.4 High Pressure Mode Sample All of the previous samples discussed in this chapter were deposited using the low pressure working mode of the ion source, but as mentioned in a previous section discussing the deposition parameters of 28Si at DC–3, it is not clear if the benefit of a significantly increased 28Si ion beam current produced in the high pressure mode is offset by a lower mass resolving power. Figure 5.5 showed that the total 28Si current 225 of the high pressure mode was typically a factor of five larger than the currents produced in the low pressure mode. An increased current is beneficial because it allows for faster growth rates which enables production of thicker samples in a typical deposition, and it reduces the relative adsorption rate of gaseous species into the 28Si film, including SiH4. However, Fig. 5.5 also showed that the mass resolving power was lower than had been observed using the low pressure mode with a measurable ion current signal between the 28 u peak and the 29 u peaks of approximately 8 nA. This current level may result in a lower geometric selectivity and thus lower realized enrichments. Additionally, the high pressure mode requires higher pressures in the ion source that result in a background pressure during deposition three times higher than the highest pressures experienced by samples deposited using the low pressure mode. Higher partial pressures of SiH4 may also decrease the enrichment level in the samples. In order to examine the viability of the high pressure mode as a source of highly enrich 28Si and compare its effectiveness to the low pressure mode, SIMS was used to measure the isotope fractions of a sample deposited with a substrate temperature of approximately 421 ◦C at DC–3 using the high pressure mode of the ion source. A SIMS depth profile of the area of highest enrichment of this sample is shown in Fig. 5.16. The 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. the sputter depth into the sample. Beyond an initial signal due to a small amount of surface contamination, the 29Si and 30Si isotope fractions show an extended minimum with very little variation in value. At a depth of around 290 nm, 226 Figure 5.16: SIMS depth profile of a 28Si sample deposited at 421 ◦C at DC–3 using the high pressure mode. 28Si (circles), 29Si (squares), and 30Si (triangles) isotope fractions are shown vs. sputter depth. The average 29Si isotope fraction in the film is 0.303 ppm, and the average 30Si isotope fraction is 0.103 (dashed lines). Many data in this region are zero counts in the SIMS cycles, especially for 30Si. The natural abundance values for each isotope (dotted and dashed lines) are also shown for reference. The isotope fraction values were averaged between 11 nm and 290 nm, where the 29Si and 30Si isotope fractions begin to increase to their natural abundance values. The interface between the film and the substrate (shaded region) is estimated to be at a depth of 315 nm. the isotope fractions begin to increase to their natural abundance values (dotted and dashed lines) in a fairly abruptly transition into the substrate, which is marked by the shaded region. The interface between the film and the substrate is estimated to be at a depth of approximately 315 nm, which also gives a value for the film thickness. The average measured 28Si isotope fraction in the film is 99.9999594(72) %. The average residual 29Si isotope fraction is 3.03(58)× 10−7 (0.303(58) ppm), and 227 the average 30Si isotope fraction is 1.03(43)× 10−7 (0.103(43) ppm) (dashed lines). These averages lie below where the data appears because many of the SIMS mea- surement cycles in this region contained zero counts and thus the data does not appears on this semi-log plot. This is especially true of the 30Si data. The averages were calculated from the data between 11 nm and 290 nm. The enrichment level of this sample is comparable to that of the most highly enriched samples deposited at DC–3 that appear on the enrichment progression timeline in Fig. 4.2. The 29Si isotope fraction of this sample is nearly identical to that of the 610 ◦C sample, al- though the 30Si isotope fraction of this sample was lower, which resulted in the 28Si isotope fraction for this 421 ◦C high pressure sample being slightly higher than that of the 610 ◦C sample. A second area on this sample was measured by SIMS also which resulted in an average 29Si isotope fraction between the two measurements of 0.355(41) ppm. The 28Si ion current achieved for this sample was approximately 3.00 µA, which is the highest for any sample in this work. This ion current, which was only achieved using the high pressure mode, combined with a relatively small de- position spot size of approximately 3.7 mm2 resulted in a high deposition rate of 3.94 nm/min. This high rate led to the very low 29Si and 30Si isotope fractions measured in this sample despite the higher background pressure during deposition. Additionally, the lowered mass resolving power and high overlap current between peaks in the mass spectrum do not appear to reduce the realized mass selectivity and enrichment in any significant (or at least measurable) way. This measurement showed that the high pressure mode of the ion source is not only viable for pro- 228 ducing highly enriched 28Si films, but from the perspective of enrichment, it is the preferred deposition mode because of the high deposition rates that are achiev- able while maintaining extremely high levels of 28Si enrichment, and conversely, extremely low levels of 29Si isotope fractions. These enrichments are comparable to the most highly enriched 28Si samples produced in this work and are potentially suitable for QI experiments such as a measurement of the dependance of electron coherence times on 29Si concentration in the single spin regime, which was discussed at the beginning of Chapter 4. SIMS measurements of samples deposited at DC–3 show that overall, a re- duction of more than a factor of five in the 29Si isotope fraction of the most highly enriched sample was achieved compared to that of the previous most highly en- riched sample deposited at LC–2, which can be seen in the enrichment progression timeline in Fig. 4.2. These SIMS measurements also shows that these extremely high enrichments are achievable both for samples deposited with elevated substrate temperatures, and for samples deposited using the high pressure mode of the ion source. 5.6 Epitaxial Deposition 5.6.1 Context Epitaxial deposition of Si and other semiconductors at low temperature has been extensively studied and characterized via molecular beam epitaxy (MBE) ex- periments. Eaglesham et al. showed for the first time in 1990 that there exists a 229 critical thickness for epitaxy, hepi, at a given deposition temperature [118]. Beyond this thickness, epitaxy breaks down and the film becomes amorphous. That study found that for a deposition rate of around 0.4 nm/min and a substrate tempera- ture of 200 ◦C, hepi = 25 nm for Si(100). At a deposition temperature of 300 ◦C, hepi increased to 120 nm. hepi was found to increase exponentially with increasing deposition temperature with an activation energy of about 0.4 eV, although this value is dependent on the deposition rate. The concept of a critical epitaxial thick- ness is only valid for these lower temperatures and typical deposition rates of a few nm/min. Above a temperature of 500 ◦C, solid phase epitaxy (SPE) begins to dom- inate the growth because the recrystallization front propagates faster (4.2 nm/min) than the deposition rate, which effectively extends hepi to be infinite. This regime is sometimes referred to as unlimited epitaxy. Numerous research groups later studied low temperature Si epitaxy phenom- ena, and some of these efforts were nicely summarized by Eaglesham [119]. For low temperature MBE of Si and Ge and other similar deposition methods including sputter deposition and ion assisted deposition (IAD), it is observed that an amor- phous phase eventually develops after epitaxial growth. Prior to this amorphous phase, an epitaxial but highly defective region forms in the film [109,110,120–122]. In this intermediate layer before amorphization occurs, the most apparent defects observed by TEM are stacking faults and microtwins. The epitaxial, defective, and amorphous regions of a Si film from Ref. [109] are seen in a cross-sectional TEM mi- crograph in Fig. 5.17, separated by dashed lines. This Si(100) film was deposited at 270 ◦C by IAD. The bottom of the film (I) is epitaxial and defect free, then the film 230 Figure 5.17: TEM cross-sectional micrograph of a Si(100) film deposited at 270 ◦C by IAD. Three visible regions are separated by dashed lines and include an epitaxial region (I) above the substrate (substrate not shown), a defective region (II) approxi- mately 80 nm thick with visible stacking faults, and an amorphous region (III) with nanocrystallites. The arrow indicates several stacking faults on {111} planes. Insets show magnified views of each region. (from Ref. [109]) becomes defective above that in the middle region (II). Stacking faults are visible in this region (marked by the arrow) running along the {111} planes. The film then becomes amorphous beyond hepi in the top region (III). The insets show magnified views of the three regions where the crystallinity is clearly visible. The density of stacking faults has been seen to generally increase with increasing deposition temperature [110] until hepi becomes unlimited at higher temperatures. Another morphological feature frequently seen in similar TEM micrographs of epitaxial thin films is the formation of pyramidal structures bounded by the {111} stacking faults, 231 which result in a large roughness at the interface between the defective region and the amorphous region. Locally, the transition to the amorphous region is fairly abrupt, however. There have been several proposed explanations for the existence of hepi and the breakdown of an epitaxial layer into a defective layer and finally an amorphous layer in low temperature Si deposition, which are summarized in Ref. [119]. One model attributes the formation of the expitaxial layer to H incorporation during deposition [123]. H then could segregate and accumulate at the growth surface and disrupt the epitaxy by altering the nominal bonding pattern of the lattice. The critical concentration of H for breakdown to occur is expected to be around 2× 1019 cm−3. Experimental observations of the effect of H coverage on epitaxy seem to indicate, however, that H cannot solely be responsible for the epitaxial breakdown in thin films [119,124] because epitaxy is possible on H terminated sur- faces. Another explanation for the epitaxial breakdown is the accumulation of de- fects in the depositing film until epitaxy is not sustainable [125]. This mechanism, however, requires very large defect densities, possibly as high as 1× 1014 cm−2. Es- timates of the density of extended defects in Si(100) put the number much lower at around 1× 107 cm−2 [119], which seems to rule out defect buildup as the cause of the amorphous transition. Finally, Eaglesham et al. propose that roughening of the growth surface itself is the cause of the breakdown of an epitaxial film into an amorphous one in low temperature Si epitaxy [126]. Roughening of the surface during deposition may be due to several factors including the presence of impurities, anisotropic surface diffusion, or faceting of the surface. 232 Si low temperature epitaxial deposition by MBE can be augmented using en- ergetic (i.e. hyperthermal energy) ions with tens of eV of kinetic energy, as briefly discussed in Chapter 1. The Si source target can be sputtered by ions to produce a flux of Si with hyperthermal energy, or ions such as Ar or Xe can be used to bombard the surface of the sample during Si epitaxial deposition. In addition to ion assisted deposition, mentioned above, these techniques are also sometimes re- ferred to as ion beam assisted deposition, ion enhanced deposition, or ion enhanced epitaxy, but they will be collectively referred to as IAD here. These techniques use hyperthermal energy ions to impart energy mostly in the form of momentum into the sample during deposition, which has the effect of enhancing the epitaxial quality of the film. Epitaxial deposition of Si and other elements has been demonstrated using IAD with qualitatively similar results to MBE in terms of a limiting epitaxial thickness and stacking fault formation, discussed above. However, IAD has been shown to extend hepi to larger thicknesses for a given deposition temperature and rate. Conversely, IAD can achieve the same hepi as MBE but at a lower tempera- ture [109, 120, 121, 127–129]. Experimental examples of the benefits of IAD include extending hepi to approximately 1 µm for a deposition at 300 ◦C and a growth rate of 6 nm/min [120], and lowering the temperature required for achieving unlimited epitaxy to around 390 ◦C, even at very high deposition rates of 300 nm/min [109]. While Si IAD achieves enhanced epitaxy using a flux of hyperthermal ions of a different element, ion beam epitaxy (IBE) achieves a similar enhancement of epi- taxial deposition using ions of the material being deposited. This technique, which typically generates the ions from an ion beamline, has been studied and used for de- 233 position of Si, Ge, and other materials, as discussed in Chapter 1. Research groups studying IBE of 28Si have demonstrated epitaxial deposition of varying crystalline quality on Si(100) with similar results to IAD. Al-Bayati et al. demonstrated IBE of 28Si with 50 eV ions [42]. Films deposited using a substrate temperature of 400 ◦C were shown by TEM to be epitaxial but defective with stacking faults and twins. Tsubouchi et al. used 40 eV ions and a deposition temperature of 600 ◦C [43] to demonstrate epitaxy. TEM analysis showed that the films were epitaxial but likely contained a high density of dislocations. Rabalais et al. extensively studied the relationship between the epitaxial qual- ity of 28Si films and the ion energy and substrate temperature [51]. They demon- strated epitaxial deposition of varying quality with deposition temperatures between 40 ◦C and 290 ◦C and ion energies between 8 eV and 50 eV. RHEED and TEM analysis showed that for 15 eV ions deposited at 160 ◦C and below, hepi was limited to less than 15 nm. 28Si films deposited at 160 ◦C but an increased ion energy of 20 eV were found to be epitaxial with no visible stacking faults. Additionally, films deposited with 15 eV ions but an increased substrate temperature of 290 ◦C were also epitaxial. Their research showed that for a given deposition temperature, high- quality epitaxial deposition occurred when using ions with energies within a certain optimal range. The 28Si ion energy that resulted in the lowest temperature epitaxial deposition was 20 eV. Other IBE experiments also show optimal epitaxial deposition with ion ener- gies around 20 eV including Matsuoka and Tohno, who observed that the highest quality epitaxial growth occurred with 25 eV Si ions deposited at 400 ◦C, accord- 234 ing to RHEED measurements [130]. 50 eV ions produced epitaxial films, but with stacking fault densities around 1× 1010 cm−2. Deposition using higher ion energies produce more defective films, even at elevated deposition temperatures. Deposit- ing Si at 740 ◦C with 200 eV ions resulted in epitaxial but highly defective films with TEM showing stacking faults and twin structures [131]. Molecular Dynam- ics simulations have also been used to predict a Si ion energy window for optimal epitaxial deposition on Si(100) [132]. These simulations showed that epitaxial depo- sition should be possible even for deposition temperatures below 200 ◦C when using ions with energies between 20 eV and 25 eV. These theoretical results along with the experimental results from Ref. [51] are represented on epitaxy phase diagrams in Fig. 5.18. The quality of the epitaxy is represented in these figures by several regions that occupy different portions of the deposition phase space of substrate temperature, T , vs. ion energy, Ei. Figure 5.18 (a) shows an epitaxy phase diagram from Ref. [132] for the be- havior of Si epitaxy as predicted my molecular dynamics simulations. Regions of unlimited epitaxy (I) and high-quality ion enhanced epitaxy (III) as well as regions where epitaxy is defective and limited (i.e. a finite hepi) by lattice registry errors (II) and vacancy formation (IV). Tepi is the temperature above which unlimited epitaxy is achievable with conventional MBE. T∗epi is the temperature above which hyper- thermal ions can enhance the films epitaxial quality, for a given deposition rate. The phase boundary between limited and enhanced epitaxy at T∗epi occurs near Ed, which is the energy threshold for lattice damage due to the ions and is about 25 eV for Si deposited on Si(100). This indicates that damage caused by ions is beneficial 235 Figure 5.18: Theoretical, (a), and experimental, (b), phase diagrams for Si ion beam epitaxy in the deposition phase space of substrate temperature, T , vs. ion energy, Ei. (a) Si epitaxy phase diagram based on molecular dynamics simulations of the qualitative behavior of the epitaxy for regions of unlimited (I) and high-quality ion enhanced epitaxy (III) as well as regions of limited, defective epitaxy (II and IV). Tepi is the temperature above which unlimited epitaxy is achievable with MBE, and T∗epi is the temperature above which hyperthermal ions enhance epitaxy. Ed is the energy threshold for lattice damage due to ions (from Ref. [132]). (b) Qualitative Si epitaxy phase diagram based on experimental results. TEM and RHEED analysis of 28Si films deposited using various ion energies and substrate temperatures were used to categorize deposition as either high-quality epitaxy (upper region) or defective growth leading to an amorphous phase (lower region) (from Ref. [51]). to the epitaxial quality. Panel (b) of Fig. 5.18 shows an epitaxy phase diagram from Ref. [51] that is similar to the one in (a) and is based on experimental results. TEM and RHEED analysis of 28Si films deposited onto Si(100) using various ion energies and substrate temperatures were used to define the phase boundary between high-quality and defective epitaxy. The epitaxial deposition was categorized as either high-quality epitaxy (upper region) or defective, limited epitaxy (lower region) with a finite hepi leading to an amorphous phase. The boundary between the two epitaxial growth 236 modes is qualitatively similar to the one in panel (a) and shows a kink where high- quality epitaxy is possible at a minimum substrate temperature of 160 ◦C and an ion energy of 20 eV. The energy deposited into the growth surface by hyperthermal ions produces several effects beneficial to epitaxy. Ions can transfer energy to the film through neu- tralization, which should be approximately the ionization potential, 8.15 eV [51]. This energy can excite nearby atoms and enhance their mobility. Molecular dy- namics simulations show that hyperthermal ions can create vacancies that facilitate adatom incorporation [133] during MBE and also suppress the formation of 3D is- lands and step pinning from impurities [127,128]. The optimal ion energy of around 20 eV, observed to lead to higher quality epitaxy, seems to match both the critical energy for defect formation in Si of 20 eV to 40 eV as well as an average Si displace- ment energy of 13 eV to 15 eV [58,134,135]. Calculations show that approximately 70 % of 10 eV Si ions impinging on Si(111) surfaces penetrate about two layers into the surface and stop in an intersticial site before diffusing to the surface to partici- pate in the film growth. The impact and transfer of momentum from the ions leads to the formation of dangling bonds and mobile defects such as Frenkel pairs. When the concentration of these defects is high enough, they can facilitate ordered recrys- tallization and epitaxial growth [136]. Simulations also show that ions with energies greater than 50 eV begin to create more permanent defects and less of the mobile, epitaxy-enhancing defects [51], which is supported by the experimental observations mentioned previously. 237 5.6.2 Morphology of Films with Deposition T > 600 ◦C 28Si samples were deposited at sample location DC–3 with a range of substrate temperatures, as was mentioned previously in this section. Initially, deposition temperatures between 610 ◦C and 1041 ◦C were chosen in order to facilitate high- quality epitaxial growth on Si(100) substrates. Depositing at these temperatures using ions with a typical average ion energy Ei ≈ 33 eV for these samples should result in a growth mode dominated by smooth layer-by-layer growth, in accordance with the epitaxy phase diagram for IBE in Fig. 5.18 (b). Ten of these initial samples, of which there were 12 in total, were deposited on boron-doped (University Wafer) substrates that were not cleaned ex situ and were loaded into the vacuum chamber with a native oxide. The final two 28Si samples deposited in this initial group were prepared ex situ using an HF etch, and one of them was deposited on a phosphorous- doped substrate. 5.6.2.1 RHEED RHEED was used as an initial check on the epitaxial nature and surface mor- phology of these samples immediately after deposition. The RHEED patterns of all of the samples deposited with a substrate temperature above 600 ◦C showed that they were crystalline and epitaxially aligned to the substrate due to the presence of Si(100) diffractions spots. RHEED can only provide a limited view of the overall epitaxial quality of the samples because it is only sensitive to the top few layers, however, it was often used intermittently throughout the deposition to monitor the 238 Figure 5.19: RHEED diffraction pattern of a 28Si sample deposited at 708 ◦C at DC–3. This image was acquired with the sample at the deposition temperature and the electron beam in the 〈110〉 direction. The presence of (1×1) bulk Si diffraction spots in this pattern indicate that the film is crystalline and aligned to the Si(100) substrate. Additionally, this pattern of spots corresponds to a 3D transmission-type pattern for diffraction from a rough surface. Faint (2×1) spots between some of the (1×1) spots are visible and are likely due to part of the electron beam diffracting from the substrate outside of the deposition spot. crystallinity as a function of deposition thickness. The same crystalline RHEED pattern was typically seen from the initial stages of deposition through to the end for these samples. A typical RHEED pattern for these higher temperature 28Si sam- ples is shown in Fig. 5.19. This sample was deposited at 708 ◦C and is about 120 nm at its thickest. The image was acquired with the substrate at the deposition tem- perature and the electron beam in the 〈110〉 direction. It is clear from the presence of (1×1) bulk Si diffraction spots that the 28Si film is crystalline and aligned with the Si(100) surface of the substrate. However, this pattern is quite different from the RHEED pattern shown in Fig. 5.1 for a (2×1) reconstructed Si(100) surface where each diffraction streak collapses into a point for a flat surface. This pattern 239 (Fig. 5.19) corresponds to a 3D transmission-type diffraction pattern where diffrac- tion occurs not at rods, but at a 3D matrix of reciprocal space points. 3D diffraction indicates that the surface of the deposited film is rough such that there is significant transmission of the electron beam through raised features such as mounds or large islands on the surface. Also seen in this RHEED pattern are faint (2×1) spots be- tween some of the (1×1) spots, which are likely due to part of the RHEED electron beam diffracting from an area of the substrate outside of the 28Si deposition spot concurrent with diffraction from the film. A SEM cross-sectional micrograph of this 708 ◦C sample was previously shown in Fig. 5.14 where the rough surface is obvi- ous. The surface roughness of these samples observed by RHEED and then by SEM, which is discussed further below, is unexpected considering the phase diagrams for smooth epitaxy in Fig. 5.18 and the comparatively high deposition temperatures (i.e. > 600 ◦C) used for this initial set of samples. RHEED also shows that the rough surfaces of these samples deposited with substrate temperatures above 600 ◦C develop higher index microfacets. Figure 5.20 shows RHEED images for two 28Si samples deposited at DC–3 with clear signs of faceting on a rough surface. The (1×1) bulk Si diffraction spots present in the patterns of both of these samples indicate that the films are crystalline and aligned to the Si(100) surface, similar to the sample represented in Fig. 5.19. The 3D transmission patterns here indicate a rough surface. Figure 5.20 (a) is the diffraction pattern for a sample deposited at 610 ◦C, and it was acquired with the sample at the deposition temperature and the RHEED electron beam in the 〈110〉 direction. This sample was measured to be approximately 162 nm thick. In the pattern, diffraction 240 Figure 5.20: RHEED diffraction patterns of two 28Si samples deposited at 610 ◦C, (a), and 705 ◦C, (b), at DC–3. Both images were acquired with the samples at the deposition temperatures and the electron beam in the 〈110〉 direction. The (1×1) bulk Si diffraction spots present in both patterns indicate that the films are crystalline and aligned to the Si(100) surface. They are also 3D transmission patterns indicating a rough surface. In (a), additional lines forming a “chevron” pattern emanate from several of the spots (solid arrows) indicating diffraction from microfacets in the {113} family of planes. These lines are parallel to the dashed arrow pointing from the (000) spot to the (113) spot, indicating the direction of the facet face. In (b), additional lines can be seen connecting diffraction spots, highlighted on the right by the dashed lines. These lines run along 〈111〉 directions, as indicated by the arrow pointing from the (000) spot to the (111) spot, and they are due to diffraction from {111} microfacets. Si(100) (2×1) spots are also seen in (b) likely due to part of the electron beam diffracting from the substrate outside of the deposition spot. 241 intensity from additional lines are observed emanating down and outward from some of the spots, marked by the solid arrows. These lines form “chevron” patterns and indicate diffraction from microfacets in the {113} family of planes. This classification is evident when noting that these lines run parallel to the dashed arrow pointing from the (000) spot to the (113) spot, indicating the plane of the facet face to be {113}. These same {113} facet “chevron” patterns are seen on a number of other RHEED patterns for samples deposited with deposition temperatures above 600 ◦C. Panel (b) of Fig. 5.20 is the diffraction pattern for a sample deposited at 705 ◦C. It was acquired with the sample at the deposition temperature and the RHEED electron beam in the 〈110〉 direction. This sample was measured to be approximately 144 nm at its thickest. Similar to the pattern in panel (a), this pattern shows diffraction intensity from additional lines connecting some of the adjacent 3D diffraction spots. These lines are visible in the left side of the image and are highlighted by the dashed lines in the right side of the image. The arrow pointing from the (000) spot to the (111) spot indicates that these diffraction lines run along 〈111〉 directions and are due to diffraction from {111} microfacets. Superimposed on the 3D pattern in this image is the nominal Si(100) (2×1) spot and rod pattern, which is likely due to part of the electron beam diffracting from an area of the substrate outside of the 28Si deposition spot. It should be noted that this sample was actually deposited later and with a different preparation procedure than the other high deposition temperature samples discussed so far in this section, but the RHEED pattern of this sample is presented here because it is more illustrative of {111} faceting than those of other samples. The presence of microfacets and their 242 orientation on the surfaces of these 28Si films are important pieces of information that provide insight on the growth mechanisms leading to the unexpected surface roughness. 5.6.2.2 STM After deposition, the surface morphology of these samples was investigated by in situ STM and ex situ SEM as a primary means of evaluating the quality of the epitaxial growth as well as confirming and measuring the extent of the surface roughness observed by RHEED. Figure 5.21 shows an example of a filled state STM topography image of a 28Si sample deposited at approximately 708 ◦C at DC–3, which had a RHEED pattern corresponding to a rough surface. The 28Si film was measured by TEM to be 155 nm at the thickest. The roughness indicated in the RHEED pattern of this and similar samples is supported by the presence of large grain-like features visible in this micrograph. These features are approximately 200 nm wide and at least 1 µm long running diagonally from the bottom left to the top right of the image. A measure of the roughness of this surface is given by the total peak-to-valley surface width, i.e. the difference in height values between the highest peak and lowest valley in the STM topography. This surface width in this image is ∆z ≈ 12.8 nm. Other measurements of the value of the surface width for this sample including those from a SEM cross-sectional micrograph show that it is as much as 60 nm. The STM derived value may be smaller because the STM tip is too large to fit in between the valleys seen in the image and thus it 243 Figure 5.21: STM topography filled state image of a 28Si sample deposited at 708 ◦C at DC–3. The image was acquired with a tip bias ≈ -1.8 V and a tunneling current ≈ 100 pA. This film is about 155 nm at its thickest and was deposited on a Si(100) substrate with no ex situ cleaning. Large grain-like features ≈ 200 nm wide are visible on the surface running diagonally from the bottom left to the top right of the image. The surface width determined from the topography is ∆z ≈ 12.8 nm. cannot accurately measure the full range between valley and peak. Alternatively, it may be that the film is thinner in the region scanned by the STM. The difficulty STM has with accurately measuring the topography of very rough surfaces limits its usefulness compared to SEM for evaluating the morphology of these samples. 5.6.2.3 SEM SEM was used ex situ to survey the surface morphology of numerous 28Si samples deposited with a substrate deposition temperature above 600 ◦C at DC–3. SEM images presented in this section were acquired in collaboration with Dr. Joshua 244 Figure 5.22: SEM tilted micrograph of the surface of a 28Si film deposited at 708 ◦C at DC–3. The tilt angle of this micrograph is 60◦. The substrate used for this sample was not cleaned ex situ and was loaded into the chamber with a native oxide. This film is almost 120 nm at its thickest. The surface morphology of the deposited film appears to be extremely rough with mounds or grain-like features that have an average length ≈ 440 nm. The longer side of some of the mounds appear to run left-to-right in the image indicating orientation with a 〈110〉 direction. Schumacher (NIST) and Dr. Vladimir Oleshko (NIST). A tilted SEM micrograph of one of the first 28Si samples deposited at DC–3 is shown in Fig. 5.22. The tilt angle of the image is 60◦. The substrate used for this deposition was boron-doped and was not cleaned ex situ before being loaded into the vacuum chamber with a native oxide and being prepared in situ in the usual manner. This sample was deposited with a substrate temperature of approximately 708 ◦C and is almost 120 nm at its thickest, as measured in cross-section. A SEM cross-sectional micrograph of this sample was previously shown in Fig. 5.14. It is obvious from Fig. 5.22 that the surface morphology of this sample is extremely rough. The surface of the film is 245 covered with large mounds or grain-like features. The rough surface indicated by the RHEED image (Fig. 5.19) is clearly confirmed and understood to be due to these tall mounds visible in the micrograph. The tilt of the image makes it difficult to estimate the height of the mounds and their size in the direction running top- to-bottom in the image, but their estimated average size in the direction running left-to-right is approximately 440 nm. The longer side of some of the mounds appear to directly run left-to-right in the image possibly indicating orientation with a 〈110〉 direction. The crystallographic directions in SEM micrographs presented here are determined from the positioning of the samples in the SEM. This orientation of mounds or grains on the surface of the films is consistent with the indication from the RHEED images (Fig. 5.20) that {111} and {113} microfacets are present on the surface. All of the 28Si samples produced in this initial batch with substrate tem- peratures above 600 ◦C which were inspected with SEM showed similarly rough surfaces with mound formation, although with slightly varying morphologies. SEM micrographs of six of these samples deposited at various temperatures are shown in Fig. 5.23. The deposition temperatures for the samples in panels (a)–(f) were 610 ◦C, 708 ◦C, 804 ◦C, 812 ◦C, 920 ◦C, and 1041 ◦C respectively. All substrates used for these samples were boron-doped except for the sample in panel (c), which was phosphorous-doped. No ex situ cleaning was performed on the substrates used for these samples except for the sample in panel (c), which was etched with HF prior to being loaded into the vacuum chamber. Large mounds are visible on the surface of the 610 ◦C sample in the top-down micrograph in panel (a). The thickness of 246 Figure 5.23: SEM top-down micrographs of the surface morphology of six 28Si films deposited at DC–3 at 610 ◦C, (a), 708 ◦C, (b), 804 ◦C, (c), 812 ◦C, (d), 920 ◦C, (e), and 1041 ◦C, (f). No ex situ cleaning was performed on these substrates except for (c), which was etched with HF. (a) Mounds are apparent on a very rough surface with an average length ≈ 320 nm and faceted sides oriented top-to-bottom or left-to- right in 〈110〉 directions. (b) Irregularly shaped mound features are apparent with an average size ≈ 450 nm. (c) Smaller grains are apparent with an average length ≈ 180 nm. (d) Mounds are visible with an average length ≈ 780 nm and align in rows running left-to-right in a 〈110〉 direction. The larger grains in (d) compared to (c) may be due to the different cleaning procedures, despite similar temperatures. (e) 60◦ tilted image of hut-like mounds with faceted sides and an average length ≈ 1105 nm. (f) Several hut-like mounds are apparent on a smooth surface with sizes between ≈ 200 nm and 500 nm. These mounds have faceted sides oriented in 〈110〉 directions. 247 this sample measured by SIMS depth profiling was determined to be approximately 162 nm. The mounds appear generally elongated and are measured to have an aver- age length ≈ 320 nm and an average width ≈ 175 nm. The length of the mounds is defined as the longer dimension here. The length and width of mounds in the SEM micrographs are determined from software analysis of the images using an autocorre- lation function for finding repeating patterns. The values determined from multiple images were then averaged together to produce the values reported here, which have a 10 % relative uncertainty. Many of the mounds in panel (a) appear to be aligned in similar directions. They also appear to have faceted sides where the edges of the microfacets run top-to-bottom or left-to-right in the image indicating orienta- tion with 〈110〉 directions. This observation of the presence of microfacets confirms the indications from the RHEED images (Fig. 5.20). The observed orientation is consistent with the microfacets being on {111} and {113} planes. Panel (b) in Fig. 5.23 is a top-down micrograph showing the morphology of the surface of the 708 ◦C sample, which consists of irregularly shaped mound features with an average size ≈ 450 nm. The maximum thickness determined from SIMS depth profiles of this sample is approximately 126 nm. Note that the area of panel (b) is roughly four times larger than the area shown in (a). The mounds in panel (b) do not seem to be oriented in any particular direction. The surface of the 804 ◦C sample is shown in the top-down micrograph in panel (c). Smaller grain-like features are seen covering the surface of the film. The scale of panel (c) is the same as that of (a) and the average length of the grains in (c) is ≈ 180 nm. As mentioned above, the substrate used for the 804 ◦C sample in panel (c) was the only one in Fig. 5.23 which 248 was etched with HF. Panel (d) shows a top-down micrograph of a sample deposited at 812 ◦C, which has a maximum thickness determined from a SIMS depth profile to be approximately 158 nm. Despite the similar deposition temperature to that used for the sample in panel (c), the surface morphology of this sample is clearly qualitatively different, possibly due to the different cleaning procedure. Much larger mounds are seen on the surface, the edges of which are not as well defined as the features of the other samples in the other micrographs. The mounds on this sample appear to run together much more giving the surface more of a wavy appearance as opposed to the granular appearance of the other micrographs. This could indicate that the surface of this film is smoother than the others, however, the SIMS depth profile of this sample (Fig. 5.12) showed the characteristic measurement artifact for rough surfaces of a gradual increase in minor isotopes throughout the profile, as determined in a previous section in this chapter. Although the boundaries of the mounds in panel (d) are not as well defined as in the other micrographs in Fig. 5.23, the autocorrelation of this image gives an average width ≈ 250 nm and an average length ≈ 780 nm. These mounds also appear to approximately align in rows running left-to-right in the image, which is consistent with alignment to a 〈110〉 direction. Figure 5.23 (e) is a 60◦ tilted micrograph showing the surface morphology of a sample deposited at 920 ◦C. Many hut-like mounds are visible on the surface with an average width ≈ 560 nm and an average length ≈ 1105 nm. These measurements were corrected for the tilt of the image. The thickness of this film was not directly measured, but the height of some of the mounds is estimated from the image to be roughly 400 nm. Many of these mounds do appear to have faceted sides giving 249 them the hut-like shape, and the long direction of several of them are clearly aligned either top-to-bottom or left-to-right in the image. It can be reasonably presumed that like in several of the other SEM micrographs shown here, the mounds in panel (e) are aligned to a 〈110〉 direction, although this cannot be confirmed because the positioning of the sample inside the SEM is not known. Finally, panel (f) is a top-down micrograph showing the surface morphology of the 1041 ◦C sample. The surface of this sample appears very different from the other samples shown here in that the film does not appear to be continuous. Several hut-like mounds are seen on an otherwise smooth surface. These mounds range in size from approximately 200 nm to 500 nm and were measured to be about 50 nm tall from a SIMS depth profile and TEM cross-sectional imaging. The mounds have clearly faceted sides with edges running from top-to-bottom and left-to-right in the image, indicating orientation with 〈110〉 directions, as expected for {111} and {113} microfacets. 5.6.2.4 Step Pinning Induced Roughness The results presented in Fig. 5.23 are not only unexpected because they show that 28Si deposition at temperatures above 600 ◦C yields films with very rough surfaces, but also because despite depositing over a range of temperatures up to 1041 ◦C, the surface roughness not only persists but appears to increase at higher temperatures. Generally, in epitaxial deposition and growth, including Si MBE and chemical vapor deposition (CVD), higher deposition temperatures result in higher quality epitaxy and smoother films. Si CVD commonly uses deposition temperatures 250 above 600 ◦C or even in excess of 1100 ◦C both to achieve high growth rates from decomposition from silane gas and to facilitate high-quality epitaxial growth of thin films [137]. For Si MBE, above a deposition temperature of approximately 500 ◦C, SPE is expected to dominate the growth [118], and IAD has been shown to produce smooth, unlimited epitaxy at a deposition temperature as low as 390 ◦C [109]. As mentioned previously, the epitaxy phase diagrams for IBE in Fig. 5.18 also predict that higher deposition temperatures should lead to higher quality and unlimited epitaxy, and IBE experiments have demonstrated epitaxial deposition of 28Si films dominated by smooth, layer-by-layer growth [51]. Smooth, epitaxial growth pro- ceeds in these cases because the dominant growth mode at higher temperatures is 2D layer-by-layer growth. The dominant growth for a given deposition flux is con- trolled by the surface diffusivity, which is a thermally activated process. A high surface diffusivity-to-flux ratio ideally leads 2D layer-by-layer or step flow growth producing a smooth surface. An explanation for the rough morphology characterized by large mounds seen on the 28Si samples deposited at high temperature is presented here and is based on the presence of contaminants interfering with smooth deposition. Contaminants such as SiC or SiO2 can act as pinning sites for step movement during deposition in a layer-by-layer growth mode at elevated temperatures. This leads to the formation of pits in the growth surface, which are a manifestation of step pinning in the early stages of thin film deposition. The continued movement of new steps around a pit will lead to local step bunching, and inevitably a high enough step density will form into larger microfacets in the film surface such as {111} and {113} microfacets. The 251 presence of strain in the substrate can also lead to increased roughness and step bunching after thermal processing. Flash annealing strained SOI (sSOI) substrates to between 900 ◦C and 1110 ◦C has been shown to produce a rough crosshatch pattern of step bunching on the surface [138]. For nominally unstrained substrates, strain can arise from the mounting of the chip in the sample holder. Strain could also arise in a deposited 28Si film due to a small lattice constant mismatch between 28Si and a natural abundance Si substrate. The lattice constant of 28Si is larger than that of natural abundance Si by a relative value of roughly 1× 10−6 [29, 139, 140]. The strain due to this difference is quite small compared to the typical strain of a sSOI wafer of approximately 1 %, and so is unlikely to result in a large amount of surface roughening. However, any lattice constant mismatch between a film and substrate will still need to be accounted for by the development of dislocation in the film. As step bunching and microfacets build up on the surface of a film, the roughness increases, and they will come to dominate the growth and morphology of the film. This process leads to the formation of mounds as raised microfacets meet forming larger structures, similar to the faceted mounds seen in the SEM micrographs of 28Si samples deposited above 600 ◦C. Numerous groups have studied the effects of roughness and faceting on the critical thickness, hepi, and epitaxial quality of films produced by Si MBE, and IBE [109, 110, 119, 120, 122]. hepi is found to be smallest on Si(111) surfaces, larger on Si(113) surfaces, and significantly larger on Si(100) surfaces [109]. These studies typically focus on the role of {111} microfacets in defect formation and transition of the growth from epitaxial to amorphous. While they do not directly address 252 the crystalline growth of mounds on a film surface, they do show the critical effect microfacets can have on the evolution of a depositing film and the important link between {111} planes and defects. General roughness on a Si(100) surface is known to develop into {111} microfacets [119], and both {111} and {113} microfacets are commonly observed in low temperature Si epitaxy [122]. {111} planes have been shown to form more easily on smaller terraces on and around islands, and here it may be that the decreased terrace sizes at step bunches have the same effect. Furthermore, incomplete filling of lattice sites by adatoms on step bunches or existing {111} microfacets can lead to the growth of {113} or {115} planes during deposition [110]. Thus, it is no surprise that {111} and {113}microfacets were found on the 28Si samples discussed above. One study showed that C contamination on a Si(100) surface can directly lead to the formation of {113} microfacets after annealing the sample to above 950 ◦C. This study concluded that impurities can effect formation of faceting, and that SiC, for example, can act as a pinning site to step motion [141]. Moreover, once pinning sites are formed from trace amounts of SiC, other C may migrate to the pinning site and form larger, more stable clusters. In general, C and O impurities can cause defects in a film during deposition that lead to pinning sites [122]. Experiments of Si growth confined in bare opening of a patterned SiO2 surface layer have shown that near boundaries, similar to step bunches but in this case the edge of the SiO2 pattern, {113} and sometimes {111} develop from the formation of rebonded double steps (DB) in the Si(100) surface [142,143]. Once {111} and {113} planes are established in the growth surface, they tend to endure and subsequently expand because they 253 are more stable, i.e. energetically favorable, than the Si(100) surface. The Si(111) surface is known to be the most stable with the lowest Si surface energy, and although the Si(113) surface has a higher surface energy than Si(100), it becomes stable when formed by the DB steps [143]. These surfaces are also more stable with the presence of C contamination. In addition to the {111} and {113} microfacets being highly stable, molecular dynamics simulations show that the surface diffusion constant of surfaces such as Si(111) is smaller than for Si(100) [144]. Similarly, total-energy calculations predict that adatoms diffusing on a {113} surface encounter energy barriers at the rebonded double steps which increases the activation energy for adatom diffusion and a lower reactivity at the DB steps [143]. The result of this may be that the growth rate of new layers on the Si(100) surfaces is higher because adatoms on the {111} and {113} surfaces will preferentially hop to a neighboring {100} surface where movement is easier and they are more likely to diffuse away from the facet boundary. The more stable {111} and {113} surfaces may also inhibit new layer formation compared to a {100} surface. A higher growth rate of the {100} surface would actually result in the shrinking of that surface and simultaneous expansion of adjacent microfacet surfaces as each new {100} layer encounters the {111} or {113} edges and adds to them, expanding those surfaces. The {111} or {113} microfacets would grow in time and come to dominate the surface morphology as the {100} surfaces shrink, producing faceted mounds like those observed on the 28Si films deposited at high temperature in this work. A similar explanation was used to describe the preferential growth of {113} microfacets on Si(100) surfaces during Si CVD, where the {113} 254 surface has a slower CVD epitaxial growth rate than the {100} surface [142]. To better understand the mechanisms driving the observed rough morphol- ogy, experiments were carried out to explore the different aspects of the proposed sequence for mound formation. The development of step bunching and roughness around pinning sites on a 28Si sample can be observed by inspecting a film thin enough to exhibit the initial formation of these features in the form of pits on the surface. A STM topography filled state image of a 28Si film deposited with a sub- strate temperature of around 709 ◦C at DC–3 which exhibits pit formation is shown in Fig. 5.24. The substrate used for this sample was prepared ex situ by an HF etch. The film is estimated to be approximately 10 nm thick based on the ion flux at the center of the deposition spot, although there is a large uncertainty on this value. Thickness estimates of particular regions of very thin films in STM images presented here may not be accurate because the ion flux across a deposition spot is not uniform and the position of the STM scan area was not well known relative to the center of the deposition spot. A large area scan of the deposited film is shown in panel (a) of Fig. 5.24 with a large number of dark, round, pit-like features in the surface. Large, flat, single terraces roughly 100 nm wide appear bounded by the pits that act as pinning sites to step movement and create step bunching. The single terraces between pits indicate a predominantly layer-by-layer growth mode. Multiple bright spots are seen within and around the pits and are probably clusters of chemical contaminants such as SiC, SiO, or Si3N4. The average areal pit density of this sample is determined to be 340 µm−2 ± 18 µm−2, and the total height scale of this image is about 2.1 nm. 255 Figure 5.24: STM topography filled state images of a 28Si film deposited at 709 ◦C at DC–3 showing pits. The images were acquired with a tip bias ≈ -1.8 V and a tunneling current ≈ 100 pA. This film was deposited after an ex situ HF etch of the substrate, and it is estimated to be 10 nm thick. (a) Large area scan of the film showing dark pits in the surface. Bounded by the pits and step bunching are large, flat, single terraces about 100 nm wide. Bright spots are seen inside the pits and are probably clusters of chemical contaminants, possibly SiC. The average areal pit density of this sample is ≈ 340 µm−2 ± 18 µm−2. The total height scale of this image is about 2.1 nm. (b) Small area scan of a pit on this sample. Numerous contaminant clusters are seen in and near the pit and step bunching can be seen around the pit, resulting in increased roughness. The height scale is about 1.5 nm. Si (2×1) dimer rows can be seen on several terraces. 