ABSTRACT Title of Dissertation: MATERIALS CHARACTERIZATION FOR SUB-MICRON SUPERCONDUCTING INTERCONNECTS IN RECIPROCAL QUANTUM LOGIC CIRCUITS Cougar Alessandro Tomas Garcia Doctor of Philosophy, 2022 Dissertation Directed by: Professor Steven M. Anlage Department of Materials Science and Engineering Today?s server based computing consumes a considerable amount of energy. Reciprocal Quantum Logic (RQL) is a classical logic family within superconducting electronics, and is a candidate for energy efficient computing technologies. Similar to the current complimentary metal-oxide semiconductor technologies, RQL interconnects are responsible for dissipating the majority of the energy. The energy dissipated in RQL interconnects comes from finite resistive losses in the superconducting wires and embedded dielectrics at radio frequencies. Therefore, material properties, processing, and performance are critical to understanding the mechanisms of loss and mitigation of power dissipation in RQL interconnects. This dissertation presents work on three aspects of materials characterization of RQL interconnects: implementing a method to deconvolve superconducting and dielectric losses, evaluating losses in three generations of RQL fabrication, and understanding the microscopic physics that determines the performance of RQL interconnects in a temperature and frequency range from 1.5-6 K and 3-12 GHz, respectively. A novel method to accurately deconvolve superconducting and dielectric losses by exploiting their frequency dependence is described. Furthermore, a finite element modeler is used to accurately extract the losses. This method is termed Dispersive Loss Deconvolution. The designed microstrip transmission line resonators are fabricated in a 0.25 ?m RQL fabrication process composed of Nb wires embedded in Tetraethyl orthosilicate (TEOS) dielectric. The Nb and TEOS losses as a function of microstrip width down to 0.25 ?m are modeled and measured. The electrical and physical material properties for 3 RQL processes over 5 wafers are evaluated. The electrical properties were evaluated by characterizing resonators in cryogenic dip probes and a dry system with ?10 mK temperature control. The physical properties were evaluated using Transmission Electron Microscopy and Energy-Dispersive Spectroscopy. Two of the processes use chemical mechanical polishing (CMP) to planarize the Nb wires, and the other using reactive ion etching (RIE) to define Nb wires. At 4.2 K, the Nb loss in the 0.25 ?m resonators between the 3 processes were surprisingly distinct. The two CMP processes yield Nb losses up to 2 times higher relative to the RIE process, and have a discernible increase in loss by as much as 20% going from 4 to 0.25 ?m microstrip widths. For the RIE process, there is no detectable upturn in Nb losses for microstrip widths down to 0.25 ?m. Most notably, the RIE process produced 0.25 ?m Nb wires with loss reaching the theoretical lower limit of intrinsic surface resistance Rs = 17 ?? at 4.2 K and 10 GHz. The superior RIE process may be linked to the incorporation of thin metal passivation layers protecting the Nb, which prevented Nb oxide from participating in additional loss mechanisms. For all 3 processes and microstrip widths from 0.25-4 ?m, the TEOS losses had negligible width dependence and varied by as much as ?20%. From the electrical characterization at 4.2 K, it was found that the Nb wires are the limiting loss mechanisms in RQL interconnects. As temperature is decreased below 4.2 K, it is well known that Nb loss will exponentially decrease and amorphous dielectrics like TEOS can have loss with a non-monotonic temperature dependence depending on the input power. This offered the opportunity to explore a possible optimum operating temperature to minimize power dissipation by the RQL interconnects. At relatively low input powers, TEOS became the limiting loss mechanism for temperatures below 3 K, and I conclude this can be attributed to losses coming from two-level system tunneling relaxation and resonant absorption processes. The work in this dissertation describes the development of methods to aid in characterization, design, and fabrication of RQL interconnects, and can be extended to potentially other Single Flux Quantum and Quantum Computing technologies. MATERIALS CHARACTERIZATION FOR SUB-MICRON SUPERCONDUCTING INTERCONNECTS IN RECIPROCAL QUANTUM LOGIC CIRCUITS by Cougar Alessandro Tomas Garcia 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 2022 Advisory Committee: Professor Steven M. Anlage, Chair/Advisor Professor Ichiro Takeuchi Professor Christopher J. Lobb Professor John Cumings Professor Christopher J. K. Richardson Industry Mentor: Dr. Vladimir V. Talanov, Northrop Grumman ?c Copyright by Cougar Alessandro Tomas Garcia 2022 Dedication TO MY WIFE, MARLANEA MY FATHER, MANUEL GARCIA MY MOTHER, DARA GARCIA MY GRANDPA & GRANDMA, ORLAN & DIAMANTE HENDRICKSON AND MY FATHER-IN-LAW, VINCE M. FRANK ii Acknowledgments ?If I have seen further it is by standing on the shoulders of Giants?. Sir Isaac Newton stated in a letter to Robert Hooke in 1675, the same decade of his pioneering work in optics and gravity. First and foremost, I would like to thank my thesis advisor Professor Steven Anlage for his guidance in my journey and his willingness to enlighten my understanding and appreciation of the physical world. In the same light, I thank Dr. Vladimir Talanov, my co-advisor and industry mentor at Northrop Grumman for his his foundational work this thesis was built upon. Additionally, I would like to thank Dr. Nicholas Rizzo for his life and technical mentorship from day one of my scientific career. Thanks also goes to Dr. Elie Track, an influence to many around the world and has played a motivational role on my journey. In like manner, thanks goes to Dr. Joshua Strong, Dr. Anna Herr, Dr. Quentin Herr for their creation of new technologies and their willingness to have me on their team, as it ultimately inaugurated my dissertation work. Special thanks goes to Dr. Daniel Queen for his guidance and fruitful conversations that ultimately helped me understand how low temperature amorphous dielectric properties contributed in Chapters 4 iii and 5. Thanks also goes to Kathy Gray, John Fusco, Dr. Aaron Martin, Andrea Yeiser, Jay Fetterman, Sean McLaughlin, and Dr. Marc Sherwin for programatically supporting my work. Thanks and appreciation also goes to Dr. Henry Luo, Dr. Ryan Dowdy, Dr. Nathan Siwak, Dr. Eric Jones, Chris Pinion, Dr. Chris Kirby, Andrew Brownfield, Nancyjane Bailey, Dr. Steven Sendelbach, Dr. Alex Sirota, Dr. Ed Kurek, Kelsey McCusker, Angelica Dayhoff, Dr. Anton Sidorov, Dr. Melissa Loving, Dr. Thomas Ambrose, Dr. Eric Seabron, Dr. Eric Goodwin, Dr. John X. Przybyz, Dr. Aaron Lee, Dr. Laura Lenard, Will Callis, Dillon Merenich, Zach McDonald, Liam Downey, Keith Crider, Dr. Phillip Fischer, Justin Goodman, Dr. Flavio Griggio, and Dr. Timothy Kohler. Furthermore, I would like to thank my father for his love and support. I would also like to give special thanks to my wife for her unconditional love, support, and unrelenting patience during this journey. Lastly, thank God! iv Table of Contents Dedication ii Acknowledgements iii Table of Contents v List of Tables viii List of Figures ix List of Abbreviations xi List of Variables xiv Chapter 1: Introduction 1 1.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1.2 Technological Challenges . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 1.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 Chapter 2: Materials for Superconducting Interconnects 11 2.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2 Superconductor Electrodynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.1 A Brief History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.2 Complex Conductivity . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 2.2.3 Surface Impedance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 2.2.4 Temperature Dependence of Surface Impedance . . . . . . . . . . . . . . 27 2.2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 2.3 Amorphous Dielectrics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.3.1 The Dielectric Loss Tangent tan ? . . . . . . . . . . . . . . . . . . . . . 30 2.3.2 Dielectric Loss Mechanisms and Phonon Scattering at T < 10K . . . . . 32 2.3.3 Temperature Dependence of tan ? . . . . . . . . . . . . . . . . . . . . . 37 2.3.4 Frequency Dependence of tan ? . . . . . . . . . . . . . . . . . . . . . . 41 2.3.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 2.4 Fabrication Processes for RQL Interconnects . . . . . . . . . . . . . . . . . . . . 45 2.4.1 Brief Background of CMOS Technologies . . . . . . . . . . . . . . . . . 45 2.4.2 SCE Fabrication Technologies . . . . . . . . . . . . . . . . . . . . . . . 47 2.4.3 Chemical Mechanical Planarization (CMP) . . . . . . . . . . . . . . . . 48 v 2.4.4 Cloisonne? Process (Dielectric CMP) . . . . . . . . . . . . . . . . . . . . 50 2.4.5 Damascene Process (Metal CMP) . . . . . . . . . . . . . . . . . . . . . 50 2.4.6 Thermal Budget . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 2.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 Chapter 3: Modeling Lossy 3D Superconductor-Dielectric 2-Port Networks 54 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 3.2 Materials Characterization of Interconnect Morphology . . . . . . . . . . . . . . 56 3.2.1 Scanning Transmission Electron Microscopy (STEM) . . . . . . . . . . . 57 3.2.2 Energy-Dispersive Spectroscopy (EDS) . . . . . . . . . . . . . . . . . . 61 3.3 Theory and Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 3.3.1 RLGC Telegrapher?s equations . . . . . . . . . . . . . . . . . . . . . . 66 3.3.2 Geometric Factors in Inhomogeneous Transmission Line . . . . . . . . . 67 3.3.3 Geometric Factors in Homogeneous Transmission Line . . . . . . . . . . 69 3.3.4 RLGC and Q-Factors From Impedance Matrix [Z] . . . . . . . . . . . . 72 3.4 HFSS Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 3.4.1 Parallel Plate Waveguide . . . . . . . . . . . . . . . . . . . . . . . . . . 77 3.4.2 Defining Solve-Inside Superconductor . . . . . . . . . . . . . . . . . . . 82 3.4.3 Solve-Inside Model Setup . . . . . . . . . . . . . . . . . . . . . . . . . . 84 3.4.4 Microstrip Transmission Line (MTL) . . . . . . . . . . . . . . . . . . . 85 3.5 MTL Geometric Factor ? Extraction . . . . . . . . . . . . . . . . . . . . . . . . 89 3.6 MTL RLGC and ? Sensitivity to Fill . . . . . . . . . . . . . . . . . . . . . . . 92 3.7 MTL ? Sensitivity to RQL Fabrication . . . . . . . . . . . . . . . . . . . . . . . 97 3.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 Chapter 4: Characterization of Microwave Loss in RQL Interconnects at 4.2 K 102 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 4.2 Resonator Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 4.2.1 Resonant Frequency . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104 4.2.2 Critical Coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 4.3 Measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 4.4 Method Demonstration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 4.4.1 Non-Dispersive tan ? Model . . . . . . . . . . . . . . . . . . . . . . . . 117 4.4.2 Homogeneous Superconductor, Dispersive tan ? Model . . . . . . . . . . 124 4.5 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 Chapter 5: Temperature Dependent Characterization of Microwave Loss in RQL Interconnects 137 5.1 Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137 5.2 Temperature Dependent Resonator Measurements . . . . . . . . . . . . . . . . . 139 5.3 Multiplexed Microstrip Transmission Line (MUX-MTL) Resonators . . . . . . . 150 5.4 Using the Dispersive Deconvolution Method to Simultaneously Measure tan ?(T ) and Rs(T ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152 5.5 TLS Absorption and Relaxation Microwave Loss in TEOS Above 1K . . . . . . 158 5.6 Dielectric Loss Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 165 vi 5.7 Temperature Dependence of the Power Dissipation in Nb and TEOS . . . . . . . 174 5.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178 Chapter 6: Summary and Conclusions 180 6.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180 6.2 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181 Appendix A: FIB Cross Sections of Process A and B 186 Appendix B: Estimation of Radiation Q-Factor 189 Appendix C: Temperature Dependent Resonator Measurements 190 C.1 Measurements System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190 Appendix D: Relevant Equations for Parallel Plate Transmission Line Resonator 192 Bibliography 196 vii List of Tables 3.1 STEM Results Summary Table . . . . . . . . . . . . . . . . . . . . . . . . . . . 58 3.2 HFSS Results Table of Fill Effect Process A . . . . . . . . . . . . . . . . . . . . 96 3.3 Nominal Fabrication and Material Properties . . . . . . . . . . . . . . . . . . . . 97 4.1 Rs and tan ? Results Summary Table - Process A, B, C . . . . . . . . . . . . . . 105 5.1 Measured TLS-Phonon Dipole Coupling in Various Materials Summary Table . . 164 5.2 MTL Resonator Temperature Dependent Results Summary Table . . . . . . . . . 171 A.1 FIB Results Summary Table . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186 viii List of Figures 1.1 RQL Interconnects - Materials Paradigm Reference Frame . . . . . . . . . . . . 4 1.2 Energy versus Delay for CMOS and RQL Interconnects . . . . . . . . . . . . . . 8 2.1 Wave Propagation into Superconductor . . . . . . . . . . . . . . . . . . . . . . . 19 2.2 Swihart TM wave in Parallel Plate Transmission Line . . . . . . . . . . . . . . . 21 2.3 Acoustic and Electronic Dielectric loss Measurement Agreement . . . . . . . . . 34 2.4 Diagram of Temperature Regimes of Energy Dissipation in Amorphous Dielectrics 39 2.5 Estimate of loss tangent tan ? plateau at 10 GHz for ??-SiO2 . . . . . . . . . . . 41 2.6 CMOS versus RQL Damascene Process . . . . . . . . . . . . . . . . . . . . . . 46 2.7 Superconducting Electronics Fabrication Process Diagram . . . . . . . . . . . . 49 3.1 STEM Images Processes Side-by-Side 0.25 ?m . . . . . . . . . . . . . . . . . . 59 3.2 STEM Images Processes Side-by-Side 1 ?m . . . . . . . . . . . . . . . . . . . . 60 3.3 STEM Images Morphology Features . . . . . . . . . . . . . . . . . . . . . . . . 61 3.4 EDS Measured Ar - Damascene Process A and B . . . . . . . . . . . . . . . . . 63 3.5 EDS Maps - Damascene Process B . . . . . . . . . . . . . . . . . . . . . . . . . 64 3.6 EDS Plot NbOx Layer Damascene Process A and B . . . . . . . . . . . . . . . . 64 3.7 EDS Maps and Line Scan Process C . . . . . . . . . . . . . . . . . . . . . . . . 65 3.8 Block Diagram of 2-Port Network Parameters . . . . . . . . . . . . . . . . . . . 74 3.9 HFSS Simulated Model Types . . . . . . . . . . . . . . . . . . . . . . . . . . . 78 3.10 HFSS Model of Parallel Plate Waveguide . . . . . . . . . . . . . . . . . . . . . 81 3.11 HFSS PPW Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82 3.12 HFSS MTL Model Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 3.13 HFSS MTL Magnetic Fringing Fields . . . . . . . . . . . . . . . . . . . . . . . 88 3.14 HFSS Results of Geometric Factor vs MTL Width . . . . . . . . . . . . . . . . . 91 3.15 HFSS STEM of Fill Process A . . . . . . . . . . . . . . . . . . . . . . . . . . . 92 3.16 HFSS Simulation of Fill Effect Process A . . . . . . . . . . . . . . . . . . . . . 94 3.17 HFSS Parametric Corner Analysis on MTL . . . . . . . . . . . . . . . . . . . . 98 3.18 HFSS Geometric Factor Variation of MTL . . . . . . . . . . . . . . . . . . . . . 100 4.1 MTL Physical Chip Layout . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 4.2 MTL Measured Qi For Varied Coupling . . . . . . . . . . . . . . . . . . . . . . 109 4.3 Wafer Map Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 4.4 Measured S21 and S?21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112 4.5 Measured Qi vs ? vs w . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 4.6 Geometric Factors ? from HFSS versus Parallel Plate and Nominal . . . . . . . . 115 ix 4.7 Geometric Factors ? from HFSS versus Parallel Plate and Nominal . . . . . . . . 118 4.8 Rs versus MTL width, Parallel plate and Inhomogeneous Superconductor Model 120 4.9 Results (Homogeneous Superconductor, non-dispersive tan ? Model) . . . . . . . 123 4.10 Dispersive tan ? estimates by Conductor Loss Subtraction . . . . . . . . . . . . . 125 4.11 Width Dependence of Dispersive tan ? estimates by Conductor Loss Subtraction . 126 4.12 Qi versus Dispersive tan ? Model Using 3-Parameter Fit . . . . . . . . . . . . . 128 4.13 MTL Width Dependence of the Dispersive tan ? 3-Parameter . . . . . . . . . . . 129 4.14 MTL Width Dependence of Rs and tan ? Using Dispersive tan ? 3-Parameter Fits 130 5.1 MTL, Process B, 4 ?m, versus Temperature . . . . . . . . . . . . . . . . . . . . 140 5.2 MTL, Power Conservation Reference Node . . . . . . . . . . . . . . . . . . . . 142 5.3 MTL Q?,Ti for -20 and -35 dBm Input Power . . . . . . . . . . . . . . . . . . . . 145 5.4 MTL Peak Current from T=1.7-4.5 K for -20 and -35 dBm Input Power . . . . . 146 5.5 Q?1i (?n) for MTL widths 0.25 - 4 ?m at 1.7 K . . . . . . . . . . . . . . . . . . . 147 5.6 Calculated Peak Currents for MTL widths 0.25 - 4 ?m at 1.7 and 4.2 K . . . . . . 149 5.7 MUX-MTL Physical Chip Layout . . . . . . . . . . . . . . . . . . . . . . . . . 151 5.8 MUX-MTL Temperature dependent S21versusFrequency . . . . . . . . . . . . 152 5.9 MUX-MTL Temperature dependent Qi versus Frequency . . . . . . . . . . . . . 153 5.10 Dispersive Loss Deconvolution Method Applied to Temperature DependentQi(T ) Using MUX-MTLs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157 5.11 Non-Monotonic Temperature Dependence of TEOS Dielectric Constant in MUX- MTL Resonators at T < 3K . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163 5.12 Frequency Dependence of TLS Dipole Coupling in TEOS . . . . . . . . . . . . . 166 5.13 Measurement of TLS Relaxation and Absorption From TEOS Loss Tangent Using MUX-MTL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169 5.14 TLS Relaxation and Absorption Fit to Q?1i Data . . . . . . . . . . . . . . . . . . 170 5.15 Stored Power and Peak Currents in MUX-MTL Resonators Below 3 K . . . . . . 173 5.16 Microwave Loss Spectroscopy of TEOS Amorphous Dielectric at T = 1.5? 3K 175 5.17 Measurements of Power Dissipation in RQL interconnects using MUX-MTLs . . 177 A.1 Process A FIB Cross Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187 A.2 Process B FIB Cross Sections . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188 C.1 Temperature Dependent RF Measurements System . . . . . . . . . . . . . . . . 191 x List of Abbreviations ABF Annular Bright Field AIST Advanced Industrial Science and Technology AQFP Adiabatic Quantum Flux Parametron BCS Bardeen, Cooper, and Schrieffer theory BF Bright Field CD Critical Dimension CS Conducting Strip CVD Chemical Vapor Deposition CMOS Complimentary Metal-Oxide-Semiconductor CMP Chemical Mechanical Polishing (or Planarization) dBm Decibel-Milliwatts DC Direct Current DLD Dispersive Loss Deconvolution EDS Energy-Dispersive Spectroscopy eSFQ Energy Efficient Single Flux Quantum FEM Finite Element Modeler FIB Focused Ion Beam xi GP Ground Plane HFSS High Frequency Structure Simulator HTS High Temperature Superconductor IBM International Business Machines Corporation IL Insertion Loss IRDS International Roadmap for Devices and Systems JJ Josephson Junction LL MIT Lincoln Laboratory MTL Microstrip Transmission Line MUX-MTL Multiplexed Microstrip Transmission Line PEC Perfect Electrical Conductor PECVD Plasma-Enhanced Chemical Vapor Deposition PPW Parallel Plate Waveguide PTL Passive Transmission Line PVD Physical Vapor Deposition QFP Quantum Flux Parametron RF Radio Frequency RIE Reactive-Ion Etching RQL Reciprocal Quantum Logic RSFQ Rapid Single Flux Quantum xii SEM Scanning Electron Microscopy SFQ Single Flux Quantum SIBC Surface Impedance Boundary Condition STEM Scanning Transmission Electron Microscopy SCE Superconducting Electronics TE Transmitted Electron TEM Transmission Electron Microscopy TEOS Tetraethyl Orthosilicate TLS Two Level System VNA Vector Network Analyzer ZC Z-Contrast ZOR Zeroth-Order Resonator xiii List of Critical Symbols ? = 8.85? 10?12 vacuum permittivity (F m?10 ), Eq. 2.18? ?wave = ?0/?0 = 120? intrinsic wave impedance (?), Eq. 2.26 ?0 = 4? ? 10?7 vacuum permeability (N A?2), Eq. 2.3 c = 3? 108 vacuum speed of light (m s?1), Eq. 2.36 h? = 1.06? 10?34 reduced Planck?s constant (J s), Eq. 2.6 k = 1.38? 10?23B Boltzmann constant (J K?1), Eq. 2.4 ? The transmission line attenuation constant (m?1), Eq. 3.19 ? ?1db ultrasonic attenuation (m or dB) ? The transmission line phase constant (m?1), Eq. 3.19 ?cm Transmission line partial geometric factor associated with losses in the m-th conductor (?/sq), Eq. 3.6 ?c0 Net conductor geometric factor at reference frequency ?0, Eq. 3.13 ?c0m Transmission line partial geometric factor at that frequency associated with loss in the mth conductor, Eq. 3.11 ?d Transmission line partial geometric factor associated with losses in that dielectric layer or tube, Eq. 3.41 ?dn Transmission line partial geometric factor associated with loss in the nth dielectric layer or tube, Eq. 3.6 ? The transmission complex propagation constant xiv of a transmission line (m?1), Eq. 2.12 ? BCS energy gap of a superconductor (J), Eq. 2.4 ??abs The change in dielectric constant due to TLS resonant absorption, Eq. 5.11 ?r relative dielectric constant, Eq. 2.26 ??rel The change in dielectric constant due to TLS relaxation, Eq. 5.10 ? magnetic penetration depth (m), Eq. 2.2 ?eff effective magnetic penetration depth (m), Eq. 2.25 ?EM the wavelength of the propagating wave (m), Eq. 2.17 ?L London penetration depth (m), Eq. 2.2 n?2 The permanent dipole coupling of TLSs having a density of states n, Eq. 5.12 n?? 2 The induced dipole coupling of TLSs having a density of states n, Eq. 5.11 ?0 microscopic coherence length of a bulk superconductor (m), Eq. 2.8 ? = ?1 ? i?2 complex conductivity (??1m?1), Eq. 2.5 ?n normal metal conductivity (??1m?1), Eq. 2.5 ?rel characteristic relaxation time (s), Eq. 2.30 Xeff effective surface reactance (?/sq), Eq. 2.23 ? angular frequency (rad s?1), Eq. 2.5 a The TLS material constant, Eq. 5.15 xv C Transmission line capacitance per unit length (F m?1), Eq. 3.5 d superconductor thickness (m), Eq. 2.22 fn Resonant frequency of the n-th mode for a microstrip transmission line resonator (Hz), Eq. 4.1 G Transmission line conductance per unit length (S m?1), Eq. 3.4 g Coupling factor of a Capacitively coupled microstrip, Eq. 4.2 Ipeak The magnitude of the peak RF current in the resonator (A) Eq. 5.5 ITLSc critical intensity for TLS in weak-strong electric field interaction (W m?2), Eq. 2.36 IL Insertion Loss (dB), Eq. 4.2 J 2c is the critical energy intensity for the saturation of TLSs (W/m ) Eq. 5.19 L Transmission line inductance per unit length (H m?1), Eq. 3.3 l Transmission line length (m), Eq. 3.18 `mfp normal electron mean free path (m), Eq. 2.8 `ph phonon mean free path (m), Eq. 2.33 `tun,res TLS resonant tunneling contribution to phonon mean free path (m), Eq. 2.34 `tun,rel TLS tunneling relaxation contribution to phonon mean free path (m), Eq. 2.34 `class,rel classical relaxation contribution to phonon mean free path (m), Eq. 2.34 xvi M? The average TLS coupling energy (eV) Eq. 5.16, Eq. 5.16 ne total electron density (m?3), Eq. 2.1 ns superfluid electron density (m?3), Eq. 2.1 Pin The incident power at the input of the resonator after cable attenuation (W ), Eq. 5.1 Pr The reflected power to the VNA (W ) 5.1 Pst The circulating (stored) power in the resonator (W ), Eq. 5.4 Pt The transmitted power (W ), Eq. 5.1 Pd The power dissipated by the resonator (W ), Eq. 5.1 Qc Partial Q-factor associated with the conductor loss, Eq. 3.1 Qd Partial Q-factor associated with the dielectric loss, Eq. 3.1 Qi Internal or intrinsic Q-factor, Eq. 3.1 R Transmission line resistance per unit length (?m?1), Eq. 3.3 Reff effective surface resistance (?/sq), Eq. 2.24 Rs intrinsic surface resistance (?/sq), Eq. 2.19 Rs0 Intrinsic surface resistance at reference frequency ?0 (?/sq), Eq. 3.13 Rs0m Transmission line intrinsic resistance at a reference frequency ?0 of the m-th conductor (?/sq), Eq. 3.11 Rsm Transmission line effective intrinsic resistance of the m-th conductor (?/sq), Eq. 3.6 |S21| The transmission coefficient magnitude (dB) Eq. 4.2 T temperature (K), Eq. 2.6 xvii T1 TLS relaxation time (s), Eq. 2.36 T2 TLS dephasing time (s), Eq. 2.36 Tc superconducting critical temperature (K), Eq. 2.1 tan ? dielectric loss tangent, Eq. 2.26 tan ?0n Transmission line loss tangent of the n-th dielectric layer or tube, Eq. 3.11 tan ?n Transmission line effective loss tangent of the n-th dielectric layer or tube, Eq. 3.6 tan ?class,rel classical relaxation contribution to the dielectric loss tangent, Eq. 2.35 tan ?tun,rel TLS tunneling relaxation contribution to the dielectric loss tangent, Eq. 2.35 tan ?tun,res TLS resonant tunneling contribution to the dielectric loss tangent, Eq. 2.35 Vpeak The magnitude of the peak RF voltage in the resonator (V) Eq. 5.6 w microstrip width (m), Eq. D.3-D.6 Xs intrinsic surface reactance (?/sq), Eq. 2.19 Y shunt admittance (??1 m?1), Eq. 2.11 Z series impedance (?m?1), Eq. 2.10 Zeff effective surface impedance (?/sq), Eq. 2.21 Z0 Characteristic impedance of the microstrip transmission line resonator (?), Eq. 3.20 Zfeed Characteristic impedance of the feed line (?), Eq. 4.4 [Zij] Impedance matrix (?), Eq. 3.22 Zin Input impedance (?), Eq. 4.4 Zs intrinsic surface impedance (?/sq), Eq. 2.20 xviii Zsub dielectric substrate bulk impedance (?/sq), Eq. 2.24 z(0) transformed impedance at film interface normalized to Zs, Eq. 2.22 z(d) transformed impedance up to the film interface normalized to Zs, Eq. 2.22 xix Chapter 1: Introduction 1.1 Overview Since the birth of the information age, there has been a relentless need for more computing power. In the 1950s-1970s mainframe computers were only accessible through terminals in close proximity. Today, high performance data centers can be accessed remotely from anywhere around the world. Data centers are now essential to almost every part of the world economy - universities, government institutions, large companies, small companies, and even individuals. Server-based computing is now becoming the primary user for complimentary metal-oxide semiconductor (CMOS) technologies, amounting to considerable energy consumption. Additionally, Moore?s Law is expected to flatten by 2025, preventing CMOS from doubling in performance every 2 years [1]. If energy demands continue to evolve across the world as a consequence of the growing usage of available computing power and Moore?s law comes to an end, the total energy consumption of data centers across the world is projected to reach 65 GW per year in 2030 (? 70% increase from today) [2], approximately equivalent to the output of 65 power plants, 31 Hoover Dams, or 6.5 billion LED Bulbs. The power consumption and performance challenges of CMOS need to be addressed, and are being studied by the International Roadmap for Devices and Systems (IRDS) [3]. Logic families within superconducting electronics (SCE) are candidates to address power dissipation. 1 Reciprocal Quantum Logic (RQL), an SCE digital logic technology, has been reported to be 300 times more energy efficient than CMOS [4], and has demonstrated the largest working circuit to date with ? 105 logic units on a single chip [5]. Additionally, RQL has a significant speed advantage over CMOS, where the upper frequency set by the plasma frequency of the Josephson junctions on the order of fp ? 250 GHz [4]. However, RQL is still a young technology as compared to CMOS having consistently ? 1010 logic units per circuit. Several engineering challenges are limiting large scale RQL-based computation. In particular, superconducting interconnects, which serve as RQL?s clocking, power delivery system, and serial data links. Moreover, unlike CMOS where the ground planes only serve as a medium for data propagation and electrical isolation between signals, RQL ground planes are engineered to mitigate magnetic flux being trapped in circuit areas that can reduce operating margins and even catastrophic circuit failure [6, 7]. Material quality and defectivity affect both power dissipation and magnetic flux trapping in RQL circuits, and the focus of this dissertation is the former. RQL interconnects are composed of superconducting wires down to submicron dimensions embedded in a dielectric. The RQL interconnect performance is governed by the power dissipation in these materials due to radio frequency (RF) resistive losses. The power dissipation in RQL interconnects is orders of magnitude lower than CMOS interconnects, but can dwarf that of the power in RQL logic, and so the superconductor and dielectric RF losses should be minimized. In CMOS circuits, static and dynamic energy dissipation have become the main limiting factors for further device scaling [8]. Specifically, at maximum scaling and density of CMOS circuits, the clock power becomes the dominant contributor to dynamic power dissipation relative to the logic. Resonator-based clocks in CMOS have been shown to reduce power dissipation 2 [9]. The 300x higher energy efficiency of RQL relative to CMOS was calculated for only the net power dissipated by the logic alone. Recently, it was found that an RQL circuit with a metamaterial resonator clock yields an energy efficiency of ? = P ?1RQL(PRQL + PRes) ? 0.3, where PRQL is the total dynamic power burned by the logic and PRes is the total static power burned by the resonator clock [10]. Therefore, the energy efficiency of RQL relative to CMOS drops down to 90x, significantly reducing its competitive advantage. Materials science and engineering plays a primary role in creating state-of-the-art computing systems and optimizing their performance. As capabilities increase, so does their complexity, making the analysis and improvement at the single component level more demanding. Grand challenges, such as the ending of Moore?s law and the world wide need for more computing power, can be addressed via the materials science tetrahedron [11]. Specifically, RQL interconnect power effeciency can be optimized by examining the materials processing (Fig. 1.1a), structure (Fig. 1.1b), properties (Fig. 1.1c), and performance (Fig. 1.1d). From this perspective, power dissipation in RQL interconnects may be alleviated with a study of fabrication processes, material morphology, electrical properties, and their affect on RQL performance. The information gleaned from such a study can be incorporated into the emerging systems materials engineering triangle [12], enabling system-level planning to facilitate maturation of superconducting electronics (SCE) technologies (e.g. RQL). This dissertation is focused on characterizing RQL superconductor-dielectric interconnect losses (see Fig. 1.1). This is accomplished by: (i) characterization of the interconnects electrical properties by developing a novel methodology to deconvolve intrinsic RF losses, (ii) evaluating the processing of interconnects using complimentary characterization techniques of the material morphology, and (iii) optimizing the performance by reducing the RF losses in superconducting 3 Figure 1.1: Addressing power dissipation with the materials science paradigm (tetrahedron). The submicron micronstrip transmission line (MTL) resonator lies at the center of the tetrahedron as the characterization ?vehicle?. a) Three chemcial mechanical planarization (CMP) processes of RQL MTLs were examined in this work. b) An example of studying the microstructure and morphology an MTL (see Chapter 3) c) MTL resonator measurements to extract the RF resistive loss propoerties of the superconudctor and dielectric materials (see Chapters 4 and 5). d) With new predictive models, improvements in the performance can be applied to increasing the energy efficiency of RQL meta material clock networks (see Chapters 4 and 5) 4 interconnects. The dissertation is organized into six chapters. In the remainder of this first chapter, I review the technological challenges of RQL interconnects. In Chapter 2, I discuss the electrodynamics of superconductors and amorphous dielectrics, and RQL fabrication processes. Chapter 3 introduces the non-idealities of RQL interconnects caused by fabrication processes, describes a method to characterize the superconductor and dielectric RF losses independently, and presents a procedure to model superconductors in HFSS in order to extract the RF losses in the superconductor. In Chapter 4, I present a resonator design and measurements of RF losses at 4.2 K in RQL interconnects down to 0.25 ?m dimensions. In this chapter, the measured intrinsic losses of the superconductor and dielectric, Rs and tan ?, will be related back to the morphology analyzed in Chapter 3. In Chapter 5, I explore the temperature dependence of RF losses in Nb and TEOS from 1.5-4.5 K to better understand if temperature plays a role in RQL interconnect performance. Finally, in Chapter 6, I provide a summary of my main results, followed by recommendations for future research. 1.2 Technological Challenges It was first theorized by Brian Josephson in 1962 that supercurrent could tunnel between two superconducting electrodes if the barrier to tunneling was thin enough [13]. Soon after, Anderson and Rowell measured this tunneling effect [14], followed by John Rowell patenting the use of the Josephson junction (JJ) as a cryogenic logic device [15]. Thus, the birth of Josephson electronics and IBM?s thrust to demonstrate a Josephson based computer in the 1960s-1980s. IBMs effort, using Pb based JJs, ended in 1983 due to only a 2.5 times speed improvement 5 relative to CMOS. The higher reliability of Nb based JJs developed by Gurvitch at Bell Labs [16] allowed for larger scale circuits and complexity, renewing SCE in early 1987. This is around the same time high-temperature superconductivity was discovered. Voltage-state gates were used during this time, where logical 0 (1) is represented by a voltage level of zero (2.5 mV), but requires a large footprint with significant logic power dissipation due to the incorporation of resistors [17]. Miraculously, Josephson microprocessors were functional and demonstrated by Fujitsu Laboratories in Japan [18]. These processors were built one-for-one with CMOS using niobium junctions having a few thousand gates and memory of a few kilobits. Although, memory, processing margins, footprint, and power consumption were the limiting factors preventing from competing with CMOS. Single flux quantum (SFQ) based logic devices, relying on the current (or phase) state to do computation, can have (i) 1-2 orders of magnitude reduction in power dissipation, (ii) 2-4 times smaller footprint compared to voltage-state Josephson logic [19], and (iii) potential for up to 50 times higher clock rates compared to CMOS [20]. SFQ based logic families are rapid single flux quantum (RSFQ) logic [21], [22], quantum flux parametron (QFP) [23], reciprocal quantum logic (RQL) [4], energy efficient single flux quantum (eSFQ) [24], and adiabatic quantum flux parametron (AQFP) [25]. The biggest advantage for SFQ technologies is computation speed, where greater than 100 GHz data rate has been demonstrated [26, 27]. Additionally, SFQ can be used as the control processor in quantum computing systems [28, 29, 30], and other niche? application specific circuits [31]. Recently, scalability and energy efficiency have been reviewed in comparison to CMOS [32, 33]. Without sacrificing computation speed, RQL logic currently has a big advantage in energy efficiency, with the largest JJ count per chip to date (a measure of the scalability) [5]. Superconducting interconnects need to be considered further. 6 Superconducting microwave transmission line (MTL) are the most common interconnects, supporting low bandwidth signals for clocks and power distribution [34, 35], as well as high bandwidth signals to propagate SFQ pulses that carry bits of information [36]. The latter is called a passive transmission line (PTL). Compared to SiO2 and Cu at room temperature, PTLs composed of SiO2 dielectric and Nb at 4.2 K are near lossless. For example, the energy dissipation in CMOS interconnects limits the clock frequency of processors to 4 GHz and forces the amount of transistors that need to be powered off at any given time to prevent the chip temperature from exceeding thermal limits [32]. Fig. 1.2 shows the energy versus delay for CMOS and RQL interconnects, adopted from the International Roadmap For Devices and Systems (IRDS) 2020 edition. Here we are not considering AQFP technology, since it does not use PTL interconnects. Accounting for the cooling energy to keep circuits at 4 K with > 10W cooling power, PTL power dissipation is 1? 10?16 J (0.3 ?W at 3 GHz) for a 1 mm long PTL (?RQL interconnect? arrow in Fig. 1.2). Therefore, RQL MTL interconnects can have 2-4 orders magnitude lower dissipation relative to CMOS. As mentioned above, to further scale RQL technology, it is desired to have global clock and power distribution with spatially equal magnitude and phase. Recently, a metamaterial clock resonator has been designed and demonstrated to deliver power to 48k JJs operating at 3.5 GHz supporting uniform power distribution across 3? 