256 Panel (b) is a small area scan of a region near a pit from another area of this 28Si sample showing numerous contaminant clusters in and near the pit. Step bunching resulting from the pinning of steps by the clusters can be seen, resulting in increased roughness. The total height scale of panel (b) is about 1.5 nm, which is mostly accounted for by the apparent height of the clusters themselves. Si(100) (2×1) dimer rows can be seen on several terraces surrounding the pit in this image. To determine if the step pinning and pit formation in the initial stages of deposition is related to certain aspects of the ion beam deposition process, a thin natural abundance Si (natSi) film was deposited using the Si electron beam evapo- rator (i.e. the EFM). Aspects of the ion beam excluded from the natSi deposition which may affect the growth include the hyperthermal energy ions, the presence of chemical contaminants in the ion beam that are transported ballistically as ions to the sample, and the presence of SiH4 or other gases which diffuse from the ion beam chamber during deposition. STM topography filled state images of a natSi thin film deposited with a substrate temperature of approximately 713 ◦C at DC–3 are shown in Fig. 5.25. Numerous dark pits are apparent on this sample, similar to the pitting on the 28Si sample in Fig. 5.24. Panel (a) of Fig. 5.25 is a large area scan of the film, which is estimated to be approximately 13 nm thick. Note that the scan area of panel (a) is about four times larger than the displayed area of the 28Si film in Fig. 5.24 (a). The pits in the natSi film are more ordered and square shaped than the pits seen in the 28Si film. Bounded by the pits and step bunching are large, flat, single terraces roughly 200 nm wide. The presence of these single terraces be- tween pits indicates a predominantly layer-by-layer growth mode. Clearly the pits 257 Figure 5.25: STM topography filled state images of a natural abundance Si film deposited at 713 ◦C at DC–3 showing pits. These images were acquired with a tip bias ≈ -2 V and a tunneling current ≈ 150 pA. This film was deposited from the natural abundance Si source after an ex situ HF etch of the substrate, and it is estimated to be 13 nm thick. (a) Large area scan showing dark, square pits in the film bordering large, flat, single terraces about 200 nm wide. Step bunching, seen around the pits, results in increased surface roughness. Contaminant clusters are seen as bright spots in several pits. The average areal pit density of this sample is ≈ 40 µm−2 ± 6 µm−2. The total height scale is about 5.6 nm. (b) Small area scan of a pit in (a). The pit resembles an inverted pyramid with sides aligned with the 〈110〉 directions. A contaminant cluster, possibly SiC, can be seen at the bottom of the pit with a height scale of about 3 nm. Si(100) (2×1) dimer rows can be seen on the terraces. 258 are located at step pinning sites where step bunching during deposition has led to increased roughness. The total height scale of this image is approximately 5.6 nm, which is accounted for mostly by the pits and gives a reference for their depth. Several of the pits in panel (a) contain brighter features at their center which are probably contaminant clusters and possibly SiC. The average areal pit density of this sample is determined to be 40 µm−2 ± 6 µm−2. This pit density is almost nine times lower than that of the 28Si sample in Fig. 5.24 indicating that the roughness observed in thicker films may be partially due to some factor inherent to the ion beam deposition process. Clearly, though, the presence of any pits on the natSi film shows that other factors not specific to the deposition sources also contribute to development of pinning sites and roughness on these films. Panel (b) in Fig. 5.25 is a small area scan of a pit in (a). The pit resembles an inverted pyramid in the surface and is larger than the pit of the 28Si film in Fig. 5.24 (b). Note that the scan area is approximately one quarter of the scan area of image showing the natSi film in Fig. 5.25 (b). The sides of the pit can be seen to align parallel or perpendicular with the Si (2×1) dimer rows that are visible on the terraces surrounding the pit, i.e. the pit sides are aligned with 〈110〉 directions. A contaminant cluster, possibly SiC, can be seen at the bottom of the pit in panel (b) and is likely the source of the step pinning that formed this pit. The total height scale of this image is about 3 nm from the bottom of the pit to the surrounding terraces, making the pit quite shallow compared to its lateral size. The pits in both the 28Si and natSi films as well as in other thin samples have a variety of depths. This indicates that while some pits likely originate at the substrate surface, other 259 pits form during the deposition at different depths in the films. This indicates that the deposition process itself can generate more pits. It seems likely that the step pinning and pit formation observed in both the thin 28Si (Fig. 5.24) and natSi (Fig. 5.25) films do lead to increased surface rough- ness and eventually large scale mound formation for deposition occurring at higher temperatures, i.e. above 600 ◦C. The step bunching and terrace formations that develop around pinning sites and pits qualitatively resemble the shape, distribution, and orientation of the mounds observed in SEM micrographs of the thicker 28Si films (Fig. 5.23). These STM images suggest that mounds may form from the con- tinued growth of the large flat terraces which separate from other nearby terraces by groupings of pits where further growth is inhibited. Groups of pits may then form the valleys between the mounds observed in SEM. It is also clear from the STM image of the natSi film in Fig. 5.24 (a) and, to a lesser extent, the image showing the 28Si film in Fig. 5.25 (a) that the edges of the large terraces are aligned with 〈110〉 directions forming rectangular sections. This is consistent with the observation that many of the large mounds visible in the SEM micrographs of 28Si samples have edges oriented in 〈110〉 directions. The features observed inside pits in STM images of both thin 28Si and natSi films, which are presumed to be clusters of contaminants, likely act as nucleation sites for the formation of the pits by pinning step motion. Several different types of contaminants may play a role in step pinning during deposition, such as SiC. It is important to identify contaminants in this system in order to better understand and ultimately ameliorate the cause of mound formation and roughness on 28Si films. 260 Carbon contamination on substrates can originate from a number of sources including the ambient environment before the substrate is loaded into the chamber, carbon-containing adsorbates produced inside the vacuum chamber, or even ex situ chemical cleaning procedures not targeted to remove organic compounds, such as HF etching. SiC contamination in the form of clusters can be observed on the surface of substrates directly by STM imaging, or it can be detected via the RHEED pattern of substrates during in situ preparation by flash annealing. While the STM can observe individual SiC clusters in very small areas on the substrate, observing the signature of SiC in RHEED generally requires a significant amount of contamination over a large area. SiC is indeed found to be present on substrates used for samples deposited at DC–3 and is most often detected by RHEED during the flashing process. An example of a RHEED image of a flashed Si(100) substrate displaying the signature of SiC contamination is shown in Fig. 5.26. This substrate was prepared ex situ by an HF etch. The RHEED image in panel (a) was acquired with the sample at 600 ◦C and the electron beam in the 〈110〉 direction. A typical Si(100) (2×1) diffraction pattern is seen consisting mainly of the five central streaks as well as bulk Si and reconstruction spots indicated by the two inner arrows. Superimposed on the Si pattern in this image is a SiC pattern. This pattern first appeared after flashing the substrate to approximately 1040 ◦C for 20 s. Additional diffraction spots are also visible just outside the bulk Si streaks, which correspond to SiC as indicated by the two outer arrows. Other spots comprising the SiC pattern are visible on the central rod as well as the far left and right edges of the image. This SiC pattern is a transmission-type pattern, which indicates the SiC exists on the surface in 3D 261 Figure 5.26: RHEED diffraction pattern of a Si(100) substrate after an ex situ HF etch and in situ flash annealing, which shows SiC contamination. (a) The RHEED image was acquired with the sample at 600 ◦C and the electron beam in the 〈110〉 direction. A Si(100) (2×1) diffraction pattern is seen consisting of streaks and bulk Si and reconstruction spots indicated by the two inner arrows. SiC diffraction spots are also visible outside the bulk Si streaks, indicated by the outer arrows, as well as on the central Si streak and the far left and right edges of the image. The transmission-type pattern of the SiC spots indicates 3D clusters. (b) A line profile of the diffraction intensity in arbitrary units is shown corresponding to the dashed line in (a). The reciprocal space mapping of the profile was calibrated from the separation of the bulk Si spots (dotted lines). The separation of the SiC spots (dashed lines) is then calculated to be about 6.5 1/nm, corresponding to the 3C-SiC lattice constant. 262 clusters aligned to the Si(100) surface. Panel (b) of Fig. 5.26 shows a line profile of the RHEED diffraction intensity in arbitrary units corresponding to the horizontal dashed line in (a). This profile is plotted in reciprocal space in order to deduce the lattice constant of the SiC clusters from the spacing of the diffraction spots. The reciprocal space mapping was calibrated using the known separation of the bulk Si spots, indicated by the dotted lines in panel (b). In reciprocal space, these spots should be separated by approximately 5.2 1/nm, which is calculated from the distance between Si lattice planes in the 〈110〉 direction and the Si lattice constant, a0 = 0.543 nm [145]. The separation of the outer SiC spots, indicated by the dashed lines is then calculated to be approximately 6.5 1/nm. Taking half of this value gives a distance of 3.25 1/nm between one SiC spot and the central streak, which is difficult to measure directly in this image. Then, inverting this value gives a real space distance between lattice planes of the SiC cluster in the 〈110〉 direction of 0.31 nm. Multiplying this spacing by 2/ √ 2 gives the 〈100〉 lattice constant, 0.44 nm. This value agrees quite well with the lattice constant of the 3C-SiC polytype, a0 = 0.436 nm [146]. 3C-SiC has a zincblende crystal structure making it able to align to the Si diamond cubic lattice, and this alignment is clear in the RHEED pattern of Fig. 5.26. SiC diffraction spots in the RHEED patterns of substrates were commonly seen during the flash annealing process of many samples in this work. This preva- lence highlights the need for ex situ cleaning procedures to mitigate C and other contaminants. Typically, indications of SiC appear after exceeding a temperature of about 1000 ◦C, which is usually the minimum temperature of the initial flash in 263 the flashing sequence after keeping the substrate at about 600 ◦C or below (see the in situ substrate preparation discussion in an previous section of this chapter). This observation matches well with the known temperatures associated with SiC cluster formation on Si surfaces [119]. SiC CVD experiments have shown that stoichiomet- ric SiC begins to form on Si(100) at substrate temperatures above 700 ◦C [147], and annealing C60 films on Si(100) substrates produces crystalline SiC between 800 ◦C and 900 ◦C [148]. Other experiments studying SiC contamination during vacuum preparation of Si surfaces have seen similar results to those observed in this work. Becker et al. used RHEED to demonstrate that SiC typically takes the form 3C-SiC when clusters form on Si(100) and Si(111) surfaces [149]. Samples were cleaned ex situ using HF and flashed in situ to between 800 ◦C and 1000 ◦C with background pressures as high as 1.3× 10−5 Pa (1.0× 10−7 Torr) resulting in SiC contamination. This study concluded that the contaminants originated from carbon-containing ad- sorbates from the vacuum chamber which dissociate to release C during flashing, including at lower pressures. Similarly, Henderson et al. showed that 3C-SiC forms on Si(111) surfaces above temperatures of 800 ◦C and likely originates from both carbon-containing adsorbates and the ex situ chemical cleaning procedure, which included an HF etch [150]. These results indicate that SiC formation is not only likely due to the flashing process, but that it possibly forms throughout a deposition occurring at higher temperatures such as the ones described here above 600 ◦C. Further, experiments by Mol et al. showed that pre-treating Si(100) substrates with HF can directly result in SiC contamination [112]. Figure 5.27 shows a STM topography filled state image from that work of a Si(100) substrate prepared ex 264 Figure 5.27: STM topography filled state image of a Si(100) substrate which was prepared ex situ by an HF etch and in situ by a 900 ◦C anneal. The image was acquired with a tip bias of -1.5 V and a tunneling current of 100 pA. Bright SiC clusters about 15 nm across are seen covering the (2×1) reconstructed surface. Step bunching is apparent in the upper left and lower right of the image. (from Ref. [112]) situ by an HF etch. The sample was loaded into the vacuum system with just the H passivation layer from the etch before being annealed at approximately 900 ◦C. Several bright SiC clusters approximately 15 nm wide can be seen on the annealed surface. Below the clusters can be seen (2×1) dimer rows on short terraces with steps appearing highly bunched in the upper right and lower left of the image, likely due to the presence of the SiC. Another sample in this study was prepared with a protective oxide layer resulting from the oxidation during ex situ cleaning with HNO3. After the same anneal to 900 ◦C to remove the oxide, the surface was found to be flat and free of contaminants. Another group has shown that IBE deposition of 28Si onto a substrate treated with HF produced a defective film with stacking 265 faults present [42]. Some of these previous experimental results regarding SiC discussed above were reproduced in this current work. Approximately 90 % of substrates prepared ex situ with an HF etch in this work showed signs of SiC contamination in the RHEED pattern after initial flashes. Most of these patterns were completely dominated by the SiC diffraction spots and showed only weak bulk Si spots or streaks, unlike the more balanced combination of patterns observed in Fig. 5.26. This suggests that the contamination was such that the surfaces of these substrates were mostly if not completely covered in SiC clusters. By contrast, approximately 32 % of the substrates which had no ex situ cleaning and were loaded into the chamber with a native oxide showed signs of SiC in the RHEED pattern. Furthermore, the RHEED patterns of these substrates typically only showed weak SiC diffraction spots with much stronger Si (2×1) spots, likely indicating only trace amounts of SiC on the surface. Another potential issue with cleaning substrates using only HF is that F atoms in the solution or left behind as residue can etch the Si surface, leading to roughening and potential pinning sites in film growth [150,151]. These observations of the prevalence of SiC on substrates as well as the results of the experiment shown in Fig. 5.27 highlight both the potential benefit of alternate chemical cleaning combined with a protective oxide layer on substrates before in situ preparation as well as the ineffectiveness of HF etches alone to mitigate contaminants. After the appearance of SiC on the surface of prepared substrates, further heat treatments in the form of higher temperature flash annealing typically results in the disappearance of the SiC diffraction pattern. For almost all of the samples where a 266 RHEED diffraction pattern corresponding to SiC appears after initial flashing to at least 1000 ◦C, flashing to between approximately 1150 ◦C and 1200 ◦C was required to remove the SiC spots and recover a nominal Si (2×1) pattern. Experimental results from other groups support this observation showing that SiC can be re- moved from Si(100) and Si(111) surfaces only by heating it to between 1100 ◦C and 1200 ◦C [150,152]. It is also suggested that upon heating to these temperatures, the C from the clusters goes into solution in the Si substrate, although it is not known how this C may affect the growth of a film. After RHEED is used to identify substrates with SiC contamination, STM imaging is used to directly observe and inspect the SiC clusters to confirm their presence and view their effect on the substrate. Additionally, STM inspection of nominally clean substrates can reveal small contaminant clusters or other partic- ulates in trace amounts below the detection capability of RHEED. These trace contaminants may still result in pinning sites during film growth. Figure 5.28 shows STM topography filled state images of two Si(100) substrates prepared in situ by flash annealing that show contaminant clusters. Both these substrates were prepared ex situ by an HF etch. These images were acquired with a tip bias ≈ -2 V and a tun- neling current ≈ 150 pA. Panel (a) shows a phosphorous-doped Si substrate flashed to approximately 1150 ◦C. This substrate was nominally clean after flashing and although the RHEED pattern showed indication of SiC initially, no SiC signal was present after the final, higher temperature flashes. Unknown contaminants appear as bright spots in a cluster approximately 30 nm across on otherwise normal (2×1) reconstructed terraces. SiC, SiO2, or other particulates are possible explanations 267 Figure 5.28: STM topography filled state images of Si(100) substrates prepared ex situ by an HF etch and in situ by flash annealing. These images were acquired with a tip bias ≈ -2 V and a tunneling current ≈ 150 pA. (a) Phosphorous-doped Si(100) substrate flashed to around 1150 ◦C with contaminants appearing as a cluster of bright spots on otherwise normal terraces. This substrate was nominally clean and free of SiC as determined by RHEED. The height scale of this image is about 1.3 nm. (b) Boron-doped Si(100) substrate flashed to around 1130 ◦C with two clusters of contaminants appearing as bright areas. The RHEED pattern of this substrate showed significant SiC contamination. The clusters act as step pinning sites causing step bunching around them, which forms small mounds. The height scale of this image is about 15 nm. 268 for clusters commonly seen during STM inspection on nominally clean substrates such as that in panel (a). These features may also be due to Si atoms stuck on the tip. During deposition of a thin film on such a surface, steps may become pinned at this cluster, causing a buildup of roughness nearby. The total height scale of this image is approximately 1.3 nm, which is mostly accounted for by the height of the cluster. Some dark dimer row defects are also seen on this surface. Panel (b) shows a boron-doped Si substrate flashed to around 1130 ◦C. The RHEED pattern of this substrate showed significant SiC contamination. Two clusters of SiC appear as bright areas approximately 25 nm across. The total height scale of this image is approximately 15 nm. These clusters clearly act as step pinning sites causing significant step bunching around them, which has resulted in the formation of small mounds that the SiC clusters sit atop. The in situ prepared Si(100) substrates used for the initial set of 28Si films deposited above 600 ◦C can not only have SiC and other clusters of contaminants and particulates on their surface, they can additionally exhibit signs of metal con- tamination. Metal atoms on or just below Si(100) surfaces produce patterns of dimer row defects observable in the STM, which are most often attributed to Ni contamination [116,153–157]. The presence of Ni or possibly other metal impurities including In, Ga, and Al on the Si(100) surface produces long chains or lines of ordered dimer defects after high temperature heat treatments such as the typical flash annealing process. Ni atoms are thought to reside just below the surface and disrupt local bonds generating surface defects. These dimer vacancy lines (DVLs) run perpendicular to the Si dimer rows on each terrace, forming (2×n) patterns 269 where typically n ≈ 8 and represents the number of normal Si dimers between the DVLs [153,154,157]. Si surfaces with DVLs present appear striped in STM imaging. The ordering of vacancies in DVLs is due to repulsive interactions between vacancies along a dimer row and attractive interactions between vacancies in adjacent dimer rows, which may be related to strain relaxation [153,158]. Figure 5.29 (a) shows an example of a STM topography filled state image of a Si(100) surface with Ni contamination from Ref. [153]. This image was acquired with a tip bias ≈ -2 V and a tunneling current ≈ 30 pA. This sample was deliberately contaminated with Ni by contacting it with stainless steel tweezers ex situ before it was loaded into the STM. That simple handling of the sample was enough to contaminate the surface with Ni. After it was flash annealed to 1150 ◦C, the sample then exhibited a highly ordered (2×n) reconstructed surface. Dark DVLs due to the Ni contamination are ubiquitous on this surface and are seen running perpendicular to the dimer row directions on each terrace. Typically, a surface coverage of < 1 % Ni can produce the (2×n) surface seen in panel (a). A similar example of metal contamination on a prepared substrate from this work is shown in Fig. 5.29 (b), which is a STM topography filled state image of a Si(100) surface prepared by flash annealing to approximately 1130 ◦C. This image was acquired with a tip bias ≈ -2 V and a tunneling current ≈ 150 pA. Although this boron-doped substrate was not intentionally contaminated with Ni or other metals, its surface appears very similar to the Si surface shown in panel (a). The surface has a (2×n) reconstruction with dark DVLs running perpendicular to the dimer row directions, although the concentration and length of DVLs is less than in panel (a). 270 Figure 5.29: STM topography filled state images of Si(100) substrates prepared in situ by flash annealing. (a) A Si(100) substrate flash annealed to 1150 ◦C has a (2×n) reconstructed surface with dark dimer vacancy lines due to Ni contamina- tion. Vacancy lines run perpendicular to the dimer row direction on each terrace. The substrate was intentionally contaminated by contacting it with stainless steel tweezers ex situ. This image was acquired with a tip bias ≈ -2 V and a tunnel- ing current ≈ 30 pA (from Ref. [153]). (b) A boron-doped Si(100) substrate flash annealed to around 1130 ◦C has a (2×n) reconstructed surface with dark dimer va- cancy lines due to metal although the substrate was not intentionally contaminated. This image was acquired with a tip bias ≈ -2 V and a tunneling current ≈ 150 pA. 271 The concentration of DVLs seen in Fig. 5.29 (b) is around the highest ob- served for substrates discussed to this point in this section, which were not chemi- cally cleaned ex situ. Much more frequently, flashed substrates will exhibit a more moderate concentration of DVLs with shorter lengths. Some substrates have had a large number of individual or small clusters of dimer vacancy defects present on their surface, although not aligned in DVLs. It is not obvious if all dimer vacancies are related to metal contamination or what causes the dimer vacancies to form lines on some substrates while remaining disordered on others. The formation of DVLs is probably related to the level of contamination as well as differences in the specific heating times and temperatures that each substrate experiences. It has been found in this work that a minimal number of flashes to around 1200 ◦C followed by a quick cool down to below 800 ◦C can eliminate DVLs from a surface. Other groups have used similar procedures to reduce or eliminate DVLs [114, 154]. It is also observed that prolonged annealing to temperatures between 600 ◦C and 800 ◦C can result in the reappearance of DVLs even after their elimination by higher temperature flashing. The solid solubility limit of Ni in Si at a temperature of around 1200 ◦C is 5.8× 1017 cm−3 [159]. So, at concentrations below this, Ni likely diffuses into the Si bulk at higher temperatures and can return to the surface during lower temperature annealing. While it is not ultimately desirable from a device point of view for 28Si films to contain Ni or other metal impurities, it is not known exactly how this type of contamination may effect the growth morphology or epitaxy of deposited 28Si films. Clearly, Ni atoms disrupt the bonding structure of the surface potentially creating strain and defects which may result in the formation of pinning sites or defects in 272 the epitaxy. In addition to SiC or other particulates and metal contamination, there are other candidates for contaminants on Si(100) substrates prepared by in situ flash annealing which may lead to step pinning sites during deposition and 28Si film growth. For the substrates which are not cleaned before they enter the vacuum chamber with a native oxide, removal of the oxide may be incomplete during flash annealing. The remaining, possibly non-stoichiometric, silicon oxide (SiOx) may form clusters similar to the observed SiC, or molecules may remain isolated being very difficult to detect using the STM. Even a small amount of SiOx on the surface of a substrate may act as a pinning site. Within the deposition chamber, N2 is one of the major residual components of the vacuum after H2 and CO2, as seen in the RGA mass spectrum in Fig. 2.22 in Chapter 2. N-containing adsorbates on a Si substrate may dissociate during thermal processes potentially leading to the formation of silicon nitride clusters (Si3N4). As will be discussed in a following section, chemical analysis shows that 28Si samples deposited after the ones discussed in this section contain a relatively high concentration of N, originating from the vacuum or the ion beam itself. Si3N4 clusters are thus a candidate for the nucleation of pinning sites during deposition. The solid solubility limits of N, C, and O in Si are shown in Fig. 5.42 and discussed in a later section. These solid solubility limits show, however, that the measured concentrations of N, C, and O in the films can lead to contaminant clusters that cause step pinning. The presence of H2 on Si surfaces is believed to lead to increased roughness during epitaxial deposition [119]. Surface interaction with H2 is nearly unavoidable in a vacuum system especially when increased amounts of H2 273 and SiH4 are introduced into the deposition chamber from the ion beam chamber. Finally, another potential contaminant coming from the vacuum during deposition is F, as seen in the deposition chamber RGA mass spectrum in Fig. 5.3. As mentioned previously, F atoms can also contaminate a substrate which was treated with HF and etch the surface when they desorb with thermal processes in the vacuum [151]. Small areas etched by F on a Si substrate and the resulting defects may lead directly to step pinning and increased roughness. Another aspect to the formation of faceted mounds is the deposition tem- perature. To restate from a previous discussion, the rough 28Si samples discussed thus far in this section were deposited with substrate temperatures between 610 ◦C and 1041 ◦C. At these temperatures, one would expect the deposition to occur in a predominantly layer-by-layer 2D growth mode. Despite the extreme rough- ness of these samples and morphology dominated by facet formation, the primary mechanism of mass transport and step motion is likely still akin to layer-by-layer growth, especially in the early stages of deposition when the surface is still fairly smooth. The high temperature layer-by-layer growth likely enhances some aspects of the mound formation. For example, the defects that form on {111} planes dur- ing Si epitaxy are found to increase in quantity as the deposition temperature is increased [120]. In general, layer-by-layer growth is much more likely to lead to the formation of microfacets, especially the rebonded DB steps required for more stable {113} planes [143]. It also leads to faster growth of {111} and {113} planes while the {100} planes diminish, as mentioned previously. Probably the most important role of predominantly layer-by-layer growth is that it allows for the flow of steps around 274 pinning sites which leads to the step bunching required for microfacet formation. If layer-by-layer growth and Si surface diffusion are in fact important driving aspects to the formation of mounds on the surface of 28Si samples, then the char- acteristics of the mounds should reflect that. Another characteristic of the mounds besides faceting which can inform on their origin is their size. Using the analysis method described previously, the length and width of mounds on several samples was analyzed from the SEM micrographs of the surface. Here, the length is defined as the longer dimension of mounds (if applicable), while the width is the shorter dimension. An autocorrelation function was used to determine these values for each sample analyzed including several of the sample micrographs shown in Fig. 5.23 as well as others. The results of this analysis are shown in Fig. 5.30. The average length (trian- gles) and width (squares) of mounds is plotted vs. the deposition temperature. Both the length and width of mounds increase with increasing temperature. The widths vary from approximately 175 nm for the 610 ◦C sample to approximately 560 nm for the 920 ◦C sample. The mound lengths vary from approximately 320 nm for the 610 ◦C sample to approximately 1105 nm for the 920 ◦C sample. The relative un- certainty in the lengths and widths is determined to be about 10 % from comparing measurements of multiple SEM images from a single sample. The uncertainty in the deposition temperature is due to uncertainty in the pyrometer readings and its calibration, as discussed in Chapter 2. The length-to-width aspect ratio of mounds formed on different samples is shown in the inset in Fig. 5.30 (open circles) to be between 1.5:1 and 3.5:1. The average aspect ratio of the mounds from all samples 275 Figure 5.30: Average mound size vs. deposition temperature for rough 28Si films deposited above 600 ◦C. The length (triangles) and width (squares) of large mounds formed on the films increase with increasing temperature as calculated from SEM micrographs of the surface morphology. The inset shows the ratios of the lengths to the widths (open circles) of the mounds for the different samples, most of which are about 2:1. analyzed in this section is found to be ≈ 2:1. This value is similar to the aspect ratio reported by Mo et al. of 2D Si islands deposited on Si(100) [160]. That study used STM imaging to determine that the aspect ratio of the islands close to their equilibrium shape was between 2:1 and 3:1, and attribute it to differences in adatom incorporation and diffusion along the island edges. The increase in mound size with increasing temperature suggests that the mounds are formed from a process or mechanism that is thermally activated, such as surface diffusion. Surface diffusion constants have the standard exponential form 276 of a thermally activated process: D = D0 exp ( − Ea kBT ) , (5.3) where D0 is the exponential prefactor, kB is the Boltzmann constant, Ea is the activation energy, and T is the substrate temperature. As the Si surface diffu- sion constant increases exponentially with increasing temperature, the mechanisms involving both the flow of steps leading to faceting and competing diffusion on dif- ferent microfacets may be enhanced causing an increase in the characteristic size of the mounds at a given temperature. Increased temperature also leads to in- creased diffusion of adatoms over steps. This may lead to merging of different facets around pinning sites at a higher rate and the formation of larger mounds at higher temperatures. It is important to note that other factors besides temperature may affect the mound size. The deposition time, deposition rate, or more likely the final film thickness may lead to different size mounds for a given temperature. However, the mound sizes do not seem to vary as a function of film thickness and most of the samples analyzed in Fig. 5.30 have similar thicknesses between approximately 110 nm and 160 nm. The thickness of the 920 ◦C sample was not measured, but it is roughly estimated to be larger than the other samples at around 400 nm, and this sample also exhibits the largest mounds. Another factor that may affect the mound size independent of the temperature is the amount of contaminants present on the substrate that would lead to pinning sites. A sample with a higher density of contaminants and pinning sites may result in smaller mounds because the area 277 between pinning sites where mounds may form would be smaller. However, most of the samples involved in this analysis were not cleaned ex situ, which should lead to similar amounts of contaminants. So, contaminants and thus the size of the mounds would not be expected to vary significantly across nominally similar samples. One sample deposited at 804 ◦C and shown in Fig. 5.23 (c) was not included in the mound size analysis because it was etched with HF, which is known to produce more SiC contamination, and it exhibited smaller grains than the other samples. This suggests that the amount of contaminants or other factors that lead to pinning do affect the size of mounds or grains that form on the surface. The link between mound size and surface diffusion is explored further by plot- ting dependance of the mound size on deposition temperature in an Arrhenius form. An effective activation energy can then be extracted that results from the activation energy of the diffusion constant. This assumes that the change in the mound size is proportional to the change in the diffusion constant over the temperature range of the data. This proportionality makes sense when considering that an increased surface diffusion would lead to an increase in adatom flux to the growing faceted sides of the mounds and thus an increased area. In order to simplify this deter- mination and not favor one particular dimension of the mounds, the mounds are treated as rectangular and the lengths and widths are combined to yield the average mound area, A. Then, taking A ∝ D and substituting A for D and a new expo- nential prefactor for the mound area, A0, for D0 in Eq. (5.3), the area data can be plotted to extract Ea. Taking the natural log of Eq. (5.3) with the aforementioned 278 Figure 5.31: Arrhenius plot of ln(A), the natural log of the average mound area (circles), vs. inverse temperature in energy units for rough 28Si samples deposited above 600 ◦C. The top axis shows the equivalent deposition temperature in ◦C. The data are fit to a line (Eq. (5.4)) whose slope gives an activation energy Ea = 0.7(3) eV. substitutions gives a linear equation, ln(A) = ln(A0)− Ea ( 1 kBT ) , (5.4) where the slope of ln(A) as function of 1 kBT is equal to Ea. In Fig. 5.31, ln(A) is plotted vs. the inverse deposition temperature, which is modified by the Boltzmann constant to have units of eV−1. The uncertainty in ln(A) is from the combined uncertainties of the lengths and widths. The data is fit to a line using Eq. (5.4), the slope of which gives a value for the activation energy for Si surface diffusion of Ea = 0.7(3) eV. The uncertainty in this value is the standard error from the fit. Again, 279 this assumes that the appropriate activation energy for diffusion would emerge from this analysis of A, the mound areas, because A changes with T in proportion to D. First-principles and experimental investigations of Si diffusion on Si(100) surfaces find activation energies of 0.6 eV and 0.67 eV respectively for diffusion along dimer rows, similar to the value reported here [161,162]. Those calculations also predict a value of Ea = 1.0 eV for diffusion perpendicular to dimer rows. However, another experiment found an average value of Ea ≈ 1.55 eV for diffusion across steps of the Si(100) surface [163], and molecular dynamics simulations of Si diffusion on Si(100) along dimer rows give a value of Ea = 0.2 eV [144]. While there is a large range of values reported in the literature, it seems reasonable that the activation energy determined here from Fig. 5.31 can be attributed to Si surface diffusion, an important component in faceted mound formation. 5.6.3 Elimination Strategies for Step Pinning Sites Several strategies were used to try to reduce the density of pinning sites on substrates, which manifest during deposition as pits and step bunching. New clean- ing protocols were established to reduce chemical contaminants, such as SiC, on or near the substrates. An ex situ CMOS cleaning procedure was implemented for preparing Si(100) substrates. This is the sample preparation cleaning procedure described previously in this chapter in Section 5.3. This CMOS clean consists of etching the substrate first with piranha solution, then etching the oxide that forms with HF, and finally using SC–2 to clean and cap the substrate with a protective oxide. This clean is designed to remove both organic impurities containing C as well 280 as metals including Ni from the substrate. An HF etch alone, while a standard Si substrate preparation method for vacuum surface science experiments, was not cho- sen because of the ample evidence presented previously that it results in a substrate surface with far more SiC contamination than one that was protected with a native oxide before in situ heat treatments. Clean Si substrates were only ever handled with teflon tweezers that were also cleaned of metals using hydrochloric acid (HCl). In addition to cleaning metal contamination off the substrates, the sample manipulator in the deposition chamber as well as the in vacuum sample holders were also cleaned. This was done to reduce cross-contamination onto the substrates of Ni and other metals, which can migrate from these parts during sample heating. While the sample holders are comprised of only Mo, several stainless steel components, which can spread Ni, were removed from the Mo section of the manipulator that gets hot during sample heating. Also, a chromel-alumel thermocouple, which is mostly Ni, was removed from the manipulator. The Mo parts were cleaned first in “base piranha”, which is a 1:1:3 mixture of ammonium hydroxide (NH4OH), hydrogen peroxide (H2O2), and water. This solution gently etches Mo surfaces removing contaminants, and it etches Group V elements. Then the Mo parts were treated with HCl to remove metals. Mo tools used for manipulating the sample holders were also cleaned in this manner. These vacuum component cleaning procedures were adopted from Richardson [164]. Despite these efforts to eliminate contaminants that cause step pinning, indi- cations of the presence of contaminants on substrates persists. The CMOS clean- ing procedure is either not sufficient at removing the offending contaminants from 281 the substrate surface, or the substrates pick up additional contaminants after the cleaning procedure. After preparing newly cleaned substrates in situ through flash annealing, inspection by RHEED shows that SiC was present more often than be- fore CMOS cleaning, although the level of contamination was still much less than that of the substrates etched with HF alone. The frequency with which SiC was ob- served depended on the source of the wafers used for substrates. Wafers that were re-claimed wafers obtained from University Wafer and prepared with the CMOS cleaning procedure exhibited signs of SiC in the RHEED pattern approximately 66 % of the time. Virginia Semiconductor is generally seen as a source of higher quality wafers (e.g. lower impurities, smoother), and approximately 35 % of those wafers exhibited signs of SiC in the RHEED pattern. That frequency of SiC for- mation is similar to that of the older substrates which were not cleaned at all ex situ. It is not clear why the samples from University Wafer exhibit more contami- nation, but it is likely a combination of surface contamination from poor handling by the company and some contaminants in solution from previous processing of the re-claimed wafers. Several samples were made to investigate any changes in contaminants, step pinning, or morphology of 28Si samples deposited above 600 ◦C after implementing the above described cleaning procedures. To view the formation of pits due to step pinning sites, thinner films were deposited from both the ion beam and the natural abundance Si EFM source on Si(100) substrates prepared by the CMOS cleaning procedure. Figure 5.32 shows STM topography filled state images of thin films of 28Si, (a) and (b), and natSi, (c) and (d), which show the formation of pits in the 282 growth surface. Both of these films were deposited with substrate temperatures of approximately 712 ◦C at DC–3, and both are estimated to be approximately 10 nm thick. Panel (a) is a large area scan of the 28Si film. Numerous square shaped pits can be seen in the growth surface surrounded by step bunching that bounds elevated areas with large, flat, single terraces roughly 50 nm wide. Single terraces between pits indicates a predominantly layer-by-layer growth mode. The pits are qualitatively different from the pits previously seen in the 28Si film in Fig. 5.24 and are more similar to the square pits seen in the natSi film in Fig. 5.25. The cause of the qualitative difference in the pits seen in the two 28Si samples is not known, although it is perhaps related to the new cleaning procedure for substrates. Also, the native step density due to the wafer miscut of the substrate used for the second 28Si sample was roughly five times higher than that of the first 28Si sample. The average areal pit density of this 28Si sample in Fig. 5.32 is determined to be 580 µm−2 ± 24 µm−2, which is significantly more than the areal density of pits in the previous 28Si sample (340 µm−2) shown in Fig. 5.24, despite the more rigorous substrate cleaning procedure used for this latter sample. Panel (b) of Fig. 5.32 shows a small area scan of a pit in panel (a). Si (2×1) dimer rows are visible on terraces around the pit, which appears as an inverted pyramid in the surface with sides aligned with the 〈110〉 directions. The total height scale in panel (b) is approximately 1.4 nm, giving an indication of the pit depth. Panel (c) in Fig. 5.32 is a large area scan of the natSi film showing six pits in the growth surface. Note that the scan area in panel (c) is roughly 25 times larger than the area shown in panel (a). Surrounding the pits are triangular shaped 283 Figure 5.32: STM topography filled state images of 28Si, (a) and (b), and natural abundance Si, (c) and (d), films deposited at 712 ◦C at DC–3. Si(100) substrates were prepared ex situ by CMOS cleaning. Both films are estimated to be about 10 nm thick. (a) Large area scan of the 28Si film showing many dark, square pits and step bunching in the surface bounding large, single terraces about 50 nm wide. The average areal pit density of this sample is ≈ 580 µm−2 ± 24 µm−2. (b) Small area scan of a pit in (a). Dimer rows are visible and the pit appears as an inverted pyramid with sides aligned with the 〈110〉 directions. Images (a) and (b) were acquired with a tip bias ≈ -1.8 V and a tunneling current ≈ 100 pA. (c) Large area scan of the natSi film showing six dark pits in the surface. Many steps < 50 nm wide are seen due to the wafer miscut, indicating step flow growth. Steps appear to flow around the pinning sites. The average areal pit density of this sample is ≈ 1.5 µm−2. (d) Small area scan of a square pit in (c). Dimer rows are visible around the pit and a bright contaminant cluster, possibly SiC, appears at the center. Images (c) and (d) were acquired with a tip bias ≈ -2 V and a tunneling current ≈ 150 pA. Si (2×1) dimer rows are visible in (b) and (d), and their total height scales are about 1.4 nm and 1.6 nm respectively. 