3mm2 area at 4.2 K [10]. This clock design is compatible with AC-powered SFQ logic families (e.g. RQL and QFP). Unfortunately, this comes at an energy cost. For the above circuit size, the resonator clock has a best case energy efficiency estimated to be ? = P ?1RQL(PRQL + PRes) ? 0.3, where PRes ? 18 ?W is the power burned by the resonator clock, and PRQL ? 40 ?W is the power burned by the logic when all logic units 7 Figure 1.2: Adopted from the International Roadmap For Devices and Systems 2020 edition (IRDS) [37]. (Yellow Box) Arrow pointing to the energy dissipated by a RQL 1 mm long passive transmission line (PTL) interconnects. (Red squares) CMOS interconnects energy dissipation and delay for comparison. 8 are saturated (producing all 1s every clock cycle). This means the resonator clock is now limiting the energy efficiency of RQL, as PRQL is much lower under normal operation of the circuit with only a small percentage of logic gates drawing power from the clock at any given time. RQL resonator clock interconnects are made of microstrip transmission line (MTL) networks [10]. For a fixed circuit size and damping ratio of the junctions, resonator clock efficiency is related to the intrinsic Q-factor of the resonator by ? ? (1 + Q?1i )?1 with nominal Q?1i = Q?1 +Q?1 ? 4? 10?3 at 10 GHz, where Q?1 and Q?1c d c d are proportional to the power dissipated and therefore RF losses in the Nb wires and SiO2 dielectric, respectively. In RQL fabrication, the SiO2 is typically made using a chemical vapor deposition process using a tetraethyl Orthosilicate (TEOS) precursor, and so SiO2 will be referred by its common precursor TEOS throughout this work. The power dissipated in TEOS and Nb are notable and comparable at 4.2 K, which contribute to the low Qi in RQL clocks. The loss mechanisms and fabrication limitations will be discussed further in Chapter 2. As a consequence, reducing RF losses in superconductor-dielectric resonant systems is critical for the scaling of RQL and AC-powered SFQ technologies. Furthermore, it is also important to fully characterize power dissipation as a function of frequency (MHz-GHz) and temperature (mK-K) to assure optimum operating points, as non-monotonic functioning are possible. The technological challenge this work will aim to address through materials science is power dissipation in RQL clock interconnects by modeling and characterizing microstrip transmission line resonators. 9 1.3 Summary In summary, there is an opportunity to increase the power efficiency of RQL resonator- based interconnects to increase the technology?s competitive advantage. This dissertation intends to remedy this state of affairs by providing novel approaches to characterize the microwave losses of Nb and TEOS. I will present techniques that serve as tools to understand both the intrinsic properties and extrinsic effects due to fabrication. These approaches include finite element modeling (e.g. HFSS), physical characterization of the material microstructure, a method to deconvolve the Nb and TEOS loss in a single measurement, and temperature dependent loss spectroscopy. The methods established can be used to predict the power efficiency of RQL interconnects for a given material system, and provide a path for future work. 10 Chapter 2: Materials for Superconducting Interconnects 2.1 Overview In this chapter, I will introduce the intrinsic material properties responsible for power dissipation in RQL interconnects. Each subsection will briefly review the history of superconductors and dielectrics. The superconductor intrinsic resistance Rs and dielectric loss tangent tan ? will be defined. Their respective temperature dependence and frequency dependence will also described, supported by models and measurements. The assumptions, concepts, and equations will be used throughout Chapters 3-5. 2.2 Superconductor Electrodynamics 2.2.1 A Brief History Superconductivity was first discovered in 1911 by Heike Kamerlingh Onnes, the pioneer of helium liquification [38], when he measured a zero resistance on a mercury sample at 4.2 K [39]. In 1934, Gorter and Casimir developed the thermodynamic two-fluid model [40], where the total electron density ne in a superconductor is split into two parallel channels: a normal electron density nn and a frictionless electron density ns. The non-dissipative electron channel ns is now commonly referred to as the superfluid electron density, coined from the discovery of 11 superfluidity in 4He (zero viscosity) by Kapitsa [41] and independently by Allen and Misener [42] in 1937. At absolute zero temperature T = 0 the superfluid electron density equals the total electron density ns/ne = 1. At a thermodynamic critical temperature Tc of a superconductor (e.g. Nb bulk or thin film), the superfluid electron density is equal to zero ns/ne = 0, having a model temperature dependent form ( ) ns ? T 4 = 1 (2.1) ne Tc In 1935, Fritz and Heinz London proposed the two constitutive relations describing the electrodynamics response of a superconductor in an electromagnetic field [43]. The second London equation paired with Ampere?s law gives ?2 1B = B (2.2) ?2L where B is the vector magnetic field inside the superconductor, and ?L is the London penetration depth ? mc2 ?L = (2.3) 4?n 2ee where ?0 is the vacuum magnetic permeability, c is the speed of light in vacuum, and m and e are the mass and charge of an electron, respectively. Note, the London penetration depth is traditionally defined in terms of the total electron density, and is not temperature dependent as shown in Eq. 2.1. The temperature dependence of ?L(T ) will be reviewed in Section 2.2.2. Eq. 2.2 suggests an applied external magnetic fieldH0 decays exponentially inside a superconductor 12 to a magnetic field of H = H0/e at a distance into the surface equal to the London penetration depth ?L ? 40 nm for bulk Nb, where e is the first power exponential. In normal metals, the decay length is associated with the commonly known skin effect and the corresponding skin depth of Cu is ? 700 nm at ?/2? = 10 GHz [44]. Also note, the skin depth of Cu is infinite at ? = 0 while ?L is finite. Jumping ahead, in 1957 the microscopic theory for superconductivity was established by Bardeen, Cooper, and Schrieffer (BCS) theory [45]. In accordance with BCS theory, the superfluid electrons have a bound (e.g. lower energy) state mediated by phonon lattice vibrations. The superfluid state consists of Cooper pairs. Using the simple electrodynamic model description, locally constricting the lattice around one paired electron in a single Cooper pair allows the other paired electron to later travel in the opposite direction with stronger lattice Coulomb interactions. The lowest-energy excitations of the superconduction states requires energy 2?, where ? is the BCS energy gap. At zero temperature T = 0, the Nb BCS energy gap, a strongly coupled superconductor, can be estimated as [46] ? ? 1.97kBTc (2.4) where kB = 86 ?eV/K is the Boltzmann constant. A strongly coupled superconductor means there is a strong electron-phonon coupling between the Cooper pairs and the phonon lattice energy h??ln, where h? is the Planck constant. For the measurements in Chapters 4 and 5 of Nb thin films, assuming Tc ? 9.2K, the stimulus frequencies f ? 10GHz are much lower than the gap frequency f  2?/h ? 750GHz. Estimates of the Nb energy gap from temperature dependent measurements of microstrip transmission line resonators will be presented in Chapter 5 using the 13 dispersive loss deconvolution (DLD) method. 2.2.2 Complex Conductivity In 1940, Heinz London measured resistive loss in superconducting Sn by means of calorimetry at high frequencies, and proposed it was due to a two-fluid model (interpenetrating mixture of normal and superconducting electrons) [47]. Later in 1947, Sir Brian Pippard measured the surface impedance of superconducting Sn and Hg at 1.2 GHz using a quarter-lambda resonator [48]. I will come back to surface impedance measurements in Section 2.2.3. The above experiments and many more led to the Mattis and Bardeen theory in 1958 that unified the anomalous skin effect for normal and superconducting metals [49]. In weak magnetic fields and in the low frequency limit h??  ?, the Mattis and Bardeen expressions for complex conductivity ? = ?1 ? i?2 of a superconductor can be approximated by [50, 51] ?1 = 2?(T ) e ?(T )/kBT ?(T )/k T 2 ln ?(T ) (2.5) ?n kBT (1+e B ) kBT ?2 = ??(T ) tanh ?(T ) (2.6) ?n h?? 2kBT where ? is the angular frequency, ?(T ) is the temperature dependent BCS superconducting gap, ?n is the normal metal conductivity, and ?1 and ?2 are the real and imaginary components of the complex conductivity at temperature T . At absolute zero temperature T = 0, the real conductivity is effectively zero ?1 = 0. Eq. 2.5 was approximated by Kautz [50] and Eq. 2.6 by Tinkham [51] for all temperatures below T < Tc. Similarly, in weak magnetic fields and in the low frequency limit h??  ?, the two-fluid 14 model yields a complex conductivity of a superconductor ? = ?1 ? i? ? ? (n ?1 ?1 ?22 n n/ne)? i?0 ? ? (2.7) where nn/ne is the normal electron density fraction, ? = 2?f is the angular frequency, and ?n is the normal metal conductivity proportional to the normal electron mean free path ?n ? `mfp [49]. Here, ?2 is London?s conductivity and ? is the resulting magnetic penetration depth of the superconductor that can deviate from the bulk London penetration depth. Using Pippard?s modification to the London theory [52], and assuming the dirty limit where the microscopic (bulk) coherence length ?0 is of the same order as the electron mean free path `mfp, the magnetic penetration depth can be approximated as ? ? ?L(?0/`mfp)1/2 (2.8) where ?0 is the characteristic coherence length of the superconductor. Eq. 2.8 suggests the screening length will effectively increase relative to the microscopic London penetration depth as a metal becomes more impure ? > ?L. Hereby the magnetic penetration depth is sensitive to material defects (e.g. impurities, grain boundaries, dislocations), where higher concentrations of defects increase scattering and therefore decreasing the `mfp. For example, bulk Nb polycrystalline material have ?L ? ?0 ? 39 nm with `mfp  ?0, whereas nanocrystalline Nb thin films have `mfp ? 7 nm due to increased defect scattering mechanisms (e.g. grain boundary scattering), and results in a magnetic penetration depth of approximately ? ? 90 nm. This is a good approximation for Nb thin films [53] and patterned submicron wires 15 [54] described in Chapters 3-5. Before the now accepted BCS theory, the two-fluid model developed by Gorter and Casimir [40] described the temperature dependent behavior of superconductors by assuming the density fraction of superconducting electrons ns/ne would drop to zero at the superconducting critical temperature Tc with dependence ?(T ) = ?(0) [1? (T/T )?]?1/2c (2.9) where ?(0) is the real (measured) magnetic penetration depth at temperature T = 0, and ? is an empirical fitting parameter. For low Tc superconductors ? ? 4 [51] (the approximation used in Eq. 2.1) and for high-temperature superconductors ? ? 2 [55]. Comparing Eqs. 2.5-2.6 and Eq. 2.7 show a relationship between the superconducting gap ? and the penentration depth ?. Abiding by the assumption that h??  ?, it is assumed here that the penetration depth ? is frequency-independent. Additionally, since all measurement temperatures are T  Tc and well below the gap frequency for Nb, it is assumed here that ?1  ?2. The complex conductivity is more accurately expressed by the BCS theory but is beyond the scope of this work, and can be found in Ref. [45] and is nicely summarized in Ref. [17]. I will refer back to the complex conductivity in the next section and use it to express the surface impedance of the superconductor. Using the complex conductivity, Nb microstrip transmission lines are modeled in Chapter 3 and simulated parameters are used to deconvolve the intrinsic surface resistance Rs from measurements in Chapters 4 and 5. 16 2.2.3 Surface Impedance The ultimate conclusion from this section will relate the complex conductivity ? into an intrinsic surface impedance Zs to obtain the mechanism of resistive power dissipation coming from the superconductor in RQL interconnect, i.e. intrinsic resistance Rs. Furthermore, the superconducting parallel plate transmission line will be introduced. It will be shown that as the magnetic penetration depth ? becomes comparable to the finite thickness and width dimensions of a superconducting microstrip transmission line, a method is needed to accurately extract Rs to compare RQL interconnects of different dimensions on equal footing and analyze extrinsic effects due to fabrication. At the end of the 19th century, engineers primarily relied on circuit theory to simplify complex circuit analysis and design without the aid of computers. Discrete components (e.g. transmission line) could be ?lumped? and connected into a network only using resistors, inductors, and capacitors. With an an electromagnetic wave applied to a combined circuit or discrete component at radio frequencies (RF), the alternating currents and voltages can be related to the circuit by an intrinsic impedance Zs introduced by Oliver Lodge in 1889 [56]. First, consider a wave propagating from vacuum to a media as shown in Fig. 2.1 with a relative dielectric constant ?r or a conductivity ? (real or complex). Fig. 2.1a is an analogy to a transverse electromagnetic (TEM) plane wave propagating along the x-direction into a perfect electrical conductor transmission line with media having a relative dielectric constant ? > 1. Using the coordinate system in Fig. 2.2, the complex voltage V and current I for a wave propagating 17 along the x-axis of a transmission line can be described by the following three equations [57] dV/dx = ?ZI (2.10) dI/dx = ?Y V (2.11) ? ? = ZY (2.12) where Z and Y are the distributed series impedance and the distributed shunt admittance of the line, respectively, and ? is the propagation constant of the line. This will be revisited in Chapter 3. The propagation of plane electromagnetic waves along the x-direction in a lossy medium shown in Fig. 2.1b can be described also by the following three equations dE/dx = ?i??H (2.13) dH/dx = ?(? + i??)H (2.14) ? ? = i??(? + i??) (2.15) where E is the electric field vector pointing along the y-direction, H is the magnetic field vector pointing along z-direction, ? is the wave angular frequency, ? = ?0?r is the permittivity of the medium, ? = ?0?r is the permeability of the medium, ?r is the relative dielectric constant of the medium, ?r is the relative permeability of the medium, ? is the medium propagation constant, and ? is the conductivity of the medium. Thus, the impedance term from transmission line theory using V and I could be extended to electromagnetic wave theory E and H, and the intrinsic impedance takes the general form and 18 Figure 2.1: Diagram showing wave propagation from vacuum into media with a relative dielectric constant ?r or conductivity ? (real or complex). a) Analogy to transmission line with media having a relative dielectric constant ? > 1 showing the Complex voltage V and current I as the wave propagates into media described by Eqs. 2.10-2.12. b) The case for a normal incidence propagating plane wave into media (see Eqs. 2.13-2.15). c) The case for a glancing incidence propagating plane wave for media with ?r > 1 with refracted angle of ? > 0. In general, the refracted wave is an in-homogeneous wave. This is a wave for which the planes of constant phase are not parallel to the planes of constant amplitude [58]. d) The case for glancing incidence propagating plane wave for media with ?r  1 with refracted angle of ? ? 0. Cases c) and d) have been adopted from [58] using Snell?s law of refraction in conductive media (see Eq. 2.17). is related to the media conductivity by ? Ey i?? ? Zs = = = (2.16) Hz ? + i?? ? + i?? 19 Fig. 2.1c and Fig. 2.1d are two cases for a wave propagating at glancing incidence relative to the vacuum-media interface. Using Snell?s law of refraction in conductive media [58], ? 2?? ?0 sin? ? sin ? ? k0? 2 sin ? ? ?/?EM (2.17) ? where ? is the incident angle for wave propagating in vacuum, ? is the refraction angle for wave propagating in the media, ? is the angular frequency, ?0 is the vacuum permittivity, ?0 is the vacuum permeability, ? is the conductivity of the medium the wave is refracted into, k0 = ? ? ?0?0 is the vacuum wave number, ?EM is the wavelength of the propagating electromagnetic wave. The left hand side approximation in Eq. 2.17 can be used for case Fig. 2.1c, where the media has an absolute relative dielectric constant greater than 1, |?r| > 1. The middle approximation in Eq. 2.17 can be used for case Fig. 2.1d for a superconductor where the relative dielectric constant is negative and its absolute value is much greater than 1 (see Section 3.4.2 where a superconductor is defined as a dielectric with negative permittivity in HFSS). The middle approximation in Eq. 2.17 assumes the conductivity of the superconductor is ? ? ?2 from Eq. 2.7 assuming ?2  ?1. Assuming a Nb magnetic penetration depth of 40 nm, ? ? 1.2? 10?5 rad. Consequently, Eq. 2.17 and Fig. 2.1d exemplify that for a normal or glancing incident wave propagating into a superconductor where |?r|  1 the refracted wave will propagate normal to the vacuum-superconductor interface. This is the basis for why a superconducting parallel plate transmission line supports a transverse magnetic (TM) propagation wave and is termed a Swihart wave [59]. In Fig. 2.2, the wave is propagating along the y-direction which is glancing incidence to the two superconductor-dielectric interfaces. At x = 0, Ey = 0 and at the superconductor 20 Figure 2.2: Diagram showing a superconducting parallel plate transmission line with a Swihart transverse magnetic (TM) wave [59] propagating along the y-axis, where looking into and out of the page the superconducting plates are infinitely long. The dielectric material between the plates has a thickness s, and each superconducting plate has a thickness d. The propagating wave is transverse magnetic because there is a small, finite y-component to the electric field at the superconductor-dielectric boundaries. The y-component of the electric field is responsible for the curvature of the electric field lines near the boundary (blue). dielectric interfaces (x = ?s/2), the Ey has a finite magnitude. It can also be seen that Fig. 2.1a and Fig. 2.2 are equivalent and Eqs. 2.10-2.12 can be used. As the dielectric thickness becomes small, and comparable to the magnetic penetration depth ?, Ey becomes significant and can greatly affect the phase velocity vp (see Appendix Eq. D.7, where fn ? vp). Assuming the superconducting plate thicknesses d and dielectric thickness s in Fig. 2.1 are much larger than the magnetic penetration depth d  ? and s  ?, respectively, the Ey magnitude is negligible to Ex magnitude, and therefore a surface impedance boundary condition (SIBC) can be enforced. The Leontovich-Shchukin SIBC approximation is derived by using the following assumptions but in the context of a good superconductor. The wavelength and magnetic penetration depth ? traveling in the media must be small relative to the material dimensions. As a consequence, the 21 material must be sufficiently thick (bulk) where reflected fields from the back surface/interface are neglected. Furthermore, the EM fields vary slowly on the surface of the medium relative to the wavelength. This constitutes a flat surface, or its smallest radius of curvature must be large relative to the magnetic penetration depth. Assuming the above, and referring to the coordinate system defined in Fig. 2.2, for a wave traveling in the y-direction along a plane surface, i.e. the plane between superconductor plates and dielectric in Fig. 2.2, the complex intrinsic wave impedance is in units of ohms per square and proportional to the ratio of the electric and the magnetic fields by ? ? ? ? Ey ? Z ? i?? ?? i??wave = ? = ? (2.18)Hz plane wave ? + i?? ???? ? where ? = ?0?r is the permeability of the medium, ? = ?0?r is the permittivity of the medium, Ex is the magnitude of the electric field pointing out-of-plane, Hy is the magnitude of the magnetic field pointing in-plane, and ? is the wave source frequency. The right approximation in Eq. 2.18 holds for good, nonmagnetic conductive media, where the relative permeability ?r and permittivity ?r are equal to unity ?r = ?r = 1, and so the permeability and permittivity are that of the vacuum ? = ?0 and ? = ?0, respectively. As it can be seen, Eq. 2.18 is not dependent on coordinates and assumes the wave impedance is constant over the plane conducting surface. In other words, the intrinsic impedance Zwave connects the electromagnetic response of conducting media with its conductivity function dependent on material properties. Eq. 2.18 can be used as a surface impedance boundary condition (SIBC), and can greatly simplify complex problems. Today, SIBC are used in finite-element modelers (FEM) to simplify geometries and numerically solve the electromagnetic fields faster. Sergey Yuferev and Nathan 22 Ida give a good overview of the different SIBCs, their history, and applications [60]. An example of using a SIBC to validate HFSS simulations can be found in Section 3.4.1. In 1938, the impedance concept was later generalized and applied to engineering applications (e.g. power engineering, transmission lines) by Sergei Schelkunoff [61] who first coined the skin effect term. In this work, assuming a superconducting medium with complex conductivity ? = ?1 ? i?2 is nonmagnetic with ? = ?0 and a good conductor with negligible displacement current ?  ??, the intrinsic impedance simplifies to ? i?0? Zs = Rs + iXs = (2.19) ?1 ? i?2 where Rs is the intrinsic surface resistance and Xs is the intrinsic surface reactance of the superconducting media. Taking the approximation for the complex conductivity in Eq. 2.7 assuming, T  Tc, ?1  ?2, and h??  ? it can be shown that the two-fluid superconductor intrinsic impedance can be written as 1 Zs = Rs + iX ? ?2 2 3s 2 0 ? ? ?1 + i?0?? (2.20) where ?0 is the vacuum magnetic permeability, ? is the stimulus angular frequency, ? is the magnetic penetration depth defined in Eq. 2.8, and ?1 is the conductivity of the normal electrons. Eq. 2.20 above models a bulk superconductor where the two-fluid model can be assumed. In this work we are concerned with measuring the intrinsic resistance Rs as it determines the power lost due to the superconductor in RQL interconnects. In this context, the intrinsic resistance is that of the effective bulk property of the superconductor and its relation to finite thickness effects will be 23 discussed later in this section. In normal metals, where ?1  ?2, the real and imaginary parts of ? the intrinsic resistance are equal, and frequency dependence Rs = Xs ? ?. For Cu at 10 GHz, the intrinsic resistance isRs ? 26m? and? 2m? at temperatures T = 300 and 10 K, respectively. For a superconductor like Nb, assuming a good isotropic superconductor with ?1  ?2, Nb has a quadratic frequency dependence of the intrinsic resistance R 2s ? ? [62, 63]. For clean bulk niobium, Rs ? 20 ? 40 ?? at 4.2 K and 10 GHz. Furthermore, for thin film Nb, the BCS Rs can be as low as 17 ?? at 4.2 K and 10 GHz when ?1 is at a theoretical optimum for the case ?0/lmfp ? 4?, where ?0 is the microscopic (bulk) coherence length [64]. This means intrinsic resistance can be lower for Nb thin films compared to bulk. To achieve minimum Rs ? 17?? in Nb, the optimal mean free path is lmfp ? 10 nm for ?0 ? 40 nm. This is conveniently close to typical Nb lmfp values for thin films having thicknesses d ? 100 ? 200 nm [53], which is the typical wiring layer thickness in RQL fabrication (see section 2.4). Using Eq. 2.8, assuming ?0 ? ?L = 40 nm, and lmfp = 10 nm yields a magnetic penetration depth of ? = 80 nm. This is within 12% of the measured value of ? = 90 nm for submicron Nb wires determined from self inductance measurements at T=4.2 K with exceptional hardware-to-model correlation [54, 65]. It is important to note that Eq. 2.20 above describes the impedance of a bulk (infinitely thick) superconductor. Here we deal with finite thin film superconductors that have been patterned into narrow strips, making the above equation incomplete. It is of vital importance in this work to ?extract? the intrinsic bulk impedance from an ?effective? impedance so we analyze extrinsic effects changing Zs as the transmission line dimensions shrink. A significant amount of work has been done to modify the impedance definition to account for non-ideal or ?non-bulk? situations, and they will be further reviewed and compared to the intrinisic impedance Zs. Starting in the 1930s, radio technology facilitated theory development of electromagnetic 24 wave propagation from a radiating antenna over the Earth?s surface. Solutions to the wave propagation can be very complex as the waves can travel in and reflect off of many media including the earth?s surface, embedded surfaces, and air over long distances where the curvature of the Earth needs to be considered. To simplify the problem, Leontovich [66, 67] and Shchukin [68] independently proposed an approach where the air region only needs to be considered by making the Earth?s surface a surface impedance boundary condition (SIBC). For the case where the magnetic penetration depth becomes comparable to the material or film-thickness ? ? 2d, modifications to the intrinsic impedance Zs need to be made. The modifications need to be made because unlike the Leontovich-Shchukin SIBC approximation, reflected fields across the bottom surface of the finite thickness superconductor cannot be neglected due to the condition ? ? 2d. Furthermore, thin-film superconductors are typically deposited on dielectric substrates, so the electromagnetic fields at the dielectric-superconducting interface as well as into the dielectric need to also be accounted for. In the 1980s high temperature superconductors (HTSs) were discovered. Klein et al. came up with the theory for the effective surface impedance Zeff in 1990 to account for finite thickness effects [69] since HTSs have very large magnetic penetration depths relative to the typical thin-film thicknesses ? ? 2d. First, assuming an infinitely thick dielectric substrate below the superconducting thin film, the effective surface impedance was calculated by using an impedance transformation method having the form Zeff = z(d)Zs (2.21) z(0) + tanh(?d) z(d) = (2.22) 1 + z(0) tanh(?d) where d is the film thickness, Zs is the intrinsic surface impedance given by Eq. 2.20, z(0) is the 25 impedance at the film-substrate interface, z(d) is the transformed impedance up to the film surface d, z(0) and z(d) are normalized to Zs of the film, and ? = ? + i? is the complex propagation constant of the superconductor. From this and accounting for the dielectric substrate, the effective surface impedance Reff of a superconducting plate can be approximated by [69, 70] Zeff = Reff + iXeff (2.23) [ ] ( ) 2 Reff = R coth (d/?) + d/? + Xs d/?s 2 2 (2.24)sinh (d/?) Zsub sinh (d/?) Xeff = ??0?eff (2.25) ( ) Zsub = ?0 1 + i tan ?1/2 (2.26) ? 2r where Reff is the effective surface resistance, Xeff is the effective surface reactance, ?eff = ? coth (d/?) is the effective penetration depth, ?0 is the permeability of free space, Zsub is the bulk impedance of the dielectric substrate, ?r and tan?? are the relative dielectric constant and loss tangent of the substrate, respectively, and ?0 ? ?0/?0 = 120? ? is the intrinsic wave impedance. It can be inferred from Eqs. 2.23-2.25 that the primary contribution to the enhancement of Zs is due to the ratio of the film thickness relative to the magnetic penetration depth d/?. As the film thickness becomes small relative to the magnetic penetration depth d/?  1, the effective impedance Zeff dramatically increases as a consequence of the current density increasing inside of the superconducting film by approximately a factor coth (d/?). Conversely, as d/?  1, Eq. 2.23 reduces to Zs of a bulk superconductor in Eq. 2.20. The second term of Reff is due 26 to power transmission into the dielectric substrate, and decreases in effect as the impedance mismatch between the dielectric substrate and and superconducting film increases Xs/|Zsub|  1 and is independent of the dielectric substrate loss tangent tan ?, assuming Xs  Rs and transmitted power is not reflected back to the film. Notice in Eq. 2.24, neglecting the substrate contribution, Reff is the intrinsic resistance multiplied by an effective geometric factor in brackets. If one wanted to compare two samples with different film thicknesses, one would need to use Eq. 2.24 to extract the intrinsic (bulk) resistance. The same concept holds for extracting the intrinsic reactance Xs from Eq. 2.25. 2.2.4 Temperature Dependence of Surface Impedance The temperature dependence of the superconductor bulk intrinsic reactance Xs (magnetic penetration depth ?) from Eq. 2.20 can be approximated by assuming two-fluid model found in Eq. 2.9. The temperature dependence of the intrinsic resistance Rs is better approximated from the temperature dependence of the BCS gap. At temperatures approximately T ? Tc/2 the intrinsic resistance scales exponentially on the ratio of the BCS gap and temperature by Rs(T ) ? e??(T )/kBT . In this temperature range, the magnetic penetration depth is assumed to be constant, having the value ? found in Eq. 2.9. Below a certain temperature the intrinsic resistance becomes temperature independent having the value Rres, and is Tres ? 1.8 K for thin film Nb [71]. With this, the temperature dependence of the intrinsic impedance Zs0 at a reference frequency ?0 at T ? Tc/2 can be approximated by ??(T ) Rs0(T ) = R (0) 1 e kBT +R A ?B(T/Tc)s0 res0 ? e + C for T ? Tc/2 (2.27)T T Xs0(T ) = ?0?0?(T ) coth (d/?(T )) (2.28) 27 ?(T ) = ?(0) [1? (T/T )?]?1/2c (2.29) where Rres0 is the residual intrinsic resistance at a reference frequency ?0 below T ? 1.8 K for thin film Nb, Rs0(0) is the BCS zero temperature intrinsic resistance, A is an empirical fitting parameter, B/Tc = ?(0)/kBTc is the gap ratio fitting parameter, C = Rres0 is the residual intrinsic resistance fitting parameter, d is the superconductor thin film thickness, and ?(0) is the magnetic penetration depth at zero temperature. Valente-Feliciano et al. found B/Tc = ?(0)/kBTc ? 1.89 for Nb thin films [71], within 5% agreement to approximation of ?(0)/kBTc ? 1.97 made in Eq. 2.4. The approximation on the right in Eq. 2.27 will be used in Chapter 5 to fit temperature dependent microstrip transmission line (MTL) data having fitted values of ?/kBTc ? 2.1. 2.2.5 Summary In summary, to meaningfully compare results for RQL interconnects having different microstrip transmission line (MTL) widths requires the determination of the intrinsic resistance Rs shown in Eq. 2.20. Thin-film superconductors screen electromagnetic fields on the length scale of the magnetic penetration depth ?. Increased defectivity in the material results in an increase in ?, causing the magnetic field penetration to drive further inside of the thin-film or MTL. Furthermore, an increase in the magnetic penetration depth means the RF currents sample defects buried deeper in the wire, and can induce an increase in Rs. The surface impedance boundary condition (SIBC) first introduced by Leontovich and Shchukin simplifies the problem making it easier to account for non-idealities relative to bulk solutions. Klein et al. approximated the SIBC condition for an infinitely wide thin film superconductor on a dielectric substrate where 28 ? is comparable or much larger than the thin film thickness ? ? 2d. Although, an additional treatment is needed to account for lossy MTLs where ? also becomes comparable to the dielectric separation thickness s between the conducting strip and the groundplane as well as the MTL width w. With the advancement and maturation of finite element modelers, it is possible to numerically solve the electromagnetic fields and extract the intrinsic impedance Zs, and the method will be carried out in Chapter 3 to extract Rs from measured data in Chapter 4 and 5. 2.3 Amorphous Dielectrics Silicon dioxide SiO2 fabricated via chemical vapor deposition using a tetraethyl orthosilicate (TEOS) precursor is the primary dielectric material used in the microstrip transmission line resonators measured in chapters 3-5. Silicon oxide in this context will be referred to by its precursor TEOS throughout this dissertation. TEOS is deposited on Si wafers and is limited to temperatures T ? 150?C (see section 2.4.6). Using common plasma-enhanced chemical vapor deposition processes similar to references [72, 73], the kinetics are such that surface diffusion is limited, preventing the growth of a crystalline lattice. The TEOS amorphous structure is confirmed by the presence of a diffuse ring in electron diffraction image similar to reference [74]. The combination of the T ? 150?C thermal budget and amorphous structure of TEOS limits the lower bound of the loss tangent to approximately tan ? ? 1?10?3 at 4.2 K. Comparatively, this is a factor of 3-4 of magnitude higher than thermal oxide grown SiO2 [75]. Therefore, this gives rise to non-ideal high power dissipation coming from the TEOS dielectric in RQL interconnects and is a motivation of this work to review the fundamental material properties and loss mechanisms. 29 The typical operating temperatures for classical superconducting logic and quantum computing circuits are T ? 4.2 K and T  1 K, respectively. At these temperatures, the power loss from the amorphous (glassy) dielectrics depends strongly on the stimulus energy h??, where ? is the applied electric field frequency. Therefore, it is important to briefly review the history and explicitly state the temperature and frequency regime carried out in measurements performed in chapters 4 and 5. First, I will introduce the fundamental dielectric loss tangent tan ? that describes how energy is dissipated in a dielectric material due to relaxation processes. Then, I will briefly review the history of low temperature measurements of tan ?. Here, I will focus the review on the difference between crystalline and amorphous dielectrics. I will also incorporate the relative dielectric constant ?r, as its temperature and frequency dependence are governed by the same mechanisms as tan ?. For a proper introduction, the two-level system model will be included in the discussion. To conclude this section, I will summarize the historical work that has been done by showing the universality of the tan ? temperature dependence and its respective mechanisms, where varying ? more or less shifts the intersection of temperature regimes. Finally, I will explicitly state the mechanism dominating tan ? in different temperature and frequency regimes used in the measurements carried out in Chapters 4 and 5. 2.3.1 The Dielectric Loss Tangent tan ? A simple RC circuit analogy can be used to describe dielectric relaxation [76]. In a linear system a dielectric material can be modeled having a resistance R due to some Joule losses in the dielectric, and capacitance C capable of holding a chargeQ. Here, we assume a seriesRC circuit with time-independent R and C. For a step-function voltage applied to the dielectric, where the 30 voltage maximum is V = V0 at time t = 0 and zero V = 0 for time t > 0, the time-dependent discharging can be described as Q = CV0e ?t/?rel (2.30) where Q reaches a fraction 1/e after a time ?rel, which is the characteristic relaxation time, or RC time constant equal to ? i?trel = RC. For a sinusoidal voltage stimulus V (t) = V0e at angular frequency ?, the charge has both in-phase and quadrature responses, and can be considered complex, and takes the form ( ) 1 ? ??relQ(?, t) = CV0 i ei?t (2.31) 1 + ?2? 2rel 1 + ? 2? 2rel Eq. 2.31 shows a time lag between the charge response Q and applied voltage V . This phase lag can be represented by the tangent of an angle ? as tan ?. Assuming tan ? << 1 implies ? ? ??rel, where ?rel is the characteristic time lag between Q and V . The time lag induces a dissipation of energy. In other words, if the sinusoidal voltage was turned off, the dielectric charge will ?relax? to a lower energy state, and that time lag, or characteristic relaxation time ?rel will dissipate some amount of energy. It can be shown that the energy dissipated W and ? are related by W ? sin ? [76]. For small angles ?, the energy dissipated in the dielectric is directly proportional to ?, hence the commonly used term loss tangent tan? and loss angle ?. For a parallel RC circuit, which is more analogous to our circuits here, the loss tangent is related to dielectric relaxation time tan ? = 1/??rel (2.32) 31 Both the phonon (quantized lattice vibrations) and polarization (charge) can be described with the above RC circuit model. For an acoustic frequency (sound wave) stimulus, the acoustic energy loss is caused by the lattice viscosity, or resistance of the vibrating atoms to displacement. Here, the resistance R is proportional to the viscosity and C is proportional to the stored energy in lattice vibrations. The time lag of lattice displacements induces energy loss. If a dielectric material can be electrically polarized by a stimulus due to some displacement of charge (e.g. doped dielectric crystals, ionic crystals, amorphous dielectrics), then charge displacement relative to its lattice vibrations under electromagnetic stimuli will induce power dissipation. Hence, phonon vibrational modes and electronic polarization are coupled and dissipate power in dielectric materials. 