284 regions of step bunching where the step motion was pinned by and flowed around the pits during deposition. A large number of steps with widths < 50 nm are seen on the surface with step edges running top-to-bottom along the hypotenuse of the pits in the image. These are a consequence of a large miscut away from {100} in the plane of the wafer surface. The presence of steps matched to the underlying substrate indicates that deposition on this sample proceeded in a step flow growth mode. The average areal pit density of this sample is determined to be 1.5 µm−2 ± 1.2 µm−2, which is less than the pit density of the previous natSi film (40 µm−2) shown in Fig. 5.25. For the approximately ten natSi thin films deposited from the EFM evaporation source that exhibited pit formation and were cleaned with the CMOS procedure, the average areal density of pits varied between approximately 1 µm−2 and 30 µm−2. Panel (d) on Fig. 5.32 is a small area scan of one of the pits in panel (a). Si (2×1) dimer rows are visible on terraces around the pit and a bright contaminant cluster approximately 10 nm across, possibly SiC, appears at the center. The flow of steps was right-to-left around the pit in this image with a high degree of step bunching seen to the right of the pit. The total height scale in panel (d) is approximately 1.6 nm, which is mostly accounted for by the height of the pit. The natSi films described above that exhibit step pinning and pits seem to show that the overall areal pit density and thus the amount of contaminants present on the substrates did decrease, if somewhat inconsistently, when implementing the CMOS and other cleaning procedures. However, the same is not true of the 28Si film, which appears to have a higher amount of contaminants. Considering these results, 285 the one to two orders of magnitude larger areal pit density of the 28Si films compared to the natSi films suggests more strongly than before that many contaminants and pinning sites are due in part or exacerbated by something intrinsic to the 28Si ion beam deposition process. Two possibilities are contaminants in the ion beam and contaminants in the gases that diffuse from the ion beam chamber to the sample. To observe surface contaminants before thin film deposition, other Si sub- strates were prepared ex situ by the CMOS cleaning procedure, flash annealed in situ, and inspected with the STM. The normal flashing procedure of heating the substrate up to a temperature of around 1200 ◦C was not used, however. Instead, substrates were flashed to a lower temperature that should form SiC but not elim- inate it from the surface, similar to the STM study of Si surface contaminants represented in Fig. 5.27 showing SiC clusters after a 900 ◦C anneal. Figure 5.33 shows STM topography filled state images of two Si(100) substrates prepared in this manner that show contaminant clusters. Panel (a) shows a phosphorous-doped Si substrate flashed to a temperature of approximately 1060 ◦C two times. The RHEED pattern of this sample did show faint signs of SiC on the surface after flash- ing. Multiple contaminant clusters, probably SiC, are seen on the surface, appearing as bright spots. Step movement during flashing caused the steps to recede around the clusters, which act as pinning sites causing step bunching nearby. The total height scale of this image is approximately 7.0 nm, which is mostly accounted for by the height of the clusters. Panel (b) shows a boron-doped Si substrate flashed to a temperature of approximately 1050 ◦C two times. It was unclear from the RHEED pattern of this sample if SiC was present or not. A bright contaminant cluster, pos- 286 Figure 5.33: STM topography filled state images of Si(100) substrates prepared ex situ by CMOS cleaning and in situ by flash annealing. These images were acquired with a tip bias ≈ -2 V and a tunneling current ≈ 150 pA. (a) Phosphorous-doped substrate flashed to around 1060 ◦C with contaminant clusters, probably SiC, ap- pearing as bright spots. Step flow due to flashing appears pinned at the clusters causing step bunching and increased roughness. The total height scale of this image is about 7.0 nm. (b) Smaller area scan of a boron-doped substrate flashed to around 1050 ◦C with a contaminant cluster appearing in the upper right corner. The height scale of this image is about 2.8 nm. 287 sibly SiC or SiOx from incomplete desorption due to the lower flash temperature, is apparent in the upper right of the image. This cluster is approximately 20 nm wide and approximately 60 nm long. The total height scale of this image is approx- imately 2.8 nm, which is accounted for mostly by the height of the cluster. Short terraces are also seen with step bunching occurring near the cluster and Si (2×1) dimer rows visible on some terraces. Finally, a thick 28Si film was deposited onto a substrate prepared with the CMOS cleaning procedure to determine any changes in roughness or surface mor- phology compared to the previous 28Si films shown in Fig. 5.23 that were not cleaned. Figure 5.34 is a top-down SEM micrograph of a thick 28Si film deposited on a clean, boron-doped Si(100) substrate with a substrate temperature of approximately 705 ◦C at DC–3. The maximum thickness of this film found by SIMS depth profiling is approximately 144 nm. The surface morphology of the deposited film appears to be very rough, similar to SEM micrographs of the previous 28Si films. Mounds cover the surface with an average width ≈ 230 nm and an average length ≈ 495 nm. The mounds have sharp, well defined, faceted edges predominantly running diagonally in the micrograph, indicating that they are oriented with the 〈110〉 directions as determined from the sample positioning in the SEM. The mounds on this sample appear qualitatively different from those of the previous 28Si sample deposited with a similar substrate temperature of 708 ◦C, which had more rounded and randomly shaped mounds as opposed to the faceted, zig-zag pattern created by the mounds in the 705 ◦C sample. It is not known if the cause of this qualitatively different morphologies is due to the updated cleaning procedures. 288 Figure 5.34: SEM top-down micrograph of the surface of a 28Si film deposited at 705 ◦C at DC–3. The substrate used for this sample was prepared ex situ by CMOS cleaning. The thickness of this film determined from SIMS depth profiling is about 144 nm. The surface morphology of the deposited film appears very rough with mounds that have an average length ≈ 495 nm. The mounds have sharp, well de- fined, faceted edges predominantly running diagonally in the micrograph, indicating they are oriented with the 〈110〉 directions. The above results show that the CMOS and other cleaning procedures im- plemented to reduce substrate contamination and thus step pinning that leads to mound formation and roughness were mostly unsuccessful. One notable aspect of the as-prepared quality of substrates that seemed improved by the new cleaning pro- cedures was the presence of Si (2×1) dimer row defects. Although DVLs similar to those in Fig. 5.29 still appeared on some 28Si samples after deposition, very few were seen on substrates prepared for deposition after flash annealing. Further, the surface density of isolated dimer row defects was reduced on prepared substrates after the new cleaning procedures. For substrates that were not cleaned ex situ, dimer row de- 289 fects were observed using STM to account for a wide range of areal densities between approximately 1 % and 34 % of the surface, with an average value of approximately 9 % being typical. These areal dimer defect densities were determined using particle detection software for the STM. By contrast, the higher quality substrates that were cleaned with the CMOS procedure and prepared after the manipulator and sample holder cleaning typically exhibited dimer row defects on approximately 2 % to 3 % of the surface. While the CMOS and other cleaning procedures mentioned above are ulti- mately believed to be an important aspect in preparing clean substrates, it is ap- parent that a potentially monumental effort would be required to sufficiently reduce contaminants on substrates in this system. Considering this and the continued influ- ence of contaminants and pinning sites on the growth morphology, surface roughness, and possibly epitaxial quality of 28Si films deposited with substrate temperatures above 600 ◦C, a strategy of lower temperature deposition was adopted to limit the effects of the pinning sites. A predominantly layer-by-layer growth mode at higher temperatures is believed to facilitate and enhance the formation of step bunching at pinning sites and faceting at step bunching, as discussed previously. Depositing with a substrate temperature below 600 ◦C in a 3D island growth mode prevents step bunching and faceting from dominating the growth. While the formation of 3D multi-layer islands represents some intrinsic roughness in the growth surface, the merging of islands in this growth mode leads to an overall smoother surface. Another strategy adopted to improve the epitaxial quality of films and heal potential defects resulting from deposition at lower temperatures is post-deposition 290 annealing. Thin film annealing has been used in both Si MBE and 28Si IBE ex- periments to recrystalize amorphous layers and extend hepi, the epitaxial critical thickness. Extending hepi for a given deposition temperature was demonstrated by annealing at a minimum temperature of 500 ◦C, which is needed to break Si-H bonds that can form at crystalline defects [165]. Another experiment found that annealing at 600 ◦C results in SPE with a recrystallization front propagation speed of approximately 60 nm/min, allowing crystalline 28Si layers to form from regions amorphized by high energy ion impacts [166]. Finally, annealing at 510 ◦C has been used to smooth thin films of Si deposited at low temperature and reduce roughness due to surface islands [167]. A post-deposition anneal of 600 ◦C for 1 h was chosen for most low temperature depositions in this work. The strategy for eliminating pinning sites and producing smooth, epitaxial thin films of 28Si consists primarily of preparing atomically clean Si substrates followed by depositing at temperatures below 600 ◦C. Although the elimination of contaminants on the substrates through ex situ cleaning is not fully experimentally realized, this strategy can be summarized as follows: 1. prepare Si substrates free of organics and metals ex situ with a CMOS cleaning procedure, 2. handle Si substrates with non-metal tools only, 3. manipulate and heat Si substrates in situ with only Mo parts cleaned of other metals, 4. flash anneal Si substrates to approximately 1200 ◦C to prepare a clean surface, verified by RHEED and STM, 5. deposit 28Si in a 3D island growth mode below 600 ◦C to reduce pinning and step bunching, 291 6. and anneal the 28Si film at 600 ◦C for 1 h. 5.6.4 Morphology of Films with Deposition T < 600 ◦C A total of 18 28Si samples were deposited at sample location DC–3 with a range of substrate temperatures following the CMOS and other strategies laid out above for producing smooth, epitaxial films. Samples were deposited with substrate tem- peratures between approximately 249 ◦C and 502 ◦C on Si(100) substrates. Typical average ion energies, Ei, between approximately 35 eV and 40 eV were used. The epitaxy phase diagram for IBE in Fig. 5.18 predicts that a deposition temperature of approximately 350 ◦C is needed with these ion energies to produce high-quality epitaxial deposition, although the exact value would vary with deposition rate and other factors. These samples were deposited on a variety of Si(100) substrates in- cluding phosphorous-doped, boron-doped, and intrinsic wafers (University Wafer) initially before transitioning to higher quality phosphorous-doped wafers (Virginia Semiconductor). 5.6.4.1 RHEED Initially, 28Si samples deposited at low temperature (i.e. below 600 ◦C) were inspected using RHEED immediately following deposition to determine the epi- taxial quality and morphology of the films. The first thicker sample produced at low temperature was a 28Si film deposited at approximately 357 ◦C at DC–3. A RHEED diffraction pattern of this sample after deposition is shown in Fig. 5.35. This image was acquired with the sample at the deposition temperature and the 292 Figure 5.35: RHEED diffraction pattern of a 28Si sample deposited at 357 ◦C at DC–3. This image was acquired with the sample at the deposition temperature and the electron beam in the 〈110〉 direction. The presence of (1×1) and weak (2×1) Si diffraction streaks indicate that the film is crystalline and aligned to the Si(100) substrate. Additionally, the streaks of this pattern corresponds to diffraction from very narrow terraces, likely due to a surface consisting of small, flat islands. A diffuse 3D transmission pattern is also visible superimposed on the streaks indicating some surface roughness. electron beam in the 〈110〉 direction. The diffraction pattern is clearly different from those of the 28Si samples deposited with substrate temperatures above 600 ◦C as in Fig. 5.19. In the pattern of this 357 ◦C sample, both (1×1) and (2×1) Si diffraction streaks are visible indicating that the film is crystalline and epitaxially aligned to the Si(100) substrate. The elongation of the nominal spots into streaks on this pattern corresponds to diffraction terraces that are narrow, e.g. about 20 nm wide, in two dimensions (parallel and perpendicular to the RHEED electron beam). This diffraction pattern is likely due to a predominantly 2D surface consisting of small, flat islands. This suggests that the low temperature deposition strategy was indeed successful at producing a smooth film free of large mounds. 293 Although the well defined 3D transmission spot pattern of the high tempera- ture samples is absent from this image, there does appear to be a diffuse transmis- sion diffraction characteristic visible superimposed on the streaks, which indicates the presence of some small amount of surface roughness, possibly due to the is- lands themselves. Additionally, there is no diffraction attributable to microfacets visible in this pattern, i.e. the “chevron” pattern or other lines connecting adja- cent diffraction rods previously seen. With Ei ≈ 46 eV for this sample, the epitaxy phase diagram for IBE in Fig. 5.18 would predict that a deposition temperature above 450 ◦C would be needed to produce the epitaxy seen in the RHEED pattern. However, this sample had a fairly slow growth rate of approximately 0.33 nm/min, which may offset the effect of a lower temperature, and it was quite thin at only approximately 50 nm as measured by SIMS depth profiling. This film may indeed have been deposited in the limited epitaxy regime as Fig. 5.18 suggests but was too thin to develop the amorphous phase at hepi, which has been shown to be as large as 1 µm for deposition occurring above 300 ◦C [120]. Nearly all of the low temperature 28Si samples had RHEED diffraction pat- terns similar to the one above in Fig. 5.35, although with varying intensities of the transmission aspects of the patterns, indicating varying degrees of roughness on the surface of the films. The most commonly used deposition temperatures for these samples were nominally 400 ◦C or 450 ◦C. A summary of the evolution of the RHEED diffraction patterns from these low temperature films to the high temper- ature films is shown for a series of eight 28Si samples in Fig. 5.36. These samples were deposited at DC–3 on Si(100) substrates with substrate temperatures of ap- 294 Figure 5.36: RHEED diffraction patterns of eight 28Si films deposited at DC–3 at 249 ◦C, (a), 349 ◦C, (b), 421 ◦C, (c), 502 ◦C, (d), 610 ◦C, (e), 708 ◦C, (f), 804 ◦C, (g), and 920 ◦C, (h) on Si(100) substrates. These images were acquired with the samples at the deposition temperatures and the electron beam in the 〈110〉 direc- tion. (a) The 249 ◦C film has a diffuse pattern with faint (1×1) bulk Si streaks indicating a partially disordered film and possibly a fully amorphous layer. (b)–(d) The patterns of films deposited between 349 ◦C and 502 ◦C have (1×1) and (2×1) Si diffraction streaks indicating crystalline films with small, flat islands on the surface. Diffuse 3D transmission spots superimposed on (c) indicates some surface rough- ness. (e)–(h) The patterns of films deposited between 610 ◦C and 920 ◦C show 3D transmission spots indicating rough, crystalline surfaces. A faint “chevron” pattern in (e) indicates diffraction from {311} microfacets. The (2×1) spots in (g) and (h) are likely due to diffraction from the substrate outside of the deposition areas. 295 proximately 249 ◦C, (a), 349 ◦C, (b), 421 ◦C, (c), 502 ◦C, (d), 610 ◦C, (e), 708 ◦C, (f), 804 ◦C, (g), and 920 ◦C, (h). These images were acquired with the samples at the deposition temperatures and the electron beam in the 〈110〉 directions. The diffraction pattern for the 249 ◦C sample in panel (a) is diffuse with faint (1×1) bulk Si streaks. This indicates that the film is at least partially disordered and possibly fully amorphous. The (1×1) streaks could be due to part of the RHEED electron beam diffracting from the substrate outside of the amorphous deposition area, although faint (2×1) streaks would probably be expected as well in that case. (1×1) streaks could also be due to diffraction of crystalline Si below a thin disor- dered layer or partial ordering within the film. This 249 ◦C sample was deposited with Ei ≈ 38 eV and was determined by SIMS depth profiling to be approximately 305 nm in the thickest area measured. Given these parameters, the implication of the RHEED pattern agrees with the epitaxy phase diagram for IBE shown in Fig. 5.18 (b), which predicts limited epitaxy transitioning to an amorphous phase. This sample differs from the 357 ◦C sample in that it had a much higher deposition rate of approximately 1.25 nm/min, and it was much thicker. The diffraction patterns of the films deposited with substrate temperatures between 349 ◦C and 502 ◦C in panels (b)–(d) of Fig. 5.36 show (1×1) and (2×1) Si diffraction streaks indicating crystalline films epitaxially aligned to the Si(100) substrates with small, flat islands on a predominantly 2D surface, similar to the pattern of the 357 ◦C sample in Fig. 5.35. Diffuse 3D transmission spots are visible superimposed on the pattern of the 421 ◦C sample in panel (c) of Fig. 5.36 and to a lesser extend on the pattern of the 502 ◦C sample in panel (d), indicates some small 296 surface roughness. For samples with deposition temperatures higher than that of the 502 ◦C sam- ple, there is an abrupt transition in the diffraction pattern from that of a 2D surface with islands, to that of a rough 3D surface, and this latter pattern persists to the highest deposition temperatures used here. The diffraction patterns of the films de- posited with substrate temperatures between 610 ◦C and 920 ◦C in panels (e)–(h) of Fig. 5.36 show 3D transmission spots indicating very rough but crystalline surfaces, which known to be covered in mounds. These patterns are similar to that of the 708 ◦C sample shown in Fig. 5.19. A faint “chevron” pattern can be seen in panel (e) of Fig. 5.36 indicating diffraction from {311} microfacets. The (2×1) spots seen in panels (g) and (h) are likely due to diffraction from the substrate outside of the deposition areas. 5.6.4.2 STM In addition to observing 2D island growth diffraction patterns in RHEED for 28Si samples deposited at lower temperature, the STM was used to inspect the sur- face morphology of these samples. The surface of a 28Si sample deposited at 357 ◦C at DC–3, corresponding to the RHEED diffraction pattern shown in Fig. 5.35, is shown in the STM topography filled state image in Fig. 5.37. This film is approxi- mately 50 nm thick as determined from SIMS depth profiling, and it was deposited on a Si(100) substrate that was prepared ex situ by the CMOS cleaning procedure. After deposition and initial imaging in the STM, the sample was annealed at approx- 297 Figure 5.37: STM topography filled state images of a 28Si sample deposited at 357 ◦C at DC–3. This film is about 50 nm thick determined from SIMS depth profiling and was deposited on a Si(100) substrate after CMOS cleaning. The sample was annealed at about 600 ◦C for 10 min. Images were acquired with a tip bias ≈ -1.8 V and a tunneling current ≈ 100 pA. (a) Large area scan of the sample showing small, multi- layer 3D islands comprising a smooth surface. Peaks and valleys are seen across the surface due to islands merging. The surface width determined from the topography is ∆z ≈ 1.6 nm or almost 12 monolayers. (b) Small area scan of an area in (a) showing multi-layer islands roughly 10 nm wide with many steps visible. Si (2×1) dimer rows are seen on the terraces making up the islands. 298 imately 600 ◦C for 10 min to form a more ordered surface before further imaging shown here. Panel (a) is a large area scan of the sample showing many small, multi-layer 3D islands comprising a relatively smooth, epitaxial surface. Peaks and valleys are seen across the surface likely the result of various islands merging during deposition. The topography of this image gives a surface width ∆z ≈ 1.6 nm for the film. The height of a single atomic layer on a Si(100) surface is approximately 0.136 nm, and so the measured surface width of this sample equates to there being approximately 13 crystalline layers exposed on the surface. Panel (b) of Fig. 5.37 is a smaller scan of an area in panel (a). Again, multi- layer islands roughly 10 nm wide are seen covering the surface with many steps and terraces. Si (2×1) dimer rows are visible on the terraces forming the islands in this image indicating epitaxial alignment with the substrate. The morphology observed here matches the surface structure determined from the RHEED pattern for this sample (Fig. 5.35). Further, the lack of mounds, large areas of step bunching, or evidence of pinning sites visible in Fig. 5.37 affirms that lower temperature deposi- tion, i.e. below 600 ◦C, does completely alter the growth morphology, resulting in a smooth film. The smooth morphology of sample is in contrast to the rough surface shown in Fig. 5.21 and the other samples with large mounds on the surface. To replicate these results, several other 28Si samples were deposited with a similar or slightly higher substrate temperature including nominal deposition tem- peratures of 400 ◦C and 450 ◦C, which were most commonly used. The smooth morphology of the 357 ◦C sample shown in Fig. 5.37 is generally found to be repro- ducible for all of these depositions, which produce films with surfaces that appear 299 qualitatively similar in the STM. Four STM topography filled state images of four different 28Si samples deposited at DC–3 with substrate temperatures of approxi- mately 349 ◦C, (a), 417 ◦C, (b), and 421 ◦C, (c) and (d), are shown in Fig. 5.38. These samples were deposited on Si(100) substrates that were prepared ex situ by the CMOS cleaning procedure. After deposition, these samples were all annealed at approximately 600 ◦C for 1 h to form more ordered surfaces. Similar to the 357 ◦C sample, all of these images (panels (a)–(d)) show multi-layer 3D islands comprising relatively smooth, epitaxial surfaces. Note that the scan areas in these four images are the same to facilitate comparisons between them. The film shown in panel (a) is approximately 206 nm thick with islands that are roughly 20 nm wide. The surface width of this film determined from the STM topography is ∆z ≈ 1.0 nm, which is approximately eight atomic layers exposed to the surface. These eight layers are visible and can be counted in the image. The film shown in panel (b) is approximately 250 nm thick with islands that are roughly 15 nm wide. The surface width of this film determined from the topography is ∆z ≈ 2.1 nm, which is approximately 16 atomic layers. The film show in panel (c) is approximately 148 nm thick with islands that are roughly 30 nm wide. The surface width of this film determined from the topography is ∆z ≈ 2.2 nm, which is approximately 17 atomic layers. Finally, the film shown in panel (d) is approximately 320 nm thick with islands that are roughly 20 nm wide. The surface width of this film determined from the topography is ∆z ≈ 1.0 nm, which is, again, approximately eight atomic layers. The surface in panel (d) also shows some dark dimer row defects on the terraces at a higher concentration than the other samples, possibly indicating 300 Figure 5.38: STM topography filled state images of four 28Si samples deposited at DC–3 at 349 ◦C, (a), 417 ◦C, (b), and 421 ◦C, (c) and (d). These samples were deposited on Si(100) substrates after CMOS cleaning and were annealed at about 600 ◦C for 1 h. Images were acquired with a tip bias ≈ -1.8 V and a tunneling current ≈ 100 pA. Multi-layer 3D islands comprising smooth surfaces are seen on all samples. Peaks and valleys are seen due to islands merging. (a) This film is about 206 nm thick with islands roughly 20 nm wide and the surface width determined from the topography is ∆z ≈ 1.0 nm. (b) This film is about 250 nm thick with islands roughly 15 nm wide and a surface width of ∆z ≈ 2.1 nm. (c) This film is about 148 nm thick with islands roughly 30 nm wide and a surface width of ∆z ≈ 2.2 nm. (d) This film is estimated to be about 320 nm thick with islands roughly 20 nm wide and a surface width of ∆z ≈ 1.0 nm. The surface in (d) also shows some dark dimer row defects on the terraces. Si (2×1) dimer rows are visible on the island terraces in (a)–(d). Film thicknesses were determined from SIMS depth profiling except for that of the sample in (d), which is an estimate. 301 the presence of impurities in the film. These film thicknesses were the maximum measured thicknesses determined from SIMS depth profiling except for that of the sample in panel (d), which is an estimate based on the measured 28Si ion beam flux during deposition. Si (2×1) dimer rows are visible on the terraces forming the island in panels (a)–(d) indicating epitaxial alignment with the substrates. The overall smoothness, i.e. the lack of extended step bunching and facet and mound formation, of these 28Si films deposited with substrate temperatures below 600 ◦C is achievable for several reasons. As discusses previously, lower tempera- ture deposition reduces step movement associated with higher temperature layer- by-layer growth. This step movement can facilitate the formation of step bunching and faceting, ultimately leading to mounds. Additionally, contaminants that cause pinning sites and step bunching may cluster or become active at higher tempera- tures. These effects are reduced during low temperature deposition in a 3D island growth mode. However, 3D island growth introduces some intrinsic roughness and step bunching around the islands, and defect formation that may act as pinning sites can result from merging islands. Despite these aspects of 3D island growth, films with a smooth morphology still occur due to the nature of island formation on Si surfaces. Si island growth is anisotropic due to the difference in the diffusion constants for adatom movement along or perpendicular to dimer rows. Anisotropic growth is believed to result in the growth surface being constrained to a limited number of atomic layers and thus a limited amount of surface roughness and step bunching [160, 168]. This appears to be the case for the 28Si films deposited in a 3D island growth mode as in Fig. 5.38. The surface width of all of these samples is 302 fairly consistent with a typical value of ∆z ≈ 2 nm, which equates to approximately 16 atomic layers being exposed on the surface at any time. The surface width of these samples does not seem to vary with deposition temperature or film thickness, although there is not enough data to more strongly support these conclusions. This possibly indicates that the measured ∆z values are indeed constant and intrinsic to Si island growth. Additionally, the hyperthermal energy of the 28Si ions may play a role in producing a smooth surface during 3D island growth. Energetic ions are known to suppress and dissociate 3D islands as well as defect clusters leading to smoother growth. Mobile adatoms formed in the break up of islands may continually fill in other inter-island trenches [120,127,169]. It is not clear if the 28Si ion energy has any effect on the smoothing of films in this work because low temperature deposition using the natural abundance Si EFM evaporator, which produces Si atoms with only thermal energies, also results in smooth films. STM inspection of these natSi films shows that they consist of small islands qualitatively similar to those of the 28Si films. While most of the 28Si films deposited with substrate temperatures below 600 ◦C had surfaces that appear similar in the STM to those of the samples in Fig. 5.38, several had surface morphologies that appeared different including two samples with deposition temperatures of 502 ◦C and 249 ◦C. STM topography filled state images of these samples and four others are shown in Fig. 5.39, which serves to summarizes the variation in film morphology over the range of deposition temperatures used in this work. 28Si samples were deposited at DC–3 with substrate 303 Figure 5.39: STM topography filled state images of six 28Si samples deposited at DC–3 at 804 ◦C, (a), 705 ◦C, (b), 502 ◦C, (c), 421 ◦C, (d), 349 ◦C, (e), and 249 ◦C, (f). Images were acquired with a tip bias≈ -1.8 V and a tunneling current≈ 100 pA. All substrates used for these samples were prepared ex situ by CMOS cleaning except for the sample in (a), which was etched with HF. (a) and (b) The morphology of films deposited above 600 ◦C appear very rough with large mounds and grains. The total height scale in (a) and (b) is about 45 nm and 50 nm respectively. (c) The 502 ◦C sample morphology appears less rough with a large mound as well as small island-like features visible. The height scale is about 13 nm. (d) and (e) The morphology of films deposited between 349 ◦C and 421 ◦C appear smooth with small, multi-layer 3D islands. The height scales in (d) and (e) are about 1.3 nm and 3.6 nm respectively. (f) The 249 ◦C sample shows a smooth surface with small features ≈ 3 nm in size. The height scale is about 1.0 nm. 304 temperatures of approximately 804 ◦C, (a), 705 ◦C, (b), 502 ◦C, (c), 421 ◦C, (d), 349 ◦C, (e), and 249 ◦C, (f). All Si(100) substrates used for these samples were prepared ex situ by the CMOS cleaning procedure except for the sample in panel (a), which was etched with HF. Note that the scan area displayed in panels (a)–(f) are all the same to facilitate comparisons between them. Panels (a) and (b) of Fig. 5.39 show the morphology of films deposited above 600 ◦C. These surfaces appear very rough with large mounds or grain-like features. The total height scale in panels (a) and (b) is approximately 45 nm and 50 nm respectively. For comparison, SEM micrographs of the 804 ◦C sample in panel (a) and the 705 ◦C sample in panel (b) were shown in Fig. 5.23 (c) and Fig. 5.34 respectively. Panel (c) of Fig. 5.39 shows the morphology of the 502 ◦C sample, which appears less rough than the samples in panels (a) and (b). This sample was annealed at approximately 600 ◦C for 1 h after deposition. A large mound is seen in the center of the image with smaller island-like features on it. The height scale of this image is approximately 13 nm. The RHEED pattern of the 502 ◦C sample (Fig. 5.36 (d)) indicated a smooth surface with small, flat islands, which does not seem to be the case from the STM image. The nature of this discrepancy is not known. Panels (d) and (e) of Fig. 5.39 show the morphology of films deposited with substrate temperatures of 349 ◦C and 421 ◦C, but they represent relatively smooth films produces at deposition temperatures between 349 ◦C and 460 ◦C in this work. These samples were annealed after deposition at approximately 600 ◦C for 1 h. The surface of these samples appear smooth with small, flat, multi-layer 3D islands. 305 Smaller area scans of these samples show Si (2×1) dimer rows on the islands in- dicating epitaxial deposition. The height scales in panels (d) and (e) are approxi- mately 1.3 nm and 3.6 nm respectively. Finally, panel (f) shows the morphology of the 249 ◦C sample, which has a smooth surface covered with small round features approximately 3 nm in size that do not appear epitaxial. The height scale of this image is approximately 1.0 nm. The RHEED pattern of this sample (Fig. 5.36 (a)) indicated an amorphous surface layer on the film, which may explain the round features on the surface and absence of larger epitaxial islands. It appears from this work that the smoothest epitaxial 28Si films are achieved with deposition temperatures between approximately 349 ◦C and 460 ◦C. However, although these smooth 28Si films were achieved with an apparent disappearance of the pinning sites seen in the higher temperature samples, the defects and contam- inants causing those pinning sites are likely still present in the film. While their effects do not manifest in the film morphology at low deposition temperatures, they may still affect the bulk crystallinity or even the electronic properties of the films. 5.7 Chemical Purity 5.7.1 Context The 28Si films produced in this work need to have a high chemical purity in order to be considered comparable to both commercially available electronics grade natural abundance Si as well as the enriched 28Si available in the QI research community. As mentioned at the beginning of this chapter, the third materials goal 306 for these 28Si films is to have chemical impurity concentrations including C and O below 2 × 1015 cm−3. Chemical contaminates are undesirable in part because of their detrimental effect on the smooth morphology and bulk epitaxy of thin films, as discussed in detail above. Additionally, as mentioned previously, chemical impurities present in the bulk of a 28Si film can act as or induce various scattering sites and charge traps. This reduces electron mobility and other electronic properties important for operating QI devices [91,92]. Further, chemical impurities can possess nuclear spin including 13C, which has a nuclear spin I = 1/2, and 14N, which has a nuclear spin I = 1. The presence of these nuclear spins will cause decoherence of qubit spins in a quantum computing device, just as 29Si does. Several sources of chemical contaminants exist within the deposition system that may contribute impurities that become incorporated into the 28Si films. Con- taminants can come from various molecular species comprising the background par- tial pressures in the deposition chamber. A residual gas mass spectrum of the base pressure of the deposition chamber acquired from the RGA is shown in Fig. 2.22 in Chapter 2 and gives insight into the contaminants present in the vacuum such as N, C, O, and F, although quantitative partial pressures are not reliable. The typical base pressure of the deposition chamber was approximately 6.7× 10−9 Pa (5.0× 10−11 Torr). Other gaseous contaminants may diffuse into the deposition chamber from the ion beam chamber during deposition. These can originate from either the background base pressure of the ion beam chamber, which was typically approximately 1.3× 10−5 Pa (1.0× 10−7 Torr), or the SiH4 gas source used during deposition. Potential contaminants from the ion beam chamber including N, C, and 307 O can be seen in the residual gas mass spectrum of the base pressure acquired from the RGA in the ion beam chamber, shown in Fig. 2.3 in Chapter 2. While the SiH4 source bottle has a purity of 99.999 % according to the gas vendor (Matheson Tri-Gas), the gas manifold used to load SiH4 into the ion source may contain much higher levels of residual gas impurities because it is only pumped out to roughly 20 mTorr. When heating samples on the manipulator during deposition, the ele- vated temperatures of the sample holder and other parts of the manipulator can cause increases in the background pressure due to outgassing. Probably the most significant source of chemical contaminants in the 28Si films, specifically N, C, and O, is the ion beam itself. Selecting for 28Si+ ions in the ion beam is the result of tuning the sector mass analyzer such that any ion with a mass-to-charge ratio ≈ 28 u/e ± 0.18 u/e (at 28 u) will pass through the mass-selecting aperture and propagate to the sample. The acceptance through the aperture of a range of masses approximately 0.36 u wide (for a singly charged ion) around 28 u is due to the width of the mass-selecting aperture (2 mm). This mass range is also similar to the mass resolution at 28 u of 0.35 u, which is the smallest mass that can be resolved by the system based on the maximum measured mass resolving power m ∆m ≈ 78. Any ions with masses within this range will be trans- ported to the sample along with 28Si. Two ionic species that match this criteria are 14N+2 and 12C16O+. The mass of 28Si is 27.97692653465(44) u, the mass of 12C16O is 27.99491461957(17) u, and the mass of 14N2 is 28.00614800886(40) u [170]. There- fore, the separation in mass between 28Si and 12C16O is approximately 0.018 u, and the separation in mass between 28Si and 14N2 is approximately 0.029 u, which are 308 both well within the range for transport past the aperture when the 28 u beam is centered on it. A mass resolving power of m ∆m ≈ 1600 would be needed to signifi- cantly separate 28Si from 12C16O and 14N2. The maximum achievable mass resolving power for this system with an ion beam width due only to the energy spread from the ion source, i.e. with no intrinsic beam width, would be approximately 350 at mass 28 for a typical energy spread of ∆E = ± 6 eV. This means that for this system, any N2 and CO ions generated in the ion source will be passed into the de- position chamber and onto the sample along with 28Si. Further, these contaminant species that have a mass separation below the mass resolution of the system are not detectable in an ion beam mass spectrum. 5.7.2 XPS XPS was used as an initial check on the chemical purity of 28Si samples de- posited at elevated temperatures at DC–3 and to compare them to the previous sample measured with XPS, which was deposited at room temperature at LC–2 (Fig. 4.17). This measurement was a search for gross chemical contaminants that may have been introduced into the samples arising from the significant experimental switch from 28Si deposition at LC–2 to DC–3 at elevated temperatures. XPS spectra shown in this section were acquired and analyzed in collaboration with Dr. Kris- ten Steffens (NIST). XPS spectra were collected after sputter-cleaning the sample with Ar to remove surface contamination from the environment. The measurement parameters for the spectra shown here are similar to those used for the previous sample (see Section 4.3.6.2). 309 Figure 5.40: XPS spectra of a 28Si sample deposited at 812 ◦C at DC–3. Count rates vs. electron binding energy for survey scans are shown for the 28Si sample after Ar sputter cleaning (upper spectrum) as well as the substrate away from the deposition area (lower spectrum). The data for the 28Si sample was shifted up for clarity. References for elemental orbital levels are included above relevant peaks. O 1s and C 1s peaks are visible with larger amplitudes in the 28Si film than the substrate. XPS spectra for the second 28Si sample deposited at DC–3 with a deposition temperature of approximately 812 ◦C is shown in Fig. 5.40 (upper spectrum). Also shown is a reference spectrum of the Si substrate away from the deposition area (lower spectrum). References for relevant elemental orbital levels are shown at the top. A C 1s peak is visible at 285.2 eV in the 28Si spectrum corresponding to an atomic fraction of approximately 3.4 %. Additionally, a O 1s peak is visible at 532.7 eV corresponding to an atomic fraction of approximately 4.5 %. No C or O peaks are visible above the noise in the spectrum for the Si substrate. A 310 small N 1s signal is also present at 400.9 eV corresponding to an atomic fraction of approximately 1.7 %. Both the 28Si and the substrate spectra show Ar peaks due to the Ar sputter cleaning process. The expected instrumental background for C and O in these scans is approximately 1 % to 2 %. This measurement is similar to the previous XPS measurement of a 28Si sample deposited at LC–2 with similar atomic fractions of C and O. The small N concentration in the sample was not previously detected. No other major contaminants were detected in this sample. Given the large background signal for these measurements, it is difficult to place a bound on the expected Si purity for this sample. 