2.3.2 Dielectric Loss Mechanisms and Phonon Scattering at T < 10K As mentioned at the beginning of section 2.3, the TEOS dielectric measured chapters 3-5 is considered amorphous. Therefore, in the context of this work, only amorphous dielectrics will be considered, but it is important to briefly summarize the differentiation between the material properties of glassy and crystalline dielectrics at cryogenic temperatures approximately T < 100 K. The surprising difference was seen in heat capacity measurements of vitreous, meaning glassy, silica and crystalline ??quartz by Zeller and Pohl in 1971 [77]. At 0.1 K, the specific heat capacity CD of vitreous silica was 2 orders of magnitude higher than quartz. Crystalline quartz follows the typical Debye dependence CD ? T 3 below its Debye temperature approximately T  TD ? 360 K [78]. At these temperatures, the phonons, or lattice vibrations, can only occupy long wavelength vibrational modes and can be treated as an elastic continuum, and so the 32 heat capacity is proportional to the thermal conductivity by 1 ?D = CDvD`ph (2.33) 3 where CD is the Debye specific heat capacity per volume of the heat carrying phonons, vD is the phonon velocity, and `ph is the phonon mean free path due to scattering by lattice defects. Referring back to the Zeller and Pohl measurements, it was found that vitreous silica, and universally for amorphous dielectric materials for T  TD, have a temperature dependent specific heat and thermal conductivity proportional to ? T and ? T 2, respectively. Early phenomenological models tried to modify the specific heat capacity alone to account for the ? T dependence, but were not consistent and are further reviewed in [79]. This led to further experimental evidence that the anomalous thermal properties arise from additional excitations inducing a change in the specific heat and suggests the phonon mean free path has a temperature dependence `ph(T ). Similarly to the thermal properties, there is a close relationship between the acoustic and dielectric attenuation/loss properties in amorphous dielectrics at temperatures approximately below 100 K [80]. For example, for amorphous dielectrics there is a quintessential maximum in loss/attenuation at a temperature Tmax. Anderson and Bommel confirmed this by plotting the temperature of this maximum loss peak as a function of stimulus frequency ?/2? and found exceptional agreement between acoustic and electrical measurements (see Fig. 2.3). An Arrhenius relationship suggests a thermally activated process and will be mentioned briefly later. Notice, Fig. 2.3 shows that Tmax will increase to higher temperatures as the measurement frequency is increased. In 1972 Anderson [82] and Phillips [83] proposed a two-level system (TLS) model to 33 Figure 2.3: This figure is adopted from reference [81] and discussed in reference [80] Fig. 6.2-6.3. 34 describe the thermal properties of amorphous dielectrics at temperatures below the crystalline Debye temperature TD. In the TLS model, atoms (or electrons, no one knows) occupy one of two minima capable of tunneling between the two states quantum mechanically. Consequently, the manifestation of tunneling between two energy states scatters phonons, prompting an additional mechanism of dissipated energy (i.e. TLS loss). With this, the TLS model universally explains the temperature dependent specific heat ? T and thermal conductivity ? T 2 of amorphous dielectrics at temperatures T < 1 K. Effectively, and to support the context of this dissertation, the anomalously low thermal conductivity for amorphous dielectrics at temperatures T  10 K arrises from low phonon scattering lengths `ph as a result of strong resonant absorption of thermal phonons by the TLSs [80]. Referring back to Fig 2.3, the dissipated loss coming from TLSs have been measured by acoustic attenuation and the dielectric loss tangent, both of which are related. The measured acoustic attenuation ?db is proportional to the inverse phonon mean free path ? ? `?1db ph . Therefore, at temperatures below T < Tmax and at a given reference frequency ?0, the temperature dependent dielectric loss tangent and phonon mean free path are approximately proportional tan ?(T, ?0) ? `?1ph (T, ?0) [84]. The frequency dependent proportionality between dieletric loss tangent and ultrasonic absorption is tan ? ? ?db/?. Note, the frequency dependence of ?db depends on which temperature below T < Tmax the measurement is taking place, and will be discussed further. Since it is established that energy dissipation measured by the loss tangent tan ? is coupled to phonon scattering at temperatures T < 10 K, I will focus on the mechanisms causing the temperature and frequency dependence of `ph(T, ?). The TLS loss can be split up into two phonon scattering mechanisms, TLS tunneling due to resonant absorption `tun,res and to relaxation 35 `tun,rel [79]. Resonant absorption occurs when TLSs are resonant with the applied electric or acoustic fields at particular energies and frequencies. Relaxation of TLSs occur when non- resonant TLSs having a dipole moment are driven out of non-equilibrium in an applied field and scatter phonons as the TLSs ?relax? in a decaying oscillatory fashion during redistribution between the energy levels. Schickfus et al. was the first to measure the energy dissipation in vitreous silica due to both TLS loss mechanisms below T  1 K [84] using the 2 cavity test setup [85]. Loss due to resonant absorption manifests in measurements at low applied electric fields, and at higher applied fields the TLSs are saturated and loss is dominated by TLS relaxation. At higher temperatures in the approximate range 10 K < Tmax ? 100 K, ?classical? relaxation dominate the phonon scattering process. Contrary to TLS relaxation via tunneling between two energy states, classical relaxation scatters phonons by the dissipation of energy from physical charge displacement (see Eq. 2.32) [76, 86]. The phonon mean free path due to classical relaxation is `class,rel. Temperatures higher T > 10 K will not be considered since it is beyond what is measured in chapter 3 and 4, but are mentioned for completeness. Therefore, the three mechanisms contributing to the total phonon free path lph(T, ?) at temperatures below T < 10K with temperature and frequency dependence are l (T, ?)?1 = `?1ph tun,res(T, ?) + ` ?1 tun,rel(T, ?) + ` ?1 class,rel(T, ?) (2.34) where the phonon scattering mean free path due to TLS resonant tunnelling, TLS tunneling relation, and classical charge (dipole) relaxation are `tunn,res, `tunn,rel, and `class,rel, respectively. Mentioned previously, the dielectric relaxation is proportional to the ultrasonic attenuation tan ?(T, ?0) ? ?db(T, ?0) [84]. The ultrasonic attenuation is inversely proportional to the phonon 36 mean free path by ?db(T, ? ?10) ? lph (T, ?0) (see [79]). Accordingly, the total dielectric loss tangent tan ? can be split up into the three loss mechanisms from Eq. 2.34 in the following way tan ?(T, ?) = tan ?tun,res(T, ?) + tan ?tun,rel(T, ?) + tan ?class,rel(T, ?) (2.35) where the dissipative loss associated TLS resonant tunnelling, TLS tunneling relaxation, and classical charge (dipole) relaxation are tan ?tunn,res, tan ?tunn,rel, and tan ?class,rel, respectively, and are all temperature and frequency dependent. 2.3.3 Temperature Dependence of tan ? Referring to Fig. 2.4 starting from the lowest temperature, depending on the applied electric field strength E, the dielectric loss is dominated by either TLS resonant tunneling (absorption) or TLS relaxation processes. The response of TLSs to applied electric fields is dependent on the coherence between their two eigenfunctions formally treated for solving the problem developed for optical saturation [87] and magnetic resonance [88]. These methods have been applied to TLSs in glasses [89] and a straighforward summary can be found in a review by Phillips et al. [79]. The treatment for TLS field strength dependence in the acoustic case can be translated to the electric case and the equations from Schickfus et al. for tan ? dependence on field strength will be presented in Chapter 5. Essentially, at very low temperatures T  1 K, the critical intensity ITLSc (in Wm ?2) is defined as the cross-over point from TLS resonant absorption (weak field) to TLS relaxation (strong field) and can be written as [79] 2 TLS 3h? ?r?0cIc = (2.36)2p20T1T2 37 where ?r is the relative dielectric constant, p0 is the electric dipole moment (inCm), c is the speed of light, and T1 and T2 are the TLS relaxation and dephasing times, respectively. For electric field intensities greater than the critical intensity ITLS  ITLSc , TLS resonant absorption is saturated and relaxation processes will dominate. For electric field intensities less than the critical intensity ITLS  ITLSc , TLS resonant absorption processes will dominate. Referring again to Fig. 2.4, below T? and for weak electric fields, tan ? is dominated by TLS resonant absorption processes and is proportional to the inverse of the phonon mean free path with temperature dependence tan ? ?1res,tun(?0) ? `res,tun(?0) ? tan h??0/kBT . Below T? , for strong electric fields, and for long relaxation minimum TLS relaxation times ??min  1, tan ? is dominated by TLS relaxation and so called ?one-phonon? processes having temperature dependence tan ?rel,tun(?0) ? `?1rel,tun(?0) ? T 3 [86]. Above T? , where relaxation times begin to decrease due to thermal excitation ??min  1, TLS relaxation still dominates but a ?plateau? is reached for a long range of temperatures yielding tan ?rel,tun(?0) ? `?1rel,tun(?0) = const. Here, it is important to note that I am not reviewing or incorporating the soft potential model first proposed by Karpov et al. and thoroughly reviewed in references [90, 91]. Essentially, the soft potential model can explain the ?plateau?? and other features in the data for measurements at higher frequency (? THz) by incorporating the existence of soft localized modes accounting for anharmonicity. For the measurements carried out in chapters 3 and 4, the temperature ranges are below T? in the frequency range ?/2? = 1 ? 20 GHz. Above T ?, classical relaxation processes dominate where charge ?hopping? dominates with temperature dependence tan ?rel,class(?0) ? `?1rel,class(?0) ? T [76, 79, 86]. At Tmax, the thermal energy is equal to the ?hopping? barrier height and thermal vibrations now dominate the system above this temperature where static spin polarizations are negligible [86]. 38 Figure 2.4: Diagram depicting the regimes of mechanisms dominating the temperature dependent tan ?. The general diagram was adopted from reference [86] for dependence of `ph and modified to include the TLS absorption regime at the lowest temperatures and lowest applied E fields, which is the most widely used and understood regime today. The loss tangent and inverse mean free path at a reference frequency ?0 have temperature dependent proportionality tan ? ? `?1ph The majority of the notations are adopted from references [90, 91, 92] that thoroughly review all of the mechanisms relating to energy dissipation, but the soft potential model is not included or reviewed here. The y-axis is plotted as the inverse quality factor or internal friction Q?1 from acoustic measurements, which is proportional to the loss tangent tan ?, the ultrasonic attenuation ?db typically measured in dB/cm, and the inverse phonon mean free path `ph by Q?1 ? tan ? ? ?db ? `?1ph . The critical temperatures T? , T ?, and Tmax are the transition temperature regimes are dominated by TLS resonant tunneling, TLS relaxation, classical relaxation processes as temperature is increased. Note, this entire plot can shift to the left in or regions can squeeze/widen in temperature depending on the relative energy scales of the various processes to the thermal energy kBT , applied frequency ?. 39 Note that the general trend in Fig 2.4 can shift left/right, or regions can squeeze/widen, in temperature depending on the relative energy scales of the various processes unique to the material such as (i) the thermal energy kBT , and (ii) applied frequency ?. As an example, internal friction data was taken from reference [93] and re-plotted to estimate T? in Fig. 2.5. Note, the inverse acoustic internal friction Q-factor is proportional to the loss tangent Q?1 ? tan ?. These plots clearly show a frequency dependence of T? . This is analogous to the frequency dependence of Tmax mentioned and confirmed by Anderson and Bommel shown in Fig. 2.3. It can be estimated that T? ? 22K at a measurement (stimulus) frequency of 10 GHz for ??SiO2. This serves as an initial estimate for measurements performed here at temperatures below 10 K and applied frequencies of 10 GHz suggesting the dominant mechanism contributing to loss is the TLS relaxation processes in Chapters 4 and 5. It is interesting to note that below T? , the temperature power law dependence of tan ? is T 2.3, which is less than T 3 for the 43 MHz measurement in Fig. 2.5. Schickfus and Hunklinger measured a tan ? temperature dependence of T 2.4 from 2-8 K in the TLS relaxation regime at 10 GHz [84]. Furthermore, Strom et al. measured a tan ? ? T 2.4 dependence in the relaxation regime for Na ?-alumina at 11.5 GHz with the transition temperature between TLS absorption and relaxation at 5 K [94]. Therefore, deviations from tan ? ? T 3 dependence in the TLS relaxation regime is not uncommon. This is most likely due to the measurements taking place at temperatures near T? or near the TLS resonant absorption regimes. Here I assume TLS relaxation processes dominate dielectric loss tan ? in measurements carried out in this work. I will not treat losses due to TLS absorption processes since measurements here are carried out above T > 1K and all TLSs are saturated above this temperature. Although, the temperature dependent data in chapter 4 suggests we see the onset of TLS absorption regime. 40 Figure 2.5: The (left) plot is data re-plotted from reference [93] Fig. 3. In this plot, i fit the low temperature regime (TLS relaxation) and the higher temperature regime (classical relaxation) of the 43 MHz data (yellow). The (right) plot is estimated T? temperatures taken at the arrows shown in (left) plot. At 10 GHz, the extrapolated temperature is T? ? 22K. With that, the model used here for the tan ? temperature dependence is tan ?(T, ?) ? tan ? 3tun,rel(T, ?) ? T (2.37) 2.3.4 Frequency Dependence of tan ? Considering only loss due to TLS relaxation with ??min  1 and T < T? , the phonon scattering length `ph ? `tun,rel is constant as a function of frequency (see Eq. 2.34). In this temperature regime, the loss tangent has a frequency dependence tan ? ? ??1, T < T? , strong E field (2.38) 41 Although, specifically in the amorphous dielectric TLS community, the frequency dependence of tan ? in Eq. 2.38 is rarely measured or has possibly never been explicitly confirmed experimen- tally. For comparison, the frequency dependence of tan ? in the temperature regime where TLS absorption dominates, the loss tangent is frequency independent by tan ?(?) = constant, T  T? , weak E field (2.39) In parallel communities the frequency dependence of dielectrics with finite electrical conductivity have been measured suggesting universality due to many body interactions. The electrical conductivity is related to the dielectric loss tangent by [95] ??? + ? tan ? = ? (2.40)? where ? is the angular frequency, ? and ?? are the real and imaginary part of the complex dielectric constant, respectively, and ? is the finite electrical conductivity in the dielectric. In 1961, Pollak et al. observed a frequency dependence of the conductivity ? in n-type Si at temperatures T ? 4.2K having the form [96] ? = As? b (2.41) where As and b are empirical fitting parameters. Pollak et al. found that b ? 0.8 fit for a range of temperatures below 10 K. The model described by Pollak et al. explaining the frequency power dependence requires low concentrations of localized conduction states allowing electron (charge) 42 to ?hop? between states giving rise to an increase in conductivity. With the application of a steady state alternating current, charge will be displaced by the ionized group-3 and group-5 impurities systematically incorporated into the Si by Pollak et al. experiment. The ?hopping? time of the charge is analogous to a relaxation time having some phase lag relative to the applied electric field (refer back to Eq. 2.31 and discussion). Therefore, energy can be dissipated through this mechanism and witnessed by measurements of the loss tangent tan ?. A host of studies found a similar frequency dependence in materials with low conductivity, predominantly dielectrics, that satisfy the presence of non-interacting dipoles free to rotate or displace in a medium. In 1977 Jonscher [97], and later updated by Ngai et al. in 1979 [98], presented the universality of dielectric frequency dependent response supported by half a century worth of data. The universal power law proposed have the forms [97] tan ? ? ?b?1, 0 < b < 1 (2.42) ? ? ?b, 0 < b < 1 (2.43) Ngai et al. refers specifically to TLS as a possible electron-phonon mechanism adhering to the power dependent universality [98]. Although, the TLS and the universal dielectric power law were not explicitly connected. Kaiser et al. argues that measurements using lumped element resonators at 4.2 K of amorphous thin film dielectrics, TLS relaxation processes dominate the measured tan ? [99]. Assuming non-interacting dipoles with equal relaxation times ?rel at GHz frequencies, Kaiser et al. makes the power exponent a free fitting parameter b to fit to Eq. 2.43 [97, 98]. Kaiser et al. suggests the parameter bd depends on the coupling of TLSs to each other, where bd = 0 is 43 non-interacting (Debye-like), bd = 0.5 nearest neighbor interactions, and bd > 0.6 is attributed to many-body interactions [98]. Although outside of the scope of this work, it is still worth noting that amorphous and crystalline metals can exhibit similar exponents with bd ? 0.8 in Pollak?s experiments of n-type Si as an example mentioned earlier. This is worth exploring further in experimentation building off the work presented in Chapters 3-5. Considering both the (i) proposed universal frequency dependence in Eq. 2.43, and (ii) TLS relaxation frequency dependence in Eq. 2.38 assuming ??min  1 and T < T? , the empirical frequency dependence of the dielectric loss tangent can be fit to tan ? = B ?b?1d , 0 < b < 1 (2.44) where Bd and b are power law free fitting parameters. For b = 0, the frequency dependence simplifies to that expected for TLS relaxation. For 0 < b < 1 the frequency dependence follows universal scaling supported by measurements at 4.2 K [99]. 2.3.5 Summary In this work, the dielectric studied is tetraethyl orthosilicate (TEOS). The thin film TEOS material fabricated (see Section 2.4 and 3.2) and measured (see Chapters 4 and 5) is considered to be an amorphous (glassy) dielectric. The resonator measurements carried out here (see Chapters 4 and 5) were done in the temperature and frequency range 1.5-10 K, 1-15 GHz, respectively, in an electric field regime between strong and weak. In this range, I assume two-level system (TLS) relaxation processes are dominating tan ?. 44 In Chapter 4, for the dielectric loss tangent tan ?, we will assume both non-dispersive (tan ? = constant) [100] and dispersive (tan ? ? ?b?1) [79, 97, 98, 99, 101] loss tangent models over a frequency range ?/2? ? 1 ? 12 GHz at T = 4.2 K. In Chapter 5, I will attempt to fit the frequency dependence of tan ? as a function of temperature, and compare tan ? temperature dependence to the expected? T 3 for TLS relaxation processes [84]. Furthermore, I will claim the non-monotonic temperature dependence of tan ? is due to TLS relaxation and resonant absorption loss and leads to TEOS being the dominate loss mechanism in RQL interconnects below 4 K. 2.4 Fabrication Processes for RQL Interconnects 2.4.1 Brief Background of CMOS Technologies Both CMOS and superconducting electronics (SCE) technologies have similar interconnect fabrication challenges, one of them being planarization. In CMOS, to achieve higher performance and lower cost per chip for successive technology generations, there is an approximate linear increase in number of devices leading to a quadratic increase in the number of interconnects required [102]. This demand is met by reducing the pitch between wires or increasing the total number of interconnect layers. This increase in density has a negative effect on performance, and has been solved by qualitative improvements of the fabrication processes. Planarization of each successive layer is paramount when increasing density, as the minimum feature sizes and pitch are set by the depth of focus in the lithography process [103]. To fabricate interconnects, current CMOS technologies use some form of a damascene Cu electroplating process for on-chip metalization [104]. The damascene process will be explained below. Electroplating is the ideal process to conformally fill high-aspect ratio dielectric trenches 45 Figure 2.6: a) SEM image adopted from reference [105] as an example of high aspect ratio trench filling using Cu electroplating in a CMOS damascene process (before metal planarization). b) RQL damascene process after planarization showing a Nb interconnect deposited using magnetron sputting at T ? 150? as an example of low aspect ratio trench filling. More characterization on this in Chapter 3. with Cu. SCE technologies are currently limited to trench aspect ratios of < 1 and resort to less conformal sputtering processes. Fig. 2.6 exemplifies the low aspect ratio filling of RQL interconnects relative to common CMOS electroplating technologies [105, 106, 107, 108, 109]. In CMOS, the push for higher density of submicron interconnects increases the overall resistivity of the Cu. This increase in resistivity induces an increase in energy dissipation and Joule heating. RQL interconnects can provide advantage over this, and was discussed in Section 1.2. 46 2.4.2 SCE Fabrication Technologies Currently, the following companies have advanced fabrication processes producing SCE logic circuits: SEEQC [110, 111, 112], National Institute of Advanced Industrial Science and Technology (AIST) in Japan [113, 114]; D-Wave Systems Inc. using Cypress semiconductor foundry in Minnesota (proprietary process unpublished, reviewed here [115]); and MIT Lincoln Laboratory (LL) in Massachusetts [115]. Parallel to semiconductor microelectronic scaling, interconnect wire width, pitch, and total number of layers determine the scalability of SFQ circuits. The LL process currently has eight wiring layers with minimum wire width and spacing (pitch) of 0.35 and 0.5 ?m, respectively [115]. Unlike CMOS where minimum dimensions are defined by the transistors, the interconnects in SCE have comparable or larger minimum dimensions relative to the logic units (Josephson junctions), therefore determining current scaling of the technology. Niobium is the most commonly used metal in SCE, as it has a relatively short penetration depth ? and long coherence length [53, 54], relatively low RF losses [71], and can be consistently fabricated having high critical current densities [54]. Silicon dioxide SiO2 is the most common interlayer dielectric used in SCE and is adopted from semiconductor technologies [116]. These dielectrics provide isolation between metal layers, and the dielectric synthesis and quality can have an impact on the metal layer performance (inductance, RF loss) as well as determining the speed and dispersion of data propagation [36], governed by the relative dielectric constant ?r and loss tangent tan ?, respectively. 47 2.4.3 Chemical Mechanical Planarization (CMP) For both CMOS and SFQ based planarized processes (e.g. [115]), interconnect fabrication cannot be made without chemical-mechanical planarization (CMP) [117, 118, 119, 120]. Pioneered at IBM, CMP is a technique to remove material with undesired topography by abrasive action in a chemical etching environment. The physical abrasion is done using a rotating polishing pad in contact with the wafer mediated by a chemical alumina slurry. There is also a pad conditioner to remove residual contaminants to maintain constant abrasion on the wafer (i.e. ?sandpaperness?). Furthermore, there are different types of endpoint detection methods to determine when the metal/dielectric has been planarized. The two most common in-situ endpoint detection methods are: ISRMTM (interferometry based) [121, 122] developed by AMAT [123], and eddy current based techniques [124]. There are also many other techniques to perform endpoint detection [125, 126]. This process can achieve a planar surfaces to sub nm levels over a 200 mm wafer, but due to the coupled physical-chemical nature of this technique, induced material defects are of primary concern. The two primary planarized processes are the cloisonne? and damascene process. In regard to availability, these planarization processes enable both high metal densities per layer (approximately 50-95%) and many layers to be used as interconnects. Although, for superconducting interconnects, as wire dimensions decrease so does the electron mean free path giving rise to higher surface resistanceRs and power dissipation (see Eq. 2.20 and Sections 2.2.2- 2.2.3). 48 Figure 2.7: Superconducting Electronics (SCE) fabrication processes showing the generalized a) cloisonne? process involving a metal etch followed by a dielectric CMP, and b) the damscene pro- cesses involving a dielectric etch followed by a metal fill and CMP. Here, the dielectric Tetraethyl orthosilicate (TEOS) and metal niobium (Nb) were used to fabricate (MTL) interconnects with both the cloisonne? and damascene process. Samples from Process A and B were fabricated using a damascene process. Samples from Process C were fabricated using a cloisonne? process. Physical characterization of the MTLs can be found in Section 3.2. Electrical characterization of the MTLs can be found in Chapter 4 and Chapter 5. 49 2.4.4 Cloisonne? Process (Dielectric CMP) Arguably, the first metal-dielectric fabrication process was Cloisonne?, which is the ancient art originating from Egypt of decorating objects with intricate metal designs encapsulated in insulating enamel to give a colorful finish. The resulting surface of the object is rough after the firing step to adhere the enamel onto the metal and is sanded down to give it a smooth finish. In microelectronics fabrication, the Cloisonne? process is started by depositing the interconnect metal (Cu, Nb) with physical vapor deposition (PVD). After the dense metal (e.g. Cu, Nb) interconnect design has been defined by photo-lithography and patterned using reactive-ion etching (RIE), the metal is embedded with insulating dielectric (e.g. SiO2, SiNx) using chemical-vapor-deposition (CVD) processes [127]. Since CVD deposits conformally over the metal, the resulting surface is planarized to make the surface planar and smooth. See Fig. 2.7. PVD and RIE processes are relatively inexpensive and produce excellent metal thickness uniformity (between 1-10%). Additionally, RIE is more scalable than the liftoff process [118]. Although, the cloisonne? process is unable to produce higher density (small wire pitch) interconnects. The RIE etch selectivity at corners of the photoresist and high metal density areas, which ultimately limit the metal thickness and induced metal sidewall angle. Typically in practice, CMOS resorts to a damascene (metal CMP) process to achieve the highest metal densities. Fortunately, there are many ?knobs? that can be turned turned in CMOS/SCE to get the cloisonne? process to meet desired requirements. 2.4.5 Damascene Process (Metal CMP) In constrast to cloisonne?, the damascene process [119, 120] starts with a uniform dielectric deposited on a planarized wafer followed by patterning the dielectric ?trenches? using photolithography 50 and oxide RIE. Metal is then deposited to fill the troughs, over filled (over burdened), and then metal is CMPed down to the dielectric surface, leaving behind only recessed metal wire. See Fig. 2.7. This process is the primary planarization process producing the highest yield and largest processing window. A large processing window means the performance of the device or product being fabricated is not sensitive to small changes in the fabrication process (e.g. small changes in temperature during metal deposition). In general, dielectrics are amiable to the process (lithography, RIE). As a consequence, this technique suffers from the poor trench filling qualities of refractory metals. Magnetron Sputtering techniques, although optimal in many ways (e.g. high deposition rates, process control, thickness uniformity), do not uniformly fill trenches because the flux concentration of atoms towards the recessed areas is hard to control [128]. Additionally, uniform coverage inside the trench facilitated by surface diffusion of adsorbed atoms is almost is extremely difficult due to magnetron sputter energy scales [129]. Because of this, trench filling at corners is especially difficult and commonly causes voiding inside of the trench due to metal deposited on the upper edge surface eventually blocking incoming flux. At the same time, this blocks atoms/plasma from leaving the trench. As one can expect, this process can induce a plethora of defects on the metal (contaminants, residual stress, voids, etc) due to trench filling and CMP. Both cloisonne? and damascene processes used in CMOS/SCE technologies give a planarized interconnect balanced by performance trade-offs. Changes in the process parameters have been shown to mitigate short comings of each process [102, 120]. For example, depsition parameters (e.g. deposition pressure, RF/DC bias, substrate temperature) have been tuned to achieve optimal adatom energies for uniform filling of trenches. Thin protective layers (e.g. Ti, Al, Ta) have been used as diffusion barriers. Trade-offs can be further mitigated to enhance performance by 51 advances in the fundamental understanding of the materials and their response to fabrication changes. 2.4.6 Thermal Budget Somewhat analogous to atomic mobility through self-diffusion in Al based wiring in CMOS due to fabrication temperatures reaching 40% of the Al melting temperature [102], the thermal budget in SFQ technologies is primarily set by the activation temperature of oxygen migration into Nb interconnects and intermixing at Josephson junction oxide interfaces. The latter could be alleviated by having the Jospehson junction integration at the end of the fabrication process. Therefore Nb interconnects are the thermal budget limiting factor where fabrication temperatures do not typically exceed T ? 150?C. As a reference, 0.1 and 1 at. % of oxygen in Nb reduces the superconducting critical temperature by approximately 0.2 and 1 K, respectively [130]. In addition, the normal state resistivity at 4.2 K has been shown to increase by a factor of 30 for 0.5 at. % oxygen in Nb. An increase in normal state resistivity will decrease the electron mean free path ` and increase the magnetic penetration depth ? (see Eq. 2.8). With this, low deposition temperatures make it difficult to fabricate high quality SiO2 and Nb thin film material. For CMOS fabrication, the typical deposition temperature is T ? 350?C for SiO2 using a TEOS precursor, which is approximately 1/3 the melting temperature of Cu. An increased thermal budget would allow for higher substrate (wafer) deposition temperatures, facilitating ideal growth kinetics of metals/dielectrics for high quality materials. 52 2.5 Summary In summary, the fabrication process has a significant impact on RQL interconnect morphology of Nb wires and the low temperature thermal budget invokes the TEOS dielectric to be amorphous in character. Consequently, the intrinsic microwave losses need to be monitored for changes in the process and the mechanisms causing extrinsic loss should be mitigated. Models of the microwave losses will be presented in Chapter 3 and demonstrated in Chapter 4 at 4.2 K. The morphology will be further explored in Chapter 3, and then compared to RF characterization in Chapter 4. Temperature dependent models will be presented and demonstrated in Chapter 5. 53 Chapter 3: Modeling Lossy 3D Superconductor-Dielectric 2-Port Networks 3.1 Introduction A resonator remains the only practical way to accurately measure intrinsic resistance Rs [62, 131] and loss tangent tan ? [100, 132] at microwave frequencies. The resonator internal Q- factor is related to the partial Q-factors associated with the conductor loss Qc and the dielectric loss Qd by Q?1i = Q?1c +Q ?1 d . There will be more discussion on the measurement techniques in Chapter 4. For an MTL embedded in a single dielectric material, the loss tangent is related to the partial dielectric Q-factor by tan ? = Q?1d and is independent of the MTL geometry. To extract the intrinsic resistance from measurements of Qc, one needs to accurately solve the geometric factor ?, where Rs ? ?/Qc. The geometric factor can be solved analytically for simple geometries or numerically for more complex geometries. The analytical modeling work to account for Zs (and ?) contribution on transmission line characteristics are accurate only when the current distributions are known and/or geometries are such that the conducting strip width w is greater than the dielectric thickness s: w/s > 1 (see Table 3.1) [133, 134, 135, 136, 137]. Sheen et al. describes the limitations in detail, presents a new method accounting for current distributions, and numerically solves for the lossy stripline parameters (e.g. ?) using a proprietary Finite Element Modeler (FEM) [138]. 54 Recently, commercially available FEM products have been used to simulate superconducting transmission lines. Analytical work applied to a variety of lossless transmission line geometries (microstrip, stripline, etc.) has been compared to a commercially available software [65], and is in good agreement with SQUID based inductance measurements [54]. Other FEM implementations have been used to simulate MTLs with lossless superconductor properties by using plane sheets with user-defined surface impedance, but is only accurate for MTLs with w  ? [139], [140]. Piccinini et al. simulated lossy normal metal CPWs for CMOS Quantum Computing systems using HFSS similar to our method for superconducting MTLs but did not solve inside the normal metal due to its negligible skin depth at microwave frequencies [141, 142]. Additionally, the superconducting microwave losses and current distributions of coplanar waveguides with strip widths down to 50 nm have been simulated and compared with measured data [143, 144]. In the Quantum Computing community, Comsol 2D simulations are used to calculate the participation ratios of varying thin dielectrics contributing to two-level system (TLS) loss and are generally not concerned with simulating the negligible superconductor loss at mK temperatures [145, 146]. In addition to the need for numerical solutions when one is dealing with nonuniform current distributions, no analytical theory can account for the common fabrication effects impacting the MTL - sidewall trench tilt, rounded edges of wires, in-homogeneous superconductor or dielectric materials due to intermixing at material interfaces, cross-sectional thickness miss- targeting, etc. Furthermore, to maintain a desired metal density ? 50% for each layer in the damascene planarization process, small superconducting fill strips on the order of 0.5?3?m need to be included above, below and beside MTLs running perpendicular or parallel to conducting strip, which again is almost impossible to account for in the analytical calculations. The expected nonuniform current distribution cannot be solved analytically for our 0.25- 55 micron fabrication process (similar to [115]), where the dielectric separation thickness s, superconductor thickness d, conducting strip cross-section width w, and magnetic penetration depth are all comparable: s ? d ? w ? ? ? 0.25 ?m. Furthermore, to extract an accurate Rs, the exact cross-sectional geometry should be measured by Scanning Electron Microscopy (SEM) or Transmission Electron Microscopy (TEM). Therefore, in this chapter, to extract Rs from measurements, we will numerically solve for ? using HFSS coupled with TEM measurements. 3.2 Materials Characterization of Interconnect Morphology The damascene fabrication process is the desired planarization process for interconnects to scale a technology. As mentioned in Section 2.4, it is a challenge to fill dielectric trenches with refractory metals (like Nb) using magnetron sputtering [128]. This creates non-ideal morphology in the metal interconnect. In this section, the morphology of damascened Nb-TEOS fabrication processes will be examined with scanning transmission electron microscopy (STEM) coupled with energy-dispersive x-ray spectroscopy (EDS). The sample preparation and measurements will be briefly presented with the focus on the primary results that can affect RF losses in the Nb and TEOS. This analysis will be referred to in Chapters 4 and 5 to help explain physical mechanisms of RF loss measured by RF electrical characterization. Reviews on STEM can be found in the following references [147, 148, 149, 150, 151] D?Alfonso et al., later confirmed by Chu et al., were the first to do spatial atomic mapping using EDS with a STEM [152, 153]. A review on this technique can be found in reference [147, 154]. 