5.7.3 SIMS SIMS was used not only to get a more accurate measurement of N, C, and O concentrations in the 28Si films than XPS can provide but to also analyze the films for a broad range of trace impurities. The SIMS measurements discussed here were performed at EAG Laboratories unless otherwise stated, and are of the major isotope of each element. A SIMS depth profile of the concentration of 22 different potential contaminants in a 28Si film deposited with a substrate temperature of approximately 460 ◦C at DC–3 is shown in Fig. 5.41. Atomic concentrations of 28Si (circles and line), 29Si (squares and line), and 30Si (triangles and line) are also shown vs. sputter depth as an indicator of the enriched film where the 29Si and 30Si values are reduced. The minimum detected 29Si and 30Si concentrations can only be taken as bounds on the enrichment of this sample because they are limited by the measurement noise floor. At a depth of approximately 145 nm they return to their 311 Figure 5.41: SIMS depth profile of the concentration of 22 atomic species in a 28Si sample deposited at 460 ◦C at DC–3. Atomic concentrations of 28Si (circles and line), 29Si (squares and line), and 30Si (triangles and line) are also shown vs. sputter depth as an indicator of the enriched film. The minimum detected 29Si and 30Si con- centrations are limited by the measurement noise floor. At a depth of about 145 nm they return to their natural abundance values in the Si(100) substrate (shaded re- gion), indicating the film thickness. Depth profiles of the atomic concentrations of many elements including light gases and metals (open and closed symbols and lines) in the enriched film and substrate are shown vs. sputter depth. N, C, O, F, and Al were detected in the film while the remaining elements were not, indicated by the box in the legend. The atomic concentration of N in the film is 7.1(1) ×1020 cm−3, the concentration of C in the film is 4.4(2) ×1019 cm−3, and the concentration of O in the film is 2.1(7) ×1019 cm−3. The signals for these elements drop to the measurement detection limit in the substrate. natural abundance values in the Si(100) substrate, which is marked by the shaded region. This indicates the film interface with the substrate and gives a value for the film thickness. Depth profiles of the atomic concentrations of many elements that are potential contaminants in the 28Si films, including light gases and metals, are shown vs. the sputter depth. N, C, O, F, and Al (solid lines) were all detected 312 in the film. The remaining elements Cl, Be, P, Na, K, Ti, V, Cr, Fe, Ni, Cu, Au, Ag, As, Sb, Ta, and W (open and closed symbols and lines) were not detected in the film or substrate down to the measurement detection limit, which is an atomic concentration of approximately 1 ×1016 cm−3 for Sb, Au, As, and P, and is an atomic concentration of approximately 5 ×1015 cm−3 for the remaining elements. The average atomic concentration of N within the film was measured to be 7.1(1) ×1020 cm−3, or 1.42(2) %. The uncertainties of the values of atomic concen- trations for the SIMS measurements discussed here are the standard deviations of the means. This value is similar to the value of the N concentration of a 28Si sample determined from XPS of approximately 1.7 %. The average atomic concentration of C in the film was measured to be 4.4(2) ×1019 cm−3, or 880(40) ppm, and the aver- age atomic concentration of O in the film was measured to be 2.1(7) ×1019 cm−3, or 420(140) ppm. These values are significantly less than the C and O values measured by XPS. This may be due in part to the high measurement background in XPS for C and O. The differences between the two samples is the ex situ cleaning procedure (none vs. CMOS) and the deposition temperatures (812 ◦C vs. 460 ◦C). It may be that an increase in the background pressure during deposition of the 812 ◦C sample due to the high temperature of the sample and holder is responsible for increased C and O concentrations. The signals for these elements drop to the detection limit of the measurement in the substrate. The structural relation between chemical impurities and the surrounding Si crystal is dictated by the solid solubility of those impurities in Si. Small changes in the atomic concentration of an impurity can greatly influence its effect on the 313 Figure 5.42: Phase diagrams for the N-Si, C-Si, and O-Si systems with low impurity concentrations near the solid solubility limit. (a) The N-Si phase diagram near the melting point shows the solid solubility of N in Si is about 4.5 ×1015 cm−3 and above this, Si3N4 forms. (b) The C-Si phase diagram near the melting point shows the solid solubility of C in Si is about 3.2 ×1017 cm−3 and above this, SiC forms. (c) The O-Si phase diagram near the melting point shows the solid solubility of O in Si is about 2.8 ×1018 cm−3 and above this, SiO forms. ((a) and (c) from Ref. [171], (b) from Ref. [172]) structural properties of the host Si. Phase diagrams for the N-Si, C-Si, and O-Si systems in the case of extremely low impurity concentrations near the solid solubility are shown in Fig. 5.42. The N-Si phase diagram from Ref. [171] in panel (a) shows that the solid solubility limit of N in Si is approximately 4.5 ×1015 cm−3. At atomic concentrations of N that are higher than this, Si3N4 crystallites form in the Si. Similarly, the C-Si phase diagram from Ref. [172] in panel (b) shows that the solid solubility limit of C in Si is approximately 3.2 ×1017 cm−3. At atomic concentrations 314 of C that are higher than this, SiC crystallites form in the Si. Finally, the O-Si phase diagram from Ref. [171] in panel (c) shows that the solid solubility limit of O in Si is approximately 2.8 ×1018 cm−3. At atomic concentrations of O that are higher than this, SiO clusters form in the Si. Based on these values, the measured atomic concentrations of N, C, and O in the 460 ◦C sample analyzed in Fig. 5.41 are well beyond their respective solid solubility limits in Si and so those contaminants likely exist as Si3N4 and SiC crystallites, and SiO clusters in the 28Si film. The average atomic concentration of F in the 460 ◦C sample analyzed in Fig. 5.41 was measured to be 3.5(3) ×1016 cm−3, or 0.70(6) ppm. This F is be- lieved to originate from the background vacuum in the deposition chamber and incorporate into the depositing film. A partial pressure of F is always observed in the RGA mass spectrum at a level larger than H2O, and can be seen in the RGA spectrum recorded while operating the ion beam in Fig. 5.3. Several instruments and apparatus components within the vacuum chamber contain teflon insulation (PTFE), which is suspected to outgas F in UHV environments and contribute to the observed partial pressure. These components include insulated wires on the manipulator that supply power for sample heating, insulated wires on the STM tip preparation tool, which heats STM tips, teflon support structures for the ion beam deceleration lenses, and Viton seals on gate valves. The STM tip preparation tool in particular was observed to significantly increase the partial pressure of F while in use. Eliminating F-containing compounds within the vacuum system can reduce the concentration of F in samples. The atomic concentration of Al detected in the film is 6.2(3) ×1015 cm−3, or 0.12(2) ppm. The origin of this Al is an Al deposition 315 source sitting above the sample location that consists of an Al wire wrapped around a tungsten heating element. The total purity of this 28Si sample calculated from these measurements is approximately 98.45(2) %, which is quite low compared to commercial Si wafers. The concentrations of N, C, and O in the sample are far too high for the material to be viable for use in quantum coherent devices or other QI related experiments. The concentration of nuclear spins due to the 14N (I = 1) in the sample is much larger than the residual 29Si isotopic concentration and is only roughly a factor of three lower than that of 29Si in natural abundance Si. It is difficult to determine if the N, C, and O detected in these films predominantly originates from the ion beam or the background vacuum. It is unlikely that the N concentration is due to adsorption from the vacuum because the partial pressure of N2 needed to account for the measured concentration (with unity sticking) is higher than the measured partial pressure of N2 in the chamber during deposition. This indicates that most, i.e. likely > 90 %, of the N in this sample was introduced through the ion beam. A second 28Si sample deposited at DC–3 was also analyzed by SIMS for chem- ical contaminants after the first sample. This second sample had a deposition tem- perature of approximately 421 ◦C. The only significant differences between the first sample and this second one are that the deposition rate of the second sample was roughly two times larger and the gas manifold was purged with Ar instead of N2 before deposition. It was thought that the previous N2 purge may have contributed to the high atomic concentration of N measured in the first sample, although this was not verified. The atomic concentrations of N, C, and O were all lower than 316 those of the first sample but not significantly, with slightly over half the total im- purities. The average atomic concentration of N in this film was measured to be 4.16(5) ×1020 cm−3, or 0.83(1) %. The average atomic concentration of C in the film was measured to be 1.81(2) ×1019 cm−3, or 363(4) ppm, and the average atomic con- centration of O in the film was measured to be 6.93(6) ×1018 cm−3, or 139(1) ppm. The total Si purity of this second sample is approximately 99.12(1) %. In total, three SIMS measurements were made on this sample in three different locations on the deposition spot. These three locations had three different measured film thicknesses, which also means that they had three different local deposition rates. The measurements of the chemical contaminants for the three locations can then be compared for different deposition rates while other variables including the background pressure are constant. Such an analysis can give insight into the source of the contaminants. For adsorption from the background vacuum, an increasing deposition rate would result in a decreasing contaminant concentration because the gas flux is decreasing relative to the ion beam flux. Contaminants incorporated from the ion beam, however, should remain constant for different deposition rates. The three deposition rates determined from the SIMS depth profiles of the three spots on this sample are approximately 0.11 nm/min, 0.32 nm/min, and 0.88 nm/min. The N concentrations were found to not decrease with increasing deposition rate and instead increased roughly 39 % from the area with the lowest deposition rate to that with the highest. This again indicates that the detected N is not significantly due to adsorption from the vacuum. C and O concentrations, however, were found to decrease with increasing deposition rate. The C concentration decreased by roughly 317 71 % from the area with the lowest deposition rate to that with the highest. This result indicates that the C and O concentrations measured in these films are due, at least partially, to adsorption from the vacuum. To determine the minimum N, C, and O contamination that is contributed to samples by adsorption and incorporation during 28Si deposition, any effects of the ion beam must be eliminated. To this end, a natural abundance Si sample was deposited in the deposition chamber at DC–3 and analyzed by SIMS for chemical contaminants. This sample was deposited by sublimating Si from a bare substrate held at approximately 1150 ◦C and positioned over the surface of the target sub- strate. The target Si(100) substrate was flash annealed before the deposition and was nominally at room temperature during it, although it may have been heated radiatively from the source substrate. SIMS depth profiles of the atomic concentrations of N, C, and Cl (solid lines) in the natSi film are shown vs. sputter depth in Fig. 5.43. An O profile did not yield usable results due to atmospheric contamination. The profiles show slightly increased values near the surface due to environmental surface contamination before leveling off through the bulk of the film. The atomic concentrations of N and C then dip just beyond 200 nm before peaking at approximately 227 nm. This peak corresponds to the interface between the film and the substrate, which is marked by the shaded region and indicates the film thickness. The N and C signals peak at the interface because adsorbates will accumulate on the surface from the vacuum before deposition begins. Initial heating of the sublimation source will also increase the partial pressures of adsorbates, increasing the contaminant concentrations in 318 Figure 5.43: SIMS depth profile of a natSi film deposited at room temperature at DC–3. This film was deposited on a Si(100) substrate by sublimating Si from a second wafer held over the surface. Atomic concentrations of N, C, and Cl in the deposited film and substrate are shown vs. sputter depth. The interface between the film and substrate (shaded region) is indicated by the peak in the N and O signals at a depth of about 227 nm, which also indicates the film thickness. The minimum atomic concentration of N in the film at about 206 nm is 1.1(1) ×1017 cm−3, and the minimum atomic concentration of C in the film is 1.43(3) ×1018 cm−3. the film. The N, C, and Cl signals in the substrate are at the detection limit of the measurement. The dip in the atomic concentrations of N and C near the interface gives an upper bound on the minimum possible N and C concentrations that would be expected in a 28Si film deposited without contamination from other sources such as the ion beam. The minimum atomic concentration of N in the natSi film was measured to be 1.1(1) ×1017 cm−3, or 2.2(2) ppm. The minimum atomic concentration of C in the film was measured to be 1.43(3) ×1018 cm−3, or 28.7(6) ppm. The higher 319 concentrations of N and C throughout the bulk of the film are likely due to increased outgassing of these elements when the target sample began to be heated by the source. While the detected atomic concentration of N in the natSi film is roughly 6.5 ×103 times lower than that measured in the first 28Si film, it is still more than 10 times higher than the 29Si isotopic concentration in the more highly enriched 28Si samples produced in this work. Additional pumping or a faster growth rate may be needed to reduce contaminants from the vacuum to concentrations that meet the purity goal for 28Si films of 2× 1015 cm−3 stated at the beginning of this section. Cl is also detected within this film with an average atomic concentration of 8.5(8) ×1016 cm−3, or 1.7(2) ppm. This Cl may be outgassing from the sample holders or other elements near the sample that get hot because there is no peak in the concentration near the surface due to build up from the vacuum. This means that Cl only appears to be present once the sample heating was turned on to start the deposition. Evidence of Cl in the vacuum chamber can be seen in the residual gas mass spectrum of the RGA. During heating of samples and sample holders for either degassing or sample flashing, several mass peaks can appear that are not normally present in the chamber, particularly peaks at masses of 31 u and 50 u. An example of a residual gas mass spectrum from the RGA that was recorded while a Si substrate was being degassed is shown in Fig. 5.44. A spectrum of the base pressure of the deposition chamber (diagonal line fill) recorded at a later date is given for reference and shows major residual gas peaks corresponding to H2 (2 u), F (19 u), CO and N2 (28 u), and CO2 (44 u). The Si substrate was degassed at approximately 600 ◦C while recording a RGA spectrum (solid fill). During degassing, the pressure 320 Figure 5.44: Residual gas mass spectra collected from the RGA in the deposition chamber while degassing a Si substrate at about 600 ◦C. The base pressure of the chamber (diagonal line fill) is shown for reference with peaks corresponding to H2, F, CO and N2, and CO2 visible. When the substrate is heated for degassing, the pressure increases and several new peaks appear (solid fill) including C and those at masses 31 u and 50 u, which may indicate Cl-containing molecules. rises from the typical base pressure, evidenced by the aforementioned peaks rising. Additionally, several other peaks appear in the spectrum that are not present or much smaller in the base pressure spectrum including a C peak as well as peaks at masses of 31 u and 50 u. There are several chemical compounds that are likely candidates for being as- sociated with these masses in a vacuum environment, some of which contain Cl. Chloromethane (CH3Cl) and related molecules Cl2F2 and CF3Cl all contribute sig- nificant signals to mass 50 u if present in a vacuum. CF3Cl also appears at mass 31. The compounds containing F are possible because F is known to be present in the 321 system, as seen in the base pressure spectra. Other possible candidates for the peak at 31 u are ethanol and propanol. The source of these alcohols as well as Cl may be the ex situ chemical cleaning procedure used on the substrates. The SC–2 solution used contains HCl and so it may be that some residue from the clean resides on the chip and is then released in the chamber during degassing and flashing procedures as well as when samples are heated during deposition of Si films. N2 and CO contamination in the 28Si ion beam can be measured in both the mass spectrum and in a deposited 28Si sample if the concentrations are high enough. A demonstration of this resulted from the case of a leak from air in the gas manifold system that delivers SiH4 to the ion source. An increase in the concentration of N2 and O2 in the ion source while it is running then leads to increased amounts of 14N+2 and 12C16O+ ions being created and extracted into the beamline. Tuning the sector mass analyzer to select for 28Si then also selects these contaminants which are measured in the SiH4 mass spectrum or deposited. 28Si samples which were deposited with these conditions are referred to here as “N-contaminated” samples. A mass spectrum of SiH4 that shows a large amount of N2 and/or CO contamination in the ion beam is shown in Fig. 5.45. The ion currents shown in this figure (open and closed circles) were recorded while sweeping the mass analyzer current, and thus the magnetic field of the analyzer (top axes). Panel (a) shows a semi-log plot of a mass spectrum used for depositing a 28Si sample at DC–3 using the low pressure plasma mode of the ion source. The typical SiH4 current peaks are seen including the 28 u peak, which is 28Si, and the 29 u peak, which is both 28SiH and 29Si. The 29 u peak is assumed to be ≈ 5 % 29Si based on previous mass spectrums. Several 322 Figure 5.45: SiH4 ion beam mass spectrum for N-contaminated 28Si samples de- posited at DC–3 using the low pressure plasma mode of the ion source. The ion currents (open and closed circles) are recorded while sweeping the mass analyzer current, and thus the magnetic field (top axes). (a) Mass spectrum for samples contaminated with large amounts of N shown on a semi-log scale with current peaks typical of SiH4. The 28 u peak is 28Si and the 29 u peak is both 28SiH and 29Si, assumed to be ≈ 5 % of the peak. Several higher order hydrides are also shown. Gaussian fits (line, Eq. (2.14)) to the 28 u and 29 u peaks are shown superimposed on the data. The 28 u peak appears higher than the 29 u peak, which is atypical. (b) Comparison between the spectrum in (a) (closed circles and line) and a nominal SiH4 spectrum (open circles and line) plotted on a linear scale. The difference in signal of the 28 u peaks of the two spectra relative to the 29 u peak is clear. The additional current in the 28 u peak beyond the nominal 28Si current is presumed to be mostly N2. 323 higher order hydride peaks are also shown. Gaussian fits (line, Eq. (2.14)) to the 28 u and 29 u peaks are shown superimposed on the data. Unlike a typical SiH4 mass spectrum for the low pressure mode such as the one acquired before depositing a 28Si sample shown in Fig. 5.4, the 28 u peak appears higher than the 29 u peak here, indicated by the horizontal dashed lines. The increased current in the 28 u peak is due to the contaminants. The difference between the spectra is highlighted in panel (b) of Fig. 5.45, which compares the contaminated mass spectrum (closed circles and line) to a nom- inal SiH4 mass spectrum (open circles and line) from Fig. 5.4 on a linear current scale. These two spectra are comparable because the current levels at the 29 u peak are nearly identical. The difference in the current peak height of the contam- inated spectrum compared to the nominal spectrum is clear here. The additional current in the 28 u peak beyond the nominal 28Si current is presumed to be mostly N2. This contamination comprises approximately 30 % of the 28 u current. For a contaminant beam of only N2, this would actually amount to approximately 46 % contamination in the sample because each N2 ion contains two N atoms. A 28Si sample was then deposited using a contaminated ion beam similar to that represented by Fig. 5.45, and it was analyzed by SIMS for chemical con- taminants. This SIMS analysis was done in collaboration with Dr. David Simons (NIST). SIMS depth profiles of the atomic concentrations of 14N, 12C, 16O, (lines) and 13C (circles and line) in the N-contaminated 28Si sample deposited with a sub- strate temperature of approximately 712 ◦C at DC–3 are shown vs. sputter depth in Fig. 5.46. Also shown is a depth profile of the 29Si isotope fractions (squares 324 Figure 5.46: SIMS depth profiles showing the atomic concentration of 14N, 12C, 16O, (lines) and 13C (circles and line) vs. sputter depth in a 28Si sample deposited at 712 ◦C at DC–3. Also shown is a depth profile of the 29Si isotope fraction (squares and line) vs. sputter depth on the right axis as an indicator of the enriched film. The average 29Si isotope fraction in the film is about 0.132(27) ppm. At a depth of about 256 nm, the 29Si isotope fraction returns to the natural abundance value in the Si(100) substrate (shaded region), which gives a value for the film thickness. The atomic concentration of N in the 28Si film is greater than 1 ×1022 cm−3. This value is not quantitatively accurate, but shows that N is present in the film at a concentration of roughly 30 %. The tail of the N profile into the substrate is an artifact of the SIMS measurement. The measurements of C and O are accurate and show that the concentration of 12C in the film is 6.1(1) ×1019 cm−3, the concentra- tion of 13C in the film is 1.46(3) ×1017 cm−3, and the concentration of 16O in the film is 2.3(1) ×1019 cm−3. The signals for these elements in the substrate are due to the measurement background. and line) vs. sputter depth, which corresponds to the right axis as an indicator of the deposited enriched film. The average measured 29Si isotope fraction in the film is 0.132(27) ppm. At a depth of approximately 256 nm, the 29Si isotope fractions return to the natural abundance value in the Si(100) substrate, which is indicated by the shaded region and gives a value for the film thickness. The atomic concen- 325 tration of 14N detected in this film is extremely high. The measurement of such a high concentration is not quantitatively accurate for the measurement conditions used here, but it appears greater than 1 ×1022 cm−3. This measurement does show that 14N is present in the film at a likely concentration of roughly 30 %. This rough SIMS value for the N contamination is supported by other measurements of simi- lar N-contaminated films including XPS and energy dispersive x-ray spectroscopy (EDX), which give similar values. With this much N present in the 28Si film, it is likely that a significant amount of silicon nitride (Si3N4) forms. The long tail of the 14N profile that extends into the substrate is an artifact of the SIMS measure- ment. 12C and 16O have concentrations much lower than that of 14N. The average atomic concentration of 16O in the film was measured to be 2.3(1) ×1019 cm−3, or 470(20) ppm. This concentration of O is only slightly higher than that of the previ- ous 28Si sample, which was not deposited with a contaminated beam. The average atomic concentration of 12C in the film was measured to be 6.1(1) ×1019 cm−3, or 1240(20) ppm, and the average atomic concentration of 13C in the film was measured to be 1.46(3) ×1017 cm−3, or 2.98(6) ppm. This concentration of C increased over that of the previous 28Si sample by roughly the same amount that the concentration of O increased, which is about 2 ×1019 cm−3. The signals for N, C, and O in the substrate are due to the background level of the measurement. The ratio of 12C to 13C in this sample can indicate its origin because only 12C is selected through the ion beam while both isotopes are present in their natural abundance coming from the vacuum. In this sample, the C ratio is measured to be 12C/13C ≈ 420. This ratio is larger than the natural abundance ratio of ap- 326 proximately 89.9, meaning that the C is being enriched and thus must be at least partially coming from the ion beam. Assuming that all the 13C originates from the background pressure in the vacuum and using the C natural abundance ratio, it is determined that the atomic concentration of 12C from the vacuum is approxi- mately 1.3 ×1019 cm−3, and the atomic concentration of 12C from the ion beam is approximately 4.8 ×1019 cm−3 in this sample. In order to produce 28Si film with higher purities, several experimental im- provements were made to the system, including the aforementioned switch to using Ar to purge the gas manifold instead of N2. First, the Al deposition source was relocated to another part of the chamber and away from the manipulator where the sample sits during deposition at DC–3. Next, some in-vacuum components that outgas F, specifically PTFE-coated wires on the manipulator and the STM tip preparation tool, were removed from the system. These wires were replaced with Kapton-coated wires. Then, the background pressure in both the deposition chamber and the ion beam chamber were reduced to minimize contaminants ad- sorbing from the vacuum during deposition. In the ion beam chamber, a gate valve that was not rated for UHV was removed from the system. This resulted in a reduction in the base pressure of the ion beam chamber from approximately 1.3 ×10−5 Pa (9.4 ×10−8 Torr) when the previously analyzed 28Si sample was de- posited (Fig. 5.41) to approximately 3.9 ×10−6 Pa (2.9 ×10−8 Torr) when the first 28Si sample was deposited after these changes were made. The base pressure in the deposition chamber was reduced by installing a new TSP to add more pumping capacity. Also, it was baked more thoroughly than what was previously done and 327 at higher temperatures of 150 ◦C to 200 ◦C. These factors resulted in a reduction in the base pressure from approximately 2.3 ×10−8 Pa (1.7 ×10−10 Torr) for the previously analyzed sample to 8.3 ×10−9 Pa (6.2 ×10−11 Torr) for the next sample deposited after these changes. Finally, for samples deposited after these experimen- tal improvements, the high pressure plasma mode of the ion source was used. This mode produces higher 28Si ion fluxes which resulted in increased growth rates of samples. Depositing with higher growth rates should reduce the concentration of contaminants adsorbed from the vacuum. Also, the higher cracking efficiency of the high pressure mode may result in cracking of some N2 molecules in the ion source, which would eliminate them from the 28 u ion beam. After enacting these changes, another 28Si sample was deposited and analyzed for chemical contaminants by SIMS. This sample was deposited with a substrate temperature of approximately 460 ◦C at DC–3 using the high pressure mode of the ion source. SIMS depth profiles of the atomic concentrations of N, C, O, F, Cl, Al, and Mo (lines) in this 28Si sample are shown vs. sputter depth in Fig. 5.47. Mo was analyzed in this sample to check for signs of contamination due to the Mo sample holder that contacts the substrate. Also shown are depth profiles of 28Si (circles and line), 29Si (squares and line), and 30Si (triangles and line), which correspond to the right axis in arbitrary units related to the count rate that are roughly aligned to the atomic concentrations on the left axis, as an indicator of the 28Si film. At a depth of approximately 293 nm, the concentrations of the Si isotopes increase and return to their natural abundance values in the Si(100) substrate, indicated by the shaded region, and giving a value for the film thickness. The minimum detected 29Si and 328 Figure 5.47: SIMS depth profiles of the concentration of contaminants in a 28Si film deposited at 460 ◦C at DC–3 using the high pressure mode of the ion source. Atomic concentrations of 28Si (circles and line), 29Si (squares and line), and 30Si (triangles and line) are also shown vs. sputter depth as an indicator of the 28Si film. The Si concentrations correspond to the right axis, displayed in arbitrary units. The minimum detected 29Si and 30Si concentrations are limited by the measurement noise floor. At a depth of about 292 nm they return to their natural abundance values in the Si(100) substrate (shaded region), indicating the film thickness. Depth profiles of the atomic concentrations of N, C, O, F, Cl, Al, and Mo (lines) are shown vs. sputter depth. N, C, O, and Cl were detected in the film. F and Al were not detected and their signals are at the measurement detection limit. The concentration of N in the film is 9.07(4) ×1018 cm−3, the concentration of C in the film is 8.27(3) ×1018 cm−3, and the concentration of O in the film is 3.09(4) ×1018 cm−3. The signals for these elements drop to the measurement detection limit in the substrate. It is unclear why Cl is detected at two different concentrations in the film. 30Si concentrations are taken as bounds on the enrichment because they are limited by the measurement noise floor, which for 29Si is approximately 20 ppm. In this 28Si sample, N, C, O, and Cl are all detected in the film, while F and Al are not detected. The signals for F and Al are at the detection limit of the measurement in the film and substrate, which is approximately 1 ×1017 cm−3 for F 329 and 3 ×1015 cm−3 for Al. The elimination of Al in this sample compared to the pre- vious sample is likely due to moving the Al deposition source away from the sample location. The elimination of F in this sample compared to the previous sample may be due in part to the removal of PTFE-coated wires in the chamber, although the partial pressure of F in the chamber was still present after removing them. It was unclear if the concentration of F in the vacuum was reduced because a different RGA was used to measure residual gases after these experimental changes were made, and it is difficult to compare the absolute values of the partial pressures to those of the previous instrument. It is unclear why Cl is detected at two different concentra- tions within the film. From the surface down to a depth of approximately 100 nm, the average atomic concentration of Cl is 4.11(4) ×1017 cm−3, or 8.24(8) ppm. The atomic concentration of Cl then drops to 8.6(3) ×1015 cm−3, or 0.172(6) ppm in the remainder of the film. This indicates that something changed roughly midway through the deposition. It may be that part of the sample holder started heating slowly during the deposition which caused increased outgassing including a Cl com- pound, as mentioned previously. Throughout most of the film the Mo signal is also at the detection limit, although there may be a slight increase between a depth of 50 nm and 100 nm to an atomic concentration of 36(8) ppb. The atomic concentrations of N, C, and O in this 28Si sample are clearly re- duced from those of the previous sample shown in Fig. 5.41. Note that the scale of the vertical axis of that figure and Fig. 5.47 are the same to facilitate easier comparisons. The average atomic concentration of N in the film was measured to be 9.07(4) ×1018 cm−3, or 181.8(8) ppm. N was reduced compared to the previ- 330 ous sample by a factor of roughly 78. The average atomic concentration of C in the film was measured to be 8.27(3) ×1018 cm−3, or 165.7(6) ppm. C was reduced compared to the previous sample by a factor of roughly 5. Finally, the average atomic concentration of O in the film was measured to be 3.09(4) ×1018 cm−3, or 61.9(8) ppm, which is reduced by almost a factor of 10 compared to the previous sample. The total purity for this 28Si sample determined from these measurements is approximately 99.96(2) %. The atomic concentrations of N and C in this sample are above their solid solubility limits, similar to the previous sample. Additionally, the atomic concentration of O in this sample is also slightly over its solid solubility limit in Si. This indicates that Si3N4, SiC, and SiO are all likely present in the 28Si film. The reduction of N in the second sample compared to the first seems likely due to changes in the ion beam. This is because the partial pressure of N2 in the deposition chamber during deposition appears to be similar for both of the sam- ples, although it is difficult to determine precisely because of the overlap of N2 and 28Si at 28 u in the residual gas mass spectrum of the RGA as well as other uncer- tainties. The deposition rate of the final sample deposited using the high pressure mode (2.21 nm/min) was roughly six times higher than that of the first sample (0.37 nm/min), which could account for some of the decrease in the N concentra- tion if the N originated from the background vacuum. However, the reduction in the N concentration is still more than an order of magnitude lower than what would be expected just due to the increased deposition rate. So, the remaining reduction may be due to the ion beam. Use of the high pressure mode of the ion source for 331 the final sample may lead to a lower concentration of N in the 28 u beam relative to 28Si. If the N2 ion flux for both the low pressure and high pressure modes is similar, then the roughly factor of five increase in 28Si beam flux (ion current) for the final sample compared to the first would result in a lower concentration of N relative to 28Si. Additionally, the possibility of N2 molecules being cracked more efficiently by the high pressure mode into atomic N and thus eliminated from the 28 u beam would also reduce the relative N concentration. The results from several samples do seem to support this hypothesis showing that the atomic concentration of N measured in the samples is inversely related to the 28Si ion beam current, although only one of those samples was deposited using the high pressure mode. This inverse relationship can be seen in Fig. 5.48 showing the atomic concentration of N as well as C and O vs. the total 28 u ion current used for deposition. The data with ion currents around 0.5 µA correspond to the first two samples deposited using the low pressure mode discussed above, while the data with an ion current of roughly 2.8 µA corresponds to the final sample discussed in this section, which was deposited using the high pressure mode. The uncertainties in the atomic concentrations (most are smaller than the data symbols) are the standard deviations of the means, and an uncertainty of 50 nA was assigned to the values of the ion current. The atomic concentration of N (open squares) appears to have a strong inverse relationship to the ion current, decreasing roughly a factor of 80 while the 28 u ion beam current increases roughly a factor of five. This trend suggests that the higher concentrations of N measured in the first samples are likely due to ballistic incorporation from the ion beam. This also indicates that both increasing the 28Si ion current and use of 332 Figure 5.48: Atomic concentrations of N (open squares), C (open circles), and O (open triangles) for three 28Si samples vs. the ion current used for deposition of the samples. The data with ion currents around 0.5 µA correspond to the samples deposited using the low pressure mode, while the data with an ion current of 2.8 µA corresponds to the sample deposited using the high pressure mode. The concentra- tions of N, C, and O all show an inverse relationship to the ion current, but N varies much more strongly than C and O, indicating that N contamination is more likely due to N2 in the ion beam. the high pressure mode do reduce the relative N2 concentration in the beam. The atomic concentrations of C (open circles) and O (open triangles) are also inversely related to the ion current, but more weakly so than N. The decrease in the concen- trations of C and O are likely mostly due to the increased deposition rates resulting from the increased ion current, similar to the result discussed above for the three measurements of the second analyzed sample. The remaining N, C, and O in the final sample are difficult to attribute to either adsorption from the vacuum or deposition from the ion beam, and their presence is likely due to a combination of both. Either way, the solution to reducing 333 them further is a reduction of the partial pressures of these elements in the vacuum because the contaminants in the ion beam ultimately originate from the vacuum in the ion source. There also seems to be a significant amount of N2 or CO entering the chamber with the SiH4 gas from the gas manifold, probably more than what is due to the base pressures of the two chambers. Reducing N, C, and O concentration in the deposition chamber, ion beam chamber, and the gas manifold is required to reliably improve the purity of 28Si films deposited with this system in order to achieve the third materials goal discussed here. 5.8 Crystallinity: Film Inspection via TEM The 28Si samples deposited at elevated temperatures at sample location DC–3 exhibit different morphologies depending on the deposition temperature, as dis- cussed previously in this chapter. Rough films produced with deposition tempera- tures above 600 ◦C and smooth films produced with deposition temperatures below 600 ◦C both appear crystalline and epitaxially aligned to the underlying substrate based on observations with RHEED and STM. While defects at the surface of a depositing film will both cause and develop from step pinning and step bunching, it is not clear what crystalline defects are formed in the bulk of the films for the two deposition temperature ranges. Here, TEM is used to image and characterize crystalline defects in 28Si films. TEM micrographs presented here were acquired in collaboration with Dr. Alline Myers and Dr. Vladimir Oleshko. TEM is a common analysis tool used for inspecting the epitaxial quality of 334 and defects in films produced by low temperature Si deposition including MBE and IBE. A common type of defect observed in Si films deposited on Si(100) substrates is {111} stacking faults. As seen in the TEM micrographs of a Si film deposited by IAD in Fig. 5.17, {111} stacking faults and defects start building up in the epitaxial layer until an amorphous phase develops. IBE experiments have demonstrated epitaxial 28Si films and used TEM to show, depending on the deposition conditions, both defective films with {111} stacking faults and microtwins, as well as higher quality epitaxial films without visible stacking faults or obvious dislocations. However, TEM inspection shows that these films likely still contain defects and strain [43,51]. Chemical contaminants can not only affect the morphology of a depositing film through surface defect formation, they can also cause structural defects and amorphization in the bulk of a film when incorporated during deposition. Several 28Si IBE experiments have been done showing the effects of chemical contaminants introduced through the ion beam during deposition on the epitaxial quality of the film. One experiment found that introducing approximately 1 % N2 into the 28Si ion beam during deposition caused the resulting film to be amorphous at a deposition temperature of 350 ◦C [169]. Others found that N2 and CO in the ion beam resulted in highly defective or amorphous films using a range of deposition temperatures [49, 50]. Finally, an experiment found that using 30Si ions to deposit a film resulted in epitaxial growth that was less defective than 28Si, and concluded that trace amounts of CO present in the 28Si beam, which are very difficult to eliminate, were the cause [134]. These experiments show the importance of reducing the concentration of N2 and CO in the ion beam, as discussed in the previous section. 335 The lattice constant mismatch between natural abundance Si and 28Si may also introduce dislocations into the film, which may be observed by TEM. As mentioned previously, however, the effect of the roughly 1× 10−6 relative difference in lattice constants is likely too small compared to other defect causing mechanisms in these films. Initially, TEM was used to inspect the bulk crystallinity of rough 28Si films deposited at higher substrate temperatures above 600 ◦C to determine the effect on the crystallinity of the surface roughness. Additionally, it was used to confirm the observations made using RHEED and STM that, despite the rough surface, the films are still crystalline and epitaxially aligned to the Si(100) substrate. TEM cross- sectional micrographs of a rough 28Si sample deposited with a substrate temperature of approximately 708 ◦C at DC–3 are shown in Fig. 5.49. The substrate used for this sample was not cleaned ex situ and was flashed annealed before deposition. The TEM specimen was prepared using a FIB, and these images were taken on the 〈110〉 zone axis. Panel (a) is a bright field image showing the 28Si film above the Si(100) substrate at the bottom of the micrograph. A protective, thin layer of C (light) and a thicker layer of Pt (dark) are seen above the film. The 28Si film consists of large mounds resulting in a very rough surface, as previously observed in the SEM micrographs of this and other rough samples (Fig. 5.22). The maximum film thickness (of the central mound) seen in this micrograph is approximately 116 nm and the minimum thickness is approximately 31 nm. The surface of the mounds comprising the film are faceted with predominately {113} and {111} microfacets visible, indicated by the arrows. The surfaces of the {113} microfacets make an 336 Figure 5.49: TEM cross-sectional micrographs of a rough 28Si sample deposited at 708 ◦C at DC–3. This image was taken on the 〈110〉 zone axis. (a) Bright field image with the Si(100) substrate is visible at the bottom of the image and the 28Si film above it, which consists of large mounds. Protective layers of C (light) and Pt (dark) were deposited on the 28Si film. The surface of the mounds are faceted with {113} and {111} microfacets visible (arrows). Within the mounds, {111} stacking faults are seen. (b) HR-TEM image of a group of {111} stacking faults and microtwins that run from the interface of the film and the substrate up through the film. The 28Si film is seen to be crystalline and epitaxially aligned to the substrate, evidenced by the continuation of 〈111〉 lattice rows across the interface. 337 angle with the surface of the substrate, or the substrate interface that matches the expected angle between the 〈113〉 and 〈100〉 planes of approximately 25.2◦. Likewise, surface of the {111} microfacets make an angle with the surface of the substrate that matches the expected angle between the 〈111〉 and 〈100〉 planes of approximately 54.7◦. These microfacets match those that were observed in the RHEED diffraction patterns for similar samples shown in Fig. 5.20. Also, within the mounds, multiple stacking faults are visible running through the film along the 〈111〉 planes. Panel (b) of Fig. 5.49 is an HR-TEM image that shows another region of the same film in (a) at a higher magnification. The substrate is seen at the bottom of the micrograph with the 28Si film above it. The 28Si film is crystalline and epitaxially aligned to the substrate, as evidenced by the continuation of 〈111〉 lattice rows across the substrate interface into the film. A group of {111} stacking faults and microtwins are seen originating at the substrate interface and running through the film. The presence of microtwins is evidenced by the dark fringes inside the stacking fault appearing with a periodicity of three times the normal 〈111〉 lattice row spacing, as is often observed for microtwins [173]. These are possibly due to defects or contaminants such as SiC present on the surface at the beginning of the growth that cause step bunching and defects to form on {111} planes. Some areas of this sample appear with fewer stacking fault defects, and the crystallinity can be inspected further and compared with that of the substrate using fast fourier transform (FFT) analysis. Figure 5.50 shows an HR-TEM cross-sectional micrograph of another area of the 708 ◦C sample analyzed using FFTs. Like with the previous micrographs of this sample, this image was taken on the 〈110〉 zone 338 Figure 5.50: HR-TEM cross-sectional micrograph of a rough 28Si sample deposited at 708 ◦C at DC–3. This image was taken on the 〈110〉 zone axis. (a) The Si(100) substrate is at the bottom of the image, and the 28Si film is above it. Several {111} stacking faults are visible running through the film. The 28Si film is seen to be crystalline and epitaxially aligned to the substrate, evidenced by the continuation of 〈111〉 lattice rows across the interface. (b) and (c) FFTs of the regions in the boxes for the 28Si film and the substrate, respectively. The FFT of the film and substrate show the same crystal pattern indicating that the film is aligned to the substrate. 339 axis. Panel (a) shows the sample with the Si(100) substrate appearing dark at the bottom of the micrograph and the 28Si film above it. Again, film appears epitaxially aligned to the substrate, as evidenced by the continuation of 〈111〉 lattice rows across the substrate interface into the film, indicated by the arrow. A few {111} stacking faults are seen in the film as well. Panels (b) and (c) are FFTs of the regions marked by the boxes in the substrate and 28Si film, respectively. These FFTs appear nearly identical showing the sample crystal pattern and orientation, indicating again that the film is epitaxially aligned to the substrate. 28Si samples made with lower deposition temperatures that were below 600 ◦C were observed to be much smoother than those with high deposition temperatures, and TEM was used to inspect these low deposition temperature samples for differ- ences in the crystallinity compared to the samples with higher deposition temper- atures. Samples with lower deposition temperatures were also prepared using the revised cleaning procedures discussed previously in the chapter, which may affect the bulk crystalline defects that develop in the film. TEM cross-sectional micro- graphs at two magnifications of a smooth 28Si sample deposited with a substrate temperature of approximately 460 ◦C at DC–3 are shown in Fig. 5.51. This sample was prepared ex situ using the CMOS cleaning procedure and was flashed annealed before being deposited using the low pressure mode of the ion source. These images were taken on the 〈110〉 zone axis. Panel (a) shows a bright field image at lower magnification with the Si(100) substrate in the lower left of the image appearing lighter and the 28Si film to the right of that. The 28Si appears quite different from the substrate with varying contrast and dark patches throughout the film. These 340 Figure 5.51: TEM cross-sectional micrographs of a smooth 28Si sample deposited at 460 ◦C at DC–3 using the low pressure mode. These images were taken on the 〈110〉 zone axis. (a) Bright field image showing the Si(100) substrate in the lower left appearing lighter and the 28Si film to the right of that with varied contrast. To the upper right of the 28Si film are protective C (bright) and Pt (dark) layers. The 28Si film thickness varies between about 105 nm and 110 nm here. (b) HR-TEM image showing the substrate in the lower left and the 28Si film in the upper right. The 28Si film is seen to be crystalline and epitaxially aligned to the substrate, evidenced by the continuation of 〈111〉 lattice rows across the interface. are likely defects causing local strain in the film. To the upper right of the 28Si are layers of C (light) and Pt (dark) which were deposited to protect the sample during preparation of the specimen, which was done using a FIB. The 28Si film thickness varies between approximately 105 nm and 110 nm in this region of the film, meaning there is a roughly 5 nm surface width here. Panel (b) shows an HR-TEM image taken at much higher magnification (790 times). The crystalline Si(100) substrate is seen in the lower left of the image and the 28Si film is in the upper right. The dashed line representing the interface is only approximate because the true interface is not clear at this magnification, and the 28Si appears similar to the substrate, un- 341 like in panel (a). The 28Si film in panel (b) is seen to be crystalline and epitaxially aligned to the substrate, evidenced by the continuation of individual 〈111〉 lattice rows throughout the image. Another HR-TEM cross-sectional micrograph of this 28Si sample from Fig. 5.51 is shown in Fig. 5.52 and the crystallinity is analyzed using FFTs. This image was taken on the 〈110〉 zone axis. Panel (a) shows the Si(100) substrate in the lower left of the micrograph with the 28Si film in the upper right. The interface between the substrate and the film is indicated by the dashed line and is less clear than in the TEM micrographs of the previous sample. The film in this sample is crystalline and epitaxially aligned to the substrate, which is evidenced by the continuation of 〈111〉 lattice rows across the interface, indicated by the arrow. No obvious stacking faults are visible in the film in this micrograph or in any other areas of this film. Several dark areas appear in the film but not the substrate, probably indicating the presence of dislocation and other defects causing local strain fields in the film. Also, the lattice rows in some areas of the film are not as clear as others or those of the substrate, but these areas do not appear amorphous. Panels (b) and (c) are FFTs of the regions marked by the boxes in the substrate and 28Si film, respectively. These FFTs appear very similar showing the sample crystal pattern and orientation, indicating that the film is, again, epitaxially aligned to the substrate. This same 460 ◦C sample was analyzed by SIMS for chemical contaminants, which was shown in Fig. 5.41, and found to have an atomic concentration of N of approximately 7.1(1) ×1020 cm−3, or 1.42(2) %. It is thus surprising that the 28Si film is not more defective given that it likely contains Si3N4 and SiC crystallites and 342 Figure 5.52: HR-TEM cross-sectional micrograph of a smooth 28Si sample deposited at 460 ◦C at DC–3 using the low pressure mode. This image was taken on the 〈110〉 zone axis. (a) The Si(100) substrate is seen in the lower left and the 28Si film in the upper right. The dashed line indicates the interface. The 28Si film is seen to be crystalline and epitaxially aligned to the substrate, evidenced by the continuation of 〈111〉 lattice rows across the interface. Several dark patches appear in the film but not the substrate. (b) and (c) FFTs of the regions in the boxes for the substrate and the 28Si film, respectively. The FFT of the film and substrate show the same crystal pattern indicating that the film is aligned to the substrate. 343 that, as mentioned previously, 1 % N contamination can lead to an amorphous Si film, although for a lower deposition temperature. The dark areas and other contrast changes observed in the 28Si film may be related the Si3N4 and SiC compounds present in the film and their related structural defects. A reduction in the amount of contaminants was then achieved in the next 460 ◦C sample analyzed in Fig. 5.47 due to having lower background pressures in the chambers and use of the high pressure mode of the ion source. A subsequent sample deposited after these experimental changes and after analyzing the second 460 ◦C sample was then inspected using TEM. This was to determine if the reduction in contaminants, particularly N and thus Si3N4 in the film, resulted in a reduction of the dark patches in the film, seen in the TEM micrograph in Fig. 5.52. An HR- TEM micrograph of this later 28Si sample deposited with a substrate temperature of approximately 421 ◦C at DC–3 is shown in Fig. 5.53. This sample was prepared ex situ using the CMOS cleaning procedure and was flashed annealed before being deposited using the high pressure mode of the ion source. The TEM specimen was prepared using a FIB, and this image was taken on the 〈110〉 zone axis. Panel (a) shows the Si(100) substrate at the bottom of the micrograph and the 28Si film at the top. The interface between the substrate and the film is roughly indicated by the dashed line. Like the previous sample, the film in this sample is crystalline and epitaxially aligned to the substrate, which is evidenced by the continuation of 〈111〉 lattice rows across the interface, indicated by the arrow. However, despite the presumed reduction in chemical contaminants in this sample compared to that shown in the previous TEM micrograph, this film appears much more defective. 344 Figure 5.53: HR-TEM cross-sectional micrograph of a smooth 28Si sample deposited at 421 ◦C at DC–3 using the high pressure mode. This image was taken on the 〈110〉 zone axis. (a) The Si(100) substrate is seen at the bottom of the image and the 28Si film is above it. The nearly horizontal dashed line roughly indicates the interface. The 28Si film is seen to be crystalline and epitaxially aligned to the substrate, evidenced by the continuation of 〈111〉 lattice rows across the interface. {111} stacking faults are seen running through the film along with several dark patches. (b) and (c) FFTs of the regions in the boxes for the substrate and the 28Si film, respectively. The FFT of the film and substrate show the same crystal pattern indicating that the film is aligned to the substrate. 345 Several {111} stacking faults are seen running through the film in this micrograph, and further inspection of different areas of the film show that there are many more stacking faults and other dislocations throughout the film. Dark patches are also seen in this film, likely due to strain from defects. Panels (b) and (c) are FFTs of the regions marked by the boxes in the substrate and 28Si film, respectively. These FFTs appear very similar showing the sample crystal pattern and orientation, indicating that the film is, again, epitaxially aligned to the substrate. The more defective structure of the 421 ◦C sample compared to the previous 460 ◦C sample may be due to the deposition rate increasing from approximately 0.37 nm/min for the 460 ◦C sample to approximately 4.56 nm/min for the later 421 ◦C sample due to use of the high pressure deposition mode. These two factors may move the quality of the film growth closer to the defective and strained region of the epitaxy phase diagram for IBE, although there is no evidence of an amorphous layer developing in this sample. In fact, none of the 28Si samples inspected with TEM showed signs of a critical thickness, hepi, or an amorphous phase developing, and they were always observed to be crystalline throughout the film including up to the top surface. While the 28Si films inspected by TEM were crystalline and epitaxially aligned to the substrate, the density of crystalline defects was too high to meet the second materials goal stated at the beginning of this chapter. A reduction in crystalline defects will likely require a reduction in chemical contaminants in these film. 346 5.9 Chapter 5 Summary The 28Si samples discussed in this chapter showed that 28Si films could be deposited in the deposition chamber at sample location DC–3 while maintaining residual 29Si and 30Si isotope fractions well below 1 ppm. Measurements also shows that extremely high enrichments are achievable for samples deposited both with elevated substrate temperatures and while using the high pressure plasma mode of the ion source. In total, 40 28Si samples were produced at DC–3 with 39 of them being deposited with elevated substrate temperatures. Additionally, seven of those samples were deposited using the high pressure mode of the ion source. The residual 29Si isotope fraction of the most highly enriched sample deposited at DC–3 was reduced by more than a factor of five compared to that of the previous most highly enriched sample deposited at LC–2, going from 0.691(74) ppm to 127(29) ppb. This progression can be seen in the enrichment progression timeline in Fig. 4.2. The isotope reduction factor of 29Si for the most highly enriched sample de- posited at DC–3 is 3.7(8)× 105. This value of the reduction factor along with the values from all the other most highly enriched 28Si samples deposited at IC–1, LC–2, and DC–3 are shown in an isotope reduction timeline, which is a progression of the Si isotope reduction factors az/( zSi/Sitot.), in Fig. 5.54. This timeline is a modified version of the enrichment progression timeline, showing the isotope reduction fac- tors vs. deposition date, where a larger reduction factor means a higher enrichment, instead. Nine 28Si samples out of a total of 61 produced in this work are represented on this timeline. As with the enrichment progression timeline, the nine samples 347 Figure 5.54: Isotope reduction timeline. A timeline of the progression of the isotope reduction factors of the lowest residual isotope fractions of 29Si (squares) and 30Si (triangles), as measured by SIMS. These were achieved for 28Si samples deposited over approximately three and one half years. These results encompass samples produced at IC–1, LC–2, and DC–3. Shown for comparison are the 29Si isotope reduction factors of the 28Si epilayers and crystals produced by Isonics and Itoh using 28SiH4 CVD (dash-double dotted line) from Ref. [36,37], the bulk 28Si material produced by the IAC using 28SiH4 CVD (dash-dotted line) from Ref. [32], and the 28Si thin films produced by Tsubouchi et al. using 28Si IBE (dotted line) from Ref. [43]. presented are those that achieved the best enrichment, or reduction factors of the minor isotopes, of any sample deposited up to that point. Both the reduction factor for 29Si (squares) and 30Si (triangles) are shown to increase in different samples over time from 16.6(1) for the initial sample deposited at LC–1 to the afore mentioned 3.7(8)× 105 for the most highly enriched sample produced at DC–3. Uncertainties in the reduction factors are derived from the uncertainties in the SIMS measurements 348 of the isotope fractions. Also shown for comparison are the 29Si isotope reduction factors of three other sources of 28Si. The 28Si epilayers and crystals produced by Isonics and Itoh using 28SiH4 CVD have a 29Si reduction factor of approximately 64 (dash-double dotted line), which is larger than only the first two 28Si samples produced here at IC–1 [36, 37]. The bulk 28Si material produced by the IAC using 28SiH4 CVD has a 29Si reduction factor of 937 (dash-dotted line) [32], and the 28Si thin films produced by Tsubouchi et al. using 28Si IBE has a 29Si reduction factor of approximately 2.9 ×103 (dotted line) [43]. Both of these values are still below the most highly enriched 28Si sample deposited at IC–1. The overall 28Si isotope fraction of this most highly enriched sample deposited at DC–3 and in this entire work was 99.9999819(35) %. These sample are more highly enriched than any other known source of 28Si, including the IAC. These samples also demonstrate enrichments sufficient to enable a robust measurement of the dependance of electron coherence time on 29Si concentration in the single spin regime and compare it to theoretical predictions (see Fig. 1.9), as proposed in Chapter 1 [12]. The achieved reduction in 29Si and 30Si isotope fractions in samples deposited at DC–3 was likely due to several factors. The deposition chamber had significantly lower background pressures than the lens chamber at LC–2 resulting in less SiH4 adsorption during deposition. Also, higher deposition rates were generally achieved for samples deposited at DC–3, including the highest rates achieved using the high pressure mode of the ion source. Depositing samples at DC–3 allowed for sample heating which was crucial in 349 achieving epitaxial deposition. Crystalline, epitaxial 28Si films were produced using elevated deposition temperatures between 349 ◦C and 1041 ◦C, although samples deposited above 600 ◦C were very rough. This was due to chemical contaminants such as SiC at the growth surface that result in step pinning sites and lead to step bunching, faceting on {111} and {113} planes, and large mound formation. 28Si samples deposited at lower temperatures between approximately 349 ◦C and 460 ◦C were found to be smooth with typical surface widths of ∆z = 2 nm. Chemical contaminants in these 28Si films were measured by SIMS, which detected N, C, O, F, Al, and Cl. The N, C, and O were detected at especially high concentrations, initially all above 1× 1019 cm−3. F and Al were eliminated from a second sample due to experimental alterations. By improving the vacuum in both the deposition chamber and the ion beam chamber, and by using a higher 28Si ion beam current generated in the high pressure mode of the ion source, N, C, and O were all able to be reduced in the final sample analyzed by SIMS. The average atomic concentration of N in the film was measured to be 9.07(4) ×1018 cm−3, or 181.8(8) ppm, the average atomic concentration of C in the film was measured to be 8.27(3) ×1018 cm−3, or 165.7(6) ppm, and the average atomic concentration of O in the film was measured to be 3.09(4) ×1018 cm−3, or 61.9(8) ppm. Based on these concentrations and the solid solubility of these elements in Si, it is likely that Si3N4, SiC, and SiO exist within the 28Si films. The resulting total best purity for a 28Si sample deposited at DC–3 and overall in this work is approximately 99.96(2) %. Finally, TEM was used to confirm that 29Si films produced with both the higher and lower deposition temperatures were crystalline and epitaxially aligned to the 350 substrates. All samples deposited with substrate temperatures above 600 ◦C were observed to have {111} stacking faults and twinning present in the films. For lower deposition temperature samples, one sample deposited using the low pressure mode did not have any visible stacking faults but probably still contained other crystalline defects. Another sample deposited with a slightly lower substrate temperature and using the high pressure mode to generate a much higher deposition rate was observed to have a lot of {111} stacking faults and likely other defects. Reduction of these crystalline defects likely requires reduction of the chemical contaminants within the 28Si films. Overall, samples deposited at DC–3 enabled improvements and new understanding regarding the second and third materials goals mentioned at the beginning of this chapter in support of the broader 28Si effort, of which this work is a part. 351 Chapter 6 Pressure and Temperature Dependent Adsorption of 29Si and 30Si During 28Si Deposition 6.1 Introduction It was shown in Chapters 4 and 5 that an extremely high level of enrich- ment was achieved for 28Si films deposited both amorphously at room temperature and epitaxially at elevated temperatures. However, despite the fact that the 28Si ion beam is well resolved and separated from the 29Si and 30Si ions, the residual 29Si and 30Si isotope fractions in these samples were not zero. Understanding this discrepancy and how the concentration of isotopic contaminants are affected by dif- ferent deposition parameters, such as substrate temperature, is necessary for the further development of 28Si ion beam deposition. The objectives of the experiments and analysis discussed in this chapter are as follows: (1) understand the source of residual 29Si and 30Si in the 28Si films, (2) determine the dependance of the residual isotope fractions on deposition tem- 352 perature, and (3) understand the mechanism by which the residual isotope fractions depend on temperature. These objectives are part of the larger goals of this work set forth in Chapter 1, which is to be able to produce 28Si samples with targeted levels of enrichment (29Si isotope fractions). Targeting specific enrichments would facilitate a study of spin coherence time as a function of 29Si concentration. Electron and nuclear T2 times of single implanted 31P measured for a range of 29Si concentrations could be compared to theoretical predictions (see Fig. 1.9), as mentioned in Chapter 1 [12]. The experiments described in this chapter were designed to test the hypothesis that the source of 29Si and 30Si, measured in the samples by SIMS, is the natural abundance SiH4 gas which diffuses from the ion source to the sample location during deposition. This diffusion results in a partial pressure of SiH4 at the surface of the 28Si sample. The SiH4 molecules (some of which are 29SiH4 and 30SiH4) will stick to the Si surface where they can be incorporated either through physisorption or in a chemisorption reaction similar to that of CVD. The sticking and growth behavior of SiH4 in Si CVD processes, described by the so-called reactive sticking coefficient, has been studied extensively. The literature on this subject, however, is quite large and diverse and results are often difficult to compare or reconcile because they depend heavily on experimental conditions such as pressure, temperature, surface condition, specific SiH4 species, and various systematic experimental uncertainties. Also, the reaction describing the conversion of gaseous SiH4 to solid incorporated Si atoms is more complex than one might naively guess because it can occur through multiple 353 decomposition channels. Several review articles by Comfort and Reif [174], Jasinski and Gates [137], and Onischuk and Panfilov [175] have given summaries of both reaction mechanisms of SiH4 CVD processes and experimental work measuring re- active sticking coefficients and the associated activation energies for these processes. The dominant reaction expected for SiH4 CVD at pressures ≤ 0.1 Pa (as is the case in this work) is described by Jasinski and Gates to be SiH4(g)→ Si(s) + 2H2(g), (6.1) where (g) represents the gaseous phase and (s) represents the solid phase [137]. This reaction is exothermic producing about 8.2 kcal/mol (0.36 eV), but energy in the form of heat is required to overcome the kinetic barriers to the decomposition, which is the activation energy. Figure 6.1 illustrates the reaction sequence and the role of dangling bond sites ( ) in CVD reactions on the Si(100) surface. SiH4 initially adsorbs on the surface at a double dangling bond site. The SiH4 then decomposes into SiH3 on one dangling bond site and H on the other. By encountering further dangling bond sites on the surface, the SiH3 decomposes further until a Si atom is left along with four H atoms, which recombine into two H2 molecules and desorb. The lone Si atom then becomes incorporated into the film. The desorption of H2 frees three dangling bond sites that are then cycled back into the reaction sequence for the decomposition of other SiH4 molecules. There are several differences between SiH4 based CVD and the sticking and/or reaction of SiH4 being incorporated into the 28Si films discussed here. These dif- ferences offer several advantages for this work over the typical experiments in the 354 Figure 6.1: SiH4 surface decomposition sequence during CVD growth for low pres- sures. The SiH4 interacts with dangling bond sites ( ) on the Si(100) surface to sequentially dissociate H which can then evaporate. (from Ref. [137]) literature, however these differences can make comparisons between the literature and this work more difficult. Unlike the data presented here for samples deposited with substrate temperatures ranging from room temperature up to 800 ◦C, CVD is not typically studied with a growth temperature below 500 ◦C to 600 ◦C because the growth rate drops dramatically below this range, depending on specific experimen- tal conditions. This is partially due to increased H coverage at lower temperatures which inhibits the CVD reaction. CVD typically uses high pressures of H2 with SiH4 partial pressures as high as 130 Pa (975 mTorr) for UHV CVD, although some experiments have used pressures as low as 1.3× 10−5 Pa (9.8× 10−8 Torr) [176]. These pressures are orders of magnitude higher than what is used for the samples described in this chapter, and so H coverage is not believed to significantly influ- ence the results reported here. The other major difference between CVD and the 355 ion beam deposited samples, is the 28Si ion beam itself, which could interact with adsorbed SiH4. The deposition rate due to the ion beam is typically much larger, i.e. > 10 times larger, than the growth rate due to the adsorbed species, therefore a SiH4 molecule will (almost) always encounter a bare Si surface. Consequently, rate limiting effects seen in CVD such as surface diffusion, SiH4 interaction with other adsorbates like H, and surface reaction rates would not occur [177, 178]. Energetic ions are also known to desorb H from Si(100) surfaces [121]. Finally, a unique fea- ture of monoisotopic ion beam deposition in this context is that very small numbers of adsorbed species can be measured. As seen in the analysis of the enrichment of 28Si samples in Chapters 4 and 5, SIMS is extremely sensitive to isotope ratios and so trace amounts of 29Si and 30Si adsorbed in the 28Si films can be easily detected. Section 6.2 of this chapter describes the experimental methods used to collect and analyze the relevant sample parameters discussed in the remainder of the chap- ter. In section 6.3, a model that describes the proposed SiH4 adsorption process is introduced, and in section 6.4, it is used to analyze the data. Sections 6.5 and 6.6 explore the role of substrate temperature on the enrichment and calculated SiH4 incorporation fractions, respectively. Section 6.7 uses the incorporation fractions to determine an activation energy for SiH4 adsorption. Section 6.8 provides a brief summary. Analysis of some of the data discussed in this chapter was previously published in Ref. [84]. 356 6.2 Experimental Methods 6.2.1 28Si Samples and Enrichment Values The general deposition parameters for 28Si samples similar to and including the samples used in the analysis of this chapter were described in detail Chapters 4 and 5. These samples were deposited on a variety of natural abundance Si(100) substrates including p-type, n-type, and undoped (intrinsic) wafers, discussed in Chapters 4 and 5. For most samples discussed here and deposited at room temperature, substrates were prepared ex situ by an HF etch to remove the native oxide and were not prepared further in vacuum. These samples were deposited at sample location LC– 2 in the lens chamber. One room temperature sample was deposited at DC–3 in the deposition chamber, and was not prepared ex situ and were loaded with a native oxide. For most of the samples deposited at elevated temperatures, substrates were cleaned ex situ using the standard CMOS cleaning procedure described in Chapter 5. Three samples deposited above 600 ◦C were not cleaned ex situ and were loaded in the vacuum chamber with a native oxide. Substrates were then prepared for deposition in situ by flash annealing them to 1200 ◦C for ≈ 10 s several times to produce a clean (2×1) reconstructed Si(100) surface on which to deposit 28Si epitaxially. These samples were all deposited at sample location DC–3 in the deposition chamber. As mentioned previously, the gas used in these experiments to generate a 28Si ion beam was natural abundance SiH4 with a purity of 99.999 % according to the gas vendor (Matheson Tri-Gas). To map out the temperature 357 dependance of the enrichment, samples were deposited with substrate temperatures including room temperature (≈ 21 ◦C), 249 ◦C, 349 ◦C, 357 ◦C, 421 ◦C, 502 ◦C, 610 ◦C, 705 ◦C, 708 ◦C, and 812 ◦C. 28Si ions were deposited onto the substrates with an average ion energy, Ei, at the sample of typically about 100 eV for the room temperature samples and ap- proximately 35 eV for the samples deposited at elevated temperatures. Typical 28Si ion beam currents, Ii, of around 500 nA were achieved over an area on the sub- strate between about 3 mm2 and 16 mm2. For one sample deposited at 421 ◦C, a higher ion beam current of approximately 3 µA was achieved. The resulting thick- nesses, d, of the deposited films were inferred from the calibration of the SIMS depth profiles and ranged from ≈ 50 nm to 350 nm. Dividing the thicknesses by the depo- sition time for each sample gives an estimate for the deposition rates, R, of around 0.32 nm/min to 3.94 nm/min. Based on these rates, the corresponding average ion flux,Fi, was then calculated for each sample. Fi varied from 2.70× 1013 cm−2 · s−1 to 3.4× 1014 cm−2 · s−1. For the samples discussed in the analysis of this chapter, the total pressure rise during deposition after subtracting the chamber base pressure ranged from approximately 9.9× 10−7 Pa to 4.9× 10−6 Pa (7.5× 10−9 Torr to 3.7× 10−8 Torr) for the samples deposited at room temperature, and it ranged from approximately 4.9× 10−7 Pa to 3.4× 10−6 Pa (3.7× 10−9 Torr to 2.5× 10−8 Torr) for the samples deposited with elevated substrate temperatures. These pressures equate to a total gas flux, F tot.g , on the surface of the sample during deposition. A more relevant flux in this analysis is the flux due to the SiH4 partial pressure, Fg, which will be 358 discussed in a later section. The 29Si and 30Si isotope fractions for these samples were measured by SIMS in collaboration with Dr. David Simons (NIST) as described in Chapter 4. Iso- tope fractions of a particular isotope of Si are defined in a SIMS measurement as the detected average counts of that isotope divided by the total average counts of the measurement and are written as zSi/Sitot. for an isotope with mass number z, as previously discussed in Chapter 4. Measurements were performed by Dr. Si- mons and the analysis presented here was done by myself. The raw measurements show that at the low end of the deposition temperature range, the 249 ◦C sample had a residual 29Si isotope fraction of 0.79(12)× 10−6 or 0.79(12) ppm. For the sample deposited at the highest temperature, 812 ◦C, the 29Si isotope fraction was 4.32(46) ppm. This increase in isotope fraction with increasing substrate tempera- ture is the focus of this discussion. The sample with the best enrichment and lowest 29Si isotope fractions in this study was deposited at 502 ◦C, as previously reported in Chapter 5 (see Fig. 5.11). The measured 28Si isotope fraction of the most highly enriched portion of this sample is 99.9999819(35) %, the average residual 29Si isotope fraction is 127(29) ppb, and the average residual 30Si isotope fraction is 55(19) ppb. The uncertainty of the isotope ratios was determined from the standard deviation of the mean of the measurements. A list of the samples discussed in this chapter, their deposition parameters, and measurement and analysis results can be found in Tables D.10 to D.13 in Appendix D. For the samples deposited at 705 ◦C, 708 ◦C, and 812 ◦C, the measured isotope fractions have to be taken as an upper bound. This is because during deposition, 359 the samples deposited above 600 ◦C developed a large amount of surface roughness on the order of the film thickness itself. As discussed in Chapter 5, this roughness had the effect of artificially inflating the isotope fractions measured by SIMS (see Fig. 5.12 and Fig. 5.14). The 610 ◦C sample is excluded from this caveat because the SIMS measurement of it was found to be more trustworthy. Nominally the SIMS depth profiles show a clear extended minimum in 29Si and 30Si isotope fractions through the thickness of the film, but instead, in these higher deposition temperature samples there is a gradual increase in isotope fractions up to the natural abundance values in the substrate. This effect is a measurement artifact caused by the SIMS sputter beam sampling the 28Si film and substrate at the sample time. The stated enrichment values for the effected samples are considered upper bounds because the measurement artifact would only increase the apparent isotope fractions but never decrease them. Care was taken to exclude data that was clearly influenced by this effect, however, it is possible that this artifact still played a small role in determining the isotope ratios of the highest temperature samples. 6.2.2 SiH4 Mass Spectrum and Mass Selectivity A key component of the analysis of SiH4 adsorption discussed in this chapter is the assumption that the ion beam is 100 % pure 28Si. This assumption can be justified in part by analysing the isolation of the 28Si ion beam as measured in the SiH4 mass spectrum that is collected during operation of the ion beam with SiH4 gas. A portion of a SiH4 mass spectrum, representative of the ion beam conditions used in Chapter 4 to deposit room temperature samples at sample location LC–2, 360 is presented in Fig. 6.2. Qualitatively similar mass spectrums were obtained for samples deposited at elevated temperatures, as can be seen in Fig. 5.4 in Chapter 5. In Fig. 6.2, the ion current peaks corresponding approximately to 28 u (28Si), 29 u (29Si and 28SiH), 30 u (predominately 28SiH2), and 31 u (predominately 28SiH3) are observed. Gaussian fits (Eq. (2.14)) to the 28 u (dashed line) and 29 u (solid line) peaks are shown superimposed on the data. The 95 % confidence bands of the two fits are also shown (dash-dotted lines). The applied current corresponding to the sweep of the magnetic field of the mass analyzer is shown on the top axis. This mass spectrum indicates a mass resolving power m ∆m ≈ 80 (measured at 10 % of the peak height). The Gaussian fits give a separation of the 28 u peak from the 29 u peak of about 11 σ (standard deviation). The Gaussian fits are used to determine the approximate geometric mass selectivity (i.e. the amount of mass separation) of the ion beam system to estimate the amount of 29Si potentially contaminating the 28Si beam. This is done by calculating the overlap of the 29 u and 28 u peaks using the parameters G28(m) and G29(m), which are the values (calculated at mass m in units of u) of the Gaussian fits to the current peaks at 28 u and 29 u respectively. The overlap of the 29 u peak on the 28 u peak is then determined from G29(28) G28(28) +G29(28) , (6.2) where m = 28 for the above parameters signifying that the values of the Gaussian fits are calculated at a mass of 28 u. The 95 % confidence band of the fit to the 28 u peak is used to calculate G28(m), and the 95 % upper confidence band of the fit to the 29 u peak is used to calculate G29(m). When also taking into account 361 Figure 6.2: SiH4 ion beam mass spectrum (circles) showing the 28Si ion current peak at 28 u, 29Si ion current at 29 u (≈ 5 % of the total current at 29 u), and two higher mass Si hydride peaks. The top axis shows the corresponding current applied to the magnetic sector mass analyzer to sweep the field. Gaussian fits (Eq. (2.14)) to the 28 u (dashed line) and 29 u (solid line) peaks along with 95 % confidence bands (dash-dotted lines) are plotted to calculate the overlap of the 29 u peak onto the 28 u peak. The centers of the 28 u and 29 u fits are separated by ≈ 11 σ. that the 29 u peak consists of approximately 5 % 29Si (95 % 28SiH), as discussed in Chapter 4, the 29Si contamination fraction of the 28 u ion current is calculated to be ≈ 5× 10−26. This extreme estimate of the contamination fraction of the 28Si peak from the 29Si peak is unphysical and is purely a measure of the geometric mass selectivity of the ion beam system. To get an estimate of the potential realized mass selectivity due to the 29Si peak overlap, a gas scattering mechanism is considered that would likely be a dominant contributing factor to the 29Si beam contamination. Inelastic scattering between ions 362 and gas molecules occurs along the flight path of the ions as they travel down the beamline. This causes an ion at mass 29 u to lose sufficient energy to be incorporated into the 28 u trajectory and pass through the mass-selecting aperture. This so-called scattering tail effect, referred to as the abundance selectivity when considering the tail contribution to an adjacent mass current peak, IS/I0, can be estimated from IS I0 ∝ P∆x ( m ∆m )n , (6.3) which was adapted from Ref. [179]. IS is the scattered ion current from a peak at mass m to a peak at mass m + ∆m, where ∆m is an integer. For the scattering contribution of an ion current peak to a current peak at an adjacent mass, ∆m = ±1. I0 is the total ion current at mass m, P is the background gas pressure in the beamline, ∆x is the width of the mass-selecting aperture, and n is a parameter corresponding to the scattering cross sections and is ≈ 1.7 for similar ion beam systems [179]. Figure 6.3 shows an experimental example of the scattering tail and abundance selectivity for a mass spectrum of ThO+ from Ref. [180]. The current of the ThO peak at mass 248 u drops quickly on either side of the peak, but then levels off, illustrating the scattering tail effect, which contributes roughly 10−6 of the 248 u peak at 247 u and 249 u. From Eq. (6.3) and literature values of the abundance selectivity for a single magnet system with an operating pressure within the beamline of approximately 1.3× 10−4 Pa (1.0× 10−6 Torr), a contribution of roughly 4× 10−6 of the higher mass peak to the lower mass peak is expected at a mass of 28 u [180, 181]. This peak tail current is not measurable in this ion beamline because the mass resolu- 363 Figure 6.3: ThO+ ion beam mass spectrum illustrating the scattering tail and abun- dance selectivity at mass 248 u. The main peak initially drops quickly on either side of the peak before leveling off as the ion current signal begins to be dominated by the scatter tail. This tail decreases more gradually and contributes approximately 10−6 of the 248 u peak at ∆m = 1 u. (from Ref. [180]) tion and current sensitivity are too low. Combining the scattering fraction with the 29Si natural abundance and the fact that the 29 u peak is typically about the same magnitude as the 28 u peak gives an estimate for an upper bound on the 29Si con- centration in the 28Si beam of roughly 2×10−7, or 200 ppb. This concentration may be significant for a few samples discussed here with the lowest 29Si isotope fractions around 200 ppb, although the true value may be lower making it less significant. Further, there is no evidence, e.g. significant and consistent attenuation of the 30Si isotope fractions compared to 29Si, that this scattering limit is having a significant effect on the measured enrichment. For example, the 502 ◦C sample with a mea- sured 29Si isotope fraction of 127(29) ppb has a 29Si/30Si ratio of 2.31 ± 0.96. This value agrees within the uncertainty with the natural abundance value of approxi- mately 1.52. Further, the expected 29Si/30Si ratio of the scatter tails from Eq. (6.3) 364 is larger than the measured ratio at ≈ 4.6. Based on this analysis, the scattering tail contribution is considered to be negligible of the purposes of the analysis discussed in this chapter, and the ion beam is assumed to be pure 28Si. For the discussion of the following sections, the difference between the expected (100 % enriched) and measured enrichment is considered by identifying only the natural abundance SiH4 gas diffusing from the ion beam into the deposition chamber as the source of 29Si and 30Si. 6.2.3 Determination of SiH4 Partial Pressures An accurate estimate of the partial pressure of SiH4, present at the sample lo- cation during 28Si deposition, is required to determine any correlations between this partial pressure and enrichment levels (i.e. 29Si and 30Si isotope fractions). Total pressure measurements were made using several different ion gauges in the system located in two of the three sample deposition locations (the lens chamber, LC–2, and the deposition chamber, DC–3) as described in Chapters 2 and 4. Comparing samples with pressure readings taken from different gauges introduces some error in the analysis because gauges may be calibrated differently. For the samples de- posited at room temperature, and some samples deposited with elevated substrate temperatures, the raw ion gauge readings from different gauges are assumed to be comparable in this analysis, although with different known uncertainties in the read- ings of different gauges, because a direct conversion was not performed at the time. Another gauge used for some of the samples deposited at elevated temperatures had a known offset which could be compared to a more accurate gauge at the same loca- 365 tion. A conversion was used to translate the pressure reading from this gauge to the equivalent readings for the more accurate gauge. Typically the readings from these types of ion gauges have a relative uncertainty of ≈ 20 %, while the more accurate gauge has a relative uncertainty of ≈ 5 %. After determining the appropriate total pressure during deposition for each sample, the base pressure at the relevant sample location immediately prior to deposition was subtracted out to get the total pressure increase due to gas diffusion from the ion source, which was typically a factor of 50 to 100 times higher than that base pressure. Pressure increases due to sample heating were also taken into account. To determine the SiH4 partial pressure component of the total pressure read- ings, the RGA in the deposition chamber was used to take partial pressure mea- surements from the residual gas mass spectrum while flowing SiH4 gas from the ion source. Because the RGA is not necessarily calibrated the same way as the ion gauges in the chamber, and the readings are influenced by the electron multiplier settings in the detector of the RGA, the absolute partial pressure readings are not reliable. Instead, the partial pressure readings of SiH4 were used to calculate the approximate fraction of the total sum of partial pressure peaks in the residual gas spectrum. Determining this SiH4 fraction is complicated by the fact that when the ion source is in operation, it cracks SiH4 into lower order Si hydrides, SiHx (1 < x < 4), and H2. The RGA filament itself further cracks SiHx. Figure 6.4 shows RGA residual gas mass spectra for the base pressure of the deposition chamber, which was about 8.3× 10−9 Pa (6.2× 10−11 Torr) (diagonal line fill), during SiH4 gas flow (horizontal line fill), and when the ion beam is on (solid line). When SiH4 366 Figure 6.4: Residual gas mass spectra collected from the RGA in the deposition chamber for the chamber base pressure (diagonal line fill), while flowing SiH4 (hor- izontal line fill), and with the ion beam in operation (solid line). The complex of SiHx peaks are observed from 28 u to 33 u during SiH4 gas flow. H2 is also observed to increase. is flowed into the chamber with the ion source off, the SiHx peaks are observed between 28 u and 33 u due to gas cracking from the RGA itself. The increased H2 signal is partially due to this process. When the ion source is operating, the SiH4 is initially cracked before diffusing to the RGA where the resulting SiHx is cracked further so that the residual gas mass spectrum does not show any significant Si hydride peaks. This is accompanied by an additional increase in H2 signal. This reduction in SiHx signal with the ion beam on precludes a direct measurement of the SiH4 partial pressure. To get an estimate of the SiH4 partial pressure during ion beam operation, the cracking efficiency of the ion source and RGA are assumed to be similar and 367 then measure the total SiH4 partial pressure fraction with the ion beam off. This assumption is supported by the fact that the ratios of Si hydride peaks seen in the ion beam mass spectrum above (Fig. 6.2) and the RGA residual gas mass spec- trum (Fig. 6.4) are similar. From these measurements and estimates for the H2 and SiH4 gas sensitivity factors [182, 183], the SiHx partial pressures is estimated to be roughly 28 % ± 5 % of the total pressure increase. This gives SiHx partial pressures, PSiHx , that range from 2.8× 10−7 Pa to 1.4× 10−6 Pa (2.1× 10−9 Torr to 1.1× 10−8 Torr) for the room temperature samples, and from 1.4× 10−7 Pa to 9.6× 10−7 Pa (1.1× 10−9 Torr to 7.2× 10−9 Torr) for the samples deposited at el- evated temperatures. The SiHx gas flux, Fg, which impinges on the sample during deposition is a critical parameter to the analysis of this chapter. Fg is derived from the SiHx partial pressures using the Hertz-Knudsen equation for gas flux [184,185], Fg = PSiHx√ 2pimkBTg , (6.4) where PSiHx is the SiHx partial pressure, m is the molecular mass of SiH4 (≈ 32 u), kB is the Boltzmann constant, and Tg is the temperature of the gas (≈ 21 ◦C). This calculation gives SiH4 gas fluxes between 7.6× 1011 cm−2 · s−1 and 3.7× 1012 cm−2 · s−1 for the room temperature samples. For the higher tempera- ture samples, the SiH4 gas fluxes are calculated to be between 3.7× 1011 cm−2 · s−1 and 2.6× 1012 cm−2 · s−1. The relative uncertainty of these estimates is ≈ 15 % to 20 %. 368 6.2.4 Substrate Temperature Calibration The temperature of the samples were carefully measured during deposition to ensure an accurate mapping of enrichment vs. temperature and determination of the temperature dependance of the incorporation fraction, s, discussed in the fol- lowing sections. As explained in Chapter 2, the samples were heated using direct current heating (DH) through the substrate for all samples except the one deposited at 249 ◦C, which was heated using the tungsten radiative back heater (RH). The substrate temperatures were measured by the two previously discussed infrared py- rometers. The Process Sensors pyrometer, which was calibrated for this system as discussed in Chapter 2, was used for most samples discussed in this chapter, but the un-calibrated Omega pyrometer was used for four of the samples. A correction was applied to the Omega pyrometer readings, which can differ from the Process Sensors readings by ≈ 25 ◦C. Including the uncertainties from the pyrometer cal- ibration, the temperature readings of the substrate are estimated to have a 5 % relative uncertainty due to fluctuations in the current used for sample heating as well as temperature gradients across the sample. The exact temperatures of the samples deposited at room temperature were not measured but instead assumed to be similar to the typical measured ambient temperature outside of the vacuum chamber, which was 21 ◦C ± 2 ◦C. For the analysis of these room temperature samples, this value was used. 369 6.3 Temperature Dependent Gas Incorporation Model To correlate the effect of SiHx partial pressure on the incorporation or ad- sorption of 29Si and 30Si during deposition of 28Si films at different temperatures, a gas sticking deposition model is formulated that describes the different contributing sources to the film deposition. The sticking of gaseous species is based simply on the idea of idealized Langmuir adsorption for a flat surface with equivalent adsorption sites [186]. This model is later compared to the sample enrichments measured by SIMS. The model describes two sources of Si atoms that contribute to deposition at the substrate: (1) the ion beam, which is presumed to be pure 28Si, deposits ions onto the sub- strate, and (2) the partial pressure of SiHx, which contains all three Si isotopes in their natural abundance, can stick and become incorporated into the film. Figure 6.5 is a cartoon representation of this two source deposition model. In this model, the isotopic concentrations measured by SIMS is the fraction of the total Si deposited that is due to 29Si or 30Si sticking from the SiHx background partial pressure. The isotope fraction of 29Si and 30Si in a sample is described by the gas sticking deposition model, cz (with z denoted as 29 for 29Si and 30 for 30Si), given by cz = Fgazs Fgs+ Fi , (6.5) 370 Figure 6.5: Cartoon illustrating the 28Si thin film deposition process described by the gas sticking deposition model. Two sources contributing Si to the deposition are the 28Si ion beam and the SiH4 gas which diffuses from the ion source and contains a natural abundance of isotopes. SiH4 can adsorb and react with the Si surface with different probabilities. where Fg is the SiHx gas flux, Fi is the 28Si ion flux, az is the isotopic abundance of 29Si or 30Si in the SiH4, which is assumed to have a natural abundance, and s is an effective incorporation fraction (or sticking coefficient) for an average SiHx species to be adsorbed into the surface. cz gives the calculated isotope fraction for a given set of deposition conditions. cz can be re-written as a function that correlates the isotope concentrations to the deposition conditions using a convenient deposition parameter; the SiH4 flux ratio, k = Fg/Fi. cz written as a function of k is then cz = azsk 1 + sk . (6.6) 371 Notice that cz increases with increasing SiH4 gas flux, and it decreases with increas- ing ion beam flux. Additionally, when Fg  Fi, then cz ∝ k, and when Fg  Fi, then cz ≈ az. Another convenient transformation of cz is to convert the isotope spe- cific model of Eq. (6.6) into a general model for SiH4 sticking by dividing by each isotope’s natural abundance so that 29Si and 30Si data can be fit together within the same model. Dividing the measured isotope fraction of a sample by its natural isotopic abundance gives an expected total adsorbed SiH4 fraction, which is then described by the gas sticking deposition model giving the calculated SiH4 fraction, ctot., for the total adsorbed SiH4 where ctot. = cz az = sk 1 + sk . (6.7) Equation (6.7) allows the full statistical weight of all the data to be used for deter- mining the incorporation fraction, s, for each sample deposited at different temper- atures and get the trend of s vs. temperature, T . Next, to describe the behavior of s vs. T , a temperature dependent incorpo- ration model, s(T ), is defined that is described by two gas sticking terms; a sticking probability resulting from physisorption, sc, and a higher temperature reactive stick- ing coefficient, sr, resulting from chemisorption. sc and sr are both expected to be activated by temperature, but sc decreases with increasing temperature, and sr increases with increasing temperature. These components are defined to be sc = 1− exp (−Ep kBT ) (6.8) and sr = Ar exp (−Ec kBT ) , (6.9) 372 where Ep is the activation energy for physisorption, Ec is the activation energy for chemisorption, kB is the Boltzmann constant, and T is the substrate temperature during deposition. The exponential prefactor Ar is left as a free parameter to account for experimental uncertainties which may introduce a constant shift in the data, and it will be discussed further in the following sections. The prefactor for sc is set to 1 because the sticking probability is expected to be close to unity as T approaches zero. The total incorporation fraction at a given temperature is then described by the temperature dependent incorporation model, which is the sum of the two sticking components, s(T ) = sc + sr = 1− exp (−Ep kBT ) + Ar exp (−Ec kBT ) . (6.10) 6.4 Correlating Enrichment to SiH4 Partial Pressure To demonstrate the correlation between the SiHx partial pressure and the 28Si sample enrichment, the raw SIMS isotope fraction data (zSi/Sitot.) for the room temperature samples are plotted as a function of the SiH4 flux ratio, k = Fg/Fi in Fig. 6.6, with 29Si (squares) and 30Si (triangles) isotope fractions plotted together. The top axis in panel (a) shows the total gas flux ratio, F tot.g /Fi, using the gas flux corresponding to the total measured pressure increase during deposition without subtracting out the estimated H2 fraction in the gas. This difference only shifts the axis laterally, and both the 29Si and 30Si isotope fractions have a strong linear correlation with k, representing the deposition conditions. The resulting Pearson 373 Figure 6.6: Correlation plot of isotope fraction vs. k for 28Si samples deposited at room temperature (≈ 21 ◦C). (a) SIMS measurements of 29Si (squares) and 30Si (triangles) are shown. The top axis shows the total gas flux during deposition after subtracting the background pressure flux. cz is fit (Eq. (6.6)) to the 29Si (solid line) and 30Si (dashed line) data giving s = 6.8(3)× 10−4. c30 is also shown calculated for two other values of s (dotted lines) to show the sensitivity of the model to s. (b) cz fits from (a) asymptote to the natural abundance values (dash-dotted lines) at large k (Fg  Fi). 