56 3.2.1 Scanning Transmission Electron Microscopy (STEM) To accurately calculate the ? from HFSS, the cross sectional geometry of an MTL needs to be accurately known. With ?, the intrinsic resistance Rs can be extracted from measurements of Qc (see Eq. 3.13). Focused Ion Beam Microscopy (FIB) (see Appendix A), and Scanning Transmission Electron Microscopy (STEM) were used to measure the MTL down to sub-nm accuracy. The STEM imaging was done as a service by EAG Laboratories [155]. First a carbon protective coating was applied prior to Focused Ion Beam (FIB) milling of the structure. A cross- section was prepared using an FEI Helios 660 Dual Beam FIB/SEM and thinned to electron transparency. Reviews on specimen preparation for TEM using FIB can be found in the following references [156, 157] Samples were imaged using Hitachi 2700 Aberration Corrected Scanning Tunnelling Electron Microscope (ACSTEM) with an accelerating voltage of 200 kV and a nominal probe size of ? 0.1 ? 1nm. The focused probe is scanned across the sample surface, and transmitted electrons are imaged. Transmitted Electron (TE) STEM images form from electrons transmitted through the samples that are un-scattered or scattered to low angles and are similar to and commonly referred to as Bright Field (BF) images (Fig. 3.15a,c). Transmitted electron (TE) is the first and standard imaging method. Annular bright field (ABF) imaging is analagous to TE, but with enhanced contrast from weakly scattered electrons from light elements by using a ?preventing disk? (see Fig 1 in [158] and reference [149]). Atomic Contrast (or Z-contrast) STEM images are formed by electrons transmitted through the sample that have scattered out to higher angles (Fig. 3.15b). These are acquired using annular dark-field (ADF) detectors. If the electrons scatter to high enough angles the image intensity is roughly proportional to the square of 57 the atomic number (Z2), hence these are termed Z-contrast (ZC) images [159, 160]. In summary, TE images are most useful in this section to analyze the dielectric material (e.g. density, porosity) while ZC images are useful to analyze the morphology of the metal wires (see Figs. 3.1 and 3.3). Below, in Table 3.1 summarizes the measured cross sectional critical dimensions. The measurements were taken from the STEM images found from Fig. 3.1 and Fig. 3.2. Table 3.1: STEM Results Summary table for all three processes measured in this work. The ?mean width? column reports the minimum/max width, pertaining to the top and bottom width of the wire (see Fig. 3.1 and Fig. 3.2). All measurements were done at EAG Laboratories [155] In Fig. 3.1 and 3.2, the bottom of all images are closest to the wafer substrate. The top of all images are closest to dielectric-air interface. For process A and C, the ground plane is below the wire, or conducting strip (CS). For process B, the ground plane is above the CS. The CS (wire) in Process A and B have an inverted trapazoid shape from the damascene process described in Section 2.4. In process C, the CS has a trapazoid shape from the cloisonne? process (see Section 2.4). The bottom metal in Fig. 3.2(a) and top metal in Fig. 3.2(b) are 6x0.4 ?m fill metal strips with 0.3 ?m spacing, and will be discussed in Section 3.6. Aside from geometries, information about the morphologies can be analyzed from STEM images. Fig. 3.3 highlights the primary features that can have an impact on the Nb and TEOS 58 Figure 3.1: STEM images taken from representative samples of 0.25 ?m MTLs fabricated by Process a) A, b) B, and c) C. The top row of images (a-c) are ZC, where bright is high Z and dark is low Z. The bottom row are TE images. The bottom of all images are closest to the wafer substrate. The top of the images are closest to dielectric-air interface. Process A and C have the ground plane (GP) below the wire (or conducting strip CS). Process B has the GP above the CS. All measurements were done at EAG Laboratories [155]. RF losses. Namely, damascene processes A and B show distinct lower metal density lines going diagonally inward from from the bottom edges of the CS. This is primarily caused during trench filling process when two growth directions (trench bottom, side wall) meet. This also gives rise to the voids, or very low density metal, found along the intersection lines (see features 1 (yellow) and 2 (orange) in Fig. 3.3). For the cloisonne? process C, the Nb CS has a relatively continuous columnar grain structure, typical for sputtered Nb. The missing metal at the top and left edges of the 0.25-um CS is from the ?undercut? during RIE (see Fig. 3.3(b) feature 6). From the ZC images, a 8-10 nm NbOx layer can be seen at the top of the planarized trench in damascene processes A and B. For process B only, there are lower density regions in the dielectric and can be see in Fig. 3.3(d). Both of these oxide features will be discussed further 59 Figure 3.2: STEM images taken from representative samples of 1 ?m MTLs fabricated by Process a) A and b) B. Both images are TE images. The bottom of all images are closest to the wafer substrate. The top of the images are closest to dielectric-air interface. Process A and C have the ground plane (GP) below the wire (or conducting strip CS). Process B has the GP above the CS. The bottom metal in a) and top metal in b) are 6x0.4 ?m fill metal strips with 0.3 ?m spacing, and will be discussed in Section 3.6. All measurements were done at EAG Laboratories [155]. in Section 3.2.2. At 0.25-micron dimensions, the electronmagnetic fringing fields become a significant fraction in cross section [134]. Additionally, the current distribution will be nonuniform taking up a significant fraction of the cross sectional area. Furthermore, the location of defects relative to the current density concentrations in cross section can have a significant impact on the Nb and TEOS RF losses, and will be discussed in detail in Section 4.5 Hence, the motivation to model the electromagnetic fields around and inside the Nb layers, as well as numerically solving ? (see Section 3.4). 60 Figure 3.3: Highlighted Morphology features from STEM images for Processes A, B and C. Note, image contrast has been enhanced to highlight features. a) Feature 1 (orange) highlights the intersection between two grain orientations. Feature 2 (yellow) highlights voids or areas of low metal concentration. b) Feature 3 (purple) highlights a continuous Nb columnar grain. Feature 6 (pink) highlights the ?undercut? of the Nb during RIE. c) Feature 4 (green) highlights the NbOx layer found in Process A and B. d) Feature 5 (blue) highlights the lower density areas (bubbling and streaks) found in Process B only. All measurements were done at EAG Laboratories [155]. 3.2.2 Energy-Dispersive Spectroscopy (EDS) Energy Dispersive X-Ray Spectroscopy (EDS) will be used to survey the spatial concentration of elements with a detection limit of > 1at% (for elements Li-Pb). Although, the measurements here were standardless, and therefore accuracy down to 1at% cannot be claimed. Rather, the analysis will only qualitatively compare between samples and processes. This tool will help 61 understand the oxide compositions, and possible contamination of elements in the Nb layers. EDS mapping was done immediately after STEM, as it only requires a different detector inside the system. A 0.1-1 nm diameter electron beam is scanned across the sample surface as X-rays are detected. Fig. 3.4 shows a detectable amount of Ar content (? 1at%) in damascene process A and B. The larger amount of noise for Process A signal in Fig. 3.4(d) is due to an overall lower number of counts relative to Process B measurement. Additionally, the Process A sample is most likely more electron transparent (thinner). Hence the reason why Process A and B are plotted together normalized to their respective Nb signal. The sputtering pressure during trench filling is typically increased for more ?conformal? coverage of the trench [161]. The mechanism is an increased Ar pressures induces higher net columnated flux of metal ions depositing into the trench. As a result, there is also higher concentration of Ar plasma able to get trapped inside the trench/Nb. Ar content was not detected in EDS measurements of Process C. Fig. 3.5 shows the EDS maps spatial maps of Nb, Si, and O for damascene Process B. The top down line scan in Fig. 3.5(a) starts with the TEOS layer at position 0 nm and suggests a stoichiometry close to SiO. This insert of Fig. 3.5(a) zooms into the TEOS-Nb oxide interface. There is a plateau of the Nb signal from 23-27 nm and then a gradualy increase (decrease) of the Nb (O) signal from 27-33 nm. The overall NbOx 10 nm (greater than a typical niobium native oxide layer), and there is a stable oxide layer of 5 nm with a stoichiometry somewhere between NbO ? NbO2. Fig. 3.6(a) shows the O/Nb ratio at the niobium oxide interface for both damscene Process A and B. The abscence of a plateau suggests Process B does not have a stable oxide layer. Oxygen content decreases to zero over the same exact length scale as Process A of 10 nm. Fig. 3.6(b) 62 Figure 3.4: EDS measurement of Ar content in damascene process A and B. EDS map of Ar signal superimposed on the STEM image for damascene process a) A and b) B. The white dashed lines signify the EDS line scans plotted in c) and d). c) The semi-quantitative atomic percent of Ar in Process A, showing there is a detectable amount of Ar. d) The Ar signal normalized to the Nb to compare process A and B. The background signal is there for reference. All measurements were done at EAG Laboratories [155] is the Tc dependence of O/Nb ratio for bulk cast alloys of niobium and oxygen adopted from Hulm et al. [162]. It can be inferred that a bulk alloy of niobium and 60% oxygen would have a Tc ? 7 K. A thin film of Nb, especially at 5 nm, will have a significantly suppressed Tc, suggesting the 5 nm, 60% rich niobium oxide may be a normal conductor as opposed to a stable oxide like Nb2O5 at these temperatures. Fig. 3.7 shows an Al and SiNx passivation layer was introduced into Process C. This Al/SiNx prevented a NbOx layer from forming. 63 Figure 3.5: EDS data on damascene process B 0.25-micron MTL showing a) semi-quatitative plot of O, Si, and Nb atomic% in a top-down line scan. b) A STEM image with two superimposed EDS maps (Nb and O) showing spatial variation and NbOx layer. c) A STEM image with a Si EDS map superimposed, showing areas with low Si concentrations. Figure 3.6: Plots of a) EDS top-down line scan showing calculated respective O/Nb ratios vs position for damascene Process A and B. b) Replotted Tc vs O/Nb ratio from [Hulm1972] [162] 64 Figure 3.7: EDS data on Process C 0.25-micron MTL showing a) semi-quatitative plot of atomic% in a top-down line scan. b) EDS maps for each element. 3.3 Theory and Method The method to deconvolve superconductor and dielectric losses will be described in this section, and is termed the Dispersive Loss Deconvolution (DLD) method. The method described here will be demonstrated in Chapters 4 and 5. 65 3.3.1 RLGC Telegrapher?s equations The internal Q-factor of a superconducting transmission line resonator can be expressed as [95] 1 1 1 R G = + = + (3.1) Qi Qc Qd ?L ?C where Qc is the partial Q-factor associated with the conductor loss, Qd is the partial Q-factor associated with the dielectric loss,R, L,G, andC are the transmission line resistance, inductance, conductance, and capacitance per unit length, respectively, ? = 2?f is the angular resonant frequency, and the radiation loss is ignored as justified in Appendix B [163]. The telegrapher?s expressions for R, L, G, and C of a superconducting transmission line, assuming ?1  ?2, embedded in an interlayer dielectric are [95, 138, 164], 2 ? R = R 2 | |2 s ?|J| ds (3.2) I S ? ?0 L = (|H|2 + ?2|J|2)ds (3.3) |I|2 S ?? ?0 G = ?r tan ?|E|2ds (3.4)|V |2 S ? ?0 C = ?r|E|2ds (3.5)|V |2 S where I is the line current, V is the potential difference between the strip and ground plane, J is the vector (longitudinal) current density, H is the (transverse) magnetic field vector, ?0 is the permeability of free space, ?0 is the permittivity of free space, Rs, ?, tan ?, and ?r are the 66 spatially dependent superconductor intrinsic resistance, magnetic penetration depth, dielectric loss tangent, and relative dielectric constant, respectively, in the lateral plane and this treatment assumes everything is homogeneous in the longitudinal direction. The integrals are over the entire cross-sectional area S of the line. 3.3.2 Geometric Factors in Inhomogeneous Transmission Line Inserting Eqs. 3.2-3.5 into Eq. 3.1 yields 1 ? Rsm ? tan ?n = + (3.6) Qi m ?cm n ?dn where Rsm is the effective intrinsic resistance of the m-th conductor, ?cm is the partial geometric factor associated with losses in that conductor, tan ?n is the effective loss tangent of the n-th dielectric layer or tube, and ?dn is the partial geometric factor associated with losses in that dielectric layer or tube. These are defined as follows ? ? 2 R = ? S Rs?|J| dsmsm ? (3.7) ? ? 2S ?|J| dsm S ?r tan ?|E|2dstan ? nn = ? ? (3.8) ? ? ? |E|2S dsn r 2 2 2 S(|?H?| + ? |J| )ds?cm = ?0? (3.9) 2 ? ??|J|2S dsm? ? S ?r|E|2ds?dn = (3.10) S ?r|E|2dsn where Sm and Sn are the cross-sectional areas of the mth wire and nth dielectric layer or tube, respectively. 67 It shall be noted that ?cm defined by Eq. 3.9 differs from the cavity geometric factor? ? definition for conductor loss (Q = R /?cavity) where ?cavityc s = ?0? V |H|2 dv/ S |H 2? | ds defined as the ratio of the volume integral due to magnetic energy stored inside the cavity to the surface integral due to energy loss into the cavity walls [63, 131] associated with the tangential fields on the conductor. In Eq. 3.9, the first addend under the numerator integral includes resonant energy stored in the magnetic field penetrating the superconductor, while the second addend is due to resonant energy stored in the kinetic energy of the superconducting carriers [138]. The ?cavity does not account for them because in voluminous high-Q superconducting resonators [70, 165, 166] those energies can be ignored due to the penetration depth (typically, nm to um) being much smaller than the resonator characteristic size (mm to m). Also, unlike ?cavity that depends on the mode index, Eq. 3.7 effectively defines ?cm per unit length of transmission line because a transmission line resonator only supports a quasi-TEM mode, making it independent of the mode index. Likewise, ?dn defined in Eq. 3.10 is the cross-sectional fractional electric field energy stored in the n-th dielectric and is equivalent to the reciprocal of both (i) the electrical energy volume filling factor described by Krupka et al. [132] and (ii) the participation ratio definition used by the Quantum Computing community [145], [146]. Due to the assumption ?1  ?2 for isotropic superconductors, low transition temperature s- wave superconductors (e.g. Nb) exhibit intrinsic resistanceRs with quadratic frequency dependence [63], which per Eq. 3.6) and Eqs. 3.7-3.10 leads to a Q ? ??1c dependence [100]. At the same time, at GHz frequencies and above, amorphous dielectrics commonly exhibit a universal frequency dependence tan ? ? ?b?1, with b = 0.6?1 [97, 98]. Thus, insertingR = R (?/? )2s s0 0 and tan ? = tan ? (?/? )b?10 0 into Eq. 3.6 and assuming frequency-independent ?dn, allows one 68 to factor out the frequency as follows ( ) 1 ? ?M R ? b?1 Ns0m ? tan ?0n = + (3.11) Qi ?0 m=1 ?c0m ?0 n=1 ?dn where Rs0m and tan ?0n are the intrinsic resistance and the loss tangent at a reference frequency ?0, ?c0m is the partial geometric factor at that frequency associated with loss in themth conductor, and ?dn is the partial geometric factor associated with loss in the nth dielectric. 3.3.3 Geometric Factors in Homogeneous Transmission Line Consequently, Eq. 3.11 implies that the experimental dependence of Q?1i on the resonant frequency ?n = 2?fn of a multi-mode transmission line resonator, with n = 1, 2, 3. . . being the mode index, can be fit to Q?1i = A? +B? b?1 (3.12) ? where the first term corresponds to the superconductor loss A = ??1 ?10 mRs0m?c0m, while the? second term corresponds to the dielectric loss B = ?1?b ?10 n tan ?0n?dn . Generally, thin film fabrication methods may cause the ground plane and conducting strip to have dissimilar superconducting properties, especially for submicron width strips [72]. Similarly, the stack may involve multiple dielectrics. In a basic case of a homogeneous superconducting resonator formed by MTL embedded into non-dispersive dielectric (?dn = 1, ? = ?0, b = 1) 69 Eq. 3.11 reduces to 1 ? Rs0 = + tan ? (3.13) Qi ?0 ?c0 where the net conductor geometric factor ?c0 is ? ? 20GP?0CS S(|H|? + ?2|J|2)ds?c0 = = ?0?0 (3.14) ? + ? 2 ?|J|20GP 0CS S ds Here ?0GP and ?0CS are the partial geometric factors of the ground plane (GP) and conducting strip (CS), respectively. We choose ?0/2? = 10 GHz because it is commonly used for scaling and comparison of the surface resistance obtained by different superconducting resonators, and is also a convenient frequency to perform HFSS simulations. Per Eq. 3.13 the experimental dependence of Q?1i on the resonant frequency can be fit with Q?1i = A? +B (3.15) where the fit y-intercept immediately yields the dielectric loss tangent as tan ? = B [100], while the fit slope coefficient A corresponds to the intrinsic resistance as follows Rs0 = A?0?c0 (3.16) where ?c0 is found either from Eq. 3.14 if the fields and currents are known, or from the HFSS simulation, as described in Section 3.4. A linear fit of Eq. 3.15 enables easy analysis of large data sets and is advantageous for engineering. Moreover, the fit y-intercept yields tan ? even for 70 the loaded Q-factor data since the capacitive coupling loss becomes zero at zero frequency. On the other hand, a dispersive model per Eq. 3.12 accounts for fundamental physical properties of dielectric material. Initially, the data will be fit to the non-dispersive loss tangent model using 3.15 and Eq. 3.16 in Chapter 4. Then, the linear model will be tested against the dispersive model and results from reference [99]. To provide a simple analytical limiting case for Eq. 3.14 where fringing fields are not a factor, consider a resonator formed by a parallel-plate waveguide with identical superconducting plates of the width w and thickness d, separated by the dielectric layer of thickness s. This geometry approximates MTL withw  swhere the fringing fields can be ignored. The geometric factor is s+ 2? coth(d/?) ?PP (?) = ??0 (3.17) 2[coth(d/?) + (d/?)csch2(d/?)] utilizing the expression for the effective surface resistance of thin superconducting films, and by Reff ? Rs[coth(d/?)+(d/?)csch2(d/?)] that ignores transmission loss through the superconducting films [69]. In the thick film limit the geometric factor reduces to ? 1PP = ??0(s + 2?) and the2 geometric factor is no longer dependent on the superconductor plate thicknesses d. For the thin film limit where ? d the full expression needs to be used in Eq. D.11 and a term including the dielectric substrate impedance may be needed [70]. Eq. 3.14 implies the resonator optimized for characterization of both losses calls for a geometric factor ??c0 = Rs/ tan ?. For Rs ? 17?? representative of high quality Nb at 4.2 K, 10 GHz [138], and tan ? ? 10?3 [99], the optimal geometric factor is ??c0 ? 17 m?. At the same time, applying Eq. D.11 to our fabrication stack in Fig. 4.1(b) yields ?PP ? 12 m?, which is 71 close to ??c0. Incidentally, this affords us the opportunity to design an MTL resonator capable of deconvolving the superconducting and dielectric losses at GHz frequencies. 3.3.4 RLGC and Q-Factors From Impedance Matrix [Z] In the last section, it was established that the net conductor geometric factor ?c0 needs to be known to extract accurate values from measured data using Eq. 3.16. The net conductor geometric factor can be approximated analytically using Eq. D.11 if a parallel plate waveguide geometry can be assumed. The right hand side of Eq. 3.14 can be solved numerically using finite element solvers like HFSS [142]. This section will describe the method to extract RLGC parameters and Q-factors from the output impedance matrix [Zij] from HFSS, where i and j refer to the i-th row and j-th column in the impedance matrix. The geometric factors can then be found from these extracted parameters and will be shown in Section 3.5. The ABCD matrix for a generalized, linear transmission line is [95] ?? ??A B ? ?? ? ? ?? ? cosh(?l) Z0 sinh(?l)?? ?? 1 + (?l) 2/2 Z0?l = ? ?? (3.18) C D sinh(?l)/Z0 cosh(?l) ?l/Z 2 0 1 + (?l) /2 where ? is the propagation constant, l is the length of transmission line, andZ0 is the characteristic impedance. The approximation on the right holds for electrically short line ?l  1 using the quadratic Taylor series expansion for hyperbolic functions. The propogation constant and characteristic impedance of a general transmission line are [57] ? ? = ? + i? = ZY (3.19) ? Z0 = Z/Y (3.20) 72 where Z and Y are the series impedance and shunt admittance per unit length of the line. Inserting Eq. 3.19 and Eq. 3.20 into approximation Eq. 3.18 and assuming the two-port network is reciprocal yields ?? ??A B ? ? ? ? ?? 1 + ZY l 2/2 Zl = ?? (3.21) C D Y l 1 + ZY l2/2 [Z ? ? ? ?(Y l) 1 + (Zl)/2 (Y l)?1 ? ij] = ? ? (3.22) (Y l)?1 (Y l)?1 + (Zl)/2 Z = 2(Z11 ? Z12)/l (3.23) Y = (Z l)?112 (3.24) where [Zij] is the 2?2 impedance matrix. See Fig. 3.8 for the simplified diagram representations of the 2-port network parameters. Without assuming a reciprocal two-port network, the series impedance and shunt admittance per unit length of the transmission line can also be found from HFSS solved [Zij] using a very simple yet intuitive differential voltage method in the following way [57]. The propagation constant can be solved in terms of series impedance Z and shunt admittance Y per unit length of ? the line ? ? ZY by taking advantage of the current at each end of the ports being approximately equal for an electrically short line. The current at one end (port 1) is equal and the opposite polarity to the current at the other end (port 2) I = I1 = ?I2. In a two-port network the relationship between current and voltage using the impedance matrix [Z] is ?? ?? ?? ?? ??V1? = ?Z11 Z12???? I ?? (3.25) V2 Z21 Z22 ?I 73 Figure 3.8: Block diagrams of 2-port network parameters showing (a) ABCD block diagram (Eq. 3.21), (b) ABCD can be transformed into the impedance matrix [Zij] (Eq. 3.22), and the 2-port network can be represented as a single series impedance Z (Eq. 3.23) or shunt admittance Y (Eq. 3.24) [95]. The differential voltage between the two ports is equal to [57] ?V = V1 ? V2 = ZI (3.26) From Eq. 3.25 and Eq. 3.26, Z can be written as Z = R + i?L ? (Z11 ? Z12 ? Z21 + Z22)/l (3.27) where Zij are the elements of the impedance matrix, and R and L are the series resistance and series inductance per unit length, respectively. Equivalently, by utilizing I = I1 = ?I2, the 74 differential current being equal to the admittance and voltage product [57] ?I = I1 ? I2 = Y V (3.28) and the relationship between current and voltage using the admittance matrix [Y ] ?? ? ? ?? ?? I1?? ??Y11 Y12= ???? V ?? (3.29) I2 Y21 Y22 ?V the admittance can be expressed as Y = G+ i?C ? (Y11 + Y12 + Y21 + Y22)/l (3.30) where Yij are the elements of the admittance matrix, G and C are the shunt conductance and shunt capacitance of the network per unit length, respectively. Now with expressions for both Z and Y ,RLGC can be separated and analyzed separately in two-port HFSS simulations. Eqs. 3.23 and 3.24 assume a reciprocal 2-port network, and Eqs. 3.27 and 3.30 do not. For Q ? 1i ???1 [95], inserting Eq. 3.23 and Eq. 3.24 into Eq. 3.19 and Eq. 3.20, and2 ? ? utilizing the identity Im[ z]/Re[ z] = tan [(arg z)/2] for a reciprocal 2-port network, one can simplify Qi in terms of the impedance matrix Eq. 3.22 by [ ] 1 1 Qi = tan arg(2Z ?1 2 2 21 (Z11 ? Z21)) (3.31) Eqs. 3.27, 3.30, and 3.31 will be used in Section 3.5 to extract numerically solved geometric factors of microstrip transmission line resonators measured in Chapters 4 and 5. The above 75 method can be applied to any electrically short, 2-port transmission line geometry. 3.4 HFSS Modeling In this section I describe how to model superconductors embedded in dielectrics by using impedance boundary conditions or by a Solve-Inside method. These distinctly different model types to be simulated are shown in Fig. 3.9, and are evaluated here to show (i) the limitations of Fig. 3.9(a) due to assuming boundary conditions, and (ii) the capability of the Solve-Inside method Fig. 3.9(b) to extract geometric factors analytically for any arbitrary geometry. Here we start by modeling a parallel plate waveguide with superconducting impedance boundary conditions. This is how superconductors have been defined in finite element modelers in the past [139, 140]. A microstrip transmission line (MTL) will be modeled using the solve-inside approach described here. From HFSS solved telegrapher?s expressions (RLGC) described above, the superconducting and dielectric net geometric factors ?c0 and ?d0, respectively, are calculated. Analytically solving these factors can be difficult as dimensions become comparable to ? and geometries increase in complexity. Here we utilize this method to extract the intrinsic resistance Rs, and determine RLGC and ? sensitivities to fabrication variation. Modeling dielectric loss tangent tan ? comes standard in HFSS and is straight forward. The critical differences in modeling a superconductor in HFSS as opposed to a normal conductor are (i) the definition of the superconducting material with a real conductivity ?1 and a negative real permittivity ?r < 1 (? = ?1? i?2 or Zs = Rs+ i??0? cannot be defined directly for 3D material [142]), (ii) the use of Perfect Electrical Conductor material to connect superconductor materials to a wave port due to wave ports not being able to handle materials with ?r < 1 (see Fig. 3.12), 76 and (iii) selecting the solve-inside option for the defined 3D lossy superconducting material due to a real conductivity much smaller than normal conductors. The first critical difference will be discussed in detail, followed by the process of converting the HFSS impedance matrix into useful parameters, and results validation. This method can be used for any linear 2-port electrically short transmission line geometry, as long as the ratio of largest to smallest dimensions defined in the simulation are less than a factor of 104. Other possible geometries that can be simulated include coplanar waveguides (CPWs) and striplines. 3.4.1 Parallel Plate Waveguide A method to simulate a superconductor in HFSS is by defining the superconductor as impedance boundary conditions [140, 141, 167, 168]. To demonstrate this, a parallel plate waveguide (PPW) will be modeled and validated against analytical equations. A range of superconductor thicknesses d and penetration depths ? are simulated to show HFSS yields proper solutions for the bulk, thin film, and intermediate limits. As described in Section 2.2, the complex surface impedance of a parallel plate waveguide (or resonator) can be modeled using Eq. 2.24 and Eq. 2.25 [70]. The expressions in brackets in Eq. 2.24 and the term coth (d/?) in Eq. 2.25 are modifications to the intrinsic impedance to account for superconducting plate thicknesses on the order or less than the penetration depths d ? ?. For a system composed of bulk superconductor with thicknesses d  ?, the intrinsic impedance of each conductor simplifies to Zs0 = R ( ? )2s0 + i?0?0?, where R? s0 is the intrinsic0 resistance at reference frequency ?0. Here, a PPW system has been enforced, and we are 77 Figure 3.9: A diagram showing the different HFSS model geometries simulated in this section: a) Parallel Plate Waveguide (PPW) with superconducting complex impedance boundary conditions on the top and bottom faces labeled ?ImpBound SC? and PerfH boundary conditions on the side faces. b) Microstrip Transmission Line (MTL) with Solve-Inside ground plane (GP) and conducting strip (CS). The conducting strip has width w. These models have superconductor thickness d and interlayer dielectric thickness s. The nominal HFSS simulation frequency is set to 10 GHz and simulated lengths ranging from 10-100 ?m. For all Ls, Cs, and l simulated here meet electrically short condition ?l  1. interested in the resulting transverse electromagnetic (TEM) mode propagating along the longitudinal length l in the x direction (see Fig. 3.10). The accuracy of HFSS solutions will be tested using the following equations for RLGC by inserting the superconducting impedance definitions Eq. 2.24 and Eq. 2.25 into PPW analytical solutions [95] 2R R = eff (3.32) w L = ? s+2? coth (d/?)0 (3.33)w 78 G = ??0?r tan ? w (3.34) s C = ? w0?r (3.35)s where ?0 is the vacuum permeability, ?0 is the vacuum permittivity, ? is the angular frequency, w is the PPW width, s is the plate separation dielectric thickness, Reff is the effective resistance defined in Eq. 2.24, and tan ? is the dielectric loss tangent. The RLGC are the series resistance, series inductance, shunt conductance, and shunt capacitance per unit length of a generalized transmission line defined in Eq. 3.2-Eq. 3.5 in Section 3.3. Using Eq. 3.1, the partial conductor and dielectric Q-factors for a PPW, Qc = ?L/R and Qd = ?C/G, respectively, are Qc = ?? s+2? coth (d/?) 0 (3.36)2Reff Qd = 1/ tan ? (3.37) Notice, the PPW width w drops out in Qc, and ? drops out in Qd with no geometry dependence. Then, the only parameters needed to model the superconducting PPW, analytically and with HFSS impedance boundaries, are s, d, ?, ?, Rs, and tan ?. The HFSS Driven Modal network analysis design will be used [142] to model the PPW. The top and bottom faces of the rectangular PPW are defined as impedance boundary conditions (see Fig. 3.10). To ensure no fringing fields, the two sides are defined with Perfect H boundary conditions. When defining the HFSS wave ports, a new line of integration needs to be defined to explicitly set the expected electric field E direction for the propagating mode 1. For the PPW case, the direction is defined to be normal to the impedance boundary plates. In this section, we 79 are concerned with the first mode of propagation only, where the electric field is exactly normal to the conducting plates (impedance planes). Take note, this is enforced with the integration line, allowing HFSS to converge to the correct solution IF and ONLY IF mode 1 is the only mode present in the model defined. If other modes are present for a given geometry and simulation frequency ?/2?, then phenomena such as mode conversion, the presence of non-propogating evanescent modes, and parasitic reflections of higher-order modes can cause HFSS to converge to non-realistic (bad) results [169]. This is true when modeling any material as complex impedance boundary conditions. To facilitate proper convergence of transmission line parameters (e.g. RLGC), the default settings are the following: (a) maximum number of passes is set to 40 (b) maximum delta S is set to 0.001 (c) output variables RLGC are set to a 1% convergence criteria (d) lambda refinement and initial mesh operations are not used (e) maximum refinement per pass of mesh is set to 50% (f) solution options are set to mixed order with relative residual of 1? 10?4 (g) maximum delta Z0 set to 2% in the port options. To aid HFSS in converging to accurate solutions quickly, the simulation geometry used in Fig. 3.10 and Fig. 3.11 will be set to l = 1 ?m, w = 0.5 ?m, and nominal s = 150 nm. For impedance definitions, the nominal ? = 90 nm, R = 20 ??, tan ? = 1 ? 10?3s , and simulation frequency ?/2? = 20GHz. 80 Figure 3.10: Model of parallel plate waveguide (PPW) in HFSS using a Driven Model solution type. The model a) isometric view, b) cross section view along x-axis, and c) cross section view along y-axis. The top and bottom faces (z-axis) of the simulation box are the superconducting plates defined as impedance boundary conditions using Eq. 2.24 and Eq. 2.25. The side faces (y-axis) are Perfect H boundaries. The PPW length, width, and plate separation thickness are l ? 1 ?m, w ? 0.5 ?m, and s = 0.15 ?m, respectively. The nominal simulation frequency is ?/2? = 20GHz. For a wide range of d and ?, the RLGC and partial Q-factor errors from HFSS solutions relative to Eq. D.3-Eq. D.6 are plotted in Fig. 3.11. The d = 1 ?m (pink) at minimum ? = 1 nm represents the bulk case (d ?), whereas the d = 10nm (red) at maximum ? = 1?m represents the thin film limit (d  ?). Everything in between is considered the intermediate regime. A positive (negative) error corresponds to HFSS converging to a larger (smaller) results relative to Eq. D.3-Eq. D.6. It can be seen the errors in all parameters do not exceed 0.35% over the bulk limit, thin film limit, and intermediate regime. This validates HFSS is converging to the expected results. Modeling superconductors as impedance boundary conditions has been used in the community and has been further validated in HFSS for a parallel plate waveguide (PPW) here. Although, this 81 Figure 3.11: HFSS results of a PPW using geometry in Fig. 3.10. The (a) series inductance L, (b) series resistance R , (c) partial conductor Q-factor Qc0 , (d) shunt capacitance C , (e) shunt conductance G, and (f) dielectric partial Q-factor Qd error relative to parallel plate analytical Eq. D.3-Eq. D.6 as a function of penetration depth ? and effective plate thickness d. In HFSS the penetration depth ? (with name $lambda on the x-axis) and plate thickness d are parameterized variables. The error was calculated using the equation Xerror = 100 ? (XHFSS ?XPP )/XPP , where XHFSS is solved value by HFSS simulation and XPP is the calculated value from parallel plate model analytical Eq. D.3-Eq. D.6. The simulation frequency was ?/2? = 20GHz. is only a good approximation under the assumptions of (i) fringing fields can be ignored and (ii) current distributions inside the superconductor are known. If this cannot be assumed and a two- fluid model can still be used [40, 43], then one must resort to numerically solving inside a 3D superconductor. 3.4.2 Defining Solve-Inside Superconductor A good, isotropic superconductor (?1  ?2) can be defined as a material with bulk conductivity ?1 and negative relative permittivity [170]. Inserting the superconductor current- field constitutive relationship J = (?1 ? i?2)E into Maxwell?s curlH equation in phasor form 82 yields ?? ? ?H E 2= ?1 + i??0(?r )E (3.38) ?0? where ?sc = ?r ? ?2/?0? can be seen as the superconductor real relative permittivity, with ?r being the ordinary dielectric constant associated with displacement current. The ?r value can be estimated by approximating a superconductor as a collisionless neutral plasma with Langmuir 2 frequency ?p = c/? and dielectric function ? ? ?p = 1 p 2 = 1 ? (??/c)?2, where c is the speed? of light [171]. Comparison of ?sc and ?p implies that ?r = 1 is the superconductor permittivity in the limit ?  ?p (approached from below), and ?2 = (?? 2 ?10? ) is the London conductivity applicable at frequencies well below the gap and temperatures not too close to Tc. Therefore, a superconductor can be defined in HFSS as a material with real relative permittivity ?HFSSsc = 1??2(T, ?)??1??10 and real conductivity ?HFSS1 = ?1(T, ?), which is applicable over the entire range of temperatures and frequencies considered in this work. At frequencies below the gap, and temperatures below Tc, a superconductor relative permittivity is substantially negative [172]. According to Eq. 3.38, the superconductor complex permittivity is  = ? (1 ? ?2 ? i?1sc 0 ),?0? ?0? where the three addends are associated with displacement, superconducting, and normal current, respectively. Due to the fact that ?/?0?  1?the displacement current is negligibly small, and the intrinsic impedance general?ly defined as??0/sc reduces to a value much smaller than the vacuum wave impedance Zs = i??0/?  ?0/0 = 120? ?. Generally, ?1 and ?2 can be calculated from the Mattis-Bardeen theory [49] and the respective ?HFSS HFSS1 (?) and ? (?) are tabulated and implemented in HFSS as datasets [49]. To meet the assumption ?1  ?2 made in Sections 2.2.2 and 3.3, I will model ?1 as a Drude-like 83 conductivity, which is related to the experimental intrinsic resistance found in Eq. 3.16 via ?HFSS = 2R ??2??2??31 s0 0 0 [138]. Likewise, we will model ?2 as the London conductivity which gives ?HFSS = ?(??/c)?2. Such an approach allows one to implement ?HFSS1 and ?HFSS analytically. Forcing HFSS to solve inside the superconductor overcomes the limitations of an impedance boundary approximation [140], [167]. 3.4.3 Solve-Inside Model Setup When simulating electromagnetic fields inside highly conductive materials with ?/?0?  1, a proper modeler setup is vital for accurate results. The HFSS model and setup were optimized for accuracy (meeting condition ?lHFSS  1) and simulation time, in order to complete parametric sweeps on the order of 100 simulations in a few hours for sensitivity analysis and MTL design optimization. The network analysis driven terminal solution type was used [142] to simulate 2- port MTLs with nominal lengths lHFSS = 10?50?m, cross-section geometry shown in Fig. 3.12, simulation frequencies fHFSS = 10 ? 100 GHz, and perfect electrical conductor port length lPEC = 1 ?m (see Fig. 3.12) [36]. The MTL is surrounded by a radiation bounding box with walls sufficiently spaced to avoid impinging magnetic and electric fields shown in Fig. 3.13. Ports were de-embedded by a length lPEC to calculate only the network parameters pertaining to the superconductor MTL. After defining a superconducting material with negative real permittivity ?HFSS and real conductivity ?HFSS1 described in last section, the solve inside option is selected, to override HFSS default to solving inside when real conductivity ?1 < 105 S/m. All geometries and material properties were parametrized and the Optimetrics option was used for sweeps and sensitivity analysis. To allow for fast and accurate convergence, the analysis setup was set to 84 Maximum Delta S of 0.001 and 0.1-1 % convergence criteria for RLGC output variables. 3.4.4 Microstrip Transmission Line (MTL) The (embedded) microstrip transmission line (MTL) geometry is the one measured in Chapter 4 and Chapter 5. The MTL geometry simulated here consists of a ground plane (GP) and a conducting strip (CS) embedded in a dielectric having thickness d, and separated by a thickness s (see Fig. 3.12). The GP and CS are defined as solve-inside superconductors as explained above. To stay within the assumption ?l  1, the simulated MTL simulation lengths and frequencies are l < 100 ?m and ?/2? < 100 GHz, respectively. Unless otherwise specified, the nominal conductor and separation thicknesses are d = 200 nm and s = 150 nm, respectively. The GP width is nominally 20 ?m when simulating CS w ? 4?m, to be sufficiently far from the electromagnetic fringing fields. Similarly, the simulation box height separation from the GP and CS is roughly 5-10s to be sufficiently far from electromagnetic fields. The perfect electrical conductor (PEC) strips (yellow) connecting the GP and CS to the waveports are lpec = 1 ?m. The wave ports are surrounding the CS cross section and only touch the corner of the GP to properly act as references. Extending the wave port into the cross section of the GP will induce an error and HFSS will not be able to solve. Additionally, the GP cannot touch the edges of the simulation, as this will effectively cause a short. Lastly, all faces of the simulation box are defined as Radiation Boundary conditions. Two-port driven terminal simulations for MTL widths w = 0.25 ? 4 ?m were done and magnetic field profiles in cross section can be seen in Fig. 3.13. Up to this point, the lossy telegraph?s R and G for an MTL at these dimensions and ? have not yet been reported. For 85 the 4 ?m MTL, the magnitude of the magnetic fields (H) at the edge (fringing fields) are only a small fraction relative to the H fields underneath the CS. Therefore, a parallel plate model is a good approximation for this width (assuming w >> s). For the 0.25 ?m MTL, the fringing fields become a significant fraction. It can also be seen from Fig. 3.13 for there is a transition in fringing field character between 0.5 and 1 ?m. At 0.5 ?m, the fringing fields on each side begin to overlap. At 0.25 ?m, a significant fraction of the CS is penetrated by magnetic field, but the current is not yet uniform, making it difficult to analytically model. When considering the GP, the fraction of fields inside reduce significantly relative to the CS. This will be discussed further in the next section. Chang et al. analytically solved telegraph?s L and C acounting for a fringing factor for microstrip widths w ? 1 ?m [134]. 86 Figure 3.12: MTL geometry modeled in HFSS. The MTL conducting strip (blue) has width w and a dielectric separation thickness s from the ground plane (gray). Both the strip and ground plane have thickness d and simulation length l. In HFSS, the MTL is surrounded by a simulation box with all six faces defined with Radiation boundaries. To properly launch waves into a superconductor with negative real permittivity, small lengths of Perfect Electrical Conductors lpec ? 1 ?m (yellow) are used to connect the superconducting ground plane and conducting strip to the wave ports. To cut down on simulation time and reduce the simulation size in half, one can utilize a Symmetry boundary condition. 87 Figure 3.13: Cross section MTLs simulated using HFSS and geometry shown in Fig. 3.12. The dielectric thickness is s = 150 nm. The ground plane and conducting strip thicknesses are both d = 200 nm. The MTL width varied from 0.25? 4 ?m going from top to bottom. The inputted intrinsic resistance Rs = 20 ??, TEOS loss tangent tan ? = 1? 10?3, and the SiNx loss tangent tan ? = 1? 10?4. The magnetic penetration depth was set to ? = 90 nm. The relative dielectric constant of TEOS and SiNx relative dielectric constant were set to ?TEOS = 4.5 and ?SiNxr r = 7.5, respectively. The simulation frequency was set to ?/2? = 10GHz. 88 3.5 MTL Geometric Factor ? Extraction For an electrically short transmission line ?l  1, the propagation constant of a general ? transmission line ? = ? + i? = ZY [57], and the internal Q-factor of a transmission line resonator is Q = 1???1i [95] This can be expressed in terms of the series impedance Z and2 ? ? shunt admittance Y per unit length yielding Q 1i = Im( ZY )/Re( ZY ). Inserting here Z ?2 2l? ? ? 1(Z11?Z21) and Y ? (Z ?112l) found in Section 3.3.4, and using an identity Im( z)/Re( z) = tan[Arg(z)/2], yields [ ( )] 1 1 2(Z11 ? Z21) Qi = tan Arg (3.39) 2 2 Z21 where Arg(z) is the argument of a complex variable z, and Zij are the respective elements of the model Z matrix solved by HFSS. For a model of arbitrary length, one shall resort to a general solution for propagation factor e??l of a transmission line with mismatched ports [36]. Setting the dielectric loss to zero, tan ? = 0 throughout the entire HFSS model leads to Qi = Qc. Then, the conductor geometric factor ?c0 can be computed as ? = RHFSSQHFSSc0 s c (3.40) where RHFSSs0 is the superconductor intrinsic resistance defined in the HFSS model, and Q HFSS c is the quality factor found from the HFSS solution using Eq. 3.39, and the simulations are done at a reference frequency ?0/2?. Here ?c0 is independent of intrinsic resistance and the product QHFSSRHFSSc s of each cross-sectional geometry can be found from a single simulation 89 at any intrinsic resistance value meeting RHFSSs  ?0?0?. This simplifies the experimental data analysis by removing the need to simulate Qc as a function of Rs for every MTL width. Similarly, by setting the superconductor intrinsic resistance to zero throughout the entire HFSS model leads to Qi = Qd. Then the dielectric geometric factor ?d can be computed as ? = tan ?HFSSQHFSSd d (3.41) where tan ?HFSS is the dielectric loss tangent defined in the HFSS model, and QHFSSd is the quality factor associated with the dielectric loss. The HFSS results can be validated by comparing the HFSS solved geometric factors ?c0 using Eq. 3.40 to the analytical parallel plate model using Eq. D.11 for MTL widths much larger than the dielectric thickness w/s  1. For wHFSS = 4 ?m the partial Q-factor associated with the conductor loss solved by HFSS QHFSSc0 (4 ?m) = 586 using Eq. 3.13 and an input value of RHFSSs0 = 20 ??. Using the parallel plate model, the partial Q-factor is found to be Q PP c0 = 584, yielding agreement within 0.34%. This can also be seen directly in Fig. 3.13, where ?totc0 (green solid line) approaches ?PPc0 (dashed line) forwHFSS ? 1?m. Intuitively this can be understood by looking at Fig. 3.14. The fringing fields at MTL left/right edges of cross-section begin to overlap for wHFSS ? 1 ?m, making the parallel plate model inapplicable as it does not account for field fringing, concentration at the edges, and penetration above the conducting strip. Furthermore, it is important to note the geometric factors calculated from HFSS solutions were independent of the input value RHFSSs = 0.1? 100 ??. 90 Figure 3.14: Geometric factors for the conducting strip ?CS , ground plane ?GP , and ?totc0 c0 c0 . This is assumingRGPs0 = R CS s0 from Eq. 3.14 for all MTL widths. The parallel plate analytical calculation ?PPc0 is compared (see Eq. D.11). 91 3.6 MTL RLGC and ? Sensitivity to Fill Spatially dense areas of metal polishes (etches) faster than lower density areas in the damascene CMP process. Floating fill patterns (metal strips) are used to maintain a certain metal density, achieving uniformity across the chip and wafer, mitigating dishing effects [54]. Fig. 3.15 shows the MTL strip below the ground plane and surrounded by floating fill patterns length (left to right) and width (into the page) ? 1.5 ?m and ? 0.5 ?m, respectively, to reach minimum density requirements of ? 65%. Figure 3.15: STEM images of (a) full MTL showing the via walls and surrounding fill, (b) measurements of fill proximity to conducting strip (CS), and (c) is a zoom in of the conducting strip showing the resulting inverted trapezoid geometry. Using the measurements in Fig. 3.15, three HFSS simulations were run and compared (see Fig. 3.16 and Table 3.2). The top row in Fig. 3.16 is the nominal cross sectional geometry. The middle row is the actual cross-sectional geometry of the MTL without surrounding fill. Note, the 92 sidewall angle was introduced in the HFSS simulation to model the damascene Process A. The bottom row is the same as the middle row but with the additional Nb fill to reflect the geometry shown in Fig. 3.15(b). 93 Figure 3.16: Cross sections of three different HFSS simulations, where the (top) is the nominal designed geometry of the 250 nm wide conducting strip (CS) underneath the ground plane (GP). The (middle) row is the actual cross section geometry as measured by TEM data for wp432- TL6 with CS shaped more like a Reeses Cup. The (Bottom) row is the same MTL geometry as (middle) but with added and repeated superconducting strip fill patterns. Table 3.2 summarizes the results from simulations shown in Fig. 3.16. The RLCG values were calculated as described in Section 3.3.4. The partial geometric factors associated to the ground plane (GP), conducting strip (CS), TEOS dielectric layer, and SiNx dielectric passivation layer are calculated based off of geometry shown in Fig. 3.16. Immediately, it is important to point out that Fig. 3.15 shows the CS is in the layer below the GP (also shown in Fig. 3.1 and summarized in Table 3.1). Therefore the electric fields are masked from the SiNx. This is confirmed in Fig. 3.2, showing ?SiNx is a factor of 10 larger than ?TEOS , and is not participating. This is also true for Process A and B where the GP is below the CS. Referencing Fig. 3.13, only a small fraction of the fields of significant magnitude reach the SiNx layer (magnetic field is shown, but electric field profile is almost identical outside of the CS 94 and GP). Therefore, for MTL widths and process geometries, ?SiNx  ?TEOS, and therefore is ignored in the remainder of this work. Table 3.2 shows the effective geometric factor ?c0 decreased 5% for actual geometry and further decreased by 2% when floating fill was introduced. This change is primarily due to the large change in geometry of the GP (?GP ) having thickness half of nominal and the dielectric thickness ? 8% thinner whereas the CS geometric factor ?CS was almost unchanged. Due to? 8% change in dielectric separation thickness bothG andC changed by? 8?10%. When fill was introduced, bothG andC increased by? 36?38%, due to the removal of dielectric replaced by metal fill. BothG andC changed the same way and therefore the dielectric geometric factor ?d0 was almost unchanged. Although, the notable increase in C by about 36% and almost no change inLwill considerable decrease the data propagation velocity by about 15% where vp ? (LC)?1/2. This result is relevant for synchronous clock timing designs where the interconnect inductance and capacitance across long distances need to be meet required values, and fill patterns around the interconnect can have a notable affect. 95 Table 3.2: Table of RLCG for three different HFSS simulations shown in Fig. 3.16 using geometry measured from Fig. 3.15 fabricated using Process A (see Table 4.1). The material properties used for all three simulations areRs0 = 20??, ? = 90nm, ?TEOS = 4.5, tan ?TEOS = 1? 10?3, ?SiNx = 7.5, and tan ?SiNx = 1? 10?4. Overall, measuring RLGC sensitivities to fill showed that geometric factors are fairly insensitive and can be approximately ignored. However, propagation velocity is sensitive to fill placement and geometry and should be taken into consideration. The method described here can be used to check the change in propogation velocity for a given fill geometry. Most importantly, the largest changes in RLGC and ? were seen in the difference between nominal (expected) cross sectional geometry and as fabricated (measured by STEM). In the next section, we will do a sensitivity analysis on MTL geometries. 96 3.7 MTL ? Sensitivity to RQL Fabrication This section will discuss howRLGC parameters (see Eq. 3.2-Eq. 3.5) and resonator geometric factors (see Eq. 3.9-Eq. 3.10) in MTLs vary at parametric corners of the fabrication process (see Fig. 4.1). A simple sensitivity analysis will be presented to get a sense of (i) the precision of extracted Rs0 from measurements in Chapter 4 (ii) the application of this method to determine sensitivity to fabrication variation. This procedure can be used for any arbitrary two port transmission line cross-sectional geometry (e.g. coplanar wave guide, strip line, etc). Table 3.3(a) shows the nominal parameters for the fabrication of samples from Process A, B, and C (see STEM summary table Fig. 3.1). Table 3.3(b) are the assumed minimum to maximum variations of the fabrication/material parameters. These parameters were inserted into HFSS as parametric variables to calculate variations in ? and RLGC using methods described in Section 3.3.4 and Section 3.5. Table 3.3: Tables showing a) the nominal material properties and simulation lengths, and b) the minimum, nominal, and maximum values chosen as the parametric corners based off fabrication variation worst case scenario. Fig. 3.17 shows the percent variation of the geometric factor ?c0 is roughly ?20% for 97 all CS widths shown. The 0.25 ?m has an additional ?2% variation. These are conservative estimates for Process A, B, and C based off worst case geometry variations shown in Table 3.3(b). Aside from the ground plane thickness offset in Process A, the summary Table 3.1 from STEM measurement show the variations in thicknesses do not exceed 10%. Although, these are critical dimension (CD) measurements on single chips and are not representative of the whole wafer or lot of wafers. Thickness uniformity and line width targeting at minimum feature sizes are one of the many challenges in fabrication. Therefore, a worst case 20% variation in ?c0 will be incorporated as error in our extraction of Rs from measurements in Chapter 4. Figure 3.17: Parametric Corners to show worst case max/min error in ?c0 for MTL widths 0.25? 4 ?m. This is assuming nominal geometry from Table 3.3(a) and worst case cross section geometric variation in Table 3.3(b), based off fabrication tolerances. The variation in ?, or Qc assuming a constant Rs, can be further scrutinized by doing a sensitivity analysis on each parameter contributing to changes in RLGC. Fig. 3.18 shows RLGC % change as a function of the parameter % change for a 0.25 ?m MTL starting at 98 nominal geometry Table 3.3(a). Note,Rs and tan ? were held constant in Fig. 3.18 showsRLGC, and therefore any variations in Qc or Qd are variations in the geometric factors exclusively. Immediately, it can be seen from Fig. 3.18(f) that ?d is insensitive to all parameter variations. This is expected for an MTL embedded in a homogeneous dielectric, as the material property tan ? is geometry independent. If phase velocity sensitivity to fabrication variation is desired (v ? (LC)?1/2p ), then the L and C sensitivities can be simply multiplied from Fig. 3.18(a)-(b). Here, the discussion will be focused on ?c sensitivity to fabrication variation. The conducting strip (CS) magnetic penetration depth ?CS varies L and R by as much as ?10%, but change in the same way and therefore ?c (Qc with constant Rs) is insensitive to ?CS . Interestingly, ?c can change by 3% for -30% change in ?GP . This is most likely due to magnetic field becoming more concentrated in and around the CS (see Fig. 3.13). Evidently, this suggests ?c is relatively insensitive to changes in ?. The most surprising result is ?c is most sensitive to thickness variations in the conducting strip dCS . The Lmaximum change is?6%, whereas theR can change by as much as?12%. The lower L sensitivity to dCS is most likely due to the internal inductance term only being one of two contributions to the total inductance (see 3.3). Whereas R is completely defined by ? (current distribution inside wire), and is greatly affected when the CS (or film) thickness is on the order of the ? ? dCS , i.e. ?eff = ? coth (d/?). This argument can also explain ?c secondary sensitivity to MTL widths wbottom and wtop. Although, ?c has a slightly lower sensitivity to wbottom relative towtop. The dielectric separation thickness s can vary ?c by?5%, as this can significantly change the current distribution and fringing fields. Using HFSS simulations and the methods described here, it has been determined ?c can vary as much as ?20% for MTL widths 0.25 ? 4 ?m. A sensitivity analysis on a 0.25 ?m 99 Figure 3.18: RLGC Sensitivity Analysis on MTL minimum width 0.25 ?m as well as the % change in Qc and Qd (see Eq. 3.1 and Eqs. 3.2-3.5). The Nominal geometry shown in Fig. 3.12, where wbottom and wtop are the dimension closest and furthest from the ground plane (GP), respectively. The absolute parameter ranges can be found in Table 3.3. with trapezoidal geometry suggests ?c is most sensitive to the conducting strip thickness dCS , s, wbottom, and wtop. This analysis affords us the ability to bound our error in extracted Rs from measurement in Chapter 4 and Chapter 5. 100 3.8 Summary In this chapter, I discussed the importance of numerically solving inside a lossy superconductor, and reviewed the work that has been done thus far. I presented a method to extract the intrinsic superconductor and dielectric RF losses coupled with HFSS modeling. I discussed the morphologies (non-idealities) in Nb-TEOS 0.25 ?m and 1 ?m MTLs from Process A, B, and C were study using STEM and EDS. For the damascene Process A and B, line defects and voiding occur at the edge of the fill trenches, quintessential to the sputtered fill step in damascene processes. These defects occur in a region of highest magnetic field (current density) concentration for Process A. For Process A and B, up to 1at% Ar was detected inside of the conducting strips. No Ar was detected in MTLs from Process C. Process A shows a total Nb oxide layer approximately 10 nm with 5 nm layer of NbO ? NbO2. Process B shows the same total Nb oxide thickness of approximately 10 nm, but no stable oxide was detected. Process C shows no detectable Nb oxide layer due to the incorporated SiNx/Al passivation layer. I presented a method to deconvolve the partial Q factorsQc ? ?/Rs from the superconductor and Qd = 1/ tan ? from the dielectric by relying on their respective dispersion relationships. To extract the intrinsic resistance Rs from Qc, I presented a procedure using HFSS to calculate the geometric factors ? numerically. The HFSS procedure was validated by comparing to analytical models and other work. The geometric factor is insensitive to the surrounding fill patterns used in Process A, B, and C. The geometric factor is notably sensitive to primarily the conducting strip thickness, interlayer dielectric thickness, and the conducting strip width, but is not sensitive to the magnetic penetration depth. The conservative error in ?c will be bounded to ?20%. 101 Chapter 4: Characterization of Microwave Loss in RQL Interconnects at 4.2 K In the previous chapter, non-idealities in the morphology were measured in TEM/EDS, and could be causing added extrinsic resistive losses in Nb interconnects. The geometric factor methodology and solving numerically in HFSS was established as a means to accurately extract the intrinsic resistance Rs from measurements of the partial conductor Q-factor Qc. This enables the ability to compare Rs across processes A, B, and C for all MTL widths. In this chapter, a brief history on Q-factor measurements in different systems and assumptions will be presented in the introduction. The MTL resonator design will then be described, followed by the measurement and fitting method to measure the internal Q-factor Qi. The Qi can then be deconvolved in the dielectric Qd and conducting Qc partial Q-factors by relying on their respective dispersive relationships. The analysis of data will be presented using two different models. The non-dispersive model and analysis will show a clear difference in Rs between the three processes and a nearly constant tan ? for all samples measured. The dispersive model is difficult to confirm or deny, but there is some agreement found with previously measured results and fundamentals expectations. In the discussion section, the morphology characterization and HFSS simulations will then be reference to help explain the difference between the three pro- cesses. 102 4.1 Introduction As said before, resonator remains the only practical way to accurately measure Rs [131], [62] and tan ? [100, 132] at microwave frequencies. Because Q?1c and Q ?1 d are additive, the two losses are inseparable. Hence, existing resonant measurement techniques exploit regimes where the net Q-factor is dominated by either the superconductor or dielectric loss. The former case with Qc  Qd, favors the measurements ofRs. The latter case withQc  Qd, favors the measurement of tan ?. Tuckerman et al. take advantage of the condition Qc  Qd to measure tan ? of Nb- polyimide flexible transmission line tapes at 2 K [100]. Dielectric measurements by Krupka et al. optimize their resonant cavity design to satisfy the condition Qc  Qd and measure the dielectric loss tangent [132],[173]. Oates et al. Q-factor measurements on Nb-SiO2 sub-micron stripline resonators show they are limited by dielectric loss except for the narrowest strips, but they do not report measurements of tan ? or Rs [174]. Superconducting cavities measure Rs under circumstances where Qc  Qd [165], [175]. Taber overcomes the above limitations by varying the geometric factor (dielectric thickness) of the parallel plate resonator [136], [70], and extracts bothRs and tan ?, but his approach is impractical for characterization of superconducting integrated circuits if one wants to deconvolve losses in a single structure and measurement. To measure the frequency dependence of tan ? of thin film dielectrics commonly used as Josephson Junction barriers, Kaiser et al. neglects the metal loss by making use of lumped element LC resonators satisfying the condition Qc  Qd. In contrast with most prior work, to measure Rs and tan ?, I designed and tested microstrip transmission line (MTL) resonators having comparable superconductor and dielectric losses such that Qc ? Qd ? 1000. To be sensitive to Qc and Qd, the MTL was designed into fabrication 103 Process A, B, and C with multiple resonances and sensitive to intrinsic Q-factor Qi. In this chapter, I compare the extract Rs and tan ? from measurements of the frequency dependent internal Q-factor Qi(?) for MTL widths 0.25-4 ?m. I was most interested in analyzing Rs MTL width dependence and difference between Process A, B, and C, and seeing how Rs correlated with morphology measured by STEM and EDS in Chapter 3. 4.2 Resonator Design To implement the proposed concept, we designed a chip with five open-ended half-lambda MTL resonators shown in Fig. 4.1. The conducting strip width varies from 0.25 to 4 ?m. Each 15 mm long resonator was folded into meander shape to preserve space. Each resonator is reactively coupled to a 50? feedline via a coupling capacitor Cc. Via walls were placed about ten dielectric thicknesses (? 10s) away from the edge of the conducting strip, to reduce cross-talk between two adjacent resonator meander sections (see STEM cross-section in 3.15). The method to deconvolve tan ? and Rs0 can be found in Section 3.3. In this section, deconvolution will be demonstrated by comparing extracted RF losses for two damascened Process A and Process B) and a cloisonne? Process C. Results are summarized in Table 4.1). 4.2.1 Resonant Frequency MTL resonators support a TM00n [59] mode with the eigen frequency n fn = ? ? ? nc (4.1) 2lres LC ? 2lres ?r 1 + 2? coth(d/?)/s 104 Table 4.1: Results summary of Extracted Rs and tan ? from Process A, B, and C. The MTL resonator measurements were done at 4.2 K and are shown in Fig. 4.5. A homogeneous MTL, with ? = 90 nm, and a non-dispersive tan ? were assumed. Figure 4.1: a) Physical layout of five MTL resonators of varying widths w showing (bottom) contact pad arrangement, feedline wiring to Cc, and resonator meander geometry taking up chip area ? 0.5 ? 1mm2. The (Top) of a) is a zoom in of how the VNA makes connections to the feedline, followed by the coupling capacitor Cc connection to the MTL. This also shows the via walls that serve as isolation for the feedlines and MTL conducting strips. b) A cross-section of fabrication process showing the nominal critical dimensions of a two metal layer process: the ground plane (dark gray), conductiong strip (blue), surrounding TEOS dielectric (light gray), and SiNx passivation layer (gold). c) Representative diagram of the open ended MTL resonator connected to Cc, 50 ? feedline and VNA 105 where n = 1, 2, 3. . . is the mode index, lres is the geometrical length of the resonator, and L and C are the conducting strip series inductance and shunt capacitance per unit length defined by Eq. 3.3 and 3.5. The approximation on the right in Eq. 4.1 holds for a wide MTL where the width is much greater than the dielectric thickness (w  s) [70], where c is the speed of light, and ?r is the dielectric constant. The resonator frequency was designed based upon the nominal material properties (?r, ?) and the fabrication process thicknesses (s, d). The selection of lres was a trade-off between placing many resonators on a single 5 ? 5 mm2 chip, and a convenient frequency range to measure multiple resonant modes using a vector network analyzer (VNA) while avoiding the standing-wave ripple due to impedance mismatch. For a resonator length of lres = 15 mm the fundamental mode resonant frequency f1 ? 3.27 GHz. The average mode 1 index resonant frequency for Process C w = 4?m resonators f1 = 3.2? 0.07 GHz and is within 0.8% of wide MTL estimate using Eq.. 4.1, and this agreement supports our nominal ? ? 90 nm [54] and ?r ? 4.2 [73]. 4.2.2 Critical Coupling To overcome parasitic ripple seen in the transmission coefficient |S21| in Fig. 4.4(a), which is caused by inevitable impedance mismatch within the dip probe, we tune the coupling between the resonator and feedline to provide about 6dB insertion loss (IL) at the resonant frequency. For a reactively coupled resonator this corresponds to critical coupling, Qi = Qe, where Qe is the external (coupling) Q-factor. The feedline is directly connected to the top plate of the coupling capacitor while the MTL resonator is directly connected to the bottom plate. The Cc value is tuned to reach a critical 106 coupling g=1 using the following equation [176] R 1? |S21| Qi g = = = = 1 (4.2) 2Z0 |S21| Qe where g is the coupling factor, R and Z0 are the series resistance and characteristic impedance of the MTL resonator, |S | = 10?IL/2021 with IL being the insertion loss at resonance in dB, Qi is the resonator internal (or intrinsic) Q-factor, and Qe is the resonator external (or coupling) Q-factor. The input impedance Zin looking into a coupling capacitor towards a resonator is [95] ?Z0 Zin = ; bc = ?CcZ0 (4.3) 2Q 2ibc where ? is the angular frequency. Converting the ABCD transmission matrix for a shunt admittance Y = Z?1in into respective S21 gives [95] 2Zin S21 = (4.4) 2Zin + Zfeed where Zfeed is the feedline characteristic impedance. Assuming a parallel plate geome?try (w  ?s) and thick plates (d ?), the MTL characteristic impedance can be estimated Z0 = L/C ? ?0(s+ 2?eff )/ (?0?rs). Eq. 4.4 can be used to calculate the Cc needed to achieve the desired coupling factor g to a resonator with a given Qi, characteristic impedance Z0, and resonant frequency f0 [95] ( )? 1 2?g Cc = (4.5) 2?f0 ZfeedQiZ0 107 Using Eq. 4.5 and the following parameters expected for a 4 ?m conducting strip resonator f0 = 3 GHz, Qi = 685, Z0 = 8.6 ?, and Zfeed = 50 ? gives a Cc = 245fF to critically couple the resonator and achieve an IL ? 6 dB. The measured IL ? 6 dB for this geometry giving g ? 0.7, which means the resonator is under coupled and the measurement is sensitive to material losses. Error is most likely due to variations in the feedline and resoantor characteristic impedance due to fabrication. Fig. 4.2 shows four different coupling strengths for a 0.25 ?m MTL resonator. Each color is data averaged over 2 wafers and 9 chips (see Table 4.1. Each symbol is a different mode index of the MTL resonator (see Fig 4.4). Fig. 4.2(a) shows the extracted Qi for each coupling design (colors) overlay fairly well with the 12 GHz modes having the most spread. Fig. 4.2(b) shows Qi for calculated coupling factor g from measurement of IL (see Eq. 4.2). Similar calculations are found when taking the ratio of Qi/Qe within 15%. This data suggests the extracted Qi for all coupling designs are equal with error from wafer spread, and confirms our extraction method is valid and is fairly insensitive to coupling strength up to g = 3. Hence, the coupling capacitor design targets the following value ? 1 2? Ccrit = (4.6) 2?f1 ZfeedQiZ0 where f1 is the fundamental resonant frequency given by Eq. 4.1, Zfeed = 50? is the characteristic impedance of the feedline, Qi is the internal quality factor of the resonator that can be estimated from Eq. D.11, and Z0 is the characteristic impedance of the MTL resonator. The characteristic impedance can be estimated using Eq. D.8 assuming a parallel plate transmission line geometry or simulated in HFSS as described in Chapter 3. In this MTL design, the Cc varies from 50?250fF 108 Figure 4.2: Demonstration of the coupling design on a 0.25 ?m MTL resonator design showing fitted Qi on four resonant modes for four different couplings as a function of a) resonant frequency (or mode index), and b) the coupling factor g (see Eq. 4.2). The measurement method is described in Section 4.3. Representative S21(?) data and Qi fitting can be seen in Fig. 4.4. These measurements are on samples fabricated using Process C taken at 4.2 K for MTL w = 0.25? 4 ?m, respectively. A plaid capacitor design (see Fig. 4.1(a)) was used for the coupling capacitors. This capacitor design enabled the fabrication of large capacitors in a damascene process, where wide metal patches are disallowed to minimize under (over) etching of low (high) metal density during CMP. The capacitor is composed of multiple parallel lines (with minimum allowed spacing) in one metal layer and many more in adjacent metal layer, running perpendicular to the lines in the first layer. The two metal layers are separated by TEOS of thickness s. Every other line in each layer are connected with vias thus creating two interwoven (plaid) electrodes. While the structure complies with design rules that limit wire width and metal density it exhibits electrical performance of a parallel plate capacitor; it has low parasitic inductance and approximately the same capacitance (within ? 30%) per unit area. 109 4.3 Measurement Chips were selected within the inner 80 mm diameter of 150 mm wafers (see Fig. 4.3). The number of chips and wafers measured from each Process can be found in summary Table 4.1. Measurements were taken at 4.2 K with a cryogenic, RF dip probe outfitted with a 32 pin (contact pad) test fixture. The same test fixutre was used in previous works [35], [34], [36], [10]. To achieve RF signal integrity and avoid wire bonding for fast sample chip exchanges, a non-magnetic Cu/Au flip chip press contact technology is used. The fixture PCB connects the flip-chip bump pads to the probe .047 inch semi-rigid coaxial cables. The chip and roughly 30 cm of the cables are submerged in L-He bath during measurements. The temperature was monitored with a calibrated Lakeshore Cernox sensor. The resonators were measured by 2-port S-parameters using a Keysight N5222A vector network analyzer (VNA). The network analyzer was calibrated up to the top of the probe. Typical measurements were taken over 2? 14GHz frequency span with 200 kHz spacing with input power ? ?20 dBm to minimize possibility of non-linear effects. The probe usable bandwidth accommodates resonant modes up to n = 4. Measurements above 14 GHz are dismissed due to the standing-wave Q-factor becoming comparable to the resonance Q-factor. Fig. 4.4(a) shows the first four resonant modes for each MTL width w. From the complex S- matrix circle fits (see Fig. 4.4(b), theQi, Qe, and fn are found using the methodologies described in [177, 178], where the Q-factors are related to the loaded Q-factor by Q?1 = Q?1 +Q?1L i e in the complex plane. The first mode fitting for each width is shown in Fig. 4.4. Measurements shown in Fig. 4.4 were taken across 4 wafers and a total of 22 chips (5 MTLs per chip). All resonances from 2-14 GHz were detected, fitted, and analyzed using a custom Python script. 110 Figure 4.3: Wafer Map Diagram used for all three Processes showing (a) the 16 mask reticles (blue) and the chip (pink) fabricated with the resonator chips. (b) A zoom in of the MTL resonators chip (see Fig. 4.1) for more details on chip layout and design. All measured chips from Process A,B, and C were picked from the inner 80 mm diameter of 150 mm wafers shown here. 4.4 Method Demonstration The statistical mean and standard deviation for all modes, conducting strip widths, and fabrication processes are shown in Fig. 4.5. The increased spread in Q?1i for n = 3, 4 is due to the fitting error due to standing wave ripples and reduced measurement repeatability. A separate gauge study was performed by installing, testing, and re-installing the same chip 4 times and yielded max percent variations in extracted Qi and fn to be < 10% and < 0.1%, respectively. From Fig. 4.5, it can immediately be seen that Process A has larger slopes for all conducting strip widths relative to Process B and Process C, and most convinently seen in Fig. 4.5(f). Using Eq. 3.16, a larger slopeA yields an approximately larger intrinsic resistance assuming the slope is 111 Figure 4.4: Data for a chip from center of the wafer fabricated using process Process A. (a) Insertion loss versus frequency for modes n = 1 ? 4 on MTL widths 0.25 ? 4 ?m at 4.2 K. (b) Complex circle fitting of the fundamental mode resonant peaks from (a) using methodology from [177, 178]. exclusively and the equal geometric factors between each process. The linear dependence of Qi with frequency ? in Fig. 4.5 supports the inclination that Eq. 3.15 is valid, but will be scrutinized in section 4.4.2. At the the same time, for lines 1?m and narrower Process B has relatively higher Rs relative to Process C, while for the 2 and 4 ?m lines Process B and Process C have nearly the same superconductor loss. It is also evident from Fig. 4.5 that all MTL widths and processes yield roughly the same linear y-intercept, which is related to the dielectric loss tan ? ? 10?3. Although, this is all assuming the ground plane and conducting strip have the same intrinsic resistance Rs, a non-dispersive tan ?, and the same geometric factors between each process. The fitted linear slope using Eq. 3.15 from Fig. 4.5 as a function of process and MTL width is shown in Fig 4.6. For reference, the analytical geometric factor assuming a parallel plate model ?PP (solid black line) is also shown calculated using Eq. 3.16 with nominal dimensions shown in the parallel plate MTL diagram. Primarily for processes A and B, the slope A is approximately 112 Figure 4.5: The reciprocal internal Q-factor Q?1i at 4.2 K as a function of the resonant frequency for each of the five MTL widths and for three fabrication processes. The bottom right graph is five linewidths and processes plotted together. 113 constant for MTL widths 1 ? 4 ?m. A noticeable upturn in A can be seen for MTL widths 0.25 and 0.5 ?m for processes A and B. Additionally, the y-offset of ? 1.5 ? 10?14 Hz?1 between process A and B is a strong indication the intrinsic resistance is larger for process A with respect to process B. Furthermore, process C has a near MTL width independent slope A and ? 1 ? 10?14 Hz?1 lower A at MTL width 0.25?m with respect to process B. Empirically, it can be concluded that process C has a lowest conductive loss and does not have a width dependence. Although, accurate conclusions about the intrinsic resistance Rs cannot be drawn due to the convoluted width dependence of MTL geometric factor ?c0 found from HFSS simulations in Chapter 3. As an example of the worst case scenario, without knowing the cross-section of the MTL fabricated it can be conceivable the differences between process B and C are completely due to geometry differences as opposed to the intrinsic resistance Rs. Utilizing the methods and procedures described in Chapter 3 as well as the measurements of cross-section dimensions from TEM and FIB, the geometric factors were numerically solved in HFSS and are plotted in Fig. 4.7. Similar to Fig. 3.14 for nominal cross-sectional geometries, Fig. 4.7(a) shows ?c0 solved for the as measured cross-sections. The solid black line is referenced for the nominal geometries. Errors are calculated between measured relative to nominal and can be seen in Fig. 4.7(b). Process A and B have percent errors increasing from of approximately 3-6% as the MTL width is reduced from 4 to 0.25 ?m, respectively. Process C has an almost consistent error for MTL widths 0.5-4 ?m of -2%, but increase up to 6% for 0.25 ?m (delta of 8%). Referring to the STEM summary Table 3.1, maximum error (variation) of 8% in ?c0 correlates fairly well to the maximum dimension variation of approximately 7-9% of the dielectric thickness. From an engineering and fabrication perspective where targeting consistent critical dimensions is important, 8% variation is manageable and is a good result confirmed by HFSS. 114 Figure 4.6: Plots showing the fit slope A as a function of MTL width for all three processes (assuming non-dispersive model). (solid black line) The predicted slope A using Eq. 3.16 assuming the minimum BCS intrinsic resistance RminBCS and geometric factor for a parallel plate MTL (very wide conducting strip) using Eq. D.11 at 10 GHz and 4.2 K. 115 In other words, an 8% variation in dielectric constant only varies the ?c0 by roughly the same amount. This affords a two-fold understanding: (1) the Q-factor of the MTL resonator (RQL interconnect) related to power dissipation is fairly resilient to fabrication variation geometrically and (2) any variations in Q-factor above the fabricated geometric variation can be attributed to extrinsic loss variation (i.e. Rs and tan ?). Deconvolving the geometric factor further, the partial geometric factors of the conducting strip ?0CS and ?0GP are plotted in Fig. 4.7(c) and Fig. 4.7(d), respectively. For the conducting strip, there is about a 5% variation between all three processes at 4?m MTL width. Surprisingly, as the MTL width is reduced, ?0CS becomes comparable for all widths and almost equal at 0.25?m MTL width. The total MTL geometric factors for all processes are also equal at 0.25 ?m MTL shown in Fig. 4.7(a). Note, the microstrip geometries are notably different and diagrams are shown on the right of Fig. 4.7(b) and better seen in Fig. 3.1. This is an important conclusion as any changes of the partial Q-factor associated to the conductor Qc, or Qi since tan ? is geometry independent, can be attributed to changes in the intrinsic resistance Rs and dielectric loss tangent tan ?. Referring back to STEM summary Table 3.1 on the large discrepancy in ground plane thickness of 50% in process A, the partial geometric factor of the ground plane ?0GP has a marginal 10% maximum difference. This is most likely due to the ground plane thickness not falling below the estimated penetration depth of ? ? 90nm. Referring to Fig. 3.13, the magnetic fields are penetration the 4 ?m MTL an approximate distance of ?. If process A ground plane thickness dGP to magnetic penetration depth ratio becomes much less than one dGP/? 1, then significant change would be seen in ?GP (see Eq. 2.24). Conversely, for a 0.25 ?m MTL, the magnetic fields (currents) are more concentrated in the conducting strip (Fig. 3.13), so dGP/? 116 1 will have less of an effect, but an effect to keep track of nonetheless. The next two sections will extract Rs and tan ? more accurately by assuming one of three models: (i) non-dispersive tan ? and homogeneous superconductor, (ii) non-dispersive tan ? and in-homogeneous superconductor, and (iii) dispersive tan ? and homogeneous superconductor. 4.4.1 Non-Dispersive tan ? Model In this subsection, the non-dispersive tan ? model will be assumed throughout unless otherwise specified using Eq. 3.13-3.16 and assumptions therein. Taking the MTL geomtric factors solved by HFSS in Fig. 4.7, the intrinsic resistance Rs from process A is extracted from the fitted slopes A as a function of MTL width in Fig. 4.6. The red data set extracts Rs assuming a parallel plate model with ?PP ? 12 m? (see dashed black line in Fig. 4.7(a)). The blue data set extracts Rs assuming a homogeneous MTL where Rs0GP = Rs0CS and solved ?c0 in Fig. 4.7(a). The black data set (overlapping the blue data set) extracts Rs0 assuming a non-homogeneous MTL where Rs0GP =6 Rs0CS for MTL widths less than 4 ?m. Immediately it can be seen that the parallel plate model (red data set) has approximately 25% high Rs for 0.25 ?m MTL relative to the homogeneous model using HFSS solve geometric factors. This means that if one was to use the parallel plate analytical model alone, the extracted Rs would be an upper bound of the actual intrinsic resistance. As MTL width dimensions shrink below 0.25?m the induced error in extracted intrinsic resistance will also increase. This makes it difficult to distinguish whether changes in Q-factor or intrinsic resistance Rs are due to extrinsic effects caused by processing as the conducting strip goes to minimum dimensions or 117 Figure 4.7: Plots showing the calculated geometric factors ? from HFSS using STEM/FIB measurements. a) The total MTL geometric factor calculated by HFSS from measured MTL geometries versus MTL width. These will be used to extractRs from the slopeA found in Fig. 4.5 b) The % difference (or error) of ?c0 from measured geometry versus assuming nominal cross section dimensions. c) The partial conductor geometric factor ?0CS as a function of MTL width calculated by HFSS. d) The partial ground plane geometric factor ?0GP as a function of MTL width calculated by HFSS. HFSS simulations and method can be found in Chapter 3. STEM measurements can be found in Section 3.2. FIB cross section measurements can be found in Appendix A 118 if its completely due to geometrical effects. Going a step further, if one wants to monitor the process of the conducting strip alone and the intrinsic resistance of the ground plane Rs0GP is known with some certainty and can be assumed to not be changing, then one can employ the homogeneous model with the following equation Rs0CS = (A?0 ?Rs0GP/?0GP )?0CS (4.7) where Rs0CS is the intrinsic resistance of the conducting strip, Rs0GP is the intrinsic resistance of the conducting strip, A is the fit slope from Fig. 4.6 and Eq. 3.15, ?0GP is the partial geometric factor of the ground plane, ?0CS is the partial geometric factor of the conducting strip, and ?0/2? = 10GHz is the reference frequency. Here, for the in-homogeneous model, it is assumed at 4 ?m MTL width the ground plane and conducting strip have equal intrinsic resistances Rs0CS = Rs0GP . The coinciding of the homogeneous (blue data) and in-homgeneous (black data) for all MTL widths suggests that the Rs0CS is not significantly larger than the Rs0GP . Although, as the MTL width decreases, the geometric factor of the conducting strip ?0CS decreases and therefore dominates the conductive loss. The geometric factor can be looked at as a participation ratio similar to methods shown in the quantum community to determine which interface dielectric losses participate more [145, 146]. In this case, a large partial geometric factor for the m-th conductor ?0m, its respective intrinsic resistance Rs will participate less in the overall Q-factor dissipating less power and vice versa. For 0.25 ?m MTLs, ?0GP ? 10?0CS so the loss in the conducting strip is dominating the microstrip conductor loss. Referring now to Fig. 4.9, to compare the power loss in the MTL resonators (MTL interconnects) 119 Figure 4.8: (red) Extracted effective resistance Reff from Qc assuming Eq. D.11 for a parallel plate. (black) Extracted intrinsic resistance Rs0 from Qc using ?c0 calculated from HFSS at each MTL width. (blue) Extracted intrinsic resistance of the conducting strip Rs0CS assuming Rs0CS(4 ?m) = Rs0CS(4 ?m) using calculated ?cGP and ?cGP from HFSS at each MTL width (see Fig. 3.14). The reference frequency is ?0/2? = 10 GHz. The cross section geometries of each MTL was measured in TEM (see Section 3.2). 120 for each fabrication process, the homogeneous model with a non-dispersive loss tangent is assumed from Eq. 3.13-3.16 using the HFSS solved MTL geometric factors ?c0(w). Fig. 4.9(a) and Fig. 4.9(b) show the extracted intrinsic resistance Rs0 of Nb at reference frequency ?/2? = 10 GHz and assumed frequency independent tan ? of TEOS as a function of MTL width at 4.2 K, respectively. The error bars are?20% as worst case error of 20% variation in geometric factor across a 150 mm wafer (see Fig. 4.3). The black dashed line is the BCS minimum intrinsic resistance RminBCS as measured by Benvenuti et al. [64] for thin film Nb samples at 4.2 K and scaled to 10 GHz using quadratic frequency scaling [62]. Since a 4 ?m MTL resonator can be approximated as a parallel plate resonator, the Rs0 can be compared to parallel plate resonator measurements at 4.2 K and 10 GHz. The intrinsic resistances for a 4 ?m MTL width ranges from Rs0 ? 15 ? 25, and is in good agreement with Rs0 ? 20 measured by Taber [136] using his parallel plate resonator technique. To get accurate measurements of Rs0, Taber varied the dielectric spacer thickness in the parallel plate resonator, which requires about 1-2 days worth of measurements. Here, we extracted Rs0 in a single measurement by relying in the frequency dependence of Qc and assuming a non-dispersive loss tangent. The data will now be compared between processes in Fig. 4.9. Process A, Rs0 is relatively high for all MTL widths relative to process B and C. This was also inferred in Fig. 4.6 and discussion. There might be a noticeable upturn in Rs0 as the MTL width goes down to 0.25 ?m, but cant be claimed with the large error bars. Recall, process A has an inverted MTL geometry where the ground plane is above the microstrip and will be discussed further in the discussion section (see Fig. 3.2 and summary Table 3.1). Process B, a damascene CMP process, has an Rs0 ? 15 ?? for MTL widths 0.5 ? 4 ?m and an upturn in Rs0 ? 20 ?? for 0.25 ?m MTL. Process B falls below the BCS minimum intrinsic resistance measured by Benvenuti et al. for 121 MTL widths 0.5 ? 4 ?m within some error. This is an outstanding result for a damascene CMP processed RQL interconnect at submicron dimensions and demonstrates its possible to fabricate Nb submicron lines with the theoretically lowest power dissipation. Based on process C results for Rs0 and small error bars, a cloisonne? process with a metal etch and dielectric CMP, it can be argued that Nb 0.25 ?m wires can have Rs0 below what has been measured before. For all processes, especially process C, a dip (or minimum) in Rs0 at about 1 ?m can be seen. At this time, a robust model of this minimum at 1 ?m has not been developed and still needs to be looked into further. Although, it is worth noting that from Fig. 3.13, the magnetic fields begin to overlap on-top of the conducting strip and a redistribution of magnetic field circulating at the top of the strip rather than be concentrated at the bottom edges may induce lower loss. Due to the uncertainty in Rs0 from the assumed worst case error bars, it can be inferred that process C Rs0 does not have a width dependence for MTL widths 0.25? 4 ?m. Now focusing on the power dissipation in the TEOS dielectric, Fig. 4.9(b) shows all three processes have approximately the same loss tangent tan ? ? 1.2 ? 0.1 ? 10?3 and virtually no MTL width dependence. This is not surprising as the deposition parameters of the TEOS were unchanged for all three processes and tan ? is ideally geometry independent. In spite of this, it is plausible that the presence of a NbOx layer in process A and B (or lack thereof in process C) could participate as an additional loss mechanism if the NbOx was a dielectric. Also, the TEOS RIE etch process in processes A and B could affect tan ? differently relative to the TEOS CMP process using chemical mechanical polishing in process C (reduce or increase loss). In retrospect, this is an compelling result that TEOS is not sensitive to small changes in the process and therefore has a desirably large processing window (see Section 2.4). 122 Figure 4.9: (a) Extracted intrinsic resistance Rs0 at ?0/2? = 10 GHz, and (right) tan ? at 4.2 K as a function of MTL conducting strip width w and fabrication process recipe. The Rs0 from slope A (see Fig. 4.6) HFSS calculated ?c0 (see Fig. 4.7). The Rs0 error bars are the worst-case scenario of +/-20% using the parametric corners results in Fig. 3.17. These are conservative error bars relative to measurement error, fitting error, and linear fitting errors in Fig. 4.5. The black dashed reference on the left is the known BCS limit of the intrinsic resistance at 4.2 K and 10 GHz. The BCS Rs can be calculated using [179]. The minimum BCS Rs has been measured at 4.2 K for Nb thin films [64]. This may be the first time it has been reported for sub-micron Nb wires. 123 4.4.2 Homogeneous Superconductor, Dispersive tan ? Model In the last section, we assumed the loss tangent tan ? was constant in the frequency range ? 1?15GHz. It was found that Process C produces MTLs having an intrinsic resistanceRs near or below BCS limit of 17 ?? at temperature T=4.2 K and ?0/2? = 10 GHz reference frequency. It was also found that the frequency dependence for Qi scaled approximately linear. This affords the opportunity to assume a model for Q?1c and subtract it from Q ?1 i in order to analyze the frequency dependence of Q?1d ? tan ?. Assuming Q?1c ? ? frequency dependence, Rs0 = 17?? at 10 GHz and 4.2 K, and a nominal geometric factor for 4 ?m MTL ?c0(4 ?m) = 11.5m?, the superconductor loss slope can be calculated to be A = 2.35 ? 10?14 Hz?1 (see second term in Eq. 4.8). In Fig. 4.10(a) below, the black solid line is the model for Q?1 = ? Rc0c just described?0 ?c0 and agrees nicely for 4 ?m MTL (purple) with a constant loss tangent tan ? = 9 ? 10?4. This is more validation the method described here fits the data well. Although, it is interesting to note the 4 ?m , 3 GHz (purple) data point does not coincide with the black line, suggesting some non-linearity. To further analyze this, and compare with Kaiser et al. [99], utilizing Eq. 3.13, the frequency dependent loss tangent tan ?(?) can be inferred from Fig. 4.10(a) by ? Rc0 tan ?(?) = Qi(?)? (4.8) ?0 ?c0 where the subtracted term is the partial Q-factor associated to the conductor Q ? Rc0c = . Using?0 ?c0 Eq. 4.8, Qc was subtracted from the Qi data and can be seen in Fig. 4.10(b). The data measured here is compared with Kaiser et al. measurements of tan ?(?) for Nb2O5 and SiO2 amorphous dielectrics (dashed lines), the dielectrics most closely related or involved in our system. For 124 Figure 4.10: a) Inverse internal Q-factor Qi versus frequency for Process C taken from Fig. 4.5. The black solid line is assuming Rs0 = 17?? at 10 GHz and 4.2 K, and nominal ?c0(4 ?m) calculated by HFSS b) Qd ? tan ? inferred using Eq. 4.8 and compared to Kaiser et al. data [99]. TEOS fabricated here is non-ideal as the deposition temperature is limited to 150 C, preventing crystal like growth kinetics and the presence of voiding (see Fig. 3.3(d) feature 5). Therefore, it is expected the loss tangent of TEOS is 1-2 orders of magnitude higher than Kaiser et al. SiO2. The Process C inferred tan ?(?) for all MTL widths in Fig. 4.10(b) agree well and fall in between Nb2O5 and SiO2 loss tangents measured by Kaiser et al.. From 3 to 10 GHz, the inferred tan ?(?) decreases by a factor of 2 for 0.25?m MTL, and only decreases by about 10% for the 4 ?m MTL. In comparison, the Qc increases by a factor of 2 from 3 to 10 GHz for all MTL widths. Using Eq. 2.44, the loss tangent data is fit to a power law and plotted as a function of MTL width in Fig. 4.11(a). The fit parameter bd agrees well with Kaiser for 0.25 ? 2 ?m MTLs in the regime of 0.5-0.75. The 4 ?m MTL has a relatively higher bd = 0.86, giving a loss tangent frequency dependence of tan ? ? ??0.14. It can be argued that the 4 ?m MTL, at low frequencies ?, is the most sensitive to tan ?, as ?c0 is the largest at this width (see Fig. 4.7 and Eq. 3.13). This could explain the non-linearity in Fig. 4.10(a) mentioned eariler, and can also be seen in Fig. 4.10(b) (purple data). Additionally, 125 Figure 4.11: Results plotted from Fig. 4.10 showing the following as a function of MTL width: a) fitted exponent bd from Eq. 3.12 and 1 standard deviation error and b) fitted loss tangent tan ? at 10 GHz. Kaiser et al. measured bd ? 0.6 and tan ? ? 3? 10?4 for SiO2 at 10 GHz. the first mode for 4 ?m MTL falls on top of all other MTL widths, but then deviates at higher frequencies. From the dispersive fits in Fig. 4.10(b), the loss tangent tan ? at ?/2? = 10GHz reference frequency as a function of MTL width is plotted in Fig. 4.11(b). In comparison to extracted non- dispersive loss tangents in Fig. 4.9, the dispersive loss tangent is approximately 20-60% lower at 10 GHz. The agreement with Kaiser et al. of tan ? magnitude and power dependence as a function of frequency for similar dielectrics is a good first indication that there is dispersion the TEOS dielectrics measured here at 4.2 K. Yet, we had to assume an Rs0 and geometric factor ?c0. To scrutinize the dispersion relationship further and demonstrate the superconductor and dielectric loss deconvolution method described here, I used Eq. 3.12 to fit the Qi(?) data to 3 free fitting parameters shown in Fig. 4.12. Except the process A 1 ?m MTL, all fits (dashed lines) fall in between the error bars, which are one standard deviation of Qi variation across the wafer. 126 An interesting visual result can be seen in Fig. 4.12(e) for the 4?mMTL, which is the most sensitive to the loss tangent, shows process B is concave down and process A and C are concave up. In other light, process B and process C 4 ?m MTLs have almost the same mode 1 and mode 4 Q?1i , but modes 2 and 3 in process B are higher. This discrepancy and false concave down character in process B may just be due to the error bar for mode 4 in process B. Other than the fits for 1 ?m MTLs and process B 4 ?m MTL, all of the fits show a minimum in Qi somewhere below 2 GHz. If tan ? is expected to decrease as a function of frequency and the inverse partial conductor Q-factor Q?1c is proportional to frequency, then a minimum in Q ?1 i is expected. This minimum in the internal Q-factor Qmini (?min) is a useful parameter for application, as it is the optimum frequency to operate RQL interconnects in terms of minimum power dissipation. Using tan ?(?) for SiO2 measured by Kaiser et al. (Fig. 4.10(b)), and assuming minimum Qc(?) with RminBCS = 17?? and ?c0 = 11.5m? (Fig. 4.10(a)), the frequency for mimum power dissipation can be estimated to be ?min/2? ? 1GHz. In a future MTL design, the fundamental mode should be decreased to ideally below 1 GHz to see if there is a minimum in Q?1i . The dispersive fitting parameters Ad, Bd, and bd as a function of process and MTL width are shown in Fig. 4.13(d-f), where A = Ad is the slope associated to the conductor loss, B = Bd is ?non-dispersive? tan ? if b = 1, and b = bd is the dispersive power exponent (see Eq. 3.12). The non-dispersive fitting parameters A = And, B = Bnd, and b = bnd are shown in the first row for reference. For a 3 parameter fit on only data points, this system is nearly over-determined and could explain a lot of the scatter in Fig. 4.13(d-f) and Fig. 4.13. Similarly said before, other than 1 ?m MTLs and process B 4 ?m MTL, all dispersive exponents are less than 1 b = bd < 1. This is a notable result, as a 0.5 < b < 1 fitted from 3 fitting parameters without assuming loss in the superconductor is in agreement in Kaiser et al.. Thus, to first approximations, this demonstration 127 Figure 4.12: Internal Q-factor Qi(?) for Processes A (red), B (blue), and C (green) for MTL widths a) 0.25 ?m (crosses), b) 0.5 ?m (diamonds), c) 1 ?m (triangles), d) 2 ?m (squares), e) 4 ?m (circles) and f) all plotted together. The 3-parameter dispersive tan ? fits (dahsed lines) use Eq. 3.12. Diagrams of the MTL geometry for each process are shown in a) for convenience. 128 Figure 4.13: The top row are fitting parameters for non-dispersive tan ? model assuming b = 1 (linear fits) shown in Fig. 4.5 and using Eq. 3.12. Width dependence of linear a) slope parameter A due to conductive loss, b) assumed non-dispersive loss tangent B = tan ? where c) b = 1. The second row are fitting parameters for dispersive tan ? model showing width dependence of free fitting parameters: d) slope A proportional to conductive loss e) pre-exponent term B f) and exponent term b. Diagrams of the MTL geometry for each process are shown in d) for convenience. serves as a validation of the deconvolution method described here because 1) the Rs results are comparable to measured values found in literature, and 2) the tan ? dispersive relationship and values agree with previous findings. Aside from the 1?mMTLs and process B 4?mMTL, extractedRs0 and tan ? assuming the dispersive model in Fig. 4.14 general agrees with the non-dispersive extraction method in Fig. 4.9. By general, i mean the Rs0 magnitude is still in between 15-35 ?? using the dispersive model and tan ? is within 50% of values relative to non-dispersive fits. This is a very important result, regarding the Rs0 values falling below BCS limit in Fig. 4.9(a) being suspect RBCS = 17 ?? < Rs0. The minimal change in Rs0 values using dispersive fits here only marginally increase and 129 Figure 4.14: Extracted a) Intrinsic resistanceRs and b) tan ? versus MTL width using dispersive tan ? 3-parameter fit model support results in Fig. 4.9(a). MTL width dependence is not discernible in Fig. 4.14 due to the noise, which is most likely due to the dispersive fitting model being close to over-determined with 3 fitting parameters and only 4 data points. Ultimately, exploring the dispersive tan ? model and fitting the data as a function of MTL width and different processes, varying in loss significantly as a function of frequency, offered the proposed gleaned information: (1) TEOS tan ?(?) extracted from process C Qi data using the simple Q?1c subtractive method agrees with Kaiser et al. in both magnitude and frequency dependence for SiO2 (2) TEOS tan ?(?) extracted from a 3-parameter fit for the majority of data (process and MTL 130 width) also agrees with Kaiser et al. in both magnitude and frequency dependence for SiO2 (3) Nb Rs0 extracted from a dispersive 3-parameter fit give support of values measured below the results measured by Benvenuti et al. [64] 4.5 Discussion The previous two sections have demonstrated the method to deconvolve the losses by assuming different models. In this discussion, we will focus on the results from Fig. 4.9, where a non-dispersive tan ?, homogeneous (Rs0GP = Rs0CS) MTL model was assumed. Models to explain the intrinsic resistance Rs difference as a function of (i) width dependence) and (ii) fabrication process (A,B,C) will be proposed relying on the morphology and chemical analysis in Section 3.2. Up-turn of Rs at 0.25 ?m In reference to Fig. 3.3(a) and (c), the damascene process induces two grain orientations, sidewall horizontal grains and vertical trench grains, meeting along a diagonal line with approximate angle of 30 degrees. Here, we will call this the sidewall grain intersection. At these intersection, voiding (or lower metal concentrations) occur. The percentage of sidewall horizontal grain on both sides of the microstrip remain fairly constant in size with maximum length of 100 nm at the top of the microstrip. As the microstrip width decreases, the crossectional area fraction of the sidewall grains increases. Here, we will call this the sidewall grain fraction. For a 1 ?m microstrip, sidewall grain fraction is only about 20%, whereas for a 0.25 ?m microstrip the sidewall grain fraction is approximately 50%. This is a model that could contribute to the upturn in Rs of about 20-30% for 0.25 ?m MTL relative to 4 ?m for process A and B. 131 Another model that could attribute to the upturn in Rs is the presence of voids, or areas of lower Nb concentration (see Fig. 3.3). Voids become a more significant fraction as microstrip width decreases. The effect of voiding on microwave surface resistance has not been studied carefully. Although, there has been extensive studies on how Nb vacancies can facilitate Nb hydride precipitate formation [180]. Consequently, the Nb voids could be acting as hydride formation sites. In addition, Nb voids could have an effect on the electron mean free path, and in turn increasing both the magnetic penetration depth and microwave surface resistance (see Eqs. 2.20 and 2.8). It is worth mentioning a model of the NbOx layer as the mechanism causing the upturn in Rs at 0.25 ?m. In reference to Fig. 3.13, the fringing magnet fields do not start start to sample (overlap) a significant percentage of the top of the microstrip until below 1 ?m. In process B, the exact geometry shown in Fig. 3.13, the upturn in Rs does not happen until 0.25 ?m. It is conceivable the Rs upturn here is due to NbOx, but would have expected a more gradual increase in Rs. This argument is more unlikely when considering Process A, as the fields are sampling the NbOx layer for all MTL widths of equal fraction, and is more likely to be a different mechanism. An upturn in Rs in Process C is not visible. Rs versus MTL width offset In process A, due to the orientation of the GP relative to the fill morphology (sidewall grain orientation), the magnetic field (current density) is more concentrated at the oxide layer. The current is also concentrated where voiding occurs. Whereas in process B, the damscened surface is not at a location of highest current density. In ?morphology section? we also found from chemical mapping that the oxide layer is closer to NbO or Nb2O3, which is a bulk superconductor with Tc approximately 6 K, but at 5 nm thickness is most likely a normal metal. Therefore, the 132 oxide location relative to the current concentration is a model that can explain the Rs offset by approximately a factor of 2 between process A and B. Process C does not have a detectable oxide at the interface of highest current concentration, suggesting the offset of approximately 40% between process B and C 0.25 ?m microstrip. Surprisingly, due to the etch undercut process, the rough and irregular surface of the strip did induce Rs larger than damascene processes A and B. 133 4.6 Summary In summary, microstrip transmission line (MTL) resonators were designed into an RQL fabrication process where the Q-factors associated to conductor and dielectric loss are comparable Qc ? Qd ? 1000. The comparable Q-factors are not compatible with established techniques to analyze the dissipative loss coming from the superconductor and dielectric separately. Using the methods and HFSS numerical modeling procedures described in Chapter 3, the intrinsic resistance Rs of Nb and loss tangent tan ? of TEOS were extracted from measured intrinsic Q-factors as a function of frequency Qi(?) utilizing their unique dispersive relationships. The uncertainty in extracted Rs0 was used from the HFSS sensitivity analysis in Section 3.7 with a worst case error of ?20%. The overall uncertainty is dictated by the fabrication geometry where the microstrip and dielectric separation thicknesses have the most significant impact on geometric factor variation. The standard deviation from measured values of the internal Q-factor Qi across a wafer was much less than ?20%, suggesting a well controlled fabrication process where both the resonator geometry and the intrinsic surface resistances are not varying significantly across the wafer. Different dispersive models were proposed and analyzed to demonstrate and give validation to the Rs and tan ? extraction method from a single measurement at 4.2 K. Three fabrication processes were analyzed using this method, and found that the cloisonne? process C yields Rs0 of Nb near theorical minimums down to 0.25 ?m MTL widths. Near equivalent Rs values can be achieved in the damscene process B down to 0.25 ?m MTL widths. A NbOx layers in MTLs can induce an extrinsic increase in Rs0, and its location relative to current density concentration inside the conducting strip can exacerbate this issue. An Al capping layer has been shown to prevent NbOx formation at the conducting strip interface, and suggests this aided in the 134 minimum and near width independent Rs0 measured in cloisonne? process C. For all fabrication processes and MTL widths measured here, TEOS loss tangent tan ? is width independent, and the measured frequency dependence agrees with previous work on similar amorphous dielectrics. The materials characterization done here informs the implications on power dissipation performance of RQL interconnect operating at 4.2 K and 10 GHz. The TEOS dissipative loss is geometry independent, and only dependent upon fabrication. If the TEOS could be deposited at temperatures higher than 150 C, then it is possible to reduce its dissipative loss 1-2 orders of magnitude (Qd ? 104 ? 105). The Nb dissipative loss has arguably reached its theoretical minimum (Qc ? 3 ? 4 ? 103). The current minimum energy efficiency is ? ? 0.3 of RQL resonator clocks relative to the power dissipated by the RQL logic ? = Plogic/(Plogic + Pclock). The efficiency is proportional to the resonator internal Q-factor by ? ? 1/(1 + Q?1i ), where the power dissipated in the clock is proportional to the Q-factor by Pdiss ? Q?1i having nominal valueQi = 250 at 4.2 K and 10 GHz. Taking a resonator composed of MTL interconnect down to 0.25?mwidths fabricated by process C, the efficiency would increase to ? ? 0.6. If the dielectric loss could be reduced by a factor of 10 and 100, the effieciency would increase to ? ? 0.83 and ? 0.86, respectively. This is an exceptional decrease in power dissipation relative to the logic at maximum operation (i.e. all logic units are active in the circuit), but it is important to note that the efficiency is at its limit and cannot be pushed further unless materials other than Nb and TEOS are used or changes in the transmission lines geometries are allowable. At the current RQL clock efficiencies of ? ? 0.3, it can be estimated that the total power consumption of RQL at 4.2 K would be a factor of 90 less than equivalent CMOS circuits, which includes refrigeration costs [4]. For ? ? 0.86, RQL would consume approximately 260 times less than equivalent CMOS circuits. In the next chapter, temperature will be used as a knob to 135 explore further reduction in power dissipation in RQL interconnects. The determination of power dissipation limits and an optimum temperature of operation will also be investigated. 136 Chapter 5: Temperature Dependent Characterization of Microwave Loss in RQL Interconnects 5.1 Overview In the last chapter, I measured the internal Q-factor of microstrip transmission lines (MTLs) and delconvolved microwave losses of the superconductor and dielectric at 4.2 K. It was found that the theoretical minimum of Nb BCS intrinsic resistanceRs ? 17?? was reached for Process C down to 0.25 ?m widths at 4.2 K. The TEOS dielectric loss tangent was found to be tan ? ? 1 ? 10?3 at 4.2 K and is MTL geometry independent. It is known both Rs and tan ? decrease with decreasing temperature. Hence, to maximize energy efficiency in RQL interconnects, the temperature could be lowered to reduce the RF losses. Although, refrigeration costs increase as the operation temperature is reduced. It takes approximately 3 times more power to reduce the operating temperature from 4.2 to 2 K [181]. As mentioned in the Chapter 2, the microwave loss of superconductors and dielectrics have different scalings with temperature. As temperature is decreased below 4.2 K, The bulk and thin film Nb Rs temperature and frequency scaling is well known as temperature is decreased below 4.2 K. For amorphous dielectrics like TEOS, the temperature (below 4 K) and frequency dependence strongly depend on the input power (electric field strength) and dominant mechanisms 137 of loss (e.g. classical dielectric relaxation, TLS tunneling relaxation, and TLS tunneling absorption). Furthermore, the dominant mechanism of dielectric loss in the dielectric can change rapidly as temperature is decreased. Up until the 1990s, there had been limited microwave dielectric measurements in the 1.5-4.2 K temperature regime, as most SCE communities operate at 4.2 K (SFQ) or at mK temperatures (Quantum Computing). In this chapter, the dispersive loss deconvolution (DLD) method will be applied to temperature dependent measurements. In the 1.5-4 K temperature range, the extracted temperature dependence ofR (T ) ? e?1/Ts and tan ?(T ) ? T 3 agree with expected behavior. The measurement sensitivity of this method will be discussed. Specifically at temperatures below 3 K, the frequency dependence of the Qi is anomalous with non-monotic behavior increasing in magnitude as temperature is decreased. With only 4 modes in MTLs, it is difficult to draw conclusions for the Qi frequency dependence. This warranted the design and measurement of multiplexed resonators in the subsequent section using process C to do further microwave spectroscopy from 1.5 - 4 K. In the temperature range 1.5-3 K, it is suggested the frequency and temperature dependence of Q (T, ?)?1i ? tan ?(T, ?) is due to two-level system loss, and is supported by changes in the dielectric constant (MTL phase velocity). At the end of this chapter, a discussion of the temperature dependent superconductor and dielectric microwave loss, and concluding statements will be made on how this affects performance and power efficiency of resonator clocks. 138 5.2 Temperature Dependent Resonator Measurements Temperature dependent measurements in this section were performed with the same test fixture described in Chapter 4, but retrofitted into a system similar to the Dry ICE 1.5 K Cryostat [182]. The cryostat base temperature is approximately 1.4 K and has temperature cryostat temperature stability of ?10mK from 1.4 to 10 K. It is important to note, this is a dry system and cooling is accomplished by means of thermal conduction. See Appendix C for more information. All temperature dependent measurements in this section were performed on MTL widths 0.25 ? 4 ?m from process B. An example of the temperature dependent data for 4 ?m MTL can be found below in Fig. 5.1 Four resonant modes from 3-12 GHz were analyzed extracting Qi(T ), Qe(T ), fn(T ) using methods from Khalil et al. [177] (see sections 4.2, 4.3). 139 Figure 5.1: Temperature dependent measurements of 4?m MTL from process B using Dry ICE 1.5 K Cryostat equivalent [182] with ?10 mK temperature control. S-Parameter measurements were with an input power of Pin = ?20 dBm. It will be shown this input power is too high and is shown only as an example that no non-linearities or bifurcation of resonant peak was found. Although, this relative high input power impacted the extracted internal Q-factors Qi. a) S21 showing the temperature dependence of 4 modes were captured. b) Zoom in of mode one showing the resonant frequency shift and widening of 3 dB width as temperature is increased. c) The complex S?121 , which were fitted using [177]. d) The phase where with resonant frequency ?/2? at zero phase. 140 The nominal input power from the vector network analyzer (VNA) to the feedline is Pin = ?20 dBm = 10 ?W at 4.2 K. It will be shown that -20 dBm is a relatively high power at 1.76 K. The effective power delivered to the resonator is dependent upon (i) the attenuation from the VNA to the feed-line on chip and (ii) the resonator coupling factor g. The coupling factor is dependent upon the resonator internal Q-factor Qi and external (coupling) Q-factor Qe by g = Qi/Qe (see Eq. 4.2). Since Q-factors are temperature dependent, the coupling factor is also temperature dependent g(T ), and therefore the effective input power to the resonator is temperature dependent. If it is desired to do temperature dependent measurements, the input power needs to be taken into consideration to prevent non-linear effects. Assuming the VNA is calibrated up to the shunt connection node, the incident power can be expressed in the following way by energy conservation [176] Pin = Pr + Pt + Pd (5.1) where Pin is the incident power in Watts at the input of the resonator after cable attenuation (at the reference node in Fig. 5.2), Pr is the reflected power to the VNA, Pt is the transmitted power, and Pd is the power dissipated by the resonator. The transmitted power Pt for a reactively coupled MTL can be calculated by [176] Pt = Pin|S 221| (5.2) where Pin is the incident power in Watts at the input of the resonator after cable attenuation, |S21| 141 Figure 5.2: a) Representation diagram of the open ended MTL resonator. The reference point for the power conservation Eq. 5.1 is the node between the two 50 ? lines and the coupling capacitor Cc. b) Block diagram showing the capacitively coupled ?/2 microstrip transmission line shown in a) is effectively an impedance inverter where Y is the shunt admittance. The node pointed out in a) is the same point in b) in reference to the power conservation Eq. 5.1 is the transmission coefficient magnitude related to the insertion loss IL in dB by |S | = 10IL/2021 at the resonant frequency ?/2?. In this case of a reactively coupled microstrip resonator, the IL is -6 dB for critical coupling g ? 1. Note, the typical insertion loss off resonance ranges from -15 to -5 dB due to cable loss and impedance mismatch. This measurement is self calibrated in the sense that |S21| is the size of the resonant dip relative to the baseline loss. For example, in Fig. 5.3(b) the 3.3 GHz resonant mode has an insertion loss of?8.5dB at 1.76 K. The transmitted power to the resonator is Pt ? 14 ?W for a -20 dBm input power. Assuming constant material properties, the transmitted power can be tuned by varying the coupling capacitor Cc, temperature, and geometry of the resonator (refer to chapter 4). From the energy conservation law for microwave circuits [10, 176], the dissipated power at resonance can be found regardless of the g value and calibration. The dissipated power can be 142 calculated by [10, 176] Pd = 2Pin(|S21| ? |S21|2) (5.3) where Pd is the dissipated power in Watts. In Fig. 5.1(b) at a nominal Pin ? 10 ?W , IL ? ?8.5 dB at 1.76 K yields a dissipated power of Pd ? 50 ?W . The circulating (stored) power in the resonator can now be calculated by Pst = QiPd = 2QiPin(|S21| ? |S 221| ) (5.4) where Qi is the internal Q-factor at resonant frequency ?/2?. Fitting the 3.3 GHz resonance at 1.76 K from Fig. 5.3(b) yields a Q 3i ? 1.3? 10 . The stored power is then Pst ? 18mW . A superconductor becomes non-linear when RF current density comes close to the temperature dependent critical current density Jc(T ). In fabricated superconducting electronic circuits, the minimum critical current density at T ? 4.2 K of Nb 0.35 ? 0.5 ?m metal lines with 200 nm thickness are typically Jc ? 20MA/cm2 [32]. Considering a worst case for a 4?m and 0.25?m, at nominal 200 nm thicknesses, the MTL minimum critical currents are approximately 160 mA and 10mA, respectively. The peak RF current in the resonator Ipeak can be compared to the expected critical current Ic to estimate onset of nonlinearity. It is also important to keep track of the peak voltage Vpeak to monitor electromagnetic strengths in both the superconductor and dielectric. The temperature dependent magnitude of the peak current Ipeak(T ) and voltage Vpeak(T ) can be calculated by 143 [183] ? I (T ) = 16 Pinpeak Pd(T )Qi(T ) (5.5)n? Z0 Vpeak(T ) = Z0Ipeak(T ) (5.6) where Pin is the input power, n is the mode index of the resonance, Z0 is the characteristic impedance of the MTL assumed to be constant at T ? 4.2K, and Qi(T ) and Pd(T ) are the temperature dependent internal Q-factor and dissipated power, respectively. The rms peak current ? is Irms = 2Ipeak. Referring back to Fig. 5.3(b), assuming a characteristic impedance Z0 ? 8 ? for 4 ?m MTL, the peak current is Ipeak ? 190 mA. This peak current of 190 mA exceeds the estimated Ic ? 160mA. For completeness, a 4?m MTL from Process B was measured at two input powers Pin of -20 dBm and -35 dBm and the temperature dependent results can be found below in Fig. 5.3. At 1.7 - 1.8 K, a clear difference in Qi as a function of frequency can be seen. The 6 and 9 GHz modes at T < 1.8K show almost equivalent Q?1i for -20 and -35 dBm, whereas the 3 GHz has a much higher loss with Pin = ?20dBm. Surprisingly, the 12 GHz mode for the lower -35 dBm power shows a factor of two increase in Q?1i relative to the 12 GHz mode at -20 dBm. This will be revisited. The non-monotonic behavior in Fig. 5.3 gives motivation to analyze the presence of non- linear effects. In Fig. 5.4, the peak current Ipeak was calculated from data in Fig. 5.3. The peak current Ipeak in the 4 ?m was calculated from Eqs. 5.5, 5.3 using measured |S21| and Qi at each resonance and temperature T . It can be immediately seen that mode 1 (yellow) Ipeak 0.5Ic,min for Pin = ?20dBm at the lowest temperatures. Mode 4 has peak current Ipeak 0.15Ic,min. For 144 Figure 5.3: Extracted Qi versus resonant frequency (mode index) as a function of temperature for a) Pin = ?20 dBm and b) Pin = ?35 dBm. No visual bifurcation of the resonant peaks were witnessed. Although, particularly at 1.7-1.8 K, there is a clear different in extracted Qi as a function of mode number an Pin = ?35dBm, the peak current drops to a factor of 10 and 20 lower for mode 1 and mode 4, respectively, relative to Ic,min. Note, looking vertically at each temperature, the peak currents in the resonator has an approximate monotonic dependence with mode index n, where n=1 is the 3.27 GHz mode. This is as expected and should go as I ?1/2peak ? n . Although, for Pin = ?35 dBm, the mode 3 peak current relative to mode 2 decreases by Irf,3 ? 1.7Irf,2 at 1.7 K, where a factor of 1.2 would be expected from mode dependence alone. Whereas the mode 4 decrease relative to mode 2 at 4.2 K is as expected with Irf,3 ? 1.7Irf,2. Referring back to Fig. 5.3, the larger decrease in Ipeak,3 for mode 3 seems to be coupled with a factor of 2 increase in Q?1i,3 relative to mode 2 at 1.7 K. With mode 2 and 3 peak currents more than an order of magnitude less than the critical current Ipeak,3 < Ipeak,2  Ic for Pin = ?35 dBm, the non- monotonic mode dependence of Q?1i in Fig. 5.3(b) suggests it is related to the voltage amplitude or some other mechanism This will be revisited towards the end of the section. 145 Figure 5.4: Calculated Peak currents in 4?m MTL as a function of temperature for 4 resonances at input powers (a) -20 dBm and (b) -35 dBm. The peak currents were calculated using Eqs. 5.5, 5.3, and measured transmission coefficient magnitude |S21| and Qi for resonance at temperature T . The characteristic impedance was calculated to be Z0 = 8.2 ? from HFSS measurements and known cross-sectional geometry. The red horizontal lines are a referenced expected minimum Ic ? 160mA estimated from Jc ? 20MA/cm2 [32]. It has now been established that the peak currents for all modes in 4 ?m MTL resonator at Pin = ?35dBm are much less than the critical current Ipeak,n  Ic for temperatures T = 1.5 ? 4.2 K. Fig. 5.5 below is a plot of the inverse internal Q-factor Qi at 1.7 K for each mode index and MTLs 0.25 - 4 ?m from process B. All mode 4 Q-factors Q?1i,4 have approximately the same magnitude for all MTL widths. Surprisingly, all mode 2 Q?1i,2 have the same magnitude. Conversely, modes 1 and 3 vary significantly by as much as a factor of 5. The same peak current and voltage analysis will be performed to determine if this is driven by an abnormal stimulus dependence. Fig. 5.6 shows the Qi as a function of calculated peak currents Ipeak for process B 0.25 - 4 ?m MTLs at 1.7 (a) and 4.2 K (b). At 4.2 K, all MTLs except 0.25 ?m yield peak currents at least a factor of two less than Ic. This suggests that 0.25 ?m MTLs should be measured with 146 Figure 5.5: Qi as a function of mode index at 1.7 K and Pin = ?35 dBm for MTL widths 0.25? 4?m. Using fitting methods described reference [177], complex S21 fitting errors for Qi, Qe, and f0 were less than 5% for all data shown here. Careful analysis of the raw data was performed, and no non-linear effects (e.g. resonant peak bifurcation) were found. 147 input powers Pin ? 35 dBm. At 1.7 K, the 4 and 2 ?m MTLs have peak currents much less than the estimated critical current Ipeak,1, Ipeak,2  Ic. The 0.5 ?m MTL peak currents come close and even cross over the estimated critical current, and could explain the decrease of the Q-factor for the fourth mode Q?1i,0.5?m ? 3 ? 10?3 and peak current Ipeak ? 17 mA Similarly, for the 1 ?m MTL a similar increase in Q?1i,1?m can be seen at the same Ipeak ? 17 mA. This could also be explained by 1 ?m Ipeak within < 2Ic of its critical current, inducing an additional loss due to some weak link at the peak current nodes in the resonator. Conversely, the 0.25 ?m MTL resonant modes have peak currents residing up to 2Ic, but the Qi increases at increasing peak currents. In summary, the high peak RF currents for all MTL resonator widths cannot explain the non-monotonic mode (frequency) dependence ofQi alone. As briefly suggested, the peak currents occur at specific locations along the resonator, and could be exacerbating weak-links or material defects in the Nb wires or ground plane. This also goes for material defects in the TEOS dielectric. In light of this, the peak current/voltage technique and analysis could be used to accentuate nonlinear properties of the superconductor or dielectric losses due to spatially varying extrinsic material defects. Furthermore, the peak current analysis proved useful to determine an input power of -20 and -35 dBm was was too high to analyze the intrinsic superconductor and dielectric RF losses as a function of temperature and frequency with the current MTL resonator design. Regretfully so, these large input powers relative to the high coupling of the current MTL design was known, and measuring at lower powers posed difficulty in signal-to-noise in the current test fixture. This facilitated the design of a multi-plexed MTL design to further analyze RF losses in RQL interconnects at temperature T = 1.7 ? 4.2 K, and is the topic of the next section. 148 Figure 5.6: Calculated peak currents Ipeak at (a) 1.7 K and (b) 4.2 K for MTL widths 0.25?4?m and mode indices 1-4 with respectiveQi(?n). The vertical lines are the assumed minimum critical currents Ic based on a critical current density Jc = 20MA/cm2 of superconducting Nb wires at 4.2 K [32]. The input power was Pin = ?35 dBm for all measurements and data shown here. 149 5.3 Multiplexed Microstrip Transmission Line (MUX-MTL) Resonators As seen in the last section, there was an anomalous mode dependence of the internal Q- factor Qi. Especially at the lowest measured temperatures T < 2 K. In this section, I will introduce a different resonator design with more resonances from 4-10 GHz to investigate the frequency and temperature dependence with better certainty. Rather than measuring a single ?/2 resonator, the new design has many microstrip transmission line resonators that are capacitively coupled to a single feedline. This multi-plexed MTL resonator design will be termed MUX-MTL. Multiplexing resonators designs are typically used to measure many sensors at once, e.g. SQUID multiplexing [184, 185]. Here, the MUX-MTL is designed to measure the intrinsic RF losses in the Nb and TEOS. Conveniently, the frequency dependence of the material properties can be measured since each resonator needs to have a unique resonant frequency. A layout of the MUX-MTL resonators can be found in Fig. 5.7. Here, the MUX-MTL highlighted in Fig. 5.7c will measured. These MTLs are coupled with coupling capacitors having the lowest value of Cc ? 25 fF . This yields a resonator coupling low enough to ensure (i) a high enough signal-to-noise (resonant peak size) at an input power of ?35 dBm and (ii) transmitted powers to the resonator low enough to avoid peak current/voltage effects discussed in the last section. Fig. 5.7d shows a zoom in of a single MTL capacitively coupled (orange) to the MUX feedline (blue horizontal line running across the bottom). The feedline and MTL have a line width of w = 0.75 ?m. As opposed to MTLs in Chapter 4, each MTL here coupled to the MUX feedline are ?/4 resonators grounded at the end. A 7 mm ?/4 MTL has a fundamental mode of approximately 4 GHz. The grounding areas are enlarged to allow for current spreading. 150 Additionally, the feedline and ?/4 MTL have TEOS dielectric thicknesses of 500 nm with nominal 200 nm Nb ground plane and conductor strip thicknesses. A single MUX-MTL will be analyzed here and was fabricated using process C. The pertinent resonator information and results are summarized in Table 5.2. Figure 5.7: Physical design of a MUX-MTL resonator showing a) a diagram and the b) physical layout of a single MUX-MTL resonator. There are 21 ?/4 MTL resonators on a single MUX feedline with resonator lengths ranging from 3-7 mm with a 0.2 mm step. There are c) 8 MUX- Resonators designed on a single 5x5 mm chip. d) A zoom in of the first ?/4 MTL resonator connected to the MUX-Feedline coupled through a coupling capacitor with nominal capacitance of Cc = 25fF . The MTL and feedline width is 0.75 ?m. The end of the MTL is shorted to GND with a wire fan out and many 0.25 ?m vias to prevent the wire going normal at the current node. Via walls surround the feedline and MTL for isolation. The MUX-MTL resonators were measured in the same way MTLs were measured and analyzed in Section 5.2. The resulting temperature dependent S-parameter measurements can be 151 Figure 5.8: a) The measured S21 versus frequency as a function of temperature b) Cross-section diagram of the MUX-MTL geometry. The 21 resonances pertain are the fundamental modes of ?/4 MTLs with lengths ranging from 4-7 mm increasing in 0.2 mm increments. found in fig. 5.8(a), and (b) is a cross section diagram of the MUX-MTL resonator. All resonances in Fig. 5.8 were analyzed using an automated python script. The internal Q-factors Qi, external (coupling) Q-factors Qe, and resonant frequencies f0 were extracted from a complex circle fitting script following the procedure in reference [177] 5.4 Using the Dispersive Deconvolution Method to Simultaneously Measure tan ?(T ) and Rs(T ) In this section, I will apply the dispersive loss deconvolution (DLD) method presented in Chapter 3 and applied in Chapter 4 to extract the temperature dependent loss tangent and intrinsic resistance simultaneously. Fig. 5.9 shows the resulting fitted inverse internal Q-factorsQ?1 of MUX-MTL ?/4 resonators 152 Figure 5.9: Frequency dependence of the inverse internal Q-factor Qi at successive temperatures from 1.6-4.5 K at a constant input power of Pin = ?35 dBm. Complex S21 fitting procedure was performed on data from Fig. 5.8. The 0.75 ?m MUX-MTLs were fabricated using process C described in Chapters 3 and 4. from S-Parameter measurements in Fig. 5.8. At 1.6 K, a non-monotonic frequency dependence of Q?1i can be seen, and will be discussed in the next section. It is important to first discuss the measurement sensitivity of the DLD method before fitting the data in Fig. 5.9. For parallel plate resonator measurements, it was found by Taber et al. that the measurement sensitivity of the intrinsic resistance was Rs0 ? 5 ?? [136], and he needed to vary the resonator geometric factor to get accurate measurements of Rs0. To estimate the DLD measurement sensitivity of Rs0 and tan ? using MTL resonators, Eqs. 3.12 and 3.13 from 153 Section 3.3 can be re-written as follows Rs0(T ) Q?1i,nd(T ) = (?/?0) + tan ?0(T ) (5.7)?c0 ?1 Rs0(T )Qi,d (T ) = (?/?0) + tan ?0(T )(?/?0) b?1 (5.8) ?c0 Rlims0 (T ) ? ?c0 tan ?0(T ) (5.9) where ? is the measurement frequency, ?c0 is partial geometric factor of the MTL, b is the free fitting power exponent assuming a dispersive loss tangent, and Rs0(T ) and tan ?0(T ) are the intrinsic resistance and loss tangent at reference frequency ?0 at temperature T , respectively. Eq. 5.7 can be used to fit data when a non-dispersive loss tangent can be assumed. Eq. 5.8 is used in this work to fit data assuming a dispersive loss tangent where 0 < b < 1. Eq. 5.9 is an estimate of R = Rlims0 s0 measurement sensitivity at reference frequency ?0 from reference [136]. Similarly, the approximate measurement sensitivity for the loss tangent is tan ?lim0 ? Rs0/?c0 In other words, when the Rs0 and tan ? are comparable for a given resonator geometry, it is important to utilize the frequency dependence, temperature dependence, or some other means to deconvolve the two loss mechanisms. The latter case was demonstrated in the previous chapter. In this section, S-parameter measurements are performed down to 1.6 K, and it is expected the Nb Rs0 will exponentially decrease Rs0 ? (1/T )exp(?1/T ) (see Eq. 2.27). Assuming TLS relaxation proccesses dominate the dielectric loss below 4.2 K, the TEOS tan ?0 is expected to decreases with ? T 3 dependence (see Section 2.3). At 4.2 K, the TEOS loss tangent is tan ? ? 1 ? 10?3. The geometric factor for the 0.75?m MUX-MTL is approximately 17.3 m?. The approximate sensitivity ofRlims0 at ?0 and 4.2 K is then 17 ??. Taking advantage of the conductor 154 Q-factor frequency dependence Q?1c ? ?, the intrinsic resistance can still be deconvolved at 4.2 K using the DLD method, as the losses are still comparable. Although, at temperatures far below 4.2 K, the case Rlims0  ?c0 tan ?0 will yield Qi ? tan ?, and vice versa when temperature is much greater than 4.2 K Qi ? ?c0/Rs0. The DLD method is applied to data in Fig. 5.9 and Rs0(T ) and tan ?(T ) are deconvolved and shown in Fig. 5.10(a) and Fig. 5.10(b), respectively. The x-axis in Fig. 5.9(a) is Tc/T where at superconducting critical temperature of Tc = 9.2 K was assumed. Both a disperive (black) and non-disperive (blue) tan ? were assumed using Eq. 5.7 and 5.8, respectively. Note, the DLD method was only applied to the resonant peaks between 4.5-5.5 GHz and 8-10 GHz, ignoring the ?loss peaks? at 6.5 and 9.5 GHz. These loss peaks will be discussed in the next sections. The dashed lines in Fig. 5.9(a) are calculated using Eq. 5.9 and serve as a reference for the approximate measurement sensitivity of Rs0 as a function of temperature. At Tc/T = 3 for the non-disperive data (blue), the Rs0 is approximately a factor of two below the sensitivity limit R lim,nd lim,ds0 ? 2Rs0 . Whereas the dispersive Rs0 (black) is only about 10% below Rs0 for non-dispersive (blue) dashed line and approximately a factor of 3 above the dispersive (black) dashed line. This suggests that tan ? is frequency dependent and has a power exponent of b ? 0 yielding tan ? ? ??1. Furthermore, the dispersive model is suggested to be more sensitive to Rs0 in the approximate temperature range Tc/T = 2? 3. The Rs0 plateau at approximately 3?? is an artifact of the measurement sensitivity, and is not a residual resistance. Typical residual resistance for Nb thin films are Rres ? 1? 20 n? and do not exceed 100 n?. The expected temperature dependence of Rs0 was fit to the dispersive data using Eq. 2.27 with fitted Nb superconducting gap of ? = 19.5 ? 2.2 K. Benvenuti et al. measured a superconducting gap of ? = 19.6K for Nb thin films [64]. Assuming a superconducting critical 155 temperature of T = 9.2 K, the gap ratio is estimated to be ?/kBTc = 2.1 ? 0.2, and is within 10% of the expected gap ratio 1.97 for Nb [46]. Valente-Feliciano measures a slightly lower gap ratio of 1.89 for Nb thin films [71]. Fig. 5.9(b), the assumed dispersive (black) and non-dispersive (blue) tan ? temperature dependence was fit to a power law expected when dielectric loss is dominated by TLS relaxation (see Eq. 2.37). Using the dispersive model fit to deconvolved TEOS tan ?, the power exponent is n = 2.5. Microwave dielectric measurements of amorphous silica have found T 2.4 from 2-8 K in the TLS relaxation regime at 10 GHz [84]. TEOS deposited at low temperatures of 150 C is considered amorphous [74]. This is the first indication that TLS relaxation processes are present in TEOS at temperatures below 3 K and will be discussed further in the next section. In this section, I have demonstrated the DLD method is capable of deconvolve Rs and tan ? as a function of temperature. From this, I estimated the the superconducting gap of the Nb in the MUX-MTL in good agreement with typical Nb thin films. Additionally, from the fitted temperature dependence i found that the TEOS tan ? shows signs of TLS relaxation. 156 Figure 5.10: The deconvolved a) Rs0 and b) tan ? as a function of Tc/T using the DLD method at reference frequency ?0/2? = 10GHz. The critical temperature is assumed to be Tc = 9.2 K. Data from Fig. 5.9 were fit to Eq. 5.7 (2 parameter fit) and 5.8 (3 parameter fit) assuming a non- dispersive and dispersive (frequency dependent) loss tangent, respectively. The DLD method was only applied to the resonances between 4.5-5.5 GHz and 8-10 GHz, ignoring the ?loss peaks? at 6.5 and 9.5 GHz. The dashed lines in a) are calculated using Eq. 5.9 showing the approximate sensitivity of the intrinsic resistance for the dispersive Rlim,ds0 and non-disperive fits R lim,nd s0 . The solid lines in a) are fits to the Rs0 temperature dependence using Eq. 2.27 in the range Tc/T = 2.2 ? 2.5. The solid lines in b) are fits to the tan ? temperature dependence using Eq. 2.37 for the entire temperature range. In a) the Rs0 plateau in at approximately 3 ?? is an artifact of the measurement sensitivity, and is not a residual resistance. 157 5.5 TLS Absorption and Relaxation Microwave Loss in TEOS Above 1K In this section, I will claim signatures of TLS loss in TEOS at temperatures below 3 K supported by measurements of the the relative dielectric constant ?r and dielectric loss tangent tan ?. Furthermore, as temperature is decreased, there is a clear transition from TLS relaxation to TLS absorption dominating the RF loss at approximately 2.25 K. This temperature transition as a non-monotonic frequency dependence. Below is a summary table for the resonant frequency dependent measurements and fitting to TLS models. The electric perturbation of a two-level systems (TLSs) interacting with the phonon lattice is analogous to the acoustic case for [79, 83]. The Hamiltonian matrixH = H +H ?0 describes the system where H0 is TLS without perturbation and H ? is due to the electric field. The magnitude of the diagonal and off-diagonal elements in H ? are the permanent dipole moment ? and induced dipole moment ??, respectively. The permanent dipole moment describes the TLS relaxation process expressing the shift in energy splitting caused by the electric field and is characteristic of a TLS relaxation process. The induced dipole moment describes the resonant absorption of TLSs by the coupling between the TLSs and electric field. The latter TLS resonant absorption process can be saturated by strong electric fields and dominates at temperatures approximately below 1K. Conversely, the former relaxation process cannot be saturated by an electric field and is the dominant process at temperature approximately above 1K [79]. Consequently, depending on the temperature, applied electric field, and the angular frequency ?, a non-monotonic dependence of the relative dielectric constant ?r and loss tangent tan ? can occur (see Fig. 2.4 in Chapter 2). Note, for a weak electric field as temperature is increased from 0 K, the cross over temperature will shift to increasing (decreasing) temperatures as ? is increased (decreased), reminiscent of 158 an increased (decreased) sampling of higher energy resonant absorption TLSs prior to relaxation dominating. Both the change in dielectric constant and loss tangent data will be fit assuming absorption and relaxation TLS processes. The temperature dependent change in relative dielectric constant ? assuming the superposition of TLS absorption and relaxation contributions is typically written as [186] ??(T ) ?(T )? ?(T0) ??abs + ??rel = = (5.10) ?(T0) ?(T0) ?(T0) where ?(T ) is the temperature dependent relative dielectric constant of the material, T0 is typically the lowest measured temperature and is arbitrary, and ??abs and ??rel are the change in dielectric constant due to TLS resonant absorption and relaxation, respectively. The relative dielectric constant of TEOS is approximately ? ? 4.2 [73] and agree with estimations from measured data in Section 4.2. At low temperatures, the resonant absorption contribution to the change in dielectric constant can be fit to [ ( ) ( ) ] 1 h?? 1 h?? T ??abs(T ) = 4? n? ? 2 Re? + ?Re? + ? ln (5.11) 2 2?ikBT 2 2?ikBT0 T0 where n?? 2 is the induced dipole coupling of TLSs having a density of states n, T and T0 are the measurement and lowest arbitrary reference temperature, respectively, kB is the Boltzmann constant, h? is the reduced Planck constant, and Re? represents the real part of the digamma function [187]. Note, the first and second terms in brackets dominant at temperature  1 K and a frequency dependence can be measured. For measurements here where T > 1.5 K, 159 the natural natural logorithmic temperature dependence dominates and can be assumed to be frequency independent. Here, we assume the relaxation process is mediated by the so-called single-phonon assisted tunneling as the TLSs relax back to equilibrium. This is also commonly referred to as the one- phonon process and is in opposition to coherent unassisted tunneling motion[79]. Assuming a constant density of states n, the TLS relaxation process responsible for the change in dielectric constant can be modeled by ? 4? ? df ( )?1 ??rel = n? 2 1 + ?2? 2 dU (5.12) 3 dU 0 ??1 = a(UkBT ) 3 coth (U/2) (5.13) ( ) U ?1f = e + 1 (5.14) where n?2 is the permanent dipole coupling, f is the Fermi function, ? is the measurement frequency under an applied electric field, and ? is the TLS relaxation time commonly referred to as the radiative lifetime of TLSs T1 [79, 79]. Found from acoustic measurements the material constant a is M?2 1 a = (5.15) ?v5a 2?h? 4 M?2 = (M2/c5 + 2M2/c5)/(c?5l l t t l + 2c ?5 t ) (5.16) where M? is the average coupling energy [188], va is the average acoustic velocity of two transverse sound velocities and one longitudinal sound velocity, ? is the mass density of the dielectric, and 160 subscripts l and t in Eq. 5.16 are the longitudinal and transverse components, respectively. For vitreous silica, M? ranges from approximately 0.28 ? 0.34 eV [188], ? ? 2.2 ? 103kg/m3, and va ? 4.7 ? 103 m/s averaging the transverse and longitudinal components. Pohl et al. gives a comprehensive review of these material constants for various glassy metals and dielectrics [93]. For glassy silicon oxides, ? and vt,l can decrease by approximately an order of magnitude for high porosity. The measured resonant frequency for reactively coupled microstrip transmission lines is proportional to the relative dielectric constant by f ? ??1/20 . Using this relationship and Eq. 5.10, the temperature dependent resonant frequency data in Fig 5.9 can be converted by ??(T ) f 20 (T )? f 2= 0 (T0) (5.17) ?(T0) f 20 (T0) where f 20 (T ) is the temperature dependent resonant frequency of each resonator in the MUX- MTL, and f0(T0) resonant frequency at the lowest temperature measured T0. Fig. 5.11 shows the resulting fits to Eq. 5.10 to data from Fig 5.9. The data was fit with a Python script using the ?Powell? minimization method in SciPy package [189]. The data was fit at each data power where the measured resonant frequency and temperature were inputs. The dielectric constant at T0 was assumed to be ? = 4.2 [73] and the material constant in Eq. 5.15 was assumed to be a = 2? 1075 J?3s?1. The two fitting parameters were the permanent and induced dipole coupling constants n?2 and n?? 2 in equations Eq. 5.11 and Eq. 5.12, respectively. The dip in the data is reminiscent of a transition from TLS resonant absorption to relaxation processes dominating the dielectric constant change as temperature is increased from T0. The magnitude offset for the 7GHz (4.4 mm) resonator in the MUX-MTL is due to a slightly higher T0. Because 161 of this, the fitted data for the 7GHz resonator will be ignored due to minimization fitting artifact. In this context, the maximum change in ??(T )/?(T0) is negative showing a slight decrease of the dielectric constant of approximately 0.06%. Surprisingly, this value is in agreement with Schickus et al. of vitreous silica with 1000 ppm OH- concentration [186]. Although, his maximum change in dielectric constant occurs at a much higher temperature of approximately 7 K, whereas in Fig. 5.11 occurs at approximately 2.2 K. Starting from 3 K and decreasing in temperature and following the color scale, the relaxation regime shows almost a monomonic frequency dependence, where for increasing frequencies the dielectric constant is effectively decreasing. At and below the dip around 2.2 K, a non-monotonic frequency dependence of dielectric constant change is visible. This will be probed further in the fits of n?2 and n?? 2. Fig. 5.12 is a plot of the data fitting (solid lines) from Fig. 5.11. Again, the 7 GHz data here will be ignored due to a fitting artifact. Immediately it can be seen that the permanent dipole coupling is 3 orders of magnitude larger than the induce dipole coupling (notice scale bars on left and right y-axis). This is not understood and reported values for all permanent dipole coupling constants for glassy dielectrics do not exceed 1 ? 10?3 [89, 93, 188]. Table 5.1 is a summary of measured permanent and induced dipole coupling for various materials. From this table, it can be seen that the permanent and induced diople coupling increases in vitreous silica by approximately an order of magnitude when the concentration of hydroxide concentration is increased. This is supported by the theory where hydroxide ions are the dominant TLS system coupling to phonons and the electromagnetic waves [186]. The TLS states in borosilicate glass have a higher permanent dipole coupling to phonons relative to vitreous silica [89]. Although predominantly a crystalline solid electrolyte, it is assumed that the existence of low-energy excitations in Na? ? ? Al2O3 is a consequence of a high degree of disorder in the Na+-sublattice 162 Figure 5.11: Temperature dependence of the dielectric constant ? for 0.75?m MUX-MTLs fabricated from process C at a constant VNA input power of Pin = ?35 dBm. The data from Fig 5.9 was converted into ?? = ?(T )??(T0)/? using Eq. 5.17. The data was fit using Eq.s 5.10, where the total change in dielectric constant is the superposition of the TLS absorption and TLS relaxation contributions using Eq.s 5.11 and 5.12, respectively. assuming material constant a = 2? 1075J?3s?1 [186]. Refer to Summary Table 5.2 163 and is considered amorphous [94, 190]. Furthermore, Na? ? ? Al2O3 has the highest induced dipole coupling with the electromagnetic waves and most likely due to its ionic nature. Li3N has a much lower induced dipole coupling and the TLS low energy excitations are suggested to be coming from hydrogen ions [190]. The permanent dipole coupling for Na? ? ? Al2O3 was not calculated, but based on the high induced dipole coupling, it is possible the permanent dipole coupling can be of order 10?2 but it is unlikely the coupling is as high as 10?1, as seen by the extract values measured for TEOS in Fig. 5.12. Nevertheless, this can be further examined in future experiments at lower temperatures and power dependent measurements as performed by Schickfus et al. [84]. Table 5.1: A summary table of measured permanent dipole coupling n?2 and induced dipole coupling n??2 in various materials at the respective measurement frequency. The values were taken from references [Schickfus1976] [186], [SchickfusThesis1977] [89], [Strom1978] [94], [Baumann1980] [190]. Care was taken in the evaluation of the integral in Eq. 5.12, and the material constant a is not expected to change orders of magnitude. Furthermore, if this were real, the 3 orders of 164 magnitude increase would have to be comming from the permanent dipole moment, as the TLS density of states n cannot increase beyond approximately an order of magnitude. Aside from the magnitude difference, the results in Fig. 5.12, the data can still be discussed qualitatively. The permanent dipole coupling increases with increasing frequency, which is in agreement with the notion that higher frequencies accesses higher energy insatiable TLS relaxation processes. The induced dipole moment decreases with increasing frequency, and is also in agreement with intuition of TLS absorption in that higher energy absorption processes contribute to the change in phase velocity (dielectric constant) in the resonator. Another way to describe the decrease in induced dipole coupling is that at higher frequency, it is expected that TLS absorption can dominate at higher temperatures [89]. The magnitude discrepancy for fitted values for n?2 measured via temperature dependence resonant frequency will be investigated further and compared with temperature dependent data of the dielectric loss tangent. 5.6 Dielectric Loss Spectroscopy In this section, the dielectric loss tangent will be fit to TLS resonant absorption and relaxation models to determine the frequency dependence of the permanent and induced dipole coupling constants n?2 and n?? 2, respectively. The data fitting models will be introduced first, followed by a suggested explanation of the anomalous loss peaks in Fig. 5.9 at 6.5 and 9.5 GHz. As a consequence of the TLS tunneling and relaxation process scattering phonons, a non- monotonic temperature dependence of the dielectric loss tangent tan ? can be measured and depends on the microwave electric field strength [84]. 165 Figure 5.12: Frequency dependence of the permanent and induced dipole coupling from the data fits in Fig. 5.11. Refer to Summary Table 5.2 for assumed parameters and results comparison to tan ? fits. MUX-MTL, Process B, Q?1i (T ) 166 The superposition of TLS processes tan ? can be written as tan ? = tan ?abs + tan ?rel (5.18) where tan ?abs and tan ?rel are the resonant absorption and relaxation contributions, respectively, and here I am assuming these are the only processes involved in the microwave loss in the temperature range T = 1.5? 3K and frequency range 4? 10GHz. The measured loss tangent from resonant absorption can be fit to the following tanh h??4? tan ? = n?? 2 2kBTabs (5.19) 3 (1 + J/J 1/2c) where n?? 2 is the induced dipole coupling constant, J is the microwave energy intensity, and Jc is the critical energy intensity. For J > Jc, resonant TLSs undergoing resonant absorption are saturated and tan ?abs continually decreases as temperature is decreased. For J < Jc TLS resonant absorption processes induce loss and an upturn in tan ? is observed as temperature is decreased to absolute zero [84]. For microwave dielectric loss dominated by TLS relaxation, data can be fit to the following equation [89] [ ] T 3 M?2 ?(4, 0.5)k3 tan ?rel = n? 2 B (5.20) ? ?v5 4a 8?h? where n?2 is the permanent dipole coupling constant, ?(4, 0.5) = 97.8 is the Riemann zeta function [89], ? is the mass density of the dielectric, M?2, va are the average coupling and acoustic velocity introduced in Eqs. 5.15-5.16, respectively. 167 Before fitting to the above models and extracting n?2 and n?? 2, the power law fits to tan ? data in Fig. 5.13 will be discussed. The resulting power law fits in Fig. 5.13(b) show the power exponent is close to 3, strongly suggesting the TEOS dielectric loss is dominated by TLS relaxation processes. The power exponent measured for vitrious silica is approximately 2.4 [84]. Power exponents below 3 are typical, as the curvature of the temperature dependence can change rapidly as a consquence of either TLS absorption at lower temperatures or classical relaxation processes at higher temperatures inducing the so called ?plataeu? effect [93]. Using Eqs. 5.18-5.20 above and data in Fig. 5.13, the TEOS dielectric loss tangent is fit for MUX-MTL resonators and shown in Fig. 5.14. The fitting results of temperature dependence in Q?1i (T ) and resonant frequency f0(T ) can be found in Summary Table 5.2. Notice, the data (circles) and fits (solid lines) show non-monotonic frequency dependence looking vertically at a particular temperature below 2.5 K. Assuming an onset of TLS resonant absorption below 2.5 K, it can be inferred that different frequencies are coupling stronger to resonant TLSs. The converse argument is that each resonator is coupling non-monotonically to the feedline, inducing different power densities circulating in the resonator. The latter will be explored in the next graph followed by frequency dependence of dipole coupling constant fits from Fig. 5.14. Fig. 5.15 shows the calculated stored power and peak currents in the MUX-MTL resonators. It can be seen quantitatively that the circulating power in the resonators scale monotonically with frequency at all temperatures. This is supporting the claim that the non-monotonic dependence in Fig. 5.14 is not an artifact of the variation in resonator coupling. Although, the peak currents in the resonator for the lowest frequency resonators are less than a factor of two away from the assumed worst case critical current for these lines. It is conceivable that this could induce non- 168 Figure 5.13: a) Temperature dependent inverse internal Q-factor of MUX-MTL resonators. At these temperatures, It is assumed that the Q-factor is dominated by the TEOS dielectric loss tangent by Q?1i = tan ?. Each color is a different resonator coupled to the MUX feedline with resonant frequency listed in the legend. The dashed lines are power law fits in the temperature range 2.5 to 3.5 K. b) Resulting power law fits for each resonator in the MUX-MTL including one sigma error bars for each fit. These resonators were fabricated using process C and constant VNA input power Pin = ?35 dBm. This is the same data from Fig 5.9. 169 Figure 5.14: Temperature dependence of the inversive internal Q-factor Q?1i for each MUX- MTL resonator with unique fundamental resonant frequencies at a constant input power of Pin = ?35dBm. At these temperatures, It is assumed that the Q-factor is dominated by the TEOS dielectric loss tangent by Q?1i = tan ?. The solid lines are fits to Eqs. 5.18-5.20 with n?? 2, n?2, and J/Jc as free fitting parameters. A summary of results and assumed constants are listed in Summary Table 5.2. 170 Table 5.2: A summary table of temperature dependent fitting results to models of TLS absorption and relaxation. At these temperatures, It is assumed that the Q-factor is dominated by the TEOS dielectric loss tangent by Q?1i = tan ?. The Q ?1 i (T ) and F0(T ) data were fit with 3 and 2 free fitting parameters to Eq.s 5.18 and 5.10, respectively. The constants were assumed from references [Kim2001] [72], [Pohl2002] [93], [Jackle1976] [188], [Rocha2004] [73], [Schickfus1976] [186]. The constant a assumed in Q?1i (T ) fits was calculated from the other assumed constants in the column using Eq. 5.15. Note, the extracted values of n?2 from resonant frequency fits are 3 orders of magnitude larger than typical values of amorphous dielectrics [Phillips1981] [89] linear enhancement of quasiparticle loss in the Nb. Although, the peak current qualitatively scales mono-tonically with frequency, where the peak current is decreasing for increasing frequencies. 171 This suggests that a similar monotonic dependence would be seen in the internal Q-factor. This is by no means a rigorous argument and should be explored further by doing power dependent measurements. Although, it is also worth noting that since the peak voltage is linearly related to the peak current by Vpeak = Z0Ipeak where Z0 ? 50 ? is the characteristic impedance of the MTL resonators. The reduction in peak voltage as frequency increased could be giving rise to the non-monotonic dependence of the loss tangent in the TEOS. This is postulated to be due to the expected increase in tan ? when voltage is increased, characteristic of TLS resonant absorption processes. Admittedly, during the time of these measurements, it was not expected that TLS mechanisms would be present nor was the design peak current incorporated as a design parameter. The subject of the original design was to investigate the coupling capacitor values for the smallest capacitors, inadvertently producing weak enough coupling, and circulating power to see unsaturated TLS resonant absorption loss. The resulting 3-parameter fits in Fig. 5.14 are plotted as a function of frequency in Fig. 5.16b and Fig. 5.16c. Fig. 5.16a is re-plotted data from Fig. 5.9 for convenience. The permanent dipole coupling constant n?2 (black) increases for increasing frequencies by a factor of 3 from 4-10 GHz. This agrees qualitatively with values fitted from temperature dependent changes in the dielectric constant in Fig. 5.12. Although, the magnitude n?2 ? 3 ? 10 ? 10?3 in Fig. 5.16b is more agreeable with published data on various amorphous dielectrics. The values of n?2 for TEOS measured here are of order values measured for borosilicate glass 6 ? 10?3 [89], and approximately an order of magnitude higher than vitreous silica for the lowest concentrations of OH? [186]. Similarly, the range of extracted induced dipole coupling n?? 2 = 4 ? 8 ? 10?4 in Fig. 5.16 also are in fairly good agreement with borosilicate glass having 3 ? 10?4 found in Schickfus thesis reported in [89]. 172 Figure 5.15: Temperature dependence of a) the stored circulating power Pst b) peak current Ipeak in the MUX-MTL resonators for VNA input power of Pin = ?35dBm. The stored power in each resonator and each temperature was calculated from Eq. 5.4. The peak currents were calculated using Eq. 5.5. The solid red line in b) is the minimum critical current Ic of a 0.75 ?m wide Nb wire with a 200 nm thickness assuming critical current density Jc = 20MA/cm2 [32]. 173 It is an important distinction to note that measured dipole moment coupling in microwave dielectric experiments are much more sensitive to impurities and the chemical composition of insulating glasses relative to acoustic measurements. Therefore, it is claimed that absolute values of n?? 2 and n?2 plotted in Fig. 5.16 may be off by factors of 2-3 due to assumed constants in Table 5.2, but the the qualitative frequency variation n?? 2 is claimed here to be due to coupling of TLSs with the electric field at distinct frequencies. In other words, the peaks of n?? 2 at 6.5 and 9.5 GHz in Fig. 5.16a align with the loss peaks in Fig. 5.16b. Furthermore, these peaks align with relatively low J/Jc ratios in Fig. 5.16c, suggesting that the loss peaks are affiliated to enhanced TLS resonant absorption loss. The maximums of J/Jc at 8.5 GHz shows a saturation of circulating power to TLS resonant absorption, reducing Q?1i . An interesting data point is the 7 GHz point, where there is an increase in both the induced coupling to TLSs and the circulating power exhibiting an overall constant Q?1i relative to the neighboring frequencies. This demonstration shows a new microwave loss spectroscopy method to probe TLS loss by relying on the frequency dependence. In future work, better care can be taken with the coupling design as well as systematic input power dependence to further explore this technique and the exhibited TLS character relative to the dielectric composition and impurities. 5.7 Temperature Dependence of the Power Dissipation in Nb and TEOS The primary goal of this dissertation is to glean information on how to reduce intrinsic microwave losses in Nb and TEOS to reduce the power dissipation in RQL resonator clock networks. This chapter has focused on demonstrating techniques to better understand the dominate loss mechanisms coming from the superconductor and dielectric materials as temperature is 174 Figure 5.16: a) Frequency dependence of Q?1i for MUX MTL resonators for constant VNA input power Pin = ?35 dBm in the temperature range T = 1.6 ? 4.5 K (Re-plotted Fig. 5.9). b) Frequency dependence of the fitted permanent (black) and induced dipole (blue) coupling constants. c) Frequency dependence of the power density ratio J/Jc where J and Jc are the circulating input power density in the resonator and the critical power density for TLS resonant absorption saturation, respectively. The fitted parameters in b) and c) were fit to data in Fig. 5.14 using Eqs. 5.18. A summary of results and assumed constants are listed in Summary Table 5.2. 175 decreased below 4.2 K. This chapter will be concluded with direct measurements of power dissipation in MUX-MTL resonators that can be directly related to RQL interconnects having the same geometry and input power. Furthermore, estimates can be made from this data for arbitrary geometries and input power. Strong et al. measured the dissipated power of meta-material zeroth-order resonators (ZORs) used in RQL clock networks and found a dissipated powers on the order of 20 ?W [10]. Using Eq. 5.3, the temperature dependent dissipated power Pd of MUX-MTL resonators at Pin = ?35 dBm is plotted in Fig. 5.17. In Fig. 5.17 at approximately 4-4.5 K, the dissipated power for all resonators (frequencies) is approximately 100?200nW for VNA input power Pin = ?35dBm. For a nominal input power Pin = ?9 dBm used in ZOR resonators, the dissipated power in MUX-MTLs is approximated to be 60 ?W , and is within a factor of 3 of ZOR power dissipation. As temperature is decreased, the 4.5GHz resonator reduces in dissipated power down to Pd = 70 nW at 1.6 K. At 1.7 K, the dissipated power decreases as frequency is increased up to 10 GHz with a minimum dissipation of Pd = 26 nW . The frequency dependence of this dissipated power is due to the saturation of TLS resonant absorption loss at higher frequencies and discussed in the previous section. This is an expected and known result, and further supports the notion that the TEOS dielectric microwave loss dominates the ZOR resonator clock power dissipation for the temperature range 1.5-3.5 K. There is a subtle maximum power dissipation at low frequencies around 3.5 K where the Nb and TEOS microwave losses are comparable for this resonator geometry, and can shift depending on the material used and the RQL clock geometry. Overall, this has established the limits to power efficiency in resonator clocks due to material losses and demonstrates a method of characterization. 176 Figure 5.17: Temperature dependence of the dissipated power Pd in MUX-MTL resonators at a constant VNA input power Pin = ?35 dBm. The dissipated power was calculated from data in Figs. 5.15 and 5.8 using Eq. 5.3. 177 5.8 Summary At the beginning of this chapter, the temperature dependence of the resonant frequency fn and internal Q-factor Qi in the range 1.5-4.5 K was measured for MTL widths 0.25 ? 4?m from process B. The anomalous non-monotonic frequency dependence of Q?1i was ambiguous. Measurements of the peak currents in the resonator suggested enhancements of Q?1i when the currents in the resonator were close to the expected critical currents of the Nb strips Although, the model did not apply when currents were more than a magnitude below the critical current. This facilitated the design of multi-plexed microstrip resonators on a single feedline to have a higher density of frequency points to analyze. In the next section, a design of multiplexed 0.75 ?m MTL resonators was presented. In an approximate range of 3-4.2 K, I was able to deconvolve the Rs(T) and tan ?(T ) temperature dependence using the dispersive loss deconvolution (DLD) method. This was validated by fitting to the expected Rs and tan ?(T ) temperature dependence e??/kBT and approximately T 3, respectively. From the extracted tan ?(T) going to temperatures below 3 K, I proposed that the change from T 3 dependence at approximately 2 K is due to a transition from TLS relaxation to TLS absorption mechanisms. This was supported by a decrease in ?r of approximately 0.03- 0.04% at 2 K from fn(T ) measurements. Fits to the temperature change in resonant frequency were fit to TLS resonant absorption and relaxation models, but the fitted values for permanent dipole coupling were 3 orders of magnitude off from expected values and is not understood. Fits to the temperature dependence of Q?1i to TLS models had much better agreement to expected. From these fits, microwave loss spectroscopy analysis showed a non-monotonic dependence on the TLS induced dipole coupling. This non-monotonic dependence agrees with loss peaks inQ?1i 178 and is claimed to be a combination of saturation of TLS absorption processes and enhancements of specific TLSs coupled to impurities in the material resonant with the probed frequency. Although, this is speculative and not a rigorous analysis, and systematic power dependent measurement should be performed. Aside from the anomalous behavior, the conclusions on temperature dependence power efficiency of RQL clock resonators can still be summarized and found in the last section. Similar to the conclusion drawn from the last chapter, tan ? is the limiting loss dominating power dissipation in RQL resonators as temperature is reduced below 3 K. This is analogous to the dominant loss mechanism in superconducting Qubit systems. However, care needs to be taken on resonator clock input power requirements as temperature falls below 3 K. At high input powers, TLS resonant absorption processes are saturated, allowing for reduced power dissipation in the resonator clock. At increased powers, there are two additional considerations: (i) Input current (power) to the resonator needs to remain far below the critical current of the MTL to mitigate nonlinear effects and increased microwave loss from the superconducting wires and groundplanes. (ii) At temperatures below 3 K, especially mK temperatures, the heat lift in cryostat decreases, and heat dissipation designs may need to be implemented. At 2 K, the internal Q-factor Qi of RQL interconnect increased by an order of magnitude, yielding an RQL interconnect efficiency of ? ? 0.9. Although, an order of magnitude increase of Qi at 2 K comes at an added refrigeration cost by a factor of 3 relative to 4.2 K [181]. From this, it can be estimated that the total power consumption of RQL operating at 10 GHz is 100x less than equivalent CMOS circuits operating at 1-4 GHz (logic and interconnects) [4]. 179 Chapter 6: Summary and Conclusions 6.1 Summary The work performed in this dissertation was aimed at performing materials characterization of microwave losses in Nb and TEOS to provide information on the limitations and mitigation strategies to reduce power dissipation in RQL clock interconnects. The superconductor and dielectric losses at temperature below 4.2 K and GHz frequencies have independent frequency and temperature scalings. On one hand this makes it difficult to design a system to optimize power efficiency performance. On the other hand, the unique frequency and temperature dependence presents the opportunity to deconvolve the loss mechanisms in a single measurement. In this dissertation I discussed the power efficiency limitations dictated by the Nb and TEOS supported by direct measurements of the microwave losses using the Dispersive Loss Deconvolution method (DLD). The DLD method confirmed the initially fabricated Nb material was of low quality, as opposed to the TEOS dielectric, and helped the fabrication team rapidly fine tune the process to yield the ultra-low loss Nb interconnect material. At 4.2 K, the theoretical minimum of Nb microwave losses can be reached and heavily depends on the microstrip geometry and the fabrication process used. At temperatures below 4.2 K, the microwave losses in the superconductor become negligible and the dielectric losses dominate and is independent of the resonator geometry. It was found that TLS relaxation pro- 180 cesses dominate the dielectric loss in the temperature range 2.5 to 4.2 K. At temperatures below 2.5 K, TLS resonant absorption processes become comparable to the relaxation loss and has a significant input power dependence. At 2 K, it was found that the RQL interconnect efficiency was estimated to be ? ? 0.9 relative to the logic. Accounting for the refrigeration costs, it can be estimated that the total power consumption of RQL operating at 10 GHz is 100x less than equivalent CMOS circuits operating at 1-4 GHz (logic and interconnects) [4]. Changing the operating temperature of RQL from 4.2 to 2 K would not create significant design changes since Nb inductance decreases by less than 20%, Nb/AlOx/Nb Josephson junction critical currents increase by less than 20%, and TEOS relative dielectric constant remains virtually constant. RQL is a classical computing technology with notable energy efficiency advantages over CMOS with circuits that can operate over 100 GHz. Although, circuit density still remains as a significant limiting factor preventing the technology from competing with CMOS. Both the size of the Jospehson junctions and the interconnects need to be pushed to dimensions below 0.25 ?m. Specifically for Nb interconnects, pushing the critical dimensions to be comparable with the magnetic penetration depth ? ? 90 nm imposes strict requirements on the fabrication process to produce low loss interconnects while increasing the processing window. 6.2 Future Work The work performed in this dissertation was aimed at improving the energy efficiency of RQL superconducting interconnects and help provide a path for RQL to meet computing power demands. The power dissipation in RQL interconnects is inherently dependent on RF losses in 181 the superconductor and dielectric. Since resonator-based clock networks deliver power on the same order of the total dynamic power of the Josephson junction logic, the RF losses in the superconductor and dielectric can directly influence the efficiency and scaling of the technology. Considering the superconductor and dielectric RF losses are comparable in RQL interconnect due to the imposed transmission line geometry from the fabrication process, both RF loss channels need to be minimized. In this dissertation, through the lens of material science, I discussed some of the essential work needed to improve the processing, properties, and performance of RQL interconnect. Furthermore, the novel Dispersive Loss Deconvolution (DLD) method and resonator designs described here serve as the groundwork for future groups interested in measuring RF losses in inseparable superconductor-dielectric systems. The outlook and possible research directions include: Processing (a) Further improvement of the metal damascene fabrication process down to and below 0.25- micron dimensions such as protective buffer and capping layers (thermal oxide diffusion barriers), and deposition parameters that yield uniform trench filling (eliminate voiding and facilitate conformal coverage). (b) Tighter control of critical dimensions and material property variation across the wafer. The sensitivity analysis in Section 3.7 showed that tan ? is geometry independent with a variation of approximately 20% across the wafer. Whereas the conductor Q-factor is both sensitive to the geometric factor ? with worst case 20% variation, and the intrinsic surface resistance Rs, which seemed have a variation less than that (see Fig. 4.9). (c) Increasing the thermal budget - (i) implement junction-on-top or more thermally resilient 182 Josephson junction barrier technologies, (ii) on wiring layers to promote ideal growth kinetics if a damascene process is needed. A damascene process is used in CMOS industry because it is difficult to etch Cu (as opposed to silicon dioxide). Nb is fairly easy to etch, so it is conceivable to rely on Nb etch processes to define the wires and CMP the dielectric, which yielded the lowest loss Nb reported in Chapter 4. (d) Surveying different superconductor and dielectric materials with lower RF losses and/or more resilient to process variations. Candidate superconducting materials are NbN and NbTiN for their low RF loss properties and sputtering compatibility. On one hand, NbN and NbTiN have higher inductance and can reduce circuit sizes, but on the other hand this will decrease propagation speed relative to Nb and inductance uniformity needs to also be considered. Silicon nitride is a candidate dielectric material as it can have an order of magnitude lower loss than TEOS, but it is important to consider the change in propagation speed (relative dielectric constant difference). The relative dielectric constant for TEOS and silicon nitride are approximately 4.2 and 7.5, respectively. Therefore the propagation velocity would decrease by approximately 25% switching from TEOS to silicon nitride. Properties (a) Reduce the fundamental mode frequency of the microstrip transmission line used in Chapters 4 and 5 down to below 1 GHz to increase the frequency range to validate the superconducting frequency dependencies. (b) Continue multi-plexed resonator technique to understand spatial distribution of properties (and critical dimensions) across a chip/wafer 183 (c) Incorporate the impact of nonlinear dynamics in FEM of superconducting interconnects, as I assumed the system displays linear response throughout this thesis, but is most likely not the case since electromagnetic field concentration at the edges and will only become exacerbated as dimensions are pushed below 0.25-microns and the input power will be pushed to its limits. (d) Incorporate common nonlinear measurements (vs power, vs applied magnetic field (in- plane, out-of-plane). (e) Physical characterization techniques with higher sensitivity to trace element contamination both in interconnect cross-section as well as spatially (e.g. atom probe tomagraphy, etc). (f) Determine if there is a correlation between dielectric loss tangent tan ? between 4.2 K and mK. Conceivably, rapid measurements at 4.2 K may be more advantageous to evaluate a superconducting interconnect. (g) Use as a test vessel to measure vortex dynamic properties on wide microstrip MTLs with temerpature dependent measurements in perpendicular magnetic field [137]. Performance (a) In the nominal configuration, RQL power dissipation is less of an issue at mK temperatures due to million level Q-factors (superconductor loss is negligible and TLSs saturate at RQL powers). (c) Continue work on exploring cross-sectional RQL interconnect geometries to mitigate RF losses by engineering the geometric factor ?. 184 (d) Further explore mK-K temperature range of superconducting-dielectric RF losses in conjuction with cooling costs to optimize energy efficient computing. (e) Higher throughput test systems to get wafer statistics, a wafer prober can alleviate this [191]. 185 Appendix A: FIB Cross Sections of Process A and B To accurately calculate the geometric factor ? of microstrip transmission line (MTL) resonators (see Chapter 3), exact measurements of the MTL geometries were measured using a similar instrument as the FEI Helios 660 Dual Beam FIB/SEM. The cross section of the MTLs were exposed using a focused ion beam (FIB) and then imaged using a scanning electron microscope (SEM). Table A.1: FIB Results Summary table for processes A and B measured in this work. All of these microstrip tranmission lines were measured using a similar instrument as the FEI Helios 660 Dual Beam FIB/SEM. Cross sections were taken at about the half way point along the meandor resonator (? 7 mm). The second column are the nominal (designed) wire widths. Example FIB cross sections can be found in Figs A.1 and A.2. The third column are the measured mean widths averaged between the bottom and top of the wire. The ? values in the last three columns are the calculated one standard deviations taken from measurements of all 5 wire widths. Table A.1 is a summary of the measurements for MTLs made from process A and B. The 186 second column are the nominal (designed) wire widths. The third column are the measured mean widths averaged between the bottom and top of the wire. The ? values in the last three columns are the calculated one standard deviations taken from measurements of all 5 wire widths. Figure A.1: Example cross sections and measurements of the microstrip tranmission line resonator geometry for the 0.25?m and 4?m wire widths fabricated using process A. In this process the conducting strip (CS) is below the ground plane (GP) 187 Figure A.2: Example cross sections and measurements of the microstrip tranmission line resonator geometry for the 0.25?m and 4?m wire widths fabricated using process B. In this process the conducting strip (CS) is above the ground plane (GP) Example cross sections for 0.25?m and 4?m MTLs can be found in Fig. A.1 from process A. In this process the conducting strip (wire) is below the ground plane. 188 Example cross sections for 0.25?m and 4?m MTLs can be found in Fig. A.2 from process B. In this process the conducting strip (wire) is above the ground plane. Appendix B: Estimation of Radiation Q-Factor The radiation Q-factor in general is the ratio of the energy stored in the resonator over the power radiated from the transmission line resonator, and can be estimated using the following equation from [163] Q 4?Z0rad = 2 (B.1)480(s?) F (?eff ) ( ) 1/2 ? +1 2? (? ? +1F (? ) = eff eff?1) ln effeff (B.2)? 3/2 1/2eff 2? ? ?1 eff eff where Z0 is the characteristic impedance, s is the dielectric thickness, ? is the phase constant, and ?eff . For a Z0 = 55 ? and ? = 670m?1 found using HFSS simulation with inputs Rs = 20 ??, ? = 90nm, ?eff = 4.5, and geometry shown in Fig. 4.1, Qrad ? 230?106. For internal Q-factor Qi and external (or coupling) Q-factor Qe on the order of 1? 103, it is safe to ignore Qrad using 189 the following relation of the loaded Q-factor Q?1 = Q?1 +Q?1 +Q?1L i e rad. Appendix C: Temperature Dependent Resonator Measurements C.1 Measurements System For high quality RF measurements (up to 12 GHz) with precise temperature control (1.5 ? 9 K), I utilize a system similar to the Dry ICE 1.5 K Cryostat [182]. This system has a base temperature of 1.4 K and temperature can be set between 1.4-10 K with a temperature stability of ?10mK. The temperature is controlled using a Lakeshore 336 PID control [192] with a 25 ? heater. 190 Figure C.1: I utilized a system similar to the Dry ICE 1.5 K Cryostat [182] for temperature control for the testing and results described in Chapter 5. The test fixture is similar to the RF dip probe outfitted with a 32 pin (contact pad) described in Chapter 4. The same test fixutre was used in previous works [35], [34], [36], [10]. To achieve RF signal integrity and avoid wire bonding for fast sample chip exchanges, a non-magnetic Cu/Au flip chip press contact technology is used. The fixture PCB connects the flip-chip bump pads to the probe .047 inch semi-rigid coaxial cables. The chip and roughly 30 cm of the cables are cooled by the Dry ICE 1.5 K Cryostat during measurements. The temperature was monitored with a calibrated Lakeshore Cernox sensor. The resonators were measured by 2-port S-parameters using a Keysight N5222A vector network analyzer (VNA). The network analyzer was calibrated up to the top of the probe. 191 Appendix D: Relevant Equations for Parallel Plate Transmission Line Resonator In the following equations in this appendix section, a parallel plate transmission line geometry is assumed. In this assumption with an infinitely wide ground plane, the microstrip width is much greater than the dielectric thickness w  s. The complex propagation constant ? = ?+ i? is assumed here to be that of a Swihart wave and is a good approximation for a wide microstrip transmission line with the form [36, 59] [ ] ? 2Reff ? = + tan ? (D.1) 2 ??0s+ 2(X?eff ?Reff tan ?) ? 2(Xeff ?Reff tan ?) ? = ? ?0?r?0 1 + (D.2) ??0s whereReff is the effective surface resistance (see Eq. 2.24),Xeff is the effective surface reactance (see Eq. 2.25), s is the dielectric thickness between the microstrip and groundplane, ?r is the relative dielectric constant of the dielectric, tan ? is the loss tangent of the dielectric material, ?0 is the permittivity of free space, and ?0 is the permeability of free space. Here, a PPW system has been enforced, and we are interested in the resulting quasi- transverse electromagnetic (quasi-TEM) mode propagating along the longitudinal length l in the x direction (see Fig. 3.10). The accuracy of HFSS solutions will be tested using the following equations forRLGC by inserting the superconducting impedance definitions Eq. 2.24 and Eq. 2.25 192 into PPW analytical solutions [95] 2R R = eff (D.3) w L = ? s+2? coth (d/?)0 (D.4)w G = ?? w0?r tan ? (D.5)s C = ? ? w0 r (D.6)s where ?0 is the vacuum permeability, ?0 is the vacuum permittivity, ? is the angular frequency, w is the PPW width, s is the plate separation dielectric thickness, Reff is the effective resistance defined in Eq. 2.24, and tan ? is the dielectric loss tangent. The RLGC are the series resistance, series inductance, shunt conductance, and shunt capacitance per unit length of a generalized transmission line defined in Eq. 3.2-Eq. 3.5 in Section 3.3. MTL resonators support a TM00n [59] mode with the eigen frequency n nc fn = ? = ? (D.7) 2lres LC ? 2lres ?r 1 + 2? coth(d/?)/s where n = 1, 2, 3. . . is the mode index, lres is the geometrical length of the resonator, and L and C are the conducting strip series inductance and shunt capacitance per unit length defined by Eq. 3.3 and 3.5. The right hand side in Eq. D.7 holds for a wide MTL where the width is much greater than the dielectric thickness (w  s) [70], where c is the speed of light, and ?r is the dielectric constant. 193 Using Eqs. D.4 and D.6, the characteristic impedance can be calculated by ? ?s ?0 Z0 = L/C = (1 + 2?eff/s) (D.8) w ?0?r where ?eff = ? coth(d/?) is the effective magnetic penetration depth. Using Eq. 3.1, the partial conductor and dielectric Q-factors for a parallel plate resonator, Qc = ?L/R and Qd = ?C/G, respectively, are Q = ?? s+2? coth (d/?)c 0 (D.9)2Reff Qd = 1/ tan ? (D.10) Notice, the PPW width w drops out in Qc, and ? drops out in Qd with no geometry dependence. Then, the only parameters needed to model the superconducting PPW, analytically and with HFSS impedance boundaries, are s, d, ?, ?, Rs, and tan ?. To provide a simple analytical limiting case for Eq. 3.14 consider a resonator formed by a parallel-plate waveguide with identical superconducting plates of the width w and thickness d, separated by the dielectric layer of thickness s. This geometry approximates MTL with w  s. The geometric factor is s+ 2? coth(d/?) ?PP (?) = ??0 (D.11) 2[coth(d/?) + (d/?)csch2(d/?)] utilizing the expression for the effective surface resistance of thin superconducting films, and by R ? R [coth(d/?)+(d/?)csch2eff s (d/?)] that ignores transmission loss through the superconducting 194 films [69]. In the thick film limit the geometric factor reduces to ? 1PP = ??0(s + 2?) and the2 geometric factor is no longer dependent on the superconductor plate thicknesses d. For the thin film limit where ? d the full expression needs to be used in Eq. D.11 and a term including the dielectric substrate impedance may be needed [70]. 195 Bibliography [1] John Shalf. The future of computing beyond moore?s law. Philosophical Transactions of the Royal Society A, 378(2166):20190061, 2020. [2] Martijn Koot and Fons Wijnhoven. Usage impact on data center electricity needs: A system dynamic forecasting model. Applied Energy, 291:116798, 2021. [3] Mark Neisser. International roadmap for devices and systems lithography roadmap. Journal of Micro/Nanopatterning, Materials, and Metrology, 20(4):044601, 2021. [4] Quentin P Herr, Anna Y Herr, Oliver T Oberg, and Alexander G Ioannidis. Ultra-low- power superconductor logic. Journal of applied physics, 109(10):103903, 2011. [5] Quentin P Herr, Joshua Osborne, Micah JA Stoutimore, Harold Hearne, Ryan Selig, Jacob Vogel, Eileen Min, Vladimir V Talanov, and Anna Y Herr. Reproducible operating margins on a 72 800-device digital superconducting chip. Superconductor Science and Technology, 28(12):124003, 2015. [6] Yuri Polyakov, Supradeep Narayana, and Vasili K Semenov. Flux trapping in superconducting circuits. IEEE transactions on applied superconductivity, 17(2):520?525, 2007. [7] Vasili K Semenov, Yuri A Polyakov, and Sergey K Tolpygo. Ac-biased shift registers as fabrication process benchmark circuits and flux trapping diagnostic tool. IEEE Transactions on Applied Superconductivity, 27(4):1?9, 2017. [8] Leland Chang, David J Frank, Robert K Montoye, Steven J Koester, Brian L Ji, Paul W Coteus, Robert H Dennard, and Wilfried Haensch. Practical strategies for power-efficient computing technologies. Proceedings of the IEEE, 98(2):215?236, 2010. [9] Raj K Jana, Gregory L Snider, and Debdeep Jena. Energy-efficient clocking based on resonant switching for low-power computation. IEEE Transactions on Circuits and Systems I: Regular Papers, 61(5):1400?1408, 2014. [10] J.A. Strong. A resonant metamaterial clock distribution network for superconducting logic. Nat Electron, 2022. 196 [11] Donald R Askeland and Wendelin J Wright. Essentials of materials science and engineering. Cengage Learning, 2018. [12] Peidong Yang and Jean-Marie Tarascon. Towards systems materials engineering. Nature materials, 11(7):560?563, 2012. [13] Brian David Josephson. Possible new effects in superconductive tunnelling. Physics letters, 1(7):251?253, 1962. [14] Philip W Anderson and John M Rowell. Probable observation of the josephson superconducting tunneling effect. Physical Review Letters, 10(6):230, 1963. [15] John M Rowell. Cryogenic supercurrent tunneling devices, October 25 1966. US Patent 3,281,609. [16] M Gurvitch, MA Washington, and HA Huggins. High quality refractory josephson tunnel junctions utilizing thin aluminum layers. Applied Physics Letters, 42(5):472?474, 1983. [17] Theodore Van Duzer and Charles W. Turner. Principles of Superconductive Devices and Circuits - 2nd edition. Prentice Hall, Inc., 1999. [18] S Hasuo. Josephson microprocessors. In The New Superconducting Electronics, pages 363?399. Springer, 1993. [19] Jay Fleischman, David Feld, Peter Xiao, and Theodore Van Duzer. Evaluation of flux- based logic schemes for high-t/sub c/applications. IEEE transactions on magnetics, 27(2):2769?2772, 1991. [20] F Bedard, N Welker, GR Cotter, MA Escavage, and JT Pinkston. Superconducting technology assessment. National Security Agency, Office of Corporate Assessments, 2005. [21] KK Likharev, OA Mukhanov, and VK Semenov. Resistive single flux quantum logic for the josephson-junction digital technology. SQUID Journal, 85:1103?1108, 1985. [22] Konstantin K Likharev and Vasilii K Semenov. Rsfq logic/memory family: A new josephson-junction technology for sub-terahertz-clock-frequency digital systems. IEEE Transactions on Applied Superconductivity, 1(1):3?28, 1991. [23] Mutsumi Hosoya, Willy Hioe, Juan Casas, Ryotaro Kamikawai, Yutaka Harada, Yasou Wada, Hideaki Nakane, Reiji Suda, and Eiichi Goto. Quantum flux parametron: a single quantum flux device for josephson supercomputer. IEEE Transactions on Applied Superconductivity, 1(2):77?89, 1991. [24] Mark H Volkmann, Anubhav Sahu, Coenrad J Fourie, and Oleg A Mukhanov. Implementation of energy efficient single flux quantum digital circuits with sub-aj/bit operation. Superconductor Science and Technology, 26(1):015002, 2012. [25] Naoki Takeuchi, Dan Ozawa, Yuki Yamanashi, and Nobuyuki Yoshikawa. An adiabatic quantum flux parametron as an ultra-low-power logic device. Superconductor Science and Technology, 26(3):035010, 2013. 197 [26] Sergey K Tolpygo, Diana Tolpygo, Richard T Hunt, Supradeep Narayana, Yuri A Polyakov, and Vasili K Semenov. Wafer bumping process and inter-chip connections for ultra-high data transfer rates in multi-chip modules with superconductor integrated circuits. IEEE transactions on applied superconductivity, 19(3):598?602, 2009. [27] S Narayana, VK Semenov, YA Polyakov, V Dotsenko, and SK Tolpygo. Design and testing of high-speed interconnects for superconducting multi-chip modules. Superconductor Science and Technology, 25(10):105012, 2012. [28] Micah Stoutimore, James Medford, Quentin Herr, Ofer Naaman, Harold Hearne, Joel Strand, Anthony Przybysz, Aaron Pesetski, and John Przybysz. Demonstrated control of a transmon using a reciprocal quantum logic digital circuit-part 1. In APS March Meeting Abstracts, volume 2015, pages J39?012, 2015. [29] Micah Stoutimore, James Medford, Quentin Herr, Ofer Naaman, Harold Hearne, Joel Strand, Anthony Przybysz, Aaron Pesetski, and John Przybysz. Demonstrated control of a transmon using a reciprocal quantum logic digital circuit-part 1. In APS March Meeting Abstracts, volume 2015, pages J39?012, 2015. [30] Swamit S Tannu, Zachary A Myers, Prashant J Nair, Douglas M Carmean, and Moinuddin K Qureshi. Taming the instruction bandwidth of quantum computers via hardware-managed error correction. In Proceedings of the 50th Annual IEEE/ACM International Symposium on Microarchitecture, pages 679?691, 2017. [31] INTERNATIONAL ROADMAP FOR DEVICES and SYSTEMS. Cryogenic electronics and quantum information processing, 2020. [32] Sergey K Tolpygo. Superconductor digital electronics: Scalability and energy efficiency issues. Low Temperature Physics, 42(5):361?379, 2016. [33] Naveen Kumar Katam, Jamil Kawa, and Massoud Pedram. Challenges and the status of superconducting single flux quantum technology. In 2019 Design, Automation & Test in Europe Conference & Exhibition (DATE), pages 1781?1787. IEEE, 2019. [34] Jonathan Egan, Max Nielsen, Joshua Strong, Vladimir V Talanov, Ed Rudman, Brainton Song, Quentin Herr, and Anna Herr. Synchronous chip-to-chip communication with a multi-chip resonator clock distribution network. arXiv preprint arXiv:2109.00560, 2021. [35] Haitao Dai, Corey Kegerreis, Deepal Wehella Gamage, Jonathan Egan, Max Nielsen, Yuan Chen, David Tuckerman, Sherman Peek, Bhargav Yelamanchili, Michael Hamilton, et al. Isochronous data link across a superconducting nb flex cable with 5 femtojoules per bit. Superconductor Science and Technology, 2022. [36] Vladimir Talanov, Derek Knee, David Harms, Kieran Perkins, Andrew Urbanas, Jonathan Egan, Quentin Herr, and Anna Y Herr. Propagation of picosecond pulses on superconducting transmission line interconnects. Superconductor Science and Technology, 2022. 198 [37] INTERNATIONAL ROADMAP FOR DEVICES and SYSTEMS. Cryogenic electronics and quantum information processing, 2020. [38] H Kamerlingh Onnes. Comm. leiden nr. 108. In Proc. Roy. Acad. Amsterdam, volume 11, page 168, 1908. [39] H Kamerlingh Onnes. The resistance of pure mercury at helium temperatures. Commun. Phys. Lab. Univ. Leiden, b, 120, 1911. [40] Cornelis Jacobus Gorter and Hendrik Casimir. On supraconductivity i. Physica, 1(1- 6):306?320, 1934. [41] Pyotr Kapitza. Viscosity of liquid helium below the ?-point. Nature, 141(3558):74?74, 1938. [42] John F Allen and AD Misener. Flow of liquid helium ii. Nature, 141(3558):75?75, 1938. [43] Fritz London and Heinz London. The electromagnetic equations of the supraconductor. Proceedings of the Royal Society of London. Series A-Mathematical and Physical Sciences, 149(866):71?88, 1935. [44] Werner Buckel and Reinhold Kleiner. Superconductivity: fundamentals and applications. John Wiley & Sons, 2008. [45] J. Bardeen, L. N. Cooper, and J. R. Schrieffer. Theory of superconductivity. Phys. Rev., 108:1175?1204, Dec 1957. [46] B Mitrovic?, HG Zarate, and JP Carbotte. The ratio 2 ? 0 k b t c within eliashberg theory. Physical Review B, 29(1):184, 1984. [47] Heinz London. The high-frequency resistance of superconducting tin. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, 176(967):522? 533, 1940. [48] AB Pippard. The surface impedance of superconductors and normal metals at high frequencies i. resistance of superconducting tin and mercury at 1200 mcyc./sec. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, 191(1026):370?384, 1947. [49] Daniel C Mattis and John Bardeen. Theory of the anomalous skin effect in normal and superconducting metals. Physical Review, 111(2):412, 1958. [50] Richard L Kautz. Picosecond pulses on superconducting striplines. Journal of Applied Physics, 49(1):308?314, 1978. [51] Michael Tinkham. Introduction to superconductivity. Courier Corporation, 2004. [52] Alfred Brian Pippard and William Lawrence Bragg. An experimental and theoretical study of the relation between magnetic field and current in a superconductor. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, 216(1127):547? 568, 1953. 199 [53] Nicola Pinto, S Javad Rezvani, Andrea Perali, Luca Flammia, Milorad V Milos?evic?, Matteo Fretto, Cristina Cassiago, and Natascia De Leo. Dimensional crossover and incipient quantum size effects in superconducting niobium nanofilms. Scientific reports, 8(1):1?12, 2018. [54] Sergey K Tolpygo, Evan B Golden, Terence J Weir, and Vladimir Bolkhovsky. Inductance of superconductor integrated circuit features with sizes down to 120 nm. Superconductor Science and Technology, 34(8):085005, 2021. [55] Orest G Vendik, Irina B Vendik, and Dimitri I Kaparkov. Empirical model of the microwave