374 correlation coefficient for the 29Si data is r = 0.95, and for the 30Si data r = 0.92. This shows that an increase in SiH4 flux corresponds to an increased isotope fraction in the sample. The correlation between enrichment and k is modeled using Eq. (6.6) to get c29 and c30, which are fit to the data with s as the only free parameter. These fits are shown in Fig. 6.6 as solid and dashed lines respectively, and they are approximately linear over the range of the data with a slope proportional to s. In panel (b), the fits are plotted in an extended range to show the crossover to az at high values of k. Above a k value of about 104, cz starts to asymptote to the natural abundance values of 4.7 % for 29Si and 3.1 % for 30Si (dash-dotted lines). In other words, when Fg/Fi  1, the contribution from the 28Si ion beam becomes negligible compared to the gas flux, and so the film composition approaches the composition of the gas, which is assumed in the model and expected in reality to have a natural abundance of Si isotopes. The fits to the data give a room temperature incorporation fraction of s = 6.8(3)×10−4. The uncertainty in this value is the standard error from the fit. Also plotted for reference is c30 calculated for two other values of s (dotted lines), 2× 10−4 and 2× 10−3, which span an order of magnitude around the data. This is to illustrate the sensitivity of the fit to s. Note that when viewed as a log-log plot as in Fig. 6.6, cz does not change apparent slope as s is varied, it only changes vertical offset. To investigate this correlation over a larger range of k values and specifically a larger range of SiH4 fluxes, i.e. pressures, samples deposited in the ion beam chamber at sample location IC–1 were also analyzed. These samples, which were 375 described in Chapter 4, were also deposited at room temperature and with much higher background pressures during deposition. These additional data are shown in Fig. 6.7, which is an expanded version of Fig. 6.6. The 29Si and 30Si data for the samples deposited at IC–1 (open squares and open triangles, respectively) exhibit similar qualitative behavior to the data for the samples mostly deposited at LC–2 (closed squares and closed triangles, respectively) at lower k values from Fig. 6.6. The fits of cz (Eq. (6.6)) from Fig. 6.6 are again shown, which are only fit to the LC–2 sample 29Si and 30Si data (solid and dashed lines, respectively). The IC–1 samples generally increase in isotope fraction with increasing k, however, they tend to deviate from the fit of cz exhibiting an apparently lower incorporation fraction. This deviation can possibly be explained as due to several factors. First, because there was no gas outlet via a vacuum pump at the sample location for these samples, there is an unknown and likely large uncertainty on the total pressure at the sample and its composition in terms of H2 and SiHx. The ion beam itself delivers gaseous species to the location of the mass-selecting aperture, which is relatively close to the sample, in the form of ionized SiHx that may locally increase the pressure around the sample. This pressure increase can be roughly estimated to increase the H2 partial pressure to ≈ 85 % of the total pressure, which is an increase from 70 % in the other experimental configurations. This correction cannot totally account for the discrepancy between the data and the model fit, however. Additionally, these samples were deposited with a total background pressure increase during deposition between approximately 1.5 ×10−4 Pa and 2.1 ×10−4 Pa (1.2 ×10−6 Torr to 1.6 ×10−6 Torr), which is at least a factor of 30 times higher 376 Figure 6.7: Correlation plot of isotope fraction vs. k for 28Si samples deposited at room temperature (≈ 21 ◦C). The data from Fig. 6.6 of samples deposited mostly at LC–2 (closed squares and closed triangles) are shown as well as SIMS measurements of 29Si (open squares) and 30Si (open triangles) from samples deposited at IC–1 with higher background pressures and thus larger Fg and k values. cz is fit (Eq. (6.6)) to only the LC–2 29Si and 30Si data (solid and dashed lines, respectively). The vertical dashed line indicates the boundary where k = 1. The LC–1 data lie to the right of this where Fg > Fi in a different deposition regime than the LC–2 data. than the other room temperature samples deposited at LC–2 and at most a factor of 220 times higher. These pressures and corresponding higher gas flux results in these samples being deposited in a regime where Fg > Fi. This is represented by the vertical dashed line in Fig. 6.7. When Fg > Fi, an incident SiHx molecule may not find an area of the surface with bare Si, and instead it may encounter another adsorbed SiHx or H2 molecule, blocking further adsorption. H2 coverage is known to decrease the effective reactive sticking coefficient of SiH4 [178], although with a small sticking coefficient of ≈ 10−4 on Si, H2 coverage may not be significant 377 Figure 6.8: 29Si/30Si isotope ratios for all measured samples deposited at room temperature (circles) and elevated temperatures (triangles). The ratios of these samples agree with the natural abundance ratio of 1.52 (line) indicating that the source of 29Si and 30Si is naturally abundant, probably the SiH4 gas. here [169]. Energetic ions can also desorb H from Si(100) surfaces [121]. It is difficult to accurately account for these high pressure effects and so the samples deposited at IC–1 are excluded from the analysis of this chapter. Further strong evidence that the SiH4 partial pressure is the source of the measured residual isotope fraction of 29Si and 30Si can be derived from the measured isotope ratios, 29Si/30Si, for each sample. If these isotopes were originating from the ion beam, one would expect an attenuation of 30Si compared to 29Si which would increase the 29Si/30Si ratio above the natural value. Instead, the measured isotope ratios are found to be very close to the natural value of approximately 1.52. Figure 6.8 shows the 29Si/30Si isotope ratios for a large number of room 378 temperature (circles) and elevated temperature (triangles) samples. Also shown is a line representing the natural abundance ratio. All of the data lie close to the natural ratio within their error bars. This indicates that the source of 29Si and 30Si has a natural abundance of Si isotopes, e.g., the SiH4 source gas. Measurement number 28 and 31, which lie above a 29Si/30Si ratio of four, suffer from discrete counting noise in the SIMS measurements due to a total 30Si count < 10 through the entire enriched film, which makes the ratio highly sensitive to single count fluctuations. 6.5 Temperature Dependence of 29Si and 30Si Adsorption The previous section showed that a strong correlation exists between the rela- tive SiH4 flux at the sample and the measured enrichment. This pressure dependance indicates that the natural abundance SiH4 gas diffusing from the ion beam is the source of residual 29Si and 30Si in the samples. Next, because substrate deposition temperature is a key parameter for facilitating and controlling the quality of epi- taxial deposition as this work progresses further, it is important to understand the effect of different substrate temperatures on the correlation between enrichment and gas flux. As reported previously in this chapter, the raw SIMS data show that the 29Si isotope fractions increase rapidly in the deposition temperature range from 502 ◦C (127 ppb) to 812 ◦C (4.32 ppm). However, the room temperature correlation plot (Fig. 6.6) showed that the isotope fractions also depend on k, the SiH4 flux ratio 379 (i.e. the deposition conditions). In order to extract an accurate temperature depen- dence of the enrichment, the raw SIMS measurements are adjusted to a common set of deposition conditions (Fg and Fi). To perform this adjustment, the k value matching the sample deposited at 502 ◦C is chosen, i.e. the sample with the lowest measured isotope fractions of 29Si and 30Si. This adjustment suppresses the effect on enrichment of varying deposition conditions across samples. By using the 502 ◦C sample as a benchmark against which to compare the other samples, the change in isotope fraction is mapped against temperature for the conditions that produced the best measured enrichment. To find the adjusted isotope fractions from the raw SIMS values, Eq. (6.6) is first solved for s, using subscript T to denote the resulting value, sT , for a specific data point with deposition temperature T (in ◦C). Equation (6.6) is then used along with each calculated sT value to generate the gas sticking deposition model curve, cz(sT ), for each data point. The k value of the 502 ◦C sample, which is the reference to which the raw SIMS data will be adjusted, is denoted as k502. To get these adjusted values, cz(sT ) at k = k502 is then evaluated for each data point. This gives cz(sT , k502), which is thus the isotope fraction of a sample with deposition temperature T adjusted to the deposition conditions (SiH4 flux ratio) of the 502 ◦C sample. For example, the adjusted isotope fraction of the 812 ◦C sample is calculated as cz(s812, k502). In Fig. 6.9, the adjusted isotope fractions given by cz(sT , k502) are plotted as a function of temperature showing the temperature dependence of the enrichment, independent of variations in deposition conditions. More specifically, the values in Fig. 6.9 are the expected isotope fractions of 29Si 380 Figure 6.9: Adjusted isotope fraction, cz(sT , k502), vs. temperature for 29Si (squares) and 30Si (triangles). The raw isotope fractions are adjusted to the deposition condi- tions (Fg and Fi) of the 502 ◦C sample. The inset shows the same data on a semi-log scale to highlight the behavior of the data below 500 ◦C. (squares) and 30Si (triangles) for all samples had they been deposited with the same SiH4 partial pressure and ion beam flux as the 502 ◦C sample. The adjusted isotope fractions are shown to trend downwards slightly from the 249 ◦C sample average of about 0.76 ppm 29Si to a minimum at the 502 ◦C average of about 0.13 ppm 29Si. This implies that if the substrate temperature of a sample similar to the 502 ◦C sample is lowered to 249 ◦C during deposition, the 29Si isotope fraction is expected to increase to approximately 0.76(17) ppm. The room temperature samples do not seem to follow this trend and instead have adjusted isotope fractions that are lower than expected. The deposition pres- sure for these samples was measured using a different ion gauge configuration than 381 the later high temperature samples, which may affect the adjustment of these data. Another difference is that the room temperature samples were all grown as amor- phous films while the samples deposited at elevated temperatures were crystalline. Surface orientation and crystallinity affect the adsorption of SiH4 on Si surfaces as described by Comfort and Reif who note that for Si CVD, the growth rate on a Si(100) surface is generally higher than on a Si(111) surface, which is itself higher than the growth rate on a polycrystalline surface [174]. These effects may lead to a lower effective sticking coefficient on the amorphous samples compared to that of the crystalline samples. Above 502 ◦C, the adjusted 29Si isotope fraction sharply increases up to a value of 5.9 ppm 29Si at 812 ◦C. Again, this shows that if the substrate temperature of a sample similar to the 502 ◦C sample was increased to 812 ◦C during deposition, the 29Si isotope fraction is expected to increase to 5.9(16) ppm. This increase is posited to be due to an increase in s as a process similar to a CVD reaction becomes more active, and the reactive sticking coefficient begins to dominate the total incorpora- tion fraction. To examine this hypothesis, the values of s at each temperature need to be determined and compared to the temperature dependent incorporation model (Eq. (6.10)). 382 6.6 Temperature Dependence of the Incorporation Fraction, s The gas incorporation fraction, s, is determined at each sample temperature using Eq. (6.6) shown in Fig. 6.6. It should be noted that the incorporation frac- tions determined here are a total net sticking probability; i.e. a molecule was in- corporated into the growing film, and remained there until detected by SIMS. A convenient transformation in the analysis of this gas sticking deposition model is to convert each isotope fraction for a given sample by dividing by their respective natural abundance values, az, so both 29Si and 30Si can contribute statistical weight together when fit with the model. This conversion follows the model of Eq. (6.7) for calculating ctot. and gives the converted isotope fraction, i.e. the expected total SiH4 fraction adsorbed in the sample, (zSi/Sitot.)/az. Figure 6.10 (a) is a correlation plot of the converted isotope fractions for several deposition temperatures: 812 ◦C (dia- monds), 705 ◦C (left-pointing triangles), 249 ◦C (hexagons), 421 ◦C (right-pointing triangles), and 502 ◦C (down-pointing triangles). These data are plotted vs. the SiH4 flux ratio, k. Also plotted is ctot. (Eq. (6.7)), which is fit to each temperature set with s as the only free parameter. The fits reported in this chapter are achieved using an orthogonal distance regression method which accounts for uncertainty in both coordinate values of each datum. Uncertainties reported with the fit values are the standard error from the fits. Within the range of Fig. 6.10, ctot. is approx- imately linear with a slope equal to s. Note that when ctot. is plotted on a linear scale, it runs through the origin because zero SiH4 flux results in a calculated SiH4 383 Figure 6.10: Correlation plot of the converted isotope fractions vs. SiH4 flux ratio, k, shown on a linear scale for samples deposited at several elevated temperatures. (a) The raw SIMS isotope fractions for 29Si and 30Si are each converted to an expected SiH4 fraction using their natural abundance, az. ctot., is fit (Eq. (6.7)) to the data for each deposition temperature (solid, dashed, dotted lines) and is approximately linear over this range with a slope of s. (b) ctot. fits from (a) asymptote to unity at large values of k. 384 fraction (adsorbed SiH4) of zero. Figure 6.10 (b) illustrates the functional form of the calculated SiH4 fraction, ctot., at large values of k = Fg/Fi where the SiH4 gas flux dominates the ratio and ctot. asymptotes to unity. From the fit values of ctot. at each temperature, the temperature dependence of s is plotted in Fig. 6.11, which shows that s follows a similar trend to the 29Si and 30Si isotope fractions in Fig. 6.9. In panel (a) of Fig. 6.11, s trends downwards slightly from a value of 1.6(2)× 10−3 at 249 ◦C to a minimum of 2.9(4)× 10−4 at 502 ◦C. In this temperature range, the data appears to behave in a similar manner to the sticking probability term resulting from physisorption, sc, which decreases with increasing temperature. Then as T is increased more, s rapidly increases to 2.3(5)× 10−2 at 812 ◦C. This increase is expected qualitatively for the reactive stick- ing coefficient term resulting from chemisorption, sr, which increases with increasing temperature. A list of the samples discussed in this chapter, their deposition param- eters, and measurement and analysis results can be found in Tables D.10 to D.13 in Appendix D. These values of s are consistent with previously reported values of the reactive sticking coefficient of silane species on Si surfaces, although there is a large variation in the literature. Si CVD studies have shown sr to range from 5× 10−4 to 5× 10−3 for polycrystalline Si deposition at 600 ◦C to 800 ◦C [177], and it ranges from 1× 10−3 to 3× 10−5 for Si(111) surfaces below 500 ◦C [187,188]. Next, the data in Fig. 6.11 is fit to the temperature dependent incorporation model of Eq. (6.10), s(T ), which is the sum of the sticking terms, sc+sr. Ep, Ec, and Ar are set as free parameters in the fit, which is shown in Fig. 6.11 (a) (line). Also plotted separately are the sc (Eq. (6.8)) and sr (Eq. (6.9)) terms (dotted and dashed 385 Figure 6.11: (a) s (circles) vs. deposition temperature. The temperature dependent incorporation model, s(T ), is fit to the data (Eq. (6.10)) and plotted (line) along with the individual sticking terms calculated from the fit; sc (dotted line, Eq. (6.8)), representing physisorption, and sr (dashed line, Eq. (6.9)), representing reactive chemisorption. (b) Semi-log plot of the fits from (a) showing the crossover from s(T ) dominated by sc to sr where s increases rapidly above 600 ◦C. 386 lines, respectively), which are calculated from the fit parameters. Panel (b) shows the fit and its modeled components on a semi-log plot to illustrate the crossover from s(T ) being dominated by sc to s(T ) being dominated by sr. The fit of s(T ) matches the data fairly well at higher temperatures, while there is some deviation in the lower temperature range. The fit also matches the minimum of s in the data around 500 ◦C. The value of the activation energy for chemisorption, Ec, given by this fit of Eq. (6.10) is 1.5(2) eV. The uncertainty in this value is the standard error from the fit. The value of Ep generated by the fit, however, has a large uncertainty indicating that the fit is under-constrained for the low temperature region, and so the value is not necessarily reliable. Because the fit’s choice of Ep parameter can also affect the high temperature part of the fit, the data dominated by Ec needs to be isolated and analyzed separately, which will be discussed in the next section. 6.7 Determination of the Reactive Sticking Activation Energy, Ec To isolate the incorporation fraction data that is most sensitive to the value of Ec, the data from Fig. 6.11 is plotted in an Arrhenius form for thermally activated processes [189, 190], and the higher temperature data is fit to a line. This analysis involves taking the natural log of s(T ) (Eq. (6.10)), the temperature dependent incorporation model, giving ln(s(T )) = ln ( 1− exp (−Ep kBT ) + Ar exp (−Ec kBT )) . (6.11) 387 This equation in this form is not useful for this analysis because a linear term cannot easily be extracted. However, for the higher temperature data, the reactive sticking coefficient is expected to dominate, which means s ≈ sr, and so taking the natural log of only the sr term (Eq. (6.9)) yields ln(sr) = ln(Ar)− Ec ( 1 kBT ) , (6.12) which is the equation for a line with intercept ln(Ar) and slope Ec. Figure 6.12 is a plot of ln(s) (circles) vs. inverse deposition temperature, 1 kBT , modified by the Boltzmann constant to have units of inverse energy. Also plotted is the natural log of s(T ) (dotted line, Eq. (6.11)). The room temperature data resides outside the graph window of Fig. 6.12 for clarity and to highlight the higher temperature data, but it is included in the s(T ) fit. Using a linear fit to determine Ec is accurate within the uncertainties in the linear region of the higher temperature data where sr dominates. This linear fit (Eq. (6.12)) is also shown in Fig. 6.12 (solid line) for the data between 502 ◦C and 812 ◦C. This fit gives a value for the chemisorption activation energy of Ec = 1.1(1) eV. The uncertainty of this value is the standard error from the fit. This value differs from the value of Ec obtained in the previous section of 1.5(2) eV from the s(T ) fit. As was mentioned previously, the Arrhenius fit is an alternative fitting method that gives a comparison to the previous fit of s(T ) and is possibly more accurate because it allows analysis of Ec separate from Ep and avoids convolution with the large uncertainty in the fit value of Ep. Both of these values, however, are consistent with reported activation energies of SiH4 CVD between 600 ◦C and 800 ◦C. The literature values for SiH4 CVD activation 388 Figure 6.12: Arrhenius plot of ln(s) (circles) vs. inverse temperature in energy units. The top axis shows the equivalent deposition temperature in ◦C. The data between the 502 ◦C sample and the 812 ◦C sample are fit to a line (Eq. (6.12)) whose slope gives an activation energy, Ec = 1.1(1) eV. Also plotted is the natural log of the fit of s(T ) (dotted line, Eq. (6.11)). energy can vary from about 0.4 eV to 2.2 eV depending heavily on experimental conditions such as surface orientation, gas pressure, and hydride species [177, 187, 191–194]. An average value of the activation energies derived from data in Ref. [174] for lower pressure CVD experiments with deposition temperatures between 500 ◦C and 900 ◦C is approximately 1.1 eV. The exponential prefactor, Ar, is required in the model for sr because it ac- counts for the vertical offset of the data in the Arrhenius plot (Fig. 6.12) from the 389 expected intercept of zero. Leaving Ar as a free parameter allows for experimental uncertainties that cannot reasonably be predicted such as the method used for es- timating the SiH4 partial pressure having a systematic shift that propagates to the calculated incorporation fraction values, which would affect Ar but not Ec. The pres- ence of different silicon hydrides, (SiHx) or possibly disilane (Si2H6) which is known to contaminate commercial silane and have higher sticking coefficients [188, 195], in the background gas could also introduce errors because they may have different reactive sticking coefficients leading to an average value being measured. Gates has shown that the activation energy of Si CVD increases from around 1.17 eV to 2.04 eV for SiH3 and SiH respectively [194]. As stated previously, the prefactor for sc is set to 1 because it is more reasonable that sc approaches unity at zero temper- ature. Also, differences in reactive sticking of different molecules should not affect the physisorption term as much, making it not as critical for the prefactor to be left free. 6.8 Chapter 6 Summary In this chapter, the measured enrichment (i.e. residual 29Si and 30Si isotope fractions) of samples grown at both room temperature and with elevated substrate temperatures ranging from 249 ◦C to 812 ◦C was analyzed to understand how en- richment changes as a function of pressure and temperature due to SiH4 incorpo- ration. The 29Si and 30Si isotope fraction were found to be highly correlated to the background partial pressure of SiH4 during deposition. This showed that the 390 dominant, if not only, source of residual minor isotopes in 28Si samples is the nat- ural abundance SiH4 gas which adsorbs into samples during growth with a room temperature incorporation fraction, s = 6.8(3) × 10−4. From further analysis, the temperature dependence of the incorporation fraction is determined and modeled using two sticking terms in the temperature dependent incorporation model. A ph- ysisorption sticking probability decreases weakly with increasing temperature while a reactive sticking coefficient due to CVD-like chemisorption increases more strongly with increasing temperature from a minimum of 2.9(4)× 10−4 for the 502 ◦C sample up almost two orders of magnitude to 2.3(5)× 10−2 for the 812 ◦C sample. These competing terms lead to a minimum in residual 29Si and 30Si isotopic concentration at around 500 ◦C. The lowest 29Si isotopic concentration for a sample deposited in that range was 127(29) ppb. As an alternative to the temperature dependent incorporation model, an Arrhenius formalism was used to determine the activation energy of the reactive sticking coefficient and found that Ec = 1.1(1) eV. Under- standing the role of SiH4 gas sticking for a range of deposition temperatures allows for better prediction of, and control over, the resulting enrichment. This knowledge enables production of 28Si samples with targeted levels of enrichment (29Si isotope fractions), which could facilitate a study of T2 coherence times as a function of 29Si concentration. 391 Chapter 7 Summary of Results and Future Experiments 7.1 Summary and Conclusions In this work, a mass selected, hyperthermal energy ion beam deposition system was used to achieve in situ isotopic enrichment of a total of three 22Ne samples, three 12C samples, and 61 28Si samples. Very high levels of enrichment were achieved, especially for 28Si, which was also successfully deposited as epitaxial thin films. Chapter 3 demonstrated successful proof of principle enrichment by implanting the minor isotope 22Ne enriched with an isotope fraction of 99.455(36) % into Si as well as depositing thin films of 12C enriched to an isotope fraction of 99.9961(4) % using the mass selected ion beam system. Realized mass selectivities of 1785:1 for 22Ne and 276:1 for 12C were achieved. While these selectivities are not directly translatable to 28Si enrichment, these experiments do show that very high levels of in situ enrichment are possible in thin film deposition using this system. Chapters 4 and 5 demonstrated that 28Si films could be deposited with residual 392 29Si and 30Si isotope fractions consistently below 1 ppm with both elevated substrate temperatures and while using the high pressure plasma mode of the ion source. The enrichment values of these samples meet the enrichment materials goal laid out at the beginning of this thesis to achieve 28Si enrichments that surpass those of any other known sources of 28Si including the IAC. The isotope reduction factor of 29Si for the most highly enriched sample deposited at DC–3 is 3.7(8)× 105. This means that there is approximately 3.7× 105 times less 29Si in that sample than in natural abundance Si. This value of the reduction factor along with the reduction factor values of the excluded isotopes from all the most highly enriched samples of 28Si, 22Ne, 12C are shown in the isotope reduction timeline of isotope reduction factors in Fig. 7.1. A version of this figure was previously shown for Si samples in Fig. 5.54 in Chapter 5. Like that figure, Fig. 7.1 shows the isotope reduction factors vs. the sample deposition date for samples with the highest enrichments achieved up to that point, where a larger reduction factor is equivalent to a higher level of enrichment. As pre- viously discussed, the isotope reduction factors are defined as the natural abundance of an isotope of an element, az, divided by the measured isotope fraction of that iso- tope. Here, the isotope fractions are written generally as zX/Xtot., where zX refers to the counts of a particular isotope in the measurement of a particular element, and Xtot. is the sum of the counts of all isotopes being measured, as discussed in Chapter 1. The isotope reduction factors are therefore written here as az/( zX/Xtot.). The isotope reduction factors for 22Ne (diamonds), 12C (circle), 29Si (squares) and 30Si (triangles) in Fig. 7.1 are shown to increase for samples deposited throughout 393 Figure 7.1: Isotope reduction timeline. Timeline of the progression of the isotope re- duction factors for the lowest residual isotope fractions of 22Ne (diamonds), 12C (cir- cle), 29Si (squares) and 30Si (triangles), as measured by SIMS. These were achieved for 22Ne, 12C, and 28Si samples deposited over approximately five years. this work. Uncertainties in the reduction factors are shown for all samples and are derived from the SIMS measurements of the isotope fractions. The overall 28Si isotope fraction of the most highly enriched 28Si sample de- posited at DC–3 and in this entire work was 99.9999819(35) %. This demonstrates an improvement over the initial 28Si sample deposited at IC–1 in the form of a re- duction in 29Si of approximately 2.2 ×104 times. This and other similar samples are more highly enriched than any other known source of 28Si, including that of the IAC. These samples also demonstrate enrichments sufficient to enable a robust measurement of the dependance of electron coherence times on 29Si concentration 394 in the single spin regime and compare it to theoretical predictions [12], as proposed in Chapter 1 (see Fig. 1.9). The overall decrease in isotope fractions for isotopes rejected during production of 22Ne, 12C, and 28Si samples can be seen in the en- richment progression timeline for 20Ne (diamonds), 13C (circles), 29Si (squares) and 30Si (triangles) in Fig. 7.2. As mentioned previously, this timeline shows the SIMS measurements of the isotope fractions vs. deposition date for the samples that were the most highly enriched of any samples deposited up to that point. These are the record enrichments of samples produced in this work. As previously discussed, the isotope fractions are written generally here as zX/Xtot.. Uncertainties in the isotope fractions are shown for all samples and are derived from the SIMS measurements, as previously discussed. Also shown for comparison are the 29Si isotope fractions of the 28Si epilayers and crystals produced by Isonics and Itoh using 28SiH4 CVD (dash-double dotted line) from Ref. [36,37], the bulk 28Si material produced by the IAC using 28SiH4 CVD (dash-dotted line) from Ref. [32], and the 28Si thin films produced by Tsubouchi et al. using 28Si IBE (dotted line) from Ref. [43]. The 28Si samples produced in this work have residual 29Si and 30Si isotope fraction that are far less than those of these other sources. Some of the key experimental changes that led to the improvements in en- richment include improving the ion beam geometric mass selectivity through beam tuning, depositing in a lower background pressure of SiH4, and achieving higher de- position rates and smaller ion beam spot sizes through beam tuning. Most of these factors relate to the adsorption of natural abundance SiH4 gas from the vacuum during deposition, which lowers the resulting 28Si enrichment. 395 Figure 7.2: Enrichment progression timeline. A timeline of the progression of the best residual isotope fractions of 20Ne (diamonds), 13C (circle), 29Si (squares), and 30Si (triangles), as measured by SIMS. These were achieved for 22Ne, 12C, and 28Si samples deposited over approximately five years. Also shown for comparison are the 29Si isotope fractions of the 28Si epilayers and crystals produced by Isonics and Itoh using 28SiH4 CVD (dash-double dotted line) from Ref. [36, 37], the bulk 28Si material produced by the IAC using 28SiH4 CVD (dash-dotted line) from Ref. [32], and the 28Si thin films produced by Tsubouchi et al. using 28Si IBE (dotted line) from Ref. [43]. Chapter 5 also demonstrated that sample heating was critical for achieving epi- taxy. Crystalline, epitaxial 28Si films were produced using elevated deposition tem- peratures between 349 ◦C and 1041 ◦C, although samples deposited above 600 ◦C were very rough. This was due to chemical contaminants such as SiC at the growth surface that result in step pinning sites that lead to step bunching, faceting on {111} and {113} planes, and large mound formation. Smooth, epitaxial 28Si films with typical surface widths of ∆z = 2 nm were achieved by depositing at lower 396 temperatures between 349 ◦C and 460 ◦C. Chemical contaminants in these 28Si films were measured by SIMS, which de- tected N, C, O, F, Al, and Cl. The N, C, and O were detected at especially high concentrations, initially all above 1× 1019 cm−3. The dominant source of the N, C, and O for this sample was the ion beam. N2 and CO ions were likely introduced ballistically into the film from the 28 u beam. F and Al contamination were elimi- nated for a later sample due to experimental alterations. By improving the vacuum in both the deposition chamber and the ion beam chamber, and by using a higher 28Si ion beam current generated in the high pressure mode of the ion source, N, C, and O were reduced in the final sample. The average atomic concentration of N in the film was measured to be 9.07(4) ×1018 cm−3, or 181.8(8) ppm, the average atomic concentration of C in the film was measured to be 8.27(3) ×1018 cm−3, or 165.7(6) ppm, and the average atomic concentration of O in the film was measured to be 3.09(4) ×1018 cm−3, or 61.9(8) ppm. Based on these concentrations and the solid solubility of these elements in Si, it is likely that Si3N4, SiC, and SiO exist within the 28Si films. These remaining atomic concentrations were likely the result of both adsorption from the background vacuum and ballistic transport from the ion beam. The resulting best total chemical purity for a 28Si sample deposited at DC–3 and overall in this work is approximately 99.96(2) %. Finally, TEM was used to confirm that 29Si films produced with both the higher and lower deposition temperatures were crystalline and epitaxially aligned to the substrates. All samples deposited with substrate temperatures above 600 ◦C were observed to have {111} stacking faults and twinning present in the films. For 397 lower deposition temperature samples, one sample deposited using the low pressure mode did not have any visible stacking faults but probably still contained other crys- talline defects. Another sample deposited using the high pressure mode to generate a much higher deposition rate was observed to have a lot of {111} stacking faults and likely other defects. This increased defect concentration may have been due to the deposition rate for the second sample being over a factor of 12 larger while the deposition temperature was slightly lower. In general, the crystalline defects in these films are likely related to the presence of clusters of chemical contaminant. Reduction of these crystalline defects likely requires reduction of the chemical con- taminants within the 28Si films. The second and third materials goals for chemical purity and crystallinity described in Chapter 1 were not pursued further within the experiments depositing 28Si samples at DC–3, although improvements and new understanding regarding the source of both chemical and crystalline defects were made. Chapter 6 explored and quantified the relation between the natural abundance gas in the chamber and the resulting sample enrichments. The measured enrichment (i.e. residual 29Si and 30Si isotope fractions) of samples grown at both room tem- perature and with elevated substrate temperatures ranging from 249 ◦C to 812 ◦C was analyzed to understand how enrichment changes as a function of pressure and temperature due to SiH4 incorporation. The 29Si and 30Si isotope fractions were found to be highly correlated to the background partial pressure of SiH4 during deposition. This showed that the dominant, if not only, source of residual minor isotopes in 28Si samples is the natural abundance SiH4 gas which adsorbs into sam- 398 ples during deposition with an incorporation fraction, s. From further analysis, the temperature dependence of s is determined and modeled using two sticking terms in a temperature dependent incorporation model. A physisorption sticking probability decreases weakly with increasing temperature while a reactive sticking coefficient due to CVD-like chemisorption increases more strongly with increasing tempera- ture from a minimum of 2.9(4)× 10−4 for the 502 ◦C sample up almost two orders of magnitude to 2.3(5)× 10−2 for the 812 ◦C sample. These competing terms lead to a minimum in residual 29Si and 30Si isotopic concentrations at around 500 ◦C. The lowest 29Si isotopic concentration for a sample deposited in that range was 127(29) ppb. As an alternative to the temperature dependent incorporation model, an Arrhenius formalism was used to determine that the activation energy of the reactive sticking coefficient is Ec = 1.1(1) eV. Understanding the role of SiH4 gas sticking for a range of deposition temperatures allows for better prediction of, and control over, the resulting enrichment. This knowledge can aid in enabling produc- tion of 28Si samples with targeted levels of enrichment (29Si isotope fractions) that could facilitate a study of T2 coherence times as a function of 29Si concentration, as discussed in the next section. The main impact of this work on the field of Si-based solid state quantum computing is the demonstration of 28Si produced with extremely high levels of en- richment that have never been produced before. This was achieved in a laboratory setting as opposed to an industrial scale effort and was comparatively cheap. Ef- fectively, a new material was engineered using a particular processing method, i.e. ion beam in situ enrichment and deposition. Overall, the highest quality 28Si films 399 produced in this work, in terms of the materials goals stated in Chapter 1, were de- posited with varied deposition conditions. However, 28Si films deposited using the high pressure mode of the ion source with a deposition temperature of approximately 450 ◦C are most likely to result in the highest quality films. This temperature is optimal for both enrichment and smooth epitaxial growth, and the high ion cur- rent achieved in the high pressure mode should lead to a minimization of chemical contaminants. The high growth rate however, may lead to diminished crystalline quality. 7.2 Proposals for Future Experiments 7.2.1 Targeted Levels of Enrichment One of the advantages of enriching material in situ nearly simultaneously with deposition is the ability to more easily modulate the enrichment level throughout the sample by switching the magnetic field of the sector mass analyzer to modulate the isotope being deposited. As a demonstration of modulating the isotopic con- centration of a material throughout its thickness, an isotope heterostructure that alternates between layers of enriched 28Si and layers of near natural abundance Si was produced. This sample was deposited at room temperature at LC–2. By switch- ing the magnetic field of the sector mass analyzer from the 28 u beam to the 29 u beam (which is composed of both 29Si and 28SiH), one can control which isotope or combination of isotopes is deposited. In a typical SiH4 mass spectrum produced when using the low pressure mode of the ion source, the ion current peak height at 400 Figure 7.3: SIMS depth profile of a Si isotope heterostructure deposited at room temperature at LC–2. The isotope fractions are shown as a function of depth for 28Si (circles), 29Si (squares), and 30Si (triangles) in an isotope heterostructure. The magnetic field of the sector mass analyzer was modulated to select 28 u or 29 u at different depths, as noted on the top axis of the figure. The 29Si isotope fractions are seen to be roughly the natural abundance within the 29 u regions and much lower within the 28 u regions. The 29Si tail in 28 u regions is a measurement artifact. The peak below a depth of 300 nm is due to ion beam tuning during the beginning of the deposition. At a depth of about 351 nm, the isotope fractions return to their natural abundance values, indicating the interface with the Si(100) substrate (shaded region). 29 u is similar to the 28Si peak height, suggesting about 95 % of the 29 u peak is composed of 28SiH, as previously discussed. The 29Si concentration when depositing from the 29 u peak is then expected to be similar to the natural abundance value at roughly 5 %. No 30Si is expected whether the analyzer is set to mass 28 u or 29 u. SIMS was used to measure the isotope fractions of the isotope heterostructure. A SIMS depth profile of this sample is shown in Fig. 7.3. Isotope fractions of 28Si 401 (circles), 29Si (squares), and 30Si (triangles) are shown as a function of sputter depth into the film. The 29Si isotope fractions are seen to be roughly the natural abundance within the 29 u regions and much lower within the 28 u regions. The 29Si tail in 28 u regions is a measurement artifact, and the true 29Si isotope fractions in these regions is expected to be roughly 1× 10−6 (1 ppm), as seen at a depth of 50 nm. The 30Si isotope fraction remains below roughly 1× 10−6 (1 ppm) throughout the entire film, as expected. The peak in 29Si and 30Si below a depth of 300 nm is due to ion beam tuning during the beginning of the deposition. At a depth of about 351 nm, the isotope fractions return to their natural abundance values, indicating the interface with the Si(100) substrate, which is marked by the shaded region. The ability to switch between 28Si and 29Si can enable deposition of 28Si sam- ples with targeted enrichments, i.e. different isotopic concentrations of 29Si. To produce a desired isotopic concentration of 29Si in a sample, one could deposit 28Si normally and then during deposition, mass select for the 29 u ion beam for the ap- propriate fraction of the total deposition time, taking into account the 28SiH in the 29 u peak. A duty cycle would be chosen to ensure sufficient mixing of the two iso- topes, depending on the deposition rate. Samples with specific enrichments ranging from natural abundance down to less than 1 ppm 29Si isotopic concentrations would enable a measurement of T2 as a function of 29Si concentration for a wide range of enrichment values, as discussed in Chapter 1. These measurements would begin to address the question of “how good is good enough?” for the enrichment of 28Si in a Si-based quantum computing architecture. 402 7.2.2 Enriched Si/Ge Deposition Quantum wells in Si/SiGe heterostructures are a promising system for solid state quantum information. Although 28Si has been used by several research groups to produce quantum coherent devices, including a 28Si quantum well in SiGe [?], fully enriched Si/SiGe devices have not yet been demonstrated. Like 28Si, 74Ge has no net nuclear spin and is the most abundant of the Ge isotopes comprising approximately 37 % of natural abundance Ge. The critical isotope to be removed is 73Ge, which does have a nuclear spin of I = 9/2 and a natural abundance of approximately 8 %. The ability described previously to switch between different masses in the ion beam during deposition enables the growth of heterostructures from isotopic mate- rials. This is achieved by cycling between the constituents of compound materials where both components are enriched, e.g. 28Si/28Si74Ge. To produce Ge ions, Ge cathodes could be used in the solids-mode ion source designed for sputtering the cathode material. The source of Si could either be SiH4 or a Si cathodes with Ar being used as the working gas. Magnetic peak switching could then be used, as described previously, to control the relative concentrations of Si and Ge needed for SiGe deposition. A duty cycle would be chosen to ensure that the compound is sufficiently mixed during deposition to produced enriched 28Si74Ge. During deposition, a layer of pure 28Si would be deposited as the quantum well layer followed by a capping layer of 28Si74Ge. Control gates would then need to be patterned on the top surface of the heterostructure. 403 7.2.3 28Si sublimation In order to facilitate collaborations with other research groups interested in using enriched 28Si to fabricate quantum coherent devices, a useful experimental technique would be thermal re-deposition of 28Si from a deposited thin film onto a fabricated device as a capping layer. Quantum dot devices can benefit from not only having a 28Si substrate, onto which the device is fabricated, but also a 28Si cap to further isolate the qubit spins from 29Si. A demonstration of the re-deposition of 28Si was attempted in this work by heating a 28Si thin film sample using the DH method of heating to approximately 1100 ◦C to cause sublimation of the 28Si. Another Si(100) substrate with a thermal oxide surface layer was used as a deposition target. In order to attempt this transfer of 28Si, the sublimation rate needed to be calibrated. A calibration for three different sublimation temperatures was performed and is shown in Fig. 7.4. The measured sublimation rates (triangles) at approximately 1100 ◦C, 1150 ◦C, and 1200 ◦C were determined from ellipsometry measurements of the film thicknesses. Also shown are calculations for the expected sublimation rate based on the Si vapor pressure at different temperatures (line). Taking into account geometric factors of the deposition, a correction to the raw calculation was applied. The corrected calculation (dashed line) has values that are 25 % of the raw calculation and matches the measured rates well. The target substrate in the 28Si sublimation test was measured by SIMS to detect the transferred 28Si film. However, the presence of the thermal oxide and the fact that the film was very thin resulted 404 Figure 7.4: Si sublimation rates for a 28Si chip and calculated values. Measured rates (triangles) at 1100 ◦C, 1150 ◦C, and 1200 ◦C are compared to calculations based on Si vapor pressures (line) and a modified calculation (dashed line) whose values are 25 % of the raw calculation to account for geometric factors of the deposition. in a null measurement with no 28Si being detected. 7.2.4 Al Dopant Devices with Hydrogen Lithography STM hydrogen lithography is a promising technique for fabricating single atom 31P dopant devices for quantum information. A STM probe is used to selectively remove H atoms from a H-terminated Si surface thus defining a lithographic pattern. This surface is then exposed to phosphene gas which incorporates into the bare Si within the pattern forming a single atom thick delta layer. Using this technique, conducting wires and quantum dot islands comprised of 31P atoms, as well as single atom islands can be fabricated [5]. Instead of using 31P, which is an electron donor, it may be interesting to use an acceptor atom to create quantum coherent devices based on the physics of hole transport. One candidate acceptor atom is Al. If Al 405 Figure 7.5: SIMS depth profile of an Al delta layer in Si showing atomic concen- trations of Al, O, H, and P (lines) in the sample. Al is clearly concentrated in a layer roughly 25 nm below the surface. Above this depth is the natSi capping layer and below it is the Si(100) substrate. The peak atomic concentration of Al is about 1× 1020 cm−3. The high concentration of O at the Al layer is likely due to outgassing of the Al thermal deposition source. can be selectively deposited in a delta layer onto Si(100) with a H pattern, then an acceptor device may be possible. To demonstrate the capability to deposit an Al delta layer, Al was deposited on a Si(100) substrate with a roughly monolayer coverage and then overgrown with natural abundance Si, natSi, from the evaporation source. The composition of this sample was measured by SIMS in a depth profile, which is shown in Fig. 7.5. The atomic concentrations of Al, O, H, and P (lines) are shown vs. the sputter depth into the sample. One can see that there is a clear Al rich layer roughly 25 nm below the surface. This first 25 nm represents the natSi capping layer. It is difficult to 406 determine from this measurement if the Al resides in a true delta layer, but the Al likely spread out to some extent, possibly due to heating the sample during natSi deposition. Beyond 25 nm represents the Si(100) substrate. The peak atomic concentration of Al is approximately 1× 1020 cm−3, which is approximately 20 % of the nominal density of crystalline Si. A large concentration of O is also observed at the delta layer with a similar concentration as the Al, possibly due to the Al thermal deposition source outgassing. This successful demonstration producing a confined Al layer in Si is the first step towards patterned Al dopant devices defined by STM hydrogen lithography. Selectivity of Al atoms deposited in a H pattern would need to be tested further before production of a device. 7.2.5 Electrical Measurements and T2 in 28Si Ultimately, the quality and usefulness of 28Si material produced in this work to quantum information research will be determined by the T2 time of spins within the material. Additionally, the electrical quality of the material is important for the operation of quantum coherent devices. Several types of devices can be used to test parameters related to the electrical quality including Schottky diodes, capacitors, and transistors in Hall and Van der Pauw measurement geometries. A cartoon schematic of a capacitor structure and measurement for a 28Si sample is shown in Fig. 7.6. A 28Si film is seen on top of a natural abundance Si substrate. An isolation gate oxide is grown from the 28Si film and then an Al gate 300 nm thick is deposited on top of the oxide. Then, to perform measurements, a metal probe contacts the top gate while the substrate contacts metal on the back side of the sample, and 407 Figure 7.6: Cartoon schematic of a 28Si capacitor. An isolating gate oxide is grown on a 28Si film and an Al top gate is deposited on top of that, forming a capacitor. A voltage sweep can then be applied between a metal probe contacting the gate and a back metal contact behind the Si(100) wafer. a voltage sweep is applied to perform a capacitance-voltage, C-V, measurement. This measurement can determine the interface trap density and free charge carrier density in the 28Si oxide. Transistors can also be fabricated on 28Si and used for measuring electron mobility and carrier type within the enriched film. While the 28Si films produced in this work typically do not comprise a large enough volume of 28Si material to contain enough spins for traditional ESR mea- surements, other specialized measurements can be used to measure T2 in these films. For a sample implanted with 31P atoms, a small ESR probe can be fabricated onto the surface of a 28Si thin film. Such a µm scale probe greatly increases the sensitivity of the measurement and can measure a much smaller number of spins, potentially as few as 10s of spins. A group led by Aharon Blank has tested this technique [40] and has already made test devices on 28Si samples produced in this work. The primary impediment of these measurements at the moment is the concentrations of chemical contaminants within the 28Si samples. The concentration of N spins in 408 these samples exceeds that of the implanted 31P atoms, making a T2 measurement impossible. Therefore, to enable these measurements, the purity of 28Si samples must be improved and potentially can be by switching to a UHV compatible ion source and a high purity gas feed line. 409 Appendix A: Ion Source and Beamline: Additional Operating Parameters This appendix gives additional operating parameters and analysis for the ion source and ion beamline. A circuit diagram and schematic of the lens elements of the beamline before the magnetic sector mass analyzer showing the relationship between the controlling power supplies and measured currents is shown in Fig. A.1. Operating parameter scans of the ion source and beamline using Ne in the gas mode ion source are shown in Fig. A.2. These give the 20Ne ion current as a function of source magnetic field, gas flow, arc voltage, and extractor voltage. Finally, the operating parameters used for producing the most highly enriched 22Ne, 12C, and 28Si samples are given in Tables A.1 to A.3, respectively. 410 Figure A.1: Circuit diagram and schematic of the electrostatic lenses of the ion beamline before the magnetic sector mass analyzer. Elements of the ion source including the anode, cathodes and the source electromagnet coil are shown at the left. The lens elements that form the potential landscape of the ion beam, depicted from left to right, are the extractor, at negative voltage, VExt, a transport tube, at negative transport voltage, VT , the focus, at negative voltage, VF , another transport tube with X-Y deflectors biased positively with VXY , the skimmer, and the electron suppressor, at negative voltage, Ve, which is set to a potential about 100 V more negative than VT . VExt, VF , VXY , and Ve all float on top of VT . VA applies a positive voltage to the anode, and Varc floats on top of the anode potential to apply a negative voltage to the cathodes. The source electromagnet drives current ISM through the solenoid and it floats on top of Varc due to the magnet housing not being well isolated from the cathode housing. Iarc is the measured current flowing between the anode and cathodes. All ion current that is neutralized on these lens elements or the vacuum chamber is recorded as IT . 411 Figure A.2: Operating parameter scans of the gas mode ion source using Ne. (a) The 20Ne ion current, Ii, and the arc current, Iarc, are shown as a function of the source magnet field. The corresponding current, ISM , applied to the electromagnet is shown on the top axis. The plasma ignites above 0.05 T and two peaks are observed. (b) Ii and Iarc for 20Ne are shown as a function of the Ne gas flow, with the corresponding pressure in the ion beam chamber displayed on the top axis. The highest ion current is observed at a flow of about 6 mL/h. (c) Ii and Iarc for 20Ne are shown as a function of the arc voltage, Varc. Ii and Iarc appear inversely related. (d) Ii and the ion current on the skimmer, Isk, for 20Ne are shown as a function of the extractor voltage, VExt. Ii was optimized at each value of Varc by tuning the other lens elements. A peak in Ii at about -6.5 kV is observed. 412 22Ne Ion Beam Parameters for 110504-22Ne-OxSi on 5/4/11 Source P VA Varc VC VT (Pa) (V) (V) (V) (kV) 1.0× 10−2 501 500 0 -4.006 Sample Bias VExt VF ISM Iarc (kV) (kV) (kV) (A) (mA) -4.00 -7.0 -13.0 1.1 22 Table A.1: Ion source and beamline operating parameters for implanting 22Ne sample 110504-22Ne-OxSi on 5/4/11 using the solids-mode ion source. This sample was the most highly enriched 22Ne sample produced in this work. Ne was used as the working gas for this sample. VA is the anode voltage, and Varc is the arc voltage applied in reference to VA to get the resulting cathode voltage, VC . VT is the transport voltage. A Si(100) substrate was biased negatively for implantation. VExt is the extractor voltage and VF is the focus voltage. The ion source plasma is ignited by setting the source electromagnet current, ISM , producing an arc current, Iarc. 12C Ion Beam Parameters for 120207-12C-OxSi on 2/7/12 Source P VA Varc VC VT (Pa) (V) (kV) (V) (kV) 3.1× 10−3 608 1.147 -539 -4.00 VExt VF ISM Iarc Isk (kV) (kV) (A) (mA) (µA) -7.50 -15.0 1.45 25 32 Table A.2: Ion source and beamline operating parameters for depositing 12C sample 120207-12C-OxSi on 2/7/12 using the solids-mode ion source. This sample was the most highly enriched 12C sample produced in this work. CO2 was used as the working gas for this sample. VA is the anode voltage, and Varc is the arc voltage applied in reference to VA to get the resulting cathode voltage, VC . VT is the transport voltage. VExt is the extractor voltage and VF is the focus voltage. The ion source plasma is ignited by setting the source electromagnet current, ISM , producing an arc current, Iarc. Isk is the ion current measured on the skimmer. 413 28Si Ion Beam Parameters for 151210-KD-7-i-28Si-500C on 12/18/15 Source P VA Varc VC VT (Pa) (V) (kV) (kV) (kV) 2.1× 10−4 50 3.420 -3.370 -4.00 VExt VF ISM Iarc IT Isk (kV) (kV) (A) (mA) (mA) (µA) -11.0 -12.5 1.69 0.7 0.6 190 Deceleration Lenses A2 A3 B2 B3 B4 X (kV) (kV) (kV) (kV) (kV) (V) -2.50 -0.49 -3.04 -0.70 -1.34 -14 Table A.3: Ion source and beamline operating parameters for depositing 28Si sample 151210-KD-7-i-28Si-500C on 12/18/15 using the gas-mode ion source. This sample was the most highly enriched 28Si sample produced in this work. SiH4 was used as the working gas for this sample. VA is the anode voltage, and Varc is the arc voltage applied in reference to VA to get the resulting cathode voltage, VC . VT is the transport voltage. VExt is the extractor voltage and VF is the focus voltage. The ion source plasma is ignited by setting the source electromagnet current, ISM , producing an arc current, Iarc. IT is the ion current measured on the transport voltage line, and Isk is the ion current measured on the skimmer. A2, A3, B2, B3, B4, and X refer to deceleration lens elements. 414 Appendix B: Experimental Apparatus Photographs This appendix shows photographs of various experimental apparatus used for producing enriched samples in this work including the hyperthermal energy ion beam deposition system as a whole, the gas manifold, the two ion sources, the electrostatic and magnetic lens elements comprising the beamline, the three mass-selecting aper- tures, a gas aperture, experimental setups and sample stages for samples produced at IC–1 in the ion beam chamber, LC–2 in the deceleration lens chamber, and DC–3 in the deposition chamber, and finally the sample apertures used for monitoring the ion beam current. 415 Figure B.1: Photographs of the ion beam deposition system (a) Side view of the ion beamline and deposition system. (b) Top-down view of the ion beamline. (c) Side view of the deposition and analysis chamber showing the manipulator where samples are located at DC–3. 416 Figure B.2: Photograph of the gas manifold. The natural abundance SiH4 gas source used for depositing 28Si is connected at the left. The manifold contains two gas reservoir tanks in the middle and a vacuum pumpout at the bottom. The manifold connects to the ion source inlet out of view at the top right. Figure B.3: Photographs of the ion source elements. (a) Inlet cathode housing of the solids- mode ion source with a Cu cathode. (b) Anode housing of the solids-mode ion source with a Cu anode disk. This housing fits over the cathode in (a). (c) Inlet cathode housing of the gas-mode ion source with a Cu cathode. (d) Anode housing of the gas-mode ion source with a cylindrical steel anode. The anode housing is seen fitted over the cathode in (c). 417 Figure B.4: Ion beamline electrostatic elements and magnetic sector mass analyzer. (a) The electrostatic elements used to generate an ion beam from the source are seen with the extractor at the left, focus and X-Y deflectors in the middle, and skimmer at the right. These reside between the ion source and the mass analyzer. (b) Deceleration lenses after the mass analyzer focus ions onto the sample are seen with alumina support rods and a teflon separator on top. (c) The bottom pole piece and solenoid of the magnetic sector mass analyzer is seen with the entrance from the beamline at the left and the exit leading to the mass-selecting aperture at the bottom. There is a 90◦ bend in the vacuum pipe of the beamline. 418 Figure B.5: Photograph of the mass-selecting aperture used for 22Ne and 12C sam- ples. The aperture consists of a 5 mm diameter hole in a standard 2.75 in size (69.85 mm vacuum flange spacer 16 mm thick. Figure B.6: Photographs of the mass-selecting aperture used for 28Si samples at IC– 1 and LC–2. The aperture consists of a 1 mm wide slit that is 15.25 mm tall and 2 mm thick in a Cu gasket. (a) A new aperture is seen at the exit of the magnetic mass selecting magnet. (b) Closeup of a new aperture. (c) A used aperture had dark deposited spots from the 30 u (far left of the aperture) and 29 u (left of the aperture) ion beams. The separation between the 28Si beam centered on the aperture and the 29Si beam is about 6.2 mm. 419 Figure B.7: Photograph of the mass-selecting aperture used for 28Si samples at DC– 3. The aperture consists of a 2 mm wide slit that is 12 mm tall with a beveled edge in a Cu gasket. The beveled edge is on the back side with the side shown facing the ion beam. Figure B.8: Photograph of a gas aperture on the inlet to the deceleration lenes in the ion beam. The aperture consists of a 12.7 mm by 6.4 mm rectangular opening in a stainless steel shim. The purpose of this gas aperture is to direct gas diffusing from the ion beam chamber around the lenses to then be pumped away. 420 Figure B.9: Photographs of the experimental setup and sample stage for 28Si samples deposited at IC–1 in the ion beam chamber. (a) An electrical vacuum feedthrough at the left serves as a sample stage and is connected to an electric break at the exit of the magnetic sector mass analyzer. (b) The feedthrough flange is seen. (c) A 28Si sample is mounted on the end of the feedthrough using carbon tape. Figure B.10: Photograph of the experimental setup of the deceleration lens chamber for samples produced at LC–2. The ion beam enters the lens chamber from the right. An electrical vacuum feedthrough serves as a sample stage at the right. 421 Figure B.11: Photographs of the sample stage for producing sample at LC–2. (a) The sample stage flange with Cu electrical vacuum feedthrough for mounting sam- ples is seen in the middle. Additional electrical feedthroughs are seen to the left and right connected to the fixed sample aperture in the metal mask. (b) A stain- less steel shim makes up the mask that covers the sample on the feedthrough. The fixed sample aperture consists of a hole about 3 mm in diameter centered over the feedthrough. (c) A Si(100) substrate is seen mounted on the end of the feedthrough using carbon tape. (d) The Cu feedthrough is separated from the rest of the sample stage with a 28Si sample mounted on the end. 422 Figure B.12: Photographs of the manipulator used for heating substates and samples and depositing 28Si samples at DC–3 in the deposition chamber. (a) The manip- ulator with the sample stage is seen in the deposition chamber with the ion beam deceleration lenses behind it. (b) A closeup of the sample stage outside of the chamber with the RH contacts visible at the left, the RH tungsten wire heater in the middle below where the sample sits, and the DH contact visible at the right. 423 Figure B.13: Photographs of interchangeable sample apertures used to measure ion beam currents prior to depositing 28Si samples at DC–3. (a) The side view of an initial sample aperture consisting of a 2.2 mm diameter hole in a stainless steel shim mounted on a sample holder. (b) The aperture in (a) sits on the manipulator and faces the ion beamline to measure ion current. (c) The side view of a later sample aperture consisting of a 2.5 mm diameter hole, and a 1 mm diameter hole in a Mo shim mounted on a sample holder. (d) The aperture in (c) sits on the manipulator to measure ion current. The 2.5 mm diameter aperture is at the center of the shim and the 1 mm diameter aperture is offset to the right. Both apertures are electrically isolated from the sample holders and connect to the DH contact. 424 Appendix C: Substrate Catalog The table shown here gives the specifications and source of the various sub- strates used for producing enriched samples in this work. 425 # Material Dopant ρ Thickness Supplier Identifier (Ω · cm) (µm) (1) Si(100) 200* lightly doped (2) Ag 25 Kurt J. Lesker Company Pure Al Foil (3) Si(100) boron 1 to 10 380 University Wafer Prime lot 12/0604 (4) Si(100) SOI boron 10 to 30 0.035† Dr. Neil Zimmerman (NIST) NIST-2 (5) Si(100) phosphorous 5 to 10 600 Dr. Mike Stewart (NIST) 55 nm thermal Ox (6) Si(100) phosphorous 1 to 5 300 University Wafer 20141231-12 (7) Si(100) boron 7 to 20 340 University Wafer 4/15/15 (8) Si(100) undoped (intrinsic) 2× 104 380 University Wafer FZ SEMI Prime L849 (9) Si(100) boron 5 to 10 300 ITME 3611/110057 (10) Si(100) phosphorous 7 to 20 300 Virginia Semiconductor ± 0.05◦ 15-10973 Table C.1: Substrate catalog for substrates used for depositing enriched films. Substrate numbers are referenced in the sample catalogs. *The wafer thickness of substrate (1) is an estimate. †The thickness of substrate (4) is the thickness of the Si device layer in the SOI stack. This SOI wafer also possessed a 400 nm buried oxide layer. A 55 nm thermal oxide was grown on the surface of the wafer for substrate (5). The wafer for substrate (10) was specified to have a miscut tolerance of ± 0.05◦ with the {100} surface. 426 Appendix D: Sample Catalogs The sample catalogs presented here list all of the enriched samples produced in this work and are organized several ways. The 22Ne sample implantation parameters and measured enrichments are summarized in Table D.1. The 12C sample deposition parameters and measured enrichments are summarized in Table D.2. 28Si samples are organized by their deposition location, IC–1, the ion beam chamber, LC–2, the deceleration lens chamber, and DC–3, the deposition and analysis chamber. 28Si sample deposition parameters for samples deposited at IC–1 are summarized in Ta- ble D.3. 28Si sample deposition parameters for samples deposited at LC–2 are sum- marized in Tables D.4 and D.5. Finally, the 28Si sample deposition parameters for samples deposited at DC–3 are summarized in Tables D.6 to D.9. Enrichment mea- surements and parameters used in the modeling analysis discussed in Chapter 6 are presented for 28Si samples deposited at room temperature in Tables D.10 and D.11. Similarly, the enrichment measurements and modeling parameters for samples de- posited with elevated temperatures are presented in Tables D.12 and D.13. 427 # Sample Load Dep. Substrate Aperture Base P Implant P VA VT Ei Name Date Date Type (mm) (Pa) (Pa) (V) (kV) (eV) 1 110301-22Ne-OxSi 2/28/11 3/1/11 Si(100):Ox (1) 5 1.1× 10−5 404 -3.885 ≈ 3000 2 110418-22Ne-OxSi 3/1/11 4/18/11 Si(100):Ox (1) 5 7.6× 10−7 3.3× 10−5 500 -4.002 4460 3 110504-22Ne-OxSi 4/28/11 5/4/11 Si(100):Ox (1) 5 3.1× 10−6 501 -4.006 4460 Isotope Fraction (zNe/Netot.) # Sample Ii Ion Dose Spot Size Implant Primary 22Ne 20Ne Name (µA) (cm−2) (mm2) (nm) Measurement (%) (%) 1 110301-22Ne-OxSi 0.14* 7.7× 1016* 2* 9 SIMS 88.4(10) 15.2(10) 2 110418-22Ne-OxSi 0.14* 7.7× 1016* 5* 12.5 3 110504-22Ne-OxSi 0.14 7.7× 1016 2* 12.5 SIMS 99.455(36) 0.545(36) Table D.1: 22Ne Sample Catalog (3/1/11–5/4/11). 22Ne samples were implanted at room temperature (≈ 21 ◦C) using Ne in the solids-mode ion source at LC–2, the deceleration lens chamber. Substrates had a native oxide and were not prepared ex situ or in situ. Aperture gives the diameter of the circular mass-selecting aperture. Pressures are raw reading of the lens chamber base and with the ion source on. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. Implant refers to the peak implantation depth. Isotope fraction measurements assume no 21Ne is present. Items marked with * are estimates. 428 # Sample Load Dep. Substrate Aperture Base P Dep. P VA VT Ei Ii Name Date Date Type (mm) (Pa) (Pa) (V) (kV) (eV) (µA) 1 120125-12C-OxSi 1/24/12 1/25/12 Si(100):Ox (1) 5 4.4× 10−6 596 -4.00 543 0.55 2 120206-12C-OxSi 2/1/12 2/6/12 Si(100):Ox (1) 5 608 -4.00 554 0.40 3 120207-12C-OxSi 2/6/12 2/7/12 Si(100):Ox (1) 5 1.7× 10−6 2.1× 10−5 608 -4.00 554 0.70 Isotope Fraction (zC/Ctot.) # Sample C. I. Ion Dose Spot Size R d Film Mass Primary 12C 13C Name (C) (cm−2) (mm2) (nm/min) (nm) (µg) Measurement(s) (%) (ppm) 1 120125-12C-OxSi 0.01 2.7× 1018 2.3 0.68* 206* 0.95 SIMS 99.996107(88) 38.93(88) 2 120206-12C-OxSi 0.00012 4.7× 1016 1.6 0.77* 4* 0.01 SEM 3 120207-12C-OxSi 0.00555 2.9× 1018 1.2 0.76 100 0.52 SIMS, SEM 99.99608(13) 39.2(13) Table D.2: 12C Sample Catalog (1/25/12–2/7/12). 12C samples were deposited at room temperature (≈ 21 ◦C) using CO2 in the solids-mode ion source at LC–2, the deceleration lens chamber. Substrates had a native oxide and were not prepared ex situ or in situ. Aperture gives the diameter of the circular mass-selecting aperture. Pressures are raw reading of the lens chamber base with the ion source on. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. C. I. is the ion current integral. R is the deposition rate, and d is the film thickness. Isotope fraction measurements of sample #1 were not corrected for instrumental error. Items marked with * are estimates. 429 # Sample Load Dep. Substrate Aperture Base P Dep. P VA VT Name Date Date Type (mm) (Pa) (Pa) (V) (kV) 1 120604-28Si-OxSi 6/1/12 6/4/12 Si(100):Ox (1) 1 6.5× 10−6 1.6× 10−4 514 -4.00 2 120613-28Si-OxSi 6/4/12 6/13/12 Si(100):Ox (1) 1 1.2× 10−5 1.7× 10−4 106 -4.315 3 120618-28Si-OxSi 6/14/12 6/18/12 Si(100):Ox (1) 1 1.3× 10−5 2.3× 10−4 106 -4.314 4 120627-28Si-Ag 6/19/12 6/27/12 Ag foil (2) 1 1.3× 10−5 2.1× 10−4 106 -4.312 5 120628-28Si-OxSi 6/27/12 6/28/12 Si(100):Ox (1) 1 1.3× 10−5 1.8× 10−4 106 -4.311 # Sample Ei Ii C. I. Ion Dose Spot Size R d Film Mass Primary Name (eV) (µA) (C) (cm−2) (mm2) (nm/min) (nm) (µg) Measurement 1 120604-28Si-OxSi 455 0.95 0.0102 3.4× 1017 19 0.07 12 1.36 SIMS 2 120613-28Si-OxSi 64 1.10 0.0136 3.2× 1017 27 0.42 86 3.63 SEM 3 120618-28Si-OxSi 64 0.77 0.0114 4.8× 1017 15 0.40 92 3.04 TOF-SIMS 4 120627-28Si-Ag 64 0.95 0.0109 2.5× 1017 27 0.24* 46* 2.90 5 120628-28Si-OxSi 64 0.85 0.0075 2.9× 1017 16 0.59 86 1.98 SIMS Table D.3: 28Si Sample Catalog: IC–1 (6/4/12–6/28/12). 28Si samples were deposited at room temperature (≈ 21 ◦C) using SiH4 with the low pressure mode of the gas-mode ion source at IC–1, the ion beam chamber. Substrates had a native oxide and were not prepared ex situ or in situ. Aperture gives the width of the mass-selecting aperture slit. Pressures are raw reading of the ion beam chamber base and with the ion source on. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. C. I. is the ion current integral. R is the deposition rate, and d is the film thickness. Thicknesses and rates correspond to the thickest measured film area. Items marked with * are estimates. 430 # Sample Load Dep. Substrate Substrate Aperture Base P Dep. P VA Name Date Date Type Prep. (ex situ) (mm) (Pa) (Pa) (V) 6 130204-28Si-Si 2/1/13 2/4/13 Si(100) (1) HF etch 1 3.2× 10−7 5.9× 10−6 100 6a 151109-KD-2-P-Ox-28SiSub 11/9/15 1/1/16 Si(100):P:Ox (5) 2.3× 10−8 2.7× 10−7 7 130206-28Si-Si 2/5/13 2/6/13 Si(100) (1) HF etch 1 7.5× 10−7 7.2× 10−6 113 8 130208-28Si-Si 2/7/13 2/8/13 Si(100) (1) HF etch 1 1.3× 10−6 6.1× 10−6 150 9 130213-28Si-Si 2/11/13 2/13/13 Si(100) (1) HF etch 1 4.5× 10−7 6.4× 10−6 100 10 130215-28Si-Si 2/14/13 2/15/13 Si(100) (1) HF etch 1 7.6× 10−7 4.8× 10−6 150 11 130221-28Si-Si 2/19/13 2/21/13 Si(100) (1) HF etch 1 3.7× 10−7 4.9× 10−6 150 12 130227-28Si-Si 2/22/13 2/27/13 Si(100) (1) HF etch 1 2.0× 10−7 5.7× 10−6 100 13 130304-28Si-Si 3/1/13 3/4/13 Si(100) (1) HF etch 1 6.7× 10−7 6.0× 10−6 300 # Sample VT Ei Ii C. I. Ion Dose Spot Size R d Film Mass Primary Name (kV) (eV) (µA) (C) (cm−2) (mm2) (nm/min) (nm) (µg) Measurement 6 130204-28Si-Si -4.000 74 0.57 0.0067 2.6× 1017 16 0.56 110 1.76 SIMS 6a 151109-KD-2-P-Ox-28SiSub 22 0.56 50 SIMS 7 130206-28Si-Si -3.941 86 0.50 8 0.40* 80* 8 130208-28Si-Si -4.000 121 0.56 0.0081 7.6× 1017 7 1.49 384 1.96 SIMS 9 130213-28Si-Si -4.000 76 0.36 0.0046 2.9× 1017 10 0.17 37 1.20 TEM 10 130215-28Si-Si -3.997 122 0.36 0.0058 4.0× 1017 9 0.25* 67* 1.38 XPS 11 130221-28Si-Si -4.000 124 0.26 0.0035 3.1× 1017 7 0.23* 51* 0.83 SEM 12 130227-28Si-Si -3.998 74 0.5* 0.0021 5.4× 1017 2 1.41* 97* 0.54 13 130304-28Si-Si -4.000 160 0.65 0.0054 6.0× 1017 6 0.67 92* 1.20 SEM Table D.4: 28Si Sample Catalog: LC–2: I (2/4/13–3/4/13). 28Si samples were deposited at room temperature (≈ 21 ◦C) using SiH4 with the low pressure mode of the gas-mode ion source at LC–2, the deceleration lens chamber. Substrates were etched with HF ex situ but not prepared in situ. Aperture gives the width of the mass-selecting aperture slit. Pressures are raw reading of the lens chamber base and with the ion source on. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. C. I. is the ion current integral. R is the deposition rate, and d is the film thickness. Thicknesses and rates correspond to the thickest measured film area. Items marked with * are estimates. Sample #6a was deposited at DC–3 by sublimating sample #6 at about 1100 ◦C. 431 # Sample Load Dep. Substrate Substrate Aperture Base P Dep. P VA Name Date Date Type Prep. (ex situ) (mm) (Pa) (Pa) (V) 14 130307-28Si-OxSi 3/4/13 3/7/13 Si(100) (1) HF etch 1 2.4× 10−7 4.0× 10−6 325* 15 130311-28Si-Si 3/8/13 3/11/13 Si(100) (1) HF etch 1 2.5× 10−7 5.5× 10−6 325* 16 130328-28Si(dP)-Si 3/25/13 3/28/13 Si(100) (1) HF etch 1 2.3× 10−7 3.7× 10 −6 (low) 142* 1.9× 10−4 (high) 17 130404-28Si(dP)-Si 4/1/13 4/4/13 Si(100) (1) HF etch 1 2.4× 10−7 3.9× 10 −6 (low) 130* 3.1× 10−4 (high) 18 130412-2829Si(28SiH)-Si 4/8/13 4/12/13 Si(100):B (3) HF etch 1 1.9× 10−7 3.6× 10−6 116 19 130520-28Si-SOI 4/15/13 5/20/13 Si(100) SOI (4) HF etch 1 1.1× 10−7 3.7× 10−6 229 20 130920-28Si(dP)-SOI 8/2/13 9/20/13 Si(100) SOI (4) HF etch 1 3.6× 10−8 5.1× 10 −6 (low) 100 5.3× 10−5 (high) 21 130924-28Si-Si-550C 9/23/13 9/24/13 Si(100):B (3) HF etch 1 1.1× 10−6 2.1× 10−5 200 # Sample VT Ei Ii C. I. Ion Dose Spot Size R d Film Mass Primary Name (kV) (eV) (µA) (C) (cm−2) (mm2) (nm/min) (nm) (µg) Measurement 14 130307-28Si-OxSi -4.000 230* 0.54 0.0089 6.0× 1017 9 1.52 415 1.94 SEM 15 130311-28Si-Si -4.000 230* 0.60 0.0066 4.3× 1017 10 0.35* 64* 1.45 16 130328-28Si(dP)-Si -3.995 64* 0.42 0.0070 2.7× 1017 6 0.93 249 2.00 SIMS 17 130404-28Si(dP)-Si -3.995 55* 0.43 0.0073 8.8× 1017 5 0.62* 175* 2.10 TOF-SIMS 18 130412-2829Si(28SiH)-Si -3.993 50 0.37 0.0070 9.1× 1017 5 0.84 351 1.92 SIMS 19 130520-28Si-SOI -4.000 179* 0.48 0.0040 3.8× 1016 7 1.07 149 0.89 SIMS 20 130920-28Si(dP)-SOI -4.000 80 0.62 0.0075 8.8× 1017 5 1.41 285 1.92 SIMS 21 130924-28Si-Si-550C -4.000 75 0.41 0.0007 1.5× 1017 3 0.74 22 0.19 Table D.5: 28Si Sample Catalog: LC–2: II (3/7/13–9/24/13). 28Si samples were deposited at room temperature (≈ 21 ◦C) using SiH4 with the low pressure mode of the gas-mode ion source at LC–2, the deceleration lens chamber. Sample #21 was deposited at about 550 ◦C. Substrates were etched with HF ex situ but not prepared in situ. Aperture gives the width of the mass-selecting aperture slit. Pressures are raw reading of the lens chamber base and with the ion source on. For samples #16, #17, and #20, the pressure was modulated from low to high to low. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. C. I. is the ion current integral. R is the deposition rate, and d is the film thickness. Thicknesses and rates correspond to the thickest measured film area. Items marked with * are estimates. 432 # Sample Load Dep. Substrate Substrate Aperture Base P Dep. P VA VT Name Date Date Type Prep. (ex situ) (mm) (Pa) (Pa) (V) (kV) 22 140224-28Si-Si-750C 2/7/14 2/24/14 Si(100):B (3) 2 2.4× 10−8 2.1× 10−7 40 -4.000 23 140521-28Si-Si-800C 3/25/14 5/21/14 Si(100):B (3) 2 4.9× 10−9 2.4× 10−7 35 -4.000 24 140526-28Si-Si-700C 5/20/14 5/26/14 Si(100):B (3) 2 2.9× 10−9 2.0× 10−7 35 -4.000 25 140603-28Si-Si-700C 2/22/14 6/3/14 Si(100):B (3) 2 3.7× 10−9 2.0× 10−7 35 -4.000 26 140604-28Si-Si-600C 2/22/14 6/4/14 Si(100):B (3) 2 5.7× 10−9 2.0× 10−7 35 -4.000 27 140610-28Si-Si-900C 5/23/14 6/10/14 Si(100):B (3) 2 2.8× 10−9 2.0× 10−7 35 -4.000 28 140619-28Si-Si-700C 6/5/14 6/19/14 Si(100):B (3) 2 2.5× 10−9 2.0× 10−7 35 -4.001 29 140627-28Si-Si-1050C 5/28/14 6/27/14 Si(100):B (3) 2 2.8× 10−9 2.7× 10−7 35 -4.000 30 140710-28Si-Si-1010C 6/26/14 7/10/14 Si(100):B (3) 2 4.9× 10−9 1.5× 10−7 36 -4.000 31 140716-28Si-Si-1000C 6/26/14 7/16/14 Si(100):B (3) 2 5.1× 10−9 1.9× 10−7 50 -4.000 # Sample T Ei Ii C. I. Ion Dose Spot Size R d Film Mass Primary Name (◦C) (eV) (µA) (C) (cm−2) (mm2) (nm/min) (nm) (µg) Measurement(s) 22 140224-28Si-Si-750C 759 50 0.18 0.0012 28* 0.21* 22 SIMS, SEM 23 140521-28Si-Si-800C 812 33 0.42 0.0044 3.7× 1017 7 0.90 158 1.20 SIMS, SEM, XPS 24 140526-28Si-Si-700C 708 33 0.58 0.0054 4.1× 1017 8 0.77 120 1.50 SEM, TEM 25 140603-28Si-Si-700C 708 33 0.51 0.0032 4.5× 1017 4 1.20 126 0.90 SIMS, SEM, XPS 26 140604-28Si-Si-600C 610 33 0.50 0.0063 5.3× 1017 7 0.77 162 1.77 SIMS, SEM 27 140610-28Si-Si-900C 920 33 0.53 0.0049 8.3× 1017 4 2.60* 400* 1.38 SEM 28 140619-28Si-Si-700C 708 34 0.53 0.0057 9.9× 1017 4 0.84 150 1.60 SIMS, SEM, TEM 29 140627-28Si-Si-1050C 1085 33 0.60 0.0054 1* 1.50 30 140710-28Si-Si-1010C 1041 33 0.57 0.0060 0.29 50 1.77 SIMS, SEM, TEM 31 140716-28Si-Si-1000C 1030 31 0.55 0.0060 1.69 Table D.6: 28Si Sample Catalog: DC–3: I (2/7/14–6/26/14). 28Si samples were deposited using SiH4 with the low pressure mode of the gas-mode ion source at DC–3, the deposition chamber. Substrates were not prepared ex situ but were flash annealed in situ. Aperture gives the width of the mass-selecting aperture slit. Pressures are raw reading of the deposition chamber base and with the ion source on. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. T is the substrate deposition temperature. C. I. is the ion current integral. R is the deposition rate, and d is the film thickness. Thicknesses and rates correspond to the thickest measured film area. Items marked with * are estimates. 433 # Sample Load Dep. Substrate Substrate Aperture Base P Dep. P VA VT Name Date Date Type Prep. (ex situ) (mm) (Pa) (Pa) (V) (kV) 32 140828-28Si-Si 6/30/14 8/28/14 Si(100):B (4) 2 2.4× 10−9 1.2× 10−7 85 -4.00 33 150202-28Si-Si-800C 1/27/15 2/2/15 Si(100):P (5) HF etch 2 2.1× 10−8 1.0× 10−7 100 -4.00 34 150323-28Si-Si-700C 3/17/15 3/23/15 Si(100):B (4) HF etch 2 1.9× 10−8 3.2× 10−7 100 -4.00 35 150627-KD-7-P-28Si-600C 6/27/15 7/6/15 Si(100):P (6) CMOS clean 2 2.7× 10−8 6.5× 10−7 100 -4.00 36 150627-KD-3-P-28Si-700C 6/27/17 7/9/15 Si(100):P (6) CMOS clean 2 2.1× 10−8 5.5× 10−7 120 -4.00 37 150707-KD-1-P-28Si-800C 7/7/15 7/14/15 Si(100):P (6) CMOS clean 2 1.5× 10−8 7.6× 10−7 108 -4.00 38 150715-KD-5-P-28Si-800C 7/15/15 7/17/15 Si(100):P (6) CMOS clean 2 1.6× 10−8 1.2× 10−6 100 -4.00 39 150719-KD-2-B-28Si-700C 7/19/15 7/22/15 Si(100):B (7) CMOS clean 2 1.5× 10−8 1.0× 10−6 100 -4.00 40 150715-KD-9-B-28Si-700C 7/15/15 7/22/15 Si(100):B (7) CMOS clean 2 1.5× 10−8 1.0× 10−6 100 -4.00 41 150721-KD-5-P-28Si-700C 7/21/15 7/27/15 Si(100):P (6) CMOS clean 2 1.5× 10−8 1.0× 10−6 120 -4.00 # Sample T Ei Ii C. I. Ion Dose Spot Size R d Film Mass Primary Name (◦C) (eV) (µA) (C) (cm−2) (mm2) (nm/min) (nm) (µg) Measurement(s) 32 140828-28Si-Si 21* 38 0.25 0.0025 6.0× 1017 3 0.32 53 0.69 SIMS 33 150202-28Si-Si-800C 804 33 0.41* 2 1.4* 170* SIMS, SEM 34 150323-28Si-Si-700C 709 20 0.32 0.0002 8 0.06 35 150627-KD-7-P-28Si-600C 619 12 0.64 0.0004 0.11 36 150627-KD-3-P-28Si-700C 712 24 0.65 0.0101 1.3× 1018 5 0.54 140 2.89 SIMS, TEM, XPS 37 150707-KD-1-P-28Si-800C 808 34* 0.85 0.0071 1.1× 1018 4 1.57* 217* 1.97 SEM 38 150715-KD-5-P-28Si-800C 808 19 0.64* 0.0077* 2.3× 1018 2 0.60 120 2.13* SIMS, SEM 39 150719-KD-2-B-28Si-700C 712 18 0.72 0.0002 5* 0.05 40 150715-KD-9-B-28Si-700C 712 18 0.67 0.0001 5 0.04 41 150721-KD-5-P-28Si-700C 712 37 0.74 0.0093 1.4× 1018 4 1.22 256 2.60 SIMS Table D.7: 28Si Sample Catalog: DC–3: II (8/28/14–7/27/15). 28Si samples were deposited using SiH4 with the low pressure mode of the gas-mode ion source at DC–3, the deposition chamber. Substrates were flash annealed in situ. Aperture gives the width of the mass-selecting aperture slit. Pressures are raw reading of the deposition chamber base and with the ion source on. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. T is the substrate deposition temperature. C. I. is the ion current integral. R is the deposition rate, and d is the film thickness. Thicknesses and rates correspond to the thickest measured film area. Items marked with * are estimates. Samples #35 to #41 contain roughly 30 % N. 434 # Sample Load Dep. Substrate Substrate Aperture Base P Dep. P VA VT Name Date Date Type Prep. (ex situ) (mm) (Pa) (Pa) (V) (kV) 42 150920-KD-3-P-28Si-308C 9/20/15 10/2/15 Si(100):P (6) CMOS clean 2 1.7× 10−8 1.1× 10−6 45 -4.00 43 150926-KD-4-P-28Si-312C 9/26/15 10/5/15 Si(100):P (6) CMOS clean 2 1.6× 10−8 1.5× 10−6 50 -4.00 44 151020-KD-9-B-28Si-700C 10/20/15 10/22/15 Si(100):B (7) CMOS clean 2 2.3× 10−8 1.3× 10−6 50 -4.00 45 151019-KD-2-B-28Si-450C 10/19/15 10/22/15 Si(100):B (7) CMOS clean 2 2.3× 10−8 1.3× 10−6 50 -4.00 46 151025-KD-4-B-28Si-400C 10/25/15 11/2/15 Si(100):B (7) CMOS clean 2 1.6× 10−8 9.6× 10−7 50 -4.00 47 151028-KD-1-B-28Si-700C 10/28/15 11/20/15 Si(100):B (7) CMOS clean 2 1.5× 10−8 1.3× 10−6 50 -4.00 48 151102-KD-7-i-28Si-400C 11/2/15 11/30/15 Si(100):i (8) CMOS clean 2 1.5× 10−8 1.2× 10−6 50 -4.00 49 151202-KD-7-i-28Si-400C 12/2/15 12/7/15 Si(100):i (8) CMOS clean 2 1.5× 10−8 1.5× 10−6 50 -4.00 50 151109-KD-9-B-28Si-400C 11/9/15 12/10/15 Si(100):B (9) CMOS clean 2 1.7× 10−8 1.3× 10−6 50 -4.00 51 151109-KD-4-B-28Si-300C 11/9/15 12/14/15 Si(100):B (9) CMOS clean 2 1.5× 10−8 1.3× 10−6 50 -4.00 # Sample T Ei Ii C. I. Ion Dose Spot Size R d Film Mass Primary Name (◦C) (eV) (µA) (C) (cm−2) (mm2) (nm/min) (nm) (µg) Measurement(s) 42 150920-KD-3-P-28Si-308C 355 40 0.53 0.0003 1.14* 9* 0.07 43 150926-KD-4-P-28Si-312C 357 46 0.55 0.0050 1.1× 1018 3 0.33 50 1.36 SIMS, XPS 44 151020-KD-9-B-28Si-700C 712 43 0.54 0.0003 1 1.11* 11* 0.09 45 151019-KD-2-B-28Si-450C 460 43 0.54 0.0097 1.6× 1018 4 0.53 160 2.69 SIMS, TEM 46 151025-KD-4-B-28Si-400C 721 42 0.47* 0.0036 6.2× 1017 4 0.21 44 0.99* SIMS 47 151028-KD-1-B-28Si-700C 705 39 0.55 0.0058 4.6× 1017 8 0.82 144 1.62 SIMS, SEM 48 151102-KD-7-i-28Si-400C 417 40 0.55 0.0079 6.8× 1017 7 0.63 150 2.20 ESR 49 151202-KD-7-i-28Si-400C 417 40 0.54 0.0087 1.1× 1018 5 0.93 250 2.42 ESR 50 151109-KD-9-B-28Si-400C 421 38 0.54 0.0082 5.6× 1017* 9* 0.60 151 2.29 SIMS 51 151109-KD-4-B-28Si-300C 349 37 0.51 0.0074 8.5× 1017 5 0.87 210 2.05 SIMS Table D.8: 28Si Sample Catalog: DC–3: III (9/20/15–11/9/15). 28Si samples were deposited using SiH4 with the low pressure mode of the gas-mode ion source at DC–3, the deposition chamber. Substrates were flash annealed in situ. Aperture gives the width of the mass-selecting aperture slit. Pressures are raw reading of the deposition chamber base and with the ion source on. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. T is the substrate deposition temperature. C. I. is the ion current integral. R is the deposition rate, and d is the film thickness. Thicknesses and rates correspond to the thickest measured film area. Items marked with * are estimates. 435 # Sample Load Dep. Substrate Substrate Aperture Base P Dep. P Plasma VA Name Date Date Type Prep. (ex situ) (mm) (Pa) (Pa) Mode (V) 52 151210-KD-7-i-28Si-500C 12/10/15 12/18/15 Si(100):i (8) CMOS clean 2 1.5× 10−8 1.2× 10−6 low P 50 53 151218-KD-4-i-28Si-200C 12/18/15 12/21/15 Si(100):i (8) CMOS clean 2 1.7× 10−8 1.2× 10−6 low P 57 54 151218-KD-9-B-28Si-400C 12/18/15 12/29/15 Si(100):B (9) CMOS clean 2 1.7× 10−8 1.2× 10−6 low P 69 55 160113-KD-4-P-28Si-400C 1/13/16 2/8/16 Si(100):P (10) CMOS clean 2 2.4× 10−8 5.3× 10−6 high P 55 56 160202-KD-7-P-28Si-400C 2/2/16 2/10/16 Si(100):P (10) CMOS clean 2 2.1× 10−8 5.2× 10−6 high P 55 57 160210-KD-1-P-28Si-400C 2/10/16 2/12/16 Si(100):P (10) CMOS clean 2 2.1× 10−8 5.1× 10−6 high P 55 58 160217-KD-7-P-28Si-450C 2/17/16 2/29/16 Si(100):P (10) CMOS clean 2 1.7× 10−8 5.5× 10−6 high P 55 59 160513-KD-7-i-28Si-450C 5/13/16 6/9/16 Si(100):i (8) CMOS clean 2 8.3× 10−9 3.6× 10−6 high P 52 60 160617-KD-3-P-28Si-400C 6/17/16 6/24/16 Si(100):P (10) CMOS clean 2 7.1× 10−9 3.4× 10−6 high P 52 61 160617-KD-7-P-28Si-403C 6/17/16 6/24/16 Si(100):P (10) CMOS clean 2 7.1× 10−9 2.9× 10−6 high P 50 # Sample VT T Ei Ii C. I. Ion Dose Spot Size R d Film Mass Primary Name (kV) (◦C) (eV) (µA) (C) (cm−2) (mm2) (nm/min) (nm) (µg) Measurement(s) 52 151210-KD-7-i-28Si-500C -4.00 502 37 0.52 0.0071 7.4× 1017 6 0.95 321 1.99 SIMS 53 151218-KD-4-i-28Si-200C -4.00 249 38 0.63 0.0091 6.6× 1017 9 1.27 305 2.54 SIMS 54 151218-KD-9-B-28Si-400C -4.00 421 40 0.69 0.0110 1.0× 1018 7 0.87 232 3.06 SIMS 55 160113-KD-4-P-28Si-400C -4.70 421 35 2.29 0.0178 1.5× 1018 7 2.48* 322* 4.99 56 160202-KD-7-P-28Si-400C -4.70 421 35 2.27 0.0190 1.4× 1018 9 2.57 359* 5.33 57 160210-KD-1-P-28Si-400C -4.70 421 36 2.28 0.0178 1.3× 1018 9 2.65* 345* 4.98 58 160217-KD-7-P-28Si-450C -4.70 460 37 2.15 0.0006 2.60* 13* 0.18 59 160513-KD-7-i-28Si-450C -4.70 460 34 2.77 0.0220 1.4× 1018 10 2.21 292 6.18 SIMS 60 160617-KD-3-P-28Si-400C -4.70 421 31 3.00 0.0144 2.4× 1018 4 4.56 365 4.07 SIMS, TEM 61 160617-KD-7-P-28Si-403C -4.70 423 32 2.50 0.0150 6.2× 1018 2 2.33* 233* 4.23 Table D.9: 28Si Sample Catalog: DC–3: IV (12/18/15–6/24/16). 28Si samples were deposited using SiH4 with the gas-mode ion source at DC–3, the deposition chamber. Substrates were flash annealed in situ. Aperture gives the width of the mass-selecting aperture slit. Pressures are raw reading of the deposition chamber base and with the ion source on. VA is the anode voltage and VT is the transport voltage. Ei is the average ion energy at the sample, and Ii is the average ion current. T is the substrate deposition temperature. C. I. is the ion current integral. R is the deposition rate, and d is the film thickness. Thicknesses and rates correspond to the thickest measured film area. Items marked with * are estimates. 436 Sample Data Dep. Ei Ii R PSiHx Fi Fg Name Set Location (eV) (µA) (nm/min) (Pa) (cm−2 · s−1) (cm−2 · s−1) 120604-28Si-OxSi 1 IC–1 455 0.95 0.07 2.4 ×10−5 1.1 ×1013 6.6 ×1013 120618-28Si-OxSi Ar IC–1 64 0.77 0.40 3.4 ×10−5 3.5 ×10 13 9.2 ×1013 SF5 0.33 2.8 ×1013 120628-28Si-OxSi 4 IC–1 64 0.85 0.5 2.6 ×10−5 4.4 ×1013 7.2 ×10135 0.5 4.4 ×10 13 7 0.47 4.2 ×1013 8 0.59 5.2 ×1013 130204-28Si-Si 1 LC–2 74 0.57 0.51 1.4 ×10−6 4.6 ×10 13 3.7 ×1012 2 0.56 5.1 ×1013 130208-28Si-Si 1 LC–2 121 0.56 1.49 1.4 ×10−6 1.5 ×1014 3.7 ×1012 2 1.05 1.1 ×1014 3 0.54 5.6 ×1013 4 0.53 5.4 ×1013 5 0.94 9.6 ×1013 130328-28Si(dP)-Si low-P LC–2 60 0.33 0.93 6.8 ×10−7 8.3 ×1013 1.8 ×1012 130412-2829Si(28SiH)-Si dip/film LC–2 50 0.31 0.84 6.8 ×10−7 7.3 ×1013 1.8 ×1012 130520-28Si-SOI 1 LC–2 180 0.48 1.07 8.3 ×10−7 1.2 ×1014 2.3 ×1012 140828-28Si-Si 1 DC–3 38 0.25 0.32 2.8 ×10−7 2.7 ×1013 7.6 ×1011 Table D.10: Deposition parameters of 28Si samples deposited at room temperature (≈ 21 ◦C) for the gas sticking deposition model analysis discussed in Chapter 6. Samples deposited at IC–1 were excluded in that analysis. Some samples have multiple SIMS data sets that are grouped together. Ei is the average ion energy at the sample. Ii is the average 28Si ion beam current. R is the deposition rate. PSiHx is the SiHx partial pressure at the sample during deposition. Fi is the 28Si ion flux. Fg is the SiHx molecular flux at the sample. 437 Isotope Fraction (zSi/Sitot.) cz(sT , k502) Sample Data Dep. d 28Si 29Si 30Si 29Si 30Si Name Set Location (nm) (%) (ppm) (ppm) (ppm) (ppm) 120604-28Si-OxSi 1 IC–1 12 99.702(12) 2822(18) 156.1(38) 0.112 120618-28Si-OxSi Ar IC–1 92 99.8850(14) 1130(14) 20.3(14) 0.032 SF5 75 99.8448(21) 1534(21) 17.9(19) 0.0228 120628-28Si-OxSi 4 IC–1 73 99.99846(19) 9.5(10) 5.9(16) 0.043 0.027 5 73 99.99842(16) 10.5(13) 5.23(87) 0.047 0.024 7 69 99.99816(11) 12.3(10) 6.15(52) 0.052 0.026 8 86 99.99799(16) 12.8(10) 7.3(13) 0.068 0.039 130204-28Si-Si 1 LC–2 100 99.999657(38) 2.02(32) 1.41(20) 0.181 0.126 2 110 99.999592(36) 2.30(26) 1.78(24) 0.227 0.176 130208-28Si-Si 1 LC–2 384 99.9998308(82) 0.993(64) 0.699(51) 0.299 0.210 2 270 99.9998031(86) 1.252(72) 0.717(48) 0.265 0.152 3 140 99.999722(12) 1.724(97) 1.055(74) 0.189 0.116 4 136 99.999696(15) 1.85(11) 1.189(96) 0.197 0.127 5 241 99.999812(10) 1.122(76) 0.760(64) 0.212 0.144 130328-28Si(dP)-Si low-P LC–2 249 99.999888(10) 0.691(74) 0.432(67) 0.225 0.141 130412-2829Si(28SiH)-Si dip/film LC–2 351 99.999889(11) 0.78(10) 0.512(42) 0.223 0.147 130520-28Si-SOI 1 LC–2 149 99.999863(16) 0.77(11) 0.60(11) 0.289 0.225 140828-28Si-Si 1 DC–3 53 99.99990(11) 0.58(26) 0.44(23) 0.148 0.113 Table D.11: Enrichment measurements and gas sticking deposition model model analysis results for 28Si samples deposited at room temperature (≈ 21 ◦C) discussed in Chapter 6. Some samples have multiple SIMS data sets that are grouped together. d is the 28Si film thickness. The raw SIMS isotope fractions as well as those adjusted to the deposition conditions of the 502 ◦C sample, cz(sT , k502), are listed. 438 Sample Data T Ei Ii R PSiHx Fi Fg Name Set (◦C) (eV) (µA) (nm/min) (Pa) (cm−2 · s−1) (cm−2 · s−1) 151218-KD-4-i-28Si-200C 1 249 38 0.63 1.27 3.2 ×10−7 1.1 ×10 14 8.7 ×1011 2 1.11 9.6 ×1013 151109-KD-4-B-28Si-300C 3 349 37 0.51 0.80 3.6 ×10−7 6.9 ×10 13 9.8 ×1011 4 0.87 7.5 ×1013 150926-KD-4-P-28Si-312C 1 357 46 0.55 0.33 4.1 ×10−7 2.9 ×1013 1.1 ×1012 151109-KD-9-B-28Si-400C 1 421 38 0.54 0.43 3.7 ×10−7 3.7 ×1013 9.9 ×10114 0.36 3.1 ×1013 6 0.60 5.1 ×1013 160617-KD-3-P-28Si-400C 1 421 31 3.00 3.94 9.6 ×10−7 3.4 ×10 14 2.6 ×1012 2 3.36 2.9 ×1014 151210-KD-7-i-28Si-500C 2 502 37 0.52 0.95 3.3 ×10−7 8.1 ×1013 8.8 ×10113 0.43 3.7 ×1013 4 1.41 1.2 ×1014 140604-28Si-Si-600C 1 610 33 0.50 0.77 1.4 ×10−7 6.6 ×1013 3.7 ×1011 151028-KD-1-B-28Si-700C 1 705 39 0.55 0.82 3.5 ×10−7 7.0 ×10 13 9.5 ×1011 2 0.66 5.7 ×1013 140603-28Si-Si-700C 1 708 33 0.51 1.20 2.2 ×10−7 1.0 ×10 14 6.0 ×1011 2 0.98 8.3 ×1013 140521-28Si-Si-800C 1 812 33 0.42 0.90 1.5 ×10−7 7.7 ×1013 4.1 ×1011 Table D.12: Deposition parameters of 28Si samples deposited at elevated temperature for the gas sticking deposition model analysis discussed in Chapter 6. Some samples have multiple SIMS data sets that are grouped together. T is the substrate deposition temperature. Ei is the average ion energy at the sample. Ii is the average 28Si ion beam current. R is the deposition rate. PSiHx is the SiHx partial pressure at the sample during deposition. Fi is the 28Si ion flux. Fg is the SiHx molecular flux at the sample. 439 Isotope Fraction (zSi/Sitot.) cz(sT , k502) Sample Data T d 28Si 29Si 30Si 29Si 30Si s Name Set (◦C) (nm) (%) (ppm) (ppm) (ppm) (ppm) 151218-KD-4-i-28Si-200C 1 249 305 99.999898(13) 0.79(12) 0.229(64) 0.72 0.208 1.6 ×10−3 2 267 99.999865(18) 1.01(17) 0.341(65) 0.80 0.271 151109-KD-4-B-28Si-300C 3 349 193 99.999884(18) 0.78(16) 0.386(88) 0.40 0.197 8 ×10−4 4 210 99.999924(17) 0.53(16) 0.230(46) 0.29 0.128 150926-KD-4-P-28Si-312C 1 357 50 99.999405(93) 4.18(70) 1.77(61) 0.80 0.34 2.0 ×10−3 151109-KD-9-B-28Si-400C 1 421 108 99.999762(27) 1.68(24) 0.70(14) 0.452 0.188 4 92 99.999812(25) 1.30(22) 0.58(12) 0.298 0.133 6 151 99.999818(26) 1.46(22) 0.35(15) 0.55 0.132 8.4 ×10−4 160617-KD-3-P-28Si-400C 1 421 315 99.9999594(72) 0.303(58) 0.103(43) 0.285 0.097 2 269 99.9999446(67) 0.407(59) 0.058(32) 0.327 0.068 151210-KD-7-i-28Si-500C 2 502 216 99.9999701(57) 0.259(53) 0.040(19) 0.174 0.027 2.9 ×10−43 99 99.999940(16) 0.31(10) 0.29(12) 0.096 0.089 4 321 99.9999819(35) 0.127(29) 0.055(19) 0.127 0.055 140604-28Si-Si-600C 1 610 162 99.9999570(70) 0.300(60) 0.130(37) 0.38 0.166 9 ×10−4 151028-KD-1-B-28Si-700C 1 705 144 99.999488(48) 3.30(25) 1.82(42) 1.77 0.98 5.6 ×10−3 2 117 99.999181(83) 4.82(70) 3.38(45) 2.11 1.48 140603-28Si-Si-700C 1 708 126 99.99986(11) 0.806(97) 0.61(37) 1.00 0.76 7 ×10−3 2 102 99.99947(18) 3.09(61) 2.17(42) 3.12 2.19 140521-28Si-Si-800C 1 812 158 99.99907(25) 4.32(46) 4.96(93) 5.9 6.8 2.3 ×10−2 Table D.13: Enrichment measurements and temperature dependant gas incorporation model analysis results for 28Si samples deposited at elevated temperature discussed in Chapter 6. Some samples have multiple SIMS data sets that are grouped together. T is the substrate deposition temperature. d is the 28Si film thickness. The raw SIMS isotope fractions as well as those adjusted to the deposition conditions of the 502 ◦C sample, cz(sT , k502), are listed. s is the incorporation fraction at each deposition temperature. 440 Appendix E: SIMS Measurement Settings 22Ne: The following is a statement of the SIMS measurement settings and tech- niques used for assessing the enrichment of 22Ne samples implanted at LC–2. This statement was adapted from one provided by Dr. Dave Simons (NIST) for sample 110504-22Ne-OxSi. A Cameca IMS-1280 magnetic sector secondary ion mass spectrometer was used to make the measurements. The primary ion beam was composed of Cs+ ions with a current of 40 nA and an energy with respect to ground of 10 keV. The beam was raster-scanned over a nominal 280 µm by 430 µm in size, and the ions were only accepted for counting from a central region that represented about 40 % of the area of the crater. At implantation doses used for these samples, it is expected that the peak concentration of implanted species would occur at the surface and would be constant roughly up to the range of the implanted ions. The maximum concentration is governed by a balance between implantation and sputter removal rates. The most sensitive method to detect noble gases by SIMS is via Cs attach- ment ions CsM+, where M is the noble gas isotope. Thus positive secondary ions 441 of 133Cs20Ne+, 133Cs22Ne+, and 133Cs29Si+ were accelerated to 5 keV, separated by mass, and detected with a secondary electron multiplier operating in a pulse- counting mode. These species were detected sequentially and repetitively over 50 cycles, with the Ne species being detected for 5 s each per cycle and the Si for 1 s. The sequence of data constitutes a depth profile since the sample surface is eroded with time by ion sputtering. Depth profile data were taken both in the area designated as having been implanted, as well as far from that area to serve as a control. 12C: The following is a statement of the SIMS measurement settings and techniques used for assessing the enrichment of 12C samples deposited at LC–2. This statement was adapted from one provided by Dr. David Simons (NIST) for sample 120207- 12C-OxSi. The isotopic composition of deposited carbon films were measured with a Cameca IMS-1270 large geometry secondary ion mass spectrometer. A Cs+ primary ion beam with an impact energy of 20 keV was raster-scanned over a 150 µm by 150 µm area on the sample in the center of the deposited region. Secondary negative ion signals of 12C and 13C were extracted from a gated area of 125 µm by 125 µm into the mass spectrometer and alternately directed by magnetic peak-switching into a secondary electron multiplier where their count rates were recorded. A mass resolving power m ∆m of 5000 at 10 % of peak amplitude was set by a combination of entrance and exit slits so that a spectral interference of 12CH on 13C was effectively 442 excluded. A set of measurements were made by pre-sputtering the area for 90 s or 360 s with an 8 nA primary ion current and then recording 40 isotopic ratio pairs with a beam current of 23 pA. The count rates were corrected for the dead time of the electron multiplier. The isotope profiles are the results of 8 successive sets of measurements of 13C/12C ratios from the same area of the carbon film. These values do not take into account an instrumental mass fractionation for carbon that would make mea- sured ratios smaller by about 5 %. A stylus profilometry is used to measure the sputtered crater depth. To assess the degree of enrichment of 12C in the films, a ratio measurement of a pyrolytic graphite disk with isotopically natural carbon was made under similar analysis conditions. In this case the measured 13C/12C ratio was 1.046× 10−2. 28Si at IC–1: The following is a statement of the SIMS measurement settings and techniques used for assessing the enrichment of 28Si samples deposited at IC–1. This statement was adapted from one provided by Dr. Dave Simons (NIST) for sample 120604- 28Si-OxSi. The mass spectrometer was set to exclude 28SiH from the 29Si signal and was used to first measure isotopic ratios in the Si substrate far from the film deposit. The primary beam species was O− with an impact energy of 23 keV and a current of about 0.4 nA. The analyzed area was 25 µm by 25 µm, defined by an aperture in the ion optics. The beam was not rastered and the spot had an elliptical shape 443 about 50 µm on the major axis. The measured ratio of 29Si/28Si was 0.0497 and the measured ratio of 30Si/28Si was 0.0323 in the Si substrate. These values are smaller than the known ratios of 0.0508 for 29Si/28Si and 0.03353 for 30Si/28Si but that result is expected since SIMS is known to ionize the lighter isotope with higher efficiency. These measured values are used to calculate the degree to which the minor isotopes have been reduced in the deposited film. The film area was then analyzed in a region that appeared to be the center, assumed to have the greatest thickness. 20 sequential analytical runs were made on the same area, sputtering in total for nearly 4 h. The count rates of 28Si, 29Si and 30Si were recorded sequentially. The count rates were corrected to account for detector dead time. The crater depth was later measured with a stylus profilometer. 28Si at LC–2: The following is a statement of the SIMS measurement settings and techniques used for assessing the enrichment of 28Si samples deposited at LC–2. This statement was adapted from one provided by Dr. Dave Simons (NIST) for sample 130208-28Si- Si. Isotopic measurements were made in a CAMECA IMS-1280 large geometry secondary ion mass spectrometer. The sample was bombarded with a primary ion beam of O+2 at an impact energy of 8 keV and a current of 0.3 nA. The beam was focused to a probe size of a few micrometers diameter and it was raster-scanned over a 50 µm by 50 µm area. Positive secondary ions were accepted for detection from the central 25 µm by 25 µm portion of the rastered area. The secondary ions passed 444 through an entrance slit of 20 µm and an exit slit of 200 µm. These conditions produce a mass resolving power of about 6000 ( m ∆m at 10 % of peak maximum). This resolving power is necessary to separate the 29Si peak from the 28SiH peak that is produced during the SIMS process. The 28SiH signal is normally about 0.05 % to 0.1 % of the 28Si signal and the ratio can vary depending on the analytical conditions. For one sample the 28SiH signal within the film was about 4000 times larger than the 29Si signal demonstrating why very good separation between these peaks is needed to measure the 29Si/28Si ratio accurately. Data were taken over 100 to 360 cycles in which the 28Si was measured for 1 s, the 29Si for 10 s, 28SiH for 1 s and the 30Si for 10 s sequentially during each cycle by switching the magnetic field. The runs took between 50 min and 100 min to sputter through the film depending on the thickness of the film at the analysis location. Quantitative isotopic ratio calculations were made by averaging the cycle-by- cycle ratio measurements in the part of the profiles where the ratios were at a minimum. These values were then corrected for instrumental mass fractionation based on the differences between the measured ratios of the wafer silicon and the accepted natural values. Uncertainties were determined from the standard deviation of the mean of the measurements and were compared with a Poisson estimation based on the total number of detected counts of the minor isotopes. The depths of each crater were measured with a stylus profilometer and the film thickness was defined as the depth where the 29Si/28Si ratio had risen to half of its natural value. Carbon isotopic signals were monitored in a separate depth profile of the film. 30Si was also monitored as a marker of the film-substrate interface and as a normal- 445 ization signal for concentration estimates. The concentrations of the carbon isotopes in the film were estimated by averaging their count rates over the cycles within the film and applying a value obtained from literature of 0.007 for the relative sensitivity factor of carbon to silicon by SIMS under similar analytical conditions. The SIMS instrumental background for carbon analyzed under similar conditions is unknown but it must be at least an order of magnitude lower than what is measured since the carbon signals typically decrease by more than a factor of 10 in the Si substrate. Trace carbon is not normally analyzed by SIMS under these conditions of oxygen ion bombardment and positive ion detection but rather with cesium ion bombardment and negative ion detection. In the present case the conditions were not changed to simplify the measurement process. 28Si at DC–3: The following is a statement of the SIMS measurement settings and techniques used for assessing the enrichment of 28Si samples deposited at DC–3. This statement was adapted from one provided by Dr. Dave Simons (NIST) as a general procedure for measuring samples deposited at elevated temperatures. Isotopic measurements of Si were made in a CAMECA IMS-1270E7 large geometry secondary ion mass spectrometer. The samples were bombarded with a primary ion beam of O+2 ions at an impact energy of 8 keV and a current of 1 nA. The beam was focused to a probe size of a few micrometers diameter and it was raster-scanned over a 50 µm by 50 µm area. Positive secondary ions were accepted for detection from the central 20 µm by 20 µm portion of the rastered area as 446 defined by a field aperture in a focal plane of the mass spectrometer. The entrance and exit slits of the spectrometer were selected to produce a mass resolving power of about 6000 ( m ∆m at 10 % of peak maximum). This resolving power is necessary to separate cleanly the 29Si peak from the 28SiH peak that is produced during the SIMS process. Under these conditions we estimate that less than 10−5 of the 28SiH signal contributes to the 29Si measurement. Depth profiles of the Si isotopes 28Si, 29Si and 30Si through deposited films were acquired by monitoring 28Si for 1 s, 29Si for 10 s, 28SiH for 1 s and 30Si for 10 s in each data cycle and collecting a sufficient number of data cycles until the profile penetrated into the silicon substrate. The sputter rate as determined by measuring the final crater depths with a stylus profilometer was approximately 0.15 nm/s under these conditions. Isotope ratios of 29Si/28Si and 30Si/28Si were calculated on a cycle-by-cycle ba- sis. Average isotopic ratios for a film were calculated by averaging the cycle-by-cycle ratio measurements in the portion of a profile where the ratios were at a relatively constant minimum value. These values were then corrected for instrumental mass fractionation based on the differences between the measured ratios from a Si wafer and the accepted natural values. Uncertainties were determined from the standard deviation of the mean of the measurements and were usually similar to Poisson estimations based on the total number of detected counts of the minor isotopes. 447 Appendix F: 28Si Deposition Fun Facts (1) the total mass of 28Si deposited in this work: ≈ 112 µg (2) the total area deposited for 28Si samples: ≈ 460 mm2 (3) the total number of 28Si atoms deposited in this work: ≈ 2.6× 1018 (4) the total time depositing 28Si: ≈ 169 h ≈ 7 days (5) the resulting film thickness if all 28Si samples were deposited at once with a spot size of 7.5 mm2: ≈ 6.5 µm (6) the average distance between 28Si ions along the beamline with an energy of 4050 eV for 0.5 µA of ion current: ≈ 54 nm (7) the mean free path of SiH4 molecules at 21 ◦C for a pressure of 1.3× 10−4 Pa (1.0× 10−6 Torr): λ ≈ 200 m 448 Bibliography [1] Jonathan Wood. “The top ten advances in materials science”. Mater. Today, 11(1-2):40–45, 2008. [2] Floris A. Zwanenburg, Andrew S. Dzurak, Andrea Morello, Michelle Y. Sim- mons, Lloyd C. L. Hollenberg, Gerhard Klimeck, Sven Rogge, Susan N. Cop- persmith, and Mark A. Eriksson. “Silicon quantum electronics”. Rev. Mod. Phys., 85(3):961–1019, 2013. [3] B. E. Kane. “A silicon-based nuclear spin quantum computer”. Nature, 393(6681):133–137, 1998. [4] VeldhorstM, J. C. C. Hwang, C. H. Yang, A. W. Leenstra, B. de Ronde, J. P. Dehollain, J. T. Muhonen, F. E. Hudson, K. M. Itoh, MorelloA, and A. S. Dzurak. “An addressable quantum dot qubit with fault-tolerant control- fidelity”. Nat. Nano., 9(12):981–985, 2014. [5] M. Fuechsle, S. Mahapatra, F. A. Zwanenburg, M. Friesen, M. A. Eriksson, and M. Y. Simmons. “Spectroscopy of few-electron single-crystal silicon quan- tum dots”. Nat. Nanotechnol., 5(7):502–505, 2010. [6] M. Veldhorst, C. H. Yang, J. C. C. Hwang, W. Huang, J. P. Dehollain, J. T. Muhonen, S. Simmons, A. Laucht, F. E. Hudson, K. M. Itoh, A. Morello, and A. S. Dzurak. “A two-qubit logic gate in silicon”. Nature, 526(7573):410–414, 2015. [7] Keiichi Edamatsu. “Quantum physics: swift control of a single spin”. Nature, 456(7219):182–183, 2008. [8] John P. Ridgway. “Cardiovascular magnetic resonance physics for clinicians: part I”. J. Cardiovasc. Magn. Reson., 12(1):71, 2010. [9] M. Steger, K. Saeedi, M. L. W. Thewalt, J. J. L. Morton, H. Riemann, N. V. Abrosimov, P. Becker, and H. J. Pohl. “Quantum information storage for 449 over 180 s using donor spins in a Si-28 ‘Semiconductor Vacuum’”. Science, 336(6086):1280–1283, 2012. [10] Juha T. Muhonen, Juan P. Dehollain, Arne Laucht, Fay E. Hudson, Rach- pon Kalra, Takeharu Sekiguchi, Kohei M. Itoh, David N. Jamieson, Jeffrey C. McCallum, Andrew S. Dzurak, and Andrea Morello. “Storing quantum infor- mation for 30 seconds in a nanoelectronic device”. Nat. Nano., 9(12):986–991, 2014. [11] Eisuke Abe, Alexei M. Tyryshkin, Shinichi Tojo, John J. L. Morton, Wayne M. Witzel, Akira Fujimoto, Joel W. Ager, Eugene E. Haller, Junichi Isoya, Stephen A. Lyon, Mike L. W. Thewalt, and Kohei M. Itoh. “Electron spin coherence of phosphorus donors in silicon: Effect of environmental nuclei”. Phys. Rev. B, 82(12), 2010. [12] Wayne M. Witzel, Malcolm S. Carroll, Lukasz Cywin´ski, and S. Das Sarma. “Quantum decoherence of the central spin in a sparse system of dipolar coupled spins”. Phys. Rev. B, 86(3):35452, 2012. [13] Alexei M. Tyryshkin, Shinichi Tojo, John J. L. Morton, Helge Riemann, Nikolai V. Abrosimov, Peter Becker, Hans-Joachim Pohl, Thomas Schenkel, Michael L. W. Thewalt, Kohei M. Itoh, and S. A. Lyon. “Electron spin coher- ence exceeding seconds in high-purity silicon”. Nat. Mater., 11(2):143–147, 2012. [14] Kamyar Saeedi, Stephanie Simmons, Jeff Z. Salvail, Phillip Dluhy, Helge Riemann, Nikolai V. Abrosimov, Peter Becker, Hans-Joachim Pohl, John J. L. Morton, and Mike L. W. Thewalt. “Room-temperature quantum bit storage exceeding 39 minutes using ionized donors in silicon-28”. Science, 342(6160):830–833, 2013. [15] S. Meiboom and D. Gill. “Modified spin-echo method for measuring nuclear relaxation times”. Rev. Sci. Instrum., 29(8):688–691, 1958. [16] Terry Gullion, David B. Baker, and Mark S. Conradi. “New, compensated Carr-Purcell sequences”. J. Magn. Reson., 89(3):479–484, 1990. [17] B. Weber, S. Mahapatra, T. F. Watson, and M. Y. Simmons. “Engineering in- dependent electrostatic control of atomic-scale (similar to 4 nm) silicon double quantum dots”. Nano Letters, 12(8):4001–4006, 2012. [18] B. Weber, S. Mahapatra, H. Ryu, S. Lee, A. Fuhrer, T. C. G. Reusch, D. L. Thompson, W. C. T. Lee, G. Klimeck, L. C. L. Hollenberg, and M. Y. Sim- mons. “Ohm’s law survives to the atomic scale”. Science, 335(6064):64–67, 2012. [19] H. Buch, S. Mahapatra, R. Rahman, A. Morello, and M. Y. Simmons. “Spin readout and addressability of phosphorus-donor clusters in silicon”. Nature Commun., 4:1038, 2013. 450 [20] J. A. Miwa, J. A. Mol, J. Salfi, S. Rogge, and M. Y. Simmons. “Transport through a single donor in p-type silicon”. Appl. Phys. Lett., 103(4):043106, 2013. [21] J. J. Pla, K. Y. Tan, J. P. Dehollain, W. H. Lim, J. J. L. Morton, F. A. Zwanenburg, D. N. Jamieson, A. S. Dzurak, and A. Morello. “High-fidelity readout and control of a nuclear spin qubit in silicon”. Nature, 496(7445):334– 338, 2013. [22] C. H. Yang, A. Rossi, R. Ruskov, N. S. Lai, F. A. Mohiyaddin, S. Lee, C. Tahan, G. Klimeck, A. Morello, and A. S. Dzurak. “Spin-valley lifetimes in a silicon quantum dot with tunable valley splitting”. Nature Commun., 4:3069, 2013. [23] L. A. Tracy, T. M. Lu, N. C. Bishop, G. A. Ten Eyck, T. Pluym, J. R. Wendt, M. P. Lilly, and M. S. Carroll. “Electron spin lifetime of a single antimony donor in silicon”. Appl. Phys. Lett., 103(14):143115, 2013. [24] A. Wild, J. Kierig, J. Sailer, J. W. Ager, E. E. Haller, G. Abstreiter, S. Ludwig, and D. Bougeard. “Few electron double quantum dot in an isotopically purified Si-28 quantum well”. Appl. Phys. Lett., 100(14):143110, 2012. [25] J. Y. Li, C. T. Huang, L. P. Rokhinson, and J. C. Sturm. “Extremely high electron mobility in isotopically-enriched Si-28 two-dimensional electron gases grown by chemical vapor deposition”. Appl. Phys. Lett., 103(16):162105, 2013. [26] C. C. Lo, A. Persaud, S. Dhuey, D. Olynick, F. Borondics, M. C. Martin, H. A. Bechtel, J. Bokor, and T. Schenkel. “Device fabrication and transport measurements of FinFETs built with Si-28 SOI wafers toward donor qubits in silicon”. Semicond. Sci. Technol., 24(10):105022, 2009. [27] John J. L. Morton, Alexei M. Tyryshkin, Richard M. Brown, Shyam Shankar, Brendon W. Lovett, Arzhang Ardavan, Thomas Schenkel, Eugene E. Haller, Joel W. Ager, and S. A. Lyon. “Solid-state quantum memory using the (31)P nuclear spin”. Nature, 455(7216):1085–1088, 2008. [28] L. A. Tracy, D. R. Luhman, S. M. Carr, N. C. Bishop, G. A. Ten Eyck, T. Pluym, J. R. Wendt, M. P. Lilly, and M. S. Carroll. “Single shot spin readout using a cryogenic high-electron-mobility transistor amplifier at sub- Kelvin temperatures”. Appl. Phys. Lett., 108(6):63101, 2016. [29] A. Yang, H. J. Lian, M. L. W. Thewalt, M. Uemura, A. Sagara, K. M. Itoh, E. E. Haller, J. W. Ager, and S. A. Lyon. “Isotopic mass dependence of the lattice parameter in silicon determined by measurement of strain-induced splitting of impurity bound exciton transitions”. Phys. B-Condensed Matter, 376:54–56, 2006. 451 [30] O. E. Dial, M. D. Shulman, S. P. Harvey, H. Bluhm, V. Umansky, and A. Ya- coby. “Charge Noise Spectroscopy Using Coherent Exchange Oscillations in a Singlet-Triplet Qubit”. Phys. Rev. Lett., 110(14):146804, 2013. [31] M. L. W. Thewalt, A. Yang, M. Steger, D. Karaiskaj, M. Cardona, H. Rie- mann, N. V. Abrosimov, A. V. Gusev, A. D. Bulanov, I. D. Kovalev, A. K. Kaliteevskii, O. N. Godisov, P. Becker, H. J. Pohl, E. E. Haller, J. W. Ager, and K. M. Itoh. “Direct observation of the donor nuclear spin in a near- gap bound exciton transition: 31P in highly enriched 28Si”. J. Appl. Phys., 101(8):081724, 2007. [32] P. Becker, H. J. Pohl, H. Riemann, and N. Abrosimov. “Enrichment of sil- icon for a better kilogram”. Phys. Status Solidi a-Applications Mater. Sci., 207(1):49–66, 2010. [33] T. Ruf, R.W. Henn, M. Asen-Palmer, E. Gmelin, M. Cardona, H.-J. Pohl, G.G. Devyatych, and P.G. Sennikov. “Thermal conductivity of isotopically enriched silicon”. Solid State Commun., 115(5):243–247, 2000. [34] P. Becker, H. Friedrich, K. Fujii, W. Giardini, G. Mana, A. Picard, H. J. Pohl, H. Riemann, and S Valkiers. “The Avogadro constant determination via enriched silicon-28”. Meas. Sci. Technol., 20(9), 2009. [35] P. Becker and D. Schiel. “The Avogadro constant and a new definition of the kilogram”. Int. J. Mass Spectrometry, 349:219–226, 2013. [36] Shigeto Fukatsu, Tomonori Takahashi, Kohei M. Itoh, Masashi Uematsu, Akira Fujiwara, Hiroyuki Kageshima, Yasuo Takahashi, Kenji Shiraishi, and Ulrich Go¨sele. “Effect of the Si/SiO2 interface on self-diffusion of Si in semiconductor-grade SiO2”. Appl. Phys. Lett., 83(19):3897–3899, 2003. [37] K. Takyu, K. M. Itoh, K. Oka, N. Saito, and V. I. Ozhogin. “Growth and characterization of the isotopically enriched Si-28 bulk single crystal”. Jpn. J. Appl. Phys., 38(12B):L1493–L1495, 1999. [38] S. A. Lyon, A. M. Tyryshkin, T. Schenkel, J. Bokor, J. Chu, W. Jantsch, F. Schaffler, J. L. Truitt, S. N. Coppersmith, and M. A. Eriksson. “Electron spin coherence in Si”. Physica E, 35(2), 2006. [39] C. C. Lo, M. Urdampilleta, P. Ross, M. F. Gonzalez-Zalba, J. Mansir, S. A. Lyon, M.L.W. Thewalt, and J. J. L. Morton. “Hybrid opticalelectrical detec- tion of donor electron spins with bound excitons in silicon”. Nature Mater., 14(5):490–494, 2015. [40] Yaron Artzi, Ygal Twig, and Aharon Blank. “Induction-detection electron spin resonance with spin sensitivity of a few tens of spins”. Appl. Phys. Lett., 106(8):84104, 2015. 452 [41] A. Bienfait, J. J. Pla, Y. Kubo, M. Stern, X. Zhou, C. C. Lo, C. D. Weis, T. Schenkel, M. L. W. Thewalt, D. Vion, D. Esteve, B. Julsgaard, K. Mølmer, J. J. L. Morton, and P. Bertet. “Reaching the quantum limit of sensitivity in electron spin resonance”. Nature Nanotech., 11(3):253–257, 2015. [42] Kevin G. Orrman-Rossiter, D. R. G. Mitchell, S. E. Donnelly, C. J. Rossouw, S. R. Glanvill, P. R. Miller, Amir H. Al-Bayati, J. A. van den Berg, and D. G. Armour. “Evidence for competing growth phases in ion-beam-deposited epitaxial silicon films”. Philos. Mag. Lett., 61(6):311–318, 1990. [43] N. Tsubouchi, A. Chayahara, Y. Mokuno, A. Kinomura, and Y. Horino. “Epi- taxial growth of pure Si-28 thin films using isotopically purified ion beams”. Jpn. J. Appl. Phys. Part 2-Letters, 40(12A):L1283–L1285, 2001. [44] Richard G. Hewlett and Oscar E. Jr. Anderson. The New World 1939/1946: A History of the United States Atomic Energy Commission. The Pennsylvania State University Press, University Park, Pa, volume 1 edition, 1962. [45] Richard B. Fair. “Analysis and design of ion-beam deposition apparatus”. J. Appl. Phys., 42(8):3176, 1971. [46] Sol Aisenberg and Ronald Chabot. “Ion-Beam Deposition of Thin Films of Diamondlike Carbon”. J. Appl. Phys., 42(7):2953, 1971. [47] S. Shimizu, N. Sasaki, S. Ogata, and O. Tsukakoshi. “Ion beam decelera- tion characteristics of a high-current, mass-separated, low-energy ion beam deposition system”. Rev. Sci. Instrum., 67(10):3664–3671, 1996. [48] N. Herbots, B.R. Appleton, T.S. Noggle, R.A. Zuhr, and S.J. Pennycook. “Ion-solid interactions during ion beam deposition of 74Ge and 30Si on Si at very low ion energies (0-200 eV range)”. Nucl. Instrum. Meth. Phys. Res. Sect. B, 13(1):250–258, 1986. [49] P. C. Zalm and L. J. Beckers. “Ion-beam epitaxy of silicon on Ge and Si at temperatures of 400–K”. Appl. Phys. Lett., 41(2):167–169, 1982. [50] K. Yagi, S. Tamura, and T. Tokuyama. “Germanium and silicon film growth by low-energy ion-beam deposition”. Jpn. J. Appl. Phys., 16(2):245–251, 1977. [51] J. W. Rabalais, A. H. Albayati, K. J. Boyd, D. Marton, J. Kulik, Z. Zhang, and W. K. Chu. “Ion-energy effects in silicon ion-beam epitaxy”. Phys. Rev. B, 53(16):10781–10792, 1996. [52] A. H. Albayati, D. Marton, S. S. Todorov, K. J. Boyd, J. W. Rabalais, D. G. Armour, J. S. Gordon, and G. Duller. “Performance of mass-analyzed, low- energy, dual-ion beam system for materials research”. Rev. Sci. Instrum., 65(8):2680–2692, 1994. 453 [53] Y. Horino, N. Tsubouchi, B. Enders, C. Heck, A. Chayahara, A. Kinomura, and K. Fujii. “Development of a new ion-beam deposition technology for ultra- high-purity film fabrication”. Phys. Status Solidi a-Applied Res., 160(2):583– 589, 1997. [54] J. P. Zhao, D. X. Huang, Z. Y. Chen, W. K. Chu, B. Makarenkov, A. J. Jacobson, B. Bahrim, and J. W. Rabalais. “Amorphous Ge quantum dots embedded in SiO2 formed by low energy ion implantation”. J. Appl. Phys., 103(12):124304, 2008. [55] A. H. Al-Bayati, S. S. Todorov, K. J. Boyd, D. Marton, J. W. Rabalais, and J. Kulik. “Homoepitaxy and controlled oxidation of silicon at low temperatures using low-energy ion beams”. J. Vac. Sci. Technol. B Microelectron. Nanom. Struct., 13(4):1639, 1995. [56] N. Tsubouchi, A. Chayahara, Y. Mokuno, A. Kinomura, and Y. Horino. “3C- SiC thin epilayer formation at low temperature using ion beams”. Appl. Surf. Sci., 212:920–925, 2003. [57] N. Tsubouchi, B. Enders, A. Chayahara, A. Kinomura, C. Heck, and Y. Horino. “Optical properties of carbon and carbon nitride films prepared by mass-separated energetic negative carbon and carbon nitrogen ions”. J. Vac. Sci. Technol. A, 17(4):2384–2388, 1999. [58] J. M. Pomeroy, J. Jacobsen, C. C. Hill, B. H. Cooper, and J. P. Sethna. “Ki- netic Monte Carlo-molecular dynamics investigations of hyperthermal copper deposition on Cu(111)”. Phys. Rev. B, 66(23):235411–235412, 2002. [59] J. M. Pomeroy, A. J. Couture, M. V. R. Murty, E. N. Butler, and B. H. Cooper. “Hyperthermal ion beam system optimized for studying the effects of kinetic energy on thin-film growth”. Rev. Sci. Instrum., 73(11):3846–3852, 2002. [60] R. C. Newman. “Defects in silicon”. Rep. Prog. Phys., 45(10):1163–1210, 1982. [61] Stanley Wolf and Richard N. Tauber. Silicon Processing for the VLSI Era: Volume 1 - Process Technology. Lattice Press, 2000. [62] H. Baumann and K. Bethge. “The Frankfurt PIG ion source”. Nucl. Instrum. Meth. Phys. Res., 189(1):107–110, 1981. [63] P. Rohwer, H. Baumann, K. Bethge, and W. Schu¨tze. “Ion energy analysis of a penning ion source using a high resolution mass spectrometer”. Nucl. Instrum. Meth. Phys. Res., 204(1):245–248, 1982. [64] P. Rohwer, H. Baumann, W. Schu¨tze, and K. Bethge. “Studies of the center potential in a penning discharge”. Nucl. Instrum. Meth. Phys. Res., 211(2- 3):543–546, 1983. 454 [65] H. Baumann and K. Bethge. “PIG ion source with end extraction for multiply charged ions”. Nucl. Instrum. Meth., 122:517–525, 1974. [66] Z. Nouri, R. Li, R. A. Holt, and S. D. Rosner. “A Penning sputter ion source with very low energy spread”. Nucl. Instrum. Meth. Phys. Res. Sect. A, 614(2):174–178, 2010. [67] Bernhard H. Wolf. Handbook of Ion Sources. CRC Press, 1995. [68] C. E. S. Phillips. “The action of magnetised electrodes upon electrical dis- charge phenomena in rarefied gases. Preliminary note”. Proc. R. Soc. Lond., 64(1):172–176, 1898. [69] J. P. Biersack and W. Eckstein. “Sputtering studies with the Monte Carlo program TRIM.SP”. Appl. Phys. A Solids Surfaces, 34(2):73–94, 1984. [70] V. I. Rudakov, V. V. Ovcharov, and V. P. Prigara. “Influence of optical properties of the SOI structure on the wafer temperature during rapid thermal annealing”. Russ. Microelectron., 41(1):15–24, 2012. [71] J. Nulman, S. Antonio, and W. Blonigan. “Observation of silicon wafer emis- sivity in rapid thermal processing chambers for pyrometric temperature mon- itoring”. Appl. Phys. Lett., 56(25):2513, 1990. [72] C. Y. Kuo and C. Gau. “Vaporsolidsolid growth of crystalline silicon nanowires using anodic aluminum oxide template”. Thin Solid Films, 519(11):3603– 3607, 2011. [73] P. C. Maurer, G. Kucsko, C. Latta, L. Jiang, N. Y. Yao, S. D. Bennett, F. Pastawski, D. Hunger, N. Chisholm, M. Markham, D. J. Twitchen, J. I. Cirac, and M. D. Lukin. “Room-temperature quantum bit memory exceeding one second”. Science, 336(6086):1283–1286, 2012. [74] K. D. Jahnke, B. Naydenov, T. Teraji, S. Koizumi, T. Umeda, J. Isoya, and F. Jelezko. “Long coherence time of spin qubits in C-12 enriched polycrys- talline chemical vapor deposition diamond”. Appl. Phys. Lett., 101(1), 2012. [75] G. Balasubramanian, P. Neumann, D. Twitchen, M. Markham, R. Kolesov, N. Mizuochi, J. Isoya, J. Achard, J. Beck, J. Tissler, V. Jacques, P. R. Hem- mer, F. Jelezko, and J. Wrachtrup. “Ultralong spin coherence time in isotopi- cally engineered diamond”. Nature Mater., 8(5):383–387, 2009. [76] K. J. Dwyer, J. M. Pomeroy, and D. S. Simons. “99.996 % C-12 films isotopi- cally enriched and deposited in situ”. Appl. Phys. Lett., 102(25), 2013. [77] K. Ohno, F. J. Heremans, L. C. Bassett, B. A. Myers, D. M. Toyli, A. C. B. Jayich, C. J. Palmstrom, and D. D. Awschalom. “Engineering shallow spins in diamond with nitrogen delta-doping”. Appl. Phys. Lett., 101(8), 2012. 455 [78] T. Teraji, T. Taniguchi, S. Koizumi, K. Watanabe, M. Y. Liao, Y. Koide, and J. Isoya. “Chemical vapor deposition of C-12 isotopically enriched polycrys- talline diamond”. Jpn. J. Appl. Phys., 51(9), 2012. [79] M. L. Markham, J. M. Dodson, G. A. Scarsbrook, D. J. Twitchen, G. Bala- subramanian, F. Jelezko, and J. Wrachtrup. “CVD diamond for spintronics”. Diamond Relat. Mater., 20(2):134–139, 2011. [80] W. M. Lau, I. Bello, X. Feng, L. J. Huang, F. G. Qin, Z. Y. Yao, Z. Z. Ren, and S. T. Lee. “Direct ion-beam deposition of carbon-films on silicon in the ion energy-range of 15–500 eV”. J. Appl. Phys., 70(10):5623–5627, 1991. [81] J. S. Hodges, L. Li, M. Lu, E. H. Chen, M. E. Trusheim, S. Allegri, X. Yao, O. Gaathon, H. Bakhru, and D. Englund. “Long-lived NV-spin coherence in high-purity diamond membranes”. New J. Phys., 14, 2012. [82] F. Kuemmeth, H. O. H. Churchill, P. K. Herring, and C. M. Marcus. “Carbon nanotubes for coherent spintronics”. Materials Today, 13(3):18–26, 2010. [83] D. A. Alman and D. N. Ruzic. “Molecular dynamics calculation of car- bon/hydrocarbon reflection coefficients on a hydrogenated graphite surface”. J. Nucl. Mater., 313:182–186, 2003. [84] K. J. Dwyer, J. M. Pomeroy, D. S. Simons, K. L. Steffens, and J. W. Lau. “Enriching 28Si beyond 99.9998% for semiconductor quantum computing”. J. Phys. D. Appl. Phys., 47(34):345105, 2014. [85] J. K. Rath. “Low temperature polycrystalline silicon: a review on deposition, physical properties and solar cell applications”. Sol. Energy Mater. Sol. Cells, 76(4):431–487, 2003. [86] Siegfried Hofmann. “Atomic mixing, surface roughness and information depth in high-resolution AES depth profiling of a GaAs/AlAs superlattice structure”. Surf. Interface Anal., 21(9):673–678, 1994. [87] T. de Jong, W. A. S. Douma, L. Smit, V. V. Korablev, and F. W. Saris. “The use of pulsed laser irradiation in silicon molecular beam epitaxy: A comparative low energy electron diffraction study”. J. Vac. Sci. Technol. B, 1(4):888, 1983. [88] J. S. Williams. “Solid phase epitaxial regrowth phenomena in silicon”. Nucl. Instrum. Meth. Phys. Res., 209:219–228, 1983. [89] F. A. Stevie and R. G. Wilson. “Relative sensitivity factors for positive atomic and molecular ions sputtered from Si and GaAs”. J. Vac. Sci. Technol. A, 9(6):3064, 1991. 456 [90] J. S. Custer, Michael O. Thompson, D. C. Jacobson, J. M. Poate, S. Roorda, W. C. Sinke, and F. Spaepen. “Density of amorphous Si”. Appl. Phys. Lett., 64(4):437, 1994. [91] Donald Long and John Myers. “Ionized-impurity scattering mobility of elec- trons in silicon”. Phys. Rev., 115(5):1107–1118, 1959. [92] Dimitrie Culcer and Neil M. Zimmerman. “Dephasing of Si singlet-triplet qubits due to charge and spin defects”. Appl. Phys. Lett., 102(23):232108, 2013. [93] K. Oura, M. Katayama, A. V. Zotov, V. G. Lifshits, and A. A. Saranin. Surface Science: An Introduction. Springer, Berlin, 2003. [94] M. Volmer and A. Weber. “Keimbildung in u¨bersa¨ttigten Gebilden”. Z. Phys. Chem., 119(277):227–301, 1926. [95] F. C. Frank and J. H. van der Merwe. “One-dimensional dislocations. II. Misfitting monolayers and oriented overgrowth”. Proc. R. Soc. London A Math. Phys. Eng. Sci., 198(1053), 1949. [96] Helmut Alexander. Dislocations in Solids. Pages 115–118. North-Holland, Amsterdam, 1980. [97] R. G. Humphreys, S. Brand, and M. Jaros. “Electronic structure of the diva- cancy in silicon”. J. Phys. C Solid State Phys., 16(12):L337–L343, 1983. [98] V. V. Kveder, T. Sekiguchi, and K. Sumino. “Electronic states associated with dislocations in p-type silicon studied by means of electric-dipole spin resonance and deep-level transient spectroscopy”. Phys. Rev. B. Condens. Matter, 51(23):16721–16727, 1995. [99] C. A. Londos. “Defect states in electron-bombarded n-type silicon”. Phys. Status Solidi, 113(2):503–510, 1989. [100] Niklas Lehto. “Shallow electron states of bounded intrinsic stacking faults in silicon”. Phys. Rev. B, 55(23):15601–15607, 1997. [101] Ted Thorbeck and Neil M. Zimmerman. “Formation of strain-induced quan- tum dots in gated semiconductor nanostructures”. AIP Adv., 5(8):087107, 2015. [102] A.J. Sigillito, A.M. Tyryshkin, and S.A. Lyon. “Anisotropic Stark effect and electric-field noise suppression for phosphorus donor qubits in silicon”. Phys. Rev. Lett., 114(21):217601, 2015. [103] Giuseppe Pica, Gary Wolfowicz, Matias Urdampilleta, Mike L. W. Thewalt, Helge Riemann, Nikolai V. Abrosimov, Peter Becker, Hans-Joachim Pohl, John J. L. Morton, R. N. Bhatt, S. A. Lyon, and Brendon W. Lovett. “Hy- perfine Stark effect of shallow donors in silicon”. Phys. Rev. B, 90, 2014. 457 [104] Dimitrie Culcer, Xuedong Hu, and S. Das Sarma. “Interface roughness, valley- orbit coupling, and valley manipulation in quantum dots”. Phys. Rev. B, 82(20):205315, 2010. [105] John King Gamble, M. A. Eriksson, S. N. Coppersmith, and Mark Friesen. “Disorder-induced valley-orbit hybrid states in Si quantum dots”. Phys. Rev. B, 88(3):035310, 2013. [106] Xiaojie Hao, Rusko Ruskov, Ming Xiao, Charles Tahan, and HongWen Jiang. “Electron spin resonance and spinvalley physics in a silicon double quantum dot”. Nat. Commun., 5:344–347, 2014. [107] S. Srinivasan, G. Klimeck, and L. P. Rokhinson. “Valley splitting in Si quan- tum dots embedded in SiGe”. Appl. Phys. Lett., 93(11):112102, 2008. [108] Neil Zimmerman, Peihao Huang, and Dimitrie Culcer. “Valley phase, inter- face roughness and voltage control, and coherent manipulation in Si quantum dots”. arXiv:1608.06881, 2016. [109] M. Nerding, L. Oberbeck, T. A. Wagner, R. B. Bergmann, and H. P. Strunk. “Single to polycrystalline transition in silicon growth by ion-assisted deposition at low temperatures”. J. Appl. Phys., 93(5):2570, 2003. [110] G. Xue, H. Z. Xiao, M.-A. Hasan, J. E. Greene, and H. K. Birnbaum. “Critical epitaxial thicknesses for low-temperature (20-100 ◦C) Ge(001)2x1 growth by molecular-beam epitaxy”. J. Appl. Phys., 74(4):2512, 1993. [111] Werner Kern. Handbook of Semiconductor Wafer Cleaning Technology: Sci- ence, Technology, and Applications. Page 623. Noyes Publications, 1993. [112] J. A. Mol, S. P. C. Beentjes, and S. Rogge. “A low temperature surface preparation method for STM nano-lithography on Si(100)”. Appl. Surf. Sci., 256(16):5042–5045, 2010. [113] R. J. Hamers, R. M. Tromp, and J. E. Demuth. “Scanning tunneling mi- croscopy of Si(001)”. Phys. Rev. B, 34(8):5343–5357, 1986. [114] K. Hata, T. Kimura, S. Ozawa, and H. Shigekawa. “How to fabricate a defect free Si(001) surface”. J. Vac. Sci. Technol. A, 18(4):1933, 2000. [115] E. Kasper, M. Bauer, and M. Oehme. “Quantitative secondary ion mass spectrometry analysis of SiO2 desorption during in situ heat cleaning”. Thin Solid Films, 321(1):148–152, 1998. [116] V. A. Ukraintsev and John T. Yate. “The role of nickel in Si(001) roughening”. Surf. Sci., 346(1):31–39, 1996. [117] H. Bracht, E. E. Haller, and R. Clark-Phelps. “Silicon self-diffusion in isotope heterostructures”. Phys. Rev. Lett., 81(2):393–396, 1998. 458 [118] D. Eaglesham, H.-J. Gossmann, and M. Cerullo. “Limiting thickness hepi for epitaxial growth and room-temperature Si growth on Si(100)”. Phys. Rev. Lett., 65(10):1227–1230, 1990. [119] D. J. Eaglesham. “Semiconductor molecular-beam epitaxy at low tempera- tures”. J. Appl. Phys., 77(8):3597, 1995. [120] N.E. Lee, G. Xue, and J. E. Greene. “Epitaxial Si(001) grown at 80-750 ◦C by ion-beam sputter deposition: crystal growth, doping, and electronic prop- erties”. J. Appl. Phys., 80(2):769–780, 1996. [121] M. V. Ramana Murty, Harry A. Atwater, A. J. Kellock, and J. E. E. Baglin. “Very low temperature (400 ◦C) silicon molecular beam epitaxy: the role of low energy ion irradiation”. Appl. Phys. Lett., 62(20):2566, 1993. [122] M. V. Ramana Murty and Harry A. Atwater. “Crystal-stateamorphous- state transition in low-temperature silicon homoepitaxy”. Phys. Rev. B, 49(12):8483–8486, 1994. [123] Jack Thiesen, Howard M. Branz, and Richard S. Crandall. “Explanation of the limiting thickness observed in low-temperature silicon epitaxy”. Appl. Phys. Lett., 77(22):3589, 2000. [124] D. J. Eaglesham, G. S. Higashi, and M. Cerullo. “370◦C clean for Si molecular beam epitaxy using a HF dip”. Appl. Phys. Lett., 59(6):685, 1991. [125] H. Jorke, H.-J. Herzog, and H. Kibbel. “Kinetics of ordered growth of Si on Si(100) at low temperatures”. Phys. Rev. B, 40(3):2005–2008, 1989. [126] D.J. Eaglesham, H.-J. Gossmann, M. Cerullo, L.N. Pfeiffer, and K.W. West. “Limited thickness epitaxy of semiconductors and Si MBE down to room temperature”. J. Cryst. Growth, 111(1):833–837, 1991. [127] E. Chason, P. Bedrossian, K. M. Horn, J. Y. Tsao, and S. T. Picraux. “Ion beam enhanced epitaxial growth of Ge(001)”. Appl. Phys. Lett., 57(17):1793, 1990. [128] C.-H. Choi, R. Ai, and S. Barnett. “Suppression of three-dimensional island nucleation during GaAs growth on Si(100)”. Phys. Rev. Lett., 67(20):2826– 2829, 1991. [129] N.-E. Lee, G. A. Tomasch, and J. E. Greene. “Low-temperature Si(001) epi- taxy using low-energy (E ' 18 eV) Si atoms”. Appl. Phys. Lett., 65(25):3236, 1994. [130] Morito Matsuoka and Shun-ichi Tohno. “Ion beam epitaxy of silicon films in an ultrahigh vacuum using a sputtering-type metal ion source”. J. Vac. Sci. Technol. A, 13(2):305, 1995. 459 [131] K. Miyake and T. Tokuyama. “Germanium and silicon ion-beam deposition”. Thin Solid Films, 92(1-2):123–129, 1982. [132] K. J. Boyd, D. Marton, J. W. Rabalais, S. Uhlmann, and Th. Frauenheim. “Semiquantitative subplantation model for low energy ion interactions with solid surfaces. III. Ion beam homoepitaxy of Si”. J. Vac. Sci. Technol. A, 16(2):463, 1998. [133] M. Kitabatake, P. Fons, and J. E. Greene. “Molecular dynamics and quasidy- namics simulations of low-energy ion/surface interactions leading to decreased epitaxial temperatures and increased dopant incorporation probabilities dur- ing Si MBE”. J. Cryst. Growth, 111(1):870–875, 1991. [134] Kevin G. Orrman-Rossiter, Amir H. Al-Bayati, D.G. Armour, S.E. Donnelly, and J.A. van den Berg. “Ion beam deposited epitaxial thin silicon films”. Nucl. Instrum. Meth. Phys. Res. Sect. B, 59:197–202, 1991. [135] L. A. Miller, D. K. Brice, A. K. Prinja, and S. T. Picraux. “Displacement- threshold energies in Si calculated by molecular dynamics”. Phys. Rev. B, 49(24):16953–16964, 1994. [136] K. A. Jackson, L. Csepregi, W. K. Chu, H. Muller, J. W. Mayer, T. W. Sigmon, G.L. Olson, S.A. Kokorowski, J.A. Roth, L.D. Hess, J. S. Williams, R. G. Elliman, W. L. Brown, T. E. Seidel, J. Linnros, R. G. Elliman, W. L. Brown, Jan Linnros, W. L. Brown, R. G. Elliman, J.P. Biersack, and L.G. Haggmark. “A defect model for ion-induced crystallization and amorphization”. J. Mater. Res., 3(06):1218–1226, 1988. [137] Joseph M. Jasinski and Stephen M. Gates. “Silicon chemical vapor deposition one step at a time: fundamental studies of silicon hydride chemistry”. Acc. Chem. Res., 24(1):9–15, 1991. [138] W. C. T. Lee, N. Bishop, D. L. Thompson, K. Xue, G. Scappucci, J. G. Cederberg, J. K. Gray, S. M. Han, G. K. Celler, M. S. Carroll, and M. Y. Simmons. “Thermal processing of strained silicon-on-insulator for atomically precise silicon device fabrication”. Appl. Surf. Sci., 265:833–838, 2013. [139] E. Massa, G. Mana, L. Ferroglio, E. G. Kessler, D. Schiel, and S. Zakel. “The lattice parameter of the 28Si spheres in the determination of the Avogadro constant”. Metrologia, 48(2):S44–S49, 2011. [140] A. Yang, M. Steger, H. J. Lian, M. L. W. Thewalt, M. Uemura, A. Sagara, K. M. Itoh, E. E. Haller, J. W. Ager III, S. A. Lyon, M. Konuma, and M. Car- dona. “High-resolution photoluminescence measurement of the isotopic-mass dependence of the lattice parameter of silicon”. Phys. Rev. B, 77(11), 2008. [141] Yu-Nong Yang and Ellen D. Williams. “The role of carbon in the faceting of silicon surfaces on the (111) to (001) azimuth”. J. Vac. Sci. Technol. A, 8(3):2481, 1990. 460 [142] Hiroyuki Hirayama, Masayuki Hiroi, and Takashi Ide. “{311} facets of selec- tively grown epitaxial Si layers on SiO2–patterned Si(100) surfaces”. Phys. Rev. B, 48(23):17331–17337, 1993. [143] Atsushi Oshiyama. “Structures of steps and appearances of {311} facets on Si(100) surfaces”. Phys. Rev. Lett., 74(1):130–133, 1995. [144] P. Ashu, C. C. Matthai, and T.-H. Shen. “Dynamics of atoms on silicon substrates”. Surf. Sci., 251:955–959, 1991. [145] P. Becker, P. Scyfried, and H. Siegert. “The lattice parameter of highly pure silicon single crystals”. Zeitschrift fu¨r Phys. B Condens. Matter, 48(1):17–21, 1982. [146] O. Madelung, U. Rossler, and M. Schulz. “Silicon carbide (SiC), lattice parameters, thermal expansion”. In Gr. IV Elem. IV-IV III-V Compd. Part b - Electron. Transp. Opt. Other Prop. Pages 1–11. Springer-Verlag, Berlin/Heidelberg, 41a1b edition, 2002. [147] F. Bozso, J. T. Yates, W. J. Choyke, and L. Muehlhoff. “Studies of SiC for- mation on Si (100) by chemical vapor deposition”. J. Appl. Phys., 57(8):2771, 1985. [148] L. Moro, A. Paul, D. C. Lorents, R. Malhotra, R. S. Ruoff, P. Lazzeri, L. Vanzetti, A. Lui, and S. Subramoney. “Silicon carbide formation by an- nealing C60 films on silicon”. J. Appl. Phys., 81(9):6141, 1997. [149] James P. Becker, Robert G. Long, and John E. Mahan. “Reflection high- energy electron diffraction patterns of carbide-contaminated silicon surfaces”. J. Vac. Sci. Technol. A, 12(1):174, 1994. [150] R. C. Henderson, R. B. Marcus, and W. J. Polito. “Carbide contamination of silicon surfaces”. J. Appl. Phys., 42(3):1208, 1971. [151] J. R. Engstrom, M. M. Nelson, and T. Engel. “Thermal decomposition of a silicon-fluoride adlayer: evidence for spatially inhomogeneous removal of a single monolayer of the silicon substrate”. Phys. Rev. B, 37(11):6563–6566, 1988. [152] M. Maeder, A. Ramı´rez, and A. Zehe. “A RHEED study of as-received and atomically clean silicon surfaces in UHV-environment”. Superf. y vac´ıo, 9:242– 244, 1999. [153] Jun-Zhong Wang, Jin-Feng Jia, Hong Liu, Jian-Long Li, Xi Liu, and Qi-Kun Xue. “Spontaneous vacancy array formation on FeSi2 and CoSi2 formed on Si(100) 2xn surface”. Appl. Phys. Lett., 80, 2002. 461 [154] Ja-Yong Koo, Jae-Yel Yi, Chanyong Hwang, Dal-Hyun Kim, Sekyung Lee, and Dong-Hyuk Shin. “Dimer-vacancy defects on the Si(001)–2x1 and the Ni-contaminated Si(001)–2xn surfaces”. Phys. Rev. B, 52(24):17269–17274, 1995. [155] B. Kolditz and K. R. Roos. “Effect of ambient pressure and nickel contami- nation on the dimer-dangling-bond surface state of Si(001)2x1”. J. Vac. Sci. Technol. A, 25(4):721–725, 2007. [156] Masamichi Yoshimura, Izumi Ono, and Kazuyuki Ueda. “Initial stages of Ni reaction on Si(100) and H-terminated Si(100) surfaces”. Appl. Surf. Sci., 130:276–281, 1998. [157] W. Koczorowski, M. Bazarnik, and R. Czajka. “SPM characterization of titanium disilicide nanostructures grown on a Ni-modified Si(100) substrate”. ACTA Phys. Pol. A, 120(3):480–484, 2011. [158] H. J. W. Zandvliet, H. K. Louwsma, P. E. Hegeman, and Bene Poelsema. “Energetics of Ni-induced vacancy line defects on Si(001)”. Phys. Rev. Lett., 75(21):3890–3893, 1995. [159] Yoichiro Yamaguchi, Masayuki Yoshida, and Hiroyuki Aoki. “Solid solubility of nickel in silicon determined by use of 63Ni as a tracer”. Jpn. J. Appl. Phys., 2(11):714–718, 1963. [160] Y.-W. Mo, R. Kariotis, B. S. Swartzentruber, M. B. Webb, and M. G. Lagally. “Scanning tunneling microscopy study of diffusion, growth, and coarsening of Si on Si(001)”. J. Vac. Sci. Technol. A, 8(1):201, 1990. [161] Y. W. Mo, J. Kleiner, M. B. Webb, and M. G. Lagally. “Activation energy for surface diffusion of Si on Si(001): a scanning-tunneling-microscopy study”. Phys. Rev. Lett., 66(15):1998–2001, 1991. [162] G. Brocks, P. J. Kelly, and R. Car. “Binding and diffusion of a Si adatom on the Si(100) surface”. Phys. Rev. Lett., 66(13):1729–1732, 1991. [163] Takahisa Doi, Masakazu Ichikawa, Shigeyuki Hosoki, and Ken Ninomiya. “Anisotropic diffusion between the step-up and the step-down directions on a Si(001) surface”. Phys. Rev. B, 53(24):16609–16614, 1996. [164] Chris Richardson. Document: “MBE and UHV component etchants”. Private communication, 2015. [165] P. Asoka-Kumar, S. Szpala, B. Nielsen, Cs. Szeles, K. G. Lynn, W. A. Lanford, C. A. Shepard, and H.-J. Gossmann. “Hydrogen-induced breakdown of low- temperature molecular-beam epitaxy of Si”. Phys. Rev. B, 51(7):4630–4632, 1995. 462 [166] J. A. Roth, G. L. Olson, D. C. Jacobson, and J. M. Poate. “Kinetics of solid phase epitaxy in thick amorphous Si layers formed by MeV ion implantation”. Appl. Phys. Lett., 57(13):1340, 1990. [167] L. Oberbeck, N. J. Curson, T. Hallam, M. Y. Simmons, G. Bilger, and R. G. Clark. “Measurement of phosphorus segregation in silicon at the atomic scale using scanning tunneling microscopy”. Appl. Phys. Lett., 85(8):1359–1361, 2004. [168] R. J. Hamers, U. K. Ko¨hler, and J. E. Demuth. “Epitaxial growth of silicon on Si(001) by scanning tunneling microscopy”. J. Vac. Sci. Technol. A, 8(1):195, 1990. [169] A. H. Albayati, K. J. Boyd, D. Marton, S. S. Todorov, J. W. Rabalais, Z. H. Zhang, and W. K. Chu. “Substrate-temperature dependence of homoepitax- ial growth of Si using mass-selected ion-beam deposition”. J. Appl. Phys., 76(7):4383–4389, 1994. [170] M. Wang, G. Audi, A.H. Wapstra, F.G. Kondev, M. MacCormick, X. Xu, and B. Pfeiffer. “The Ame2012 atomic mass evaluation”. Chinese Phys. C, 36(12):1603–2014, 2012. [171] Y. Yatsurugi, N. Akiyama, Y. Endo, and T. Nozaki. “Concentration, solubil- ity, and equilibrium distribution coefficient of nitrogen and oxygen in semi- conductor silicon”. J. Electrochem. Soc., 120(7):975, 1973. [172] Tadashi Nozaki, Yoshifumi Yatsurugi, and Nobuyuki Akiyama. “Concentra- tion and behavior of carbon in semiconductor silicon”. J. Electrochem. Soc., 117(12):1566, 1970. [173] M. I. den Hertog, C. Cayron, P. Gentile, F. Dhalluin, F. Oehler, T. Baron, and J. L. Rouviere. “Hidden defects in silicon nanowires”. Nanotechnology, 23(2):025701, 2012. [174] James H. Comfort and Rafael Reif. “Chemical vapor deposition of epitaxial silicon from silane at low temperatures”. J. Electrochem. Soc., 136(8):2386, 1989. [175] Andrei A. Onischuk and Viktor N. O. Panfilov. “Mechanism of thermal de- composition of silanes”. Russ. Chem. Rev., 70(4):321–332, 2001. [176] B. A. Joyce, R. R. Bradley, and G. R. Booker. “A study of nucleation in chemically grown epitaxial silicon films using molecular beam techniques III. Nucleation rate measurements and the effect of oxygen on initial growth be- haviour”. Philos. Mag., 15(138):1167–1187, 1967. [177] R. J. Buss, P. Ho, W. G. Breiland, and M. E. Coltrin. “Reactive sticking coefficient for silane and disilane on polycrystalline silicon”. J. Appl. Phys., 63(8):2808–2819, 1988. 463 [178] M. Liehr, C. M. Greenlief, S. R. Kasi, and M. Offenberg. “Kinetics of silicon epitaxy using SiH4 in a rapid thermal chemical vapor deposition reactor”. Appl. Phys. Lett., 56(7):629, 1990. [179] F. G. Ruedenauer. “Gas scattering as a limit to trace sensitivity in analytical mass spectrometers”. Rev. Sci. Instrum., 41(10):1487, 1970. [180] Trevor R. Ireland. “Invited Review Article: Recent developments in isotope- ratio mass spectrometry for geochemistry and cosmochemistry”. Rev. Sci. Instrum., 84(1):011101, 2013. [181] N. J. Freeman, N. R. Daly, and R. E. Powell. “Retardation lens used to improve the abundance sensitivity of a mass spectrometer”. Rev. Sci. Instrum., 38(7):945, 1967. [182] R. L. Summers. “Empirical observations on the sensitivity of hot cathode ionization type vacuum gauges”. Technical report, National Aeronautics and Space Administration, NASA TN D-5285, Washington, D.C., 1969. [183] C. Y. Peng, A. A. Woodworth, and C. D. Stinespring. “Method for rapid deter- mination of ion gauge sensitivity factors”. Rev. Sci. Instrum., 78(11):113908, 2007. [184] H. Hertz. “On the evaporation of liquids, especially mercury, in vacuo”. Ann. Phys. (N. Y.), 17:177, 1882. [185] M. Knudesn. “Maximum rate of vaporization of mercury”. Ann. Phys. (N. Y.), 47:697, 1915. [186] Irving Langmuir. “The adsorption of gases on plane surfaces of glass, mica, and platinum”. J. Am. Chem. Soc., 40(9):1361–1403, 1918. [187] S. M. Gates, C. M. Greenlief, D. B. Beach, and R. R. Kunz. “Reactive stick- ing coefficient of silane on the Si(111)-(7x7) surface”. Chem. Phys. Lett., 154(6):505–510, 1989. [188] Stephen M. Gates. “Adsorption kinetics of SiH4, Si2H6 and Si3H8 on the Si(111)-(7x7) surface”. Surf. Sci., 195(1):307–329, 1988. [189] S. R. Logan. “The origin and status of the Arrhenius equation”. J. Chem. Educ., 59(4):279, 1982. [190] Keith J. Laidler. “The development of the Arrhenius equation”. J. Chem. Educ., 61(6):494, 1984. [191] B. A. Scott, R. D. Estes, and J. M. Jasinski. “The role of surface reactions in monosilane pyrolysis”. J. Chem. Phys., 89(4):2544–2549, 1988. 464 [192] Karl F. Roenigk, Klavs F. Jensen, and Robert W. Carr. “Rice-Ramsperger- Kassel-Marcus theoretical prediction of high-pressure Arrhenius parameters by nonlinear regression: application to silane and disilane decomposition”. J. Phys. Chem., 91(22):5732–5739, 1987. [193] F. Hirose. “Modeling growth in Si gas-source molecular beam epitaxy using Si2H6”. J. Cryst. Growth, 179(1):108–114, 1997. [194] S. M. Gates, C. M. Greenlief, S. K. Kulkarni, and H. H. Sawin. “Surface reactions in Si chemical vapor deposition from silane”. J. Vac. Sci. Technol. A Vacuum, Surfaces, Film., 8(3):2965, 1990. [195] D. S. Lin, E. S. Hirschorn, T. C. Chiang, R. Tsu, D. Lubben, and J. E. Greene. “Scanning-tunneling-microscopy studies of disilane adsorption and pyrolytic growth on Si(100)-(2X1)”. Phys. Rev. B, 45(7):3494–3498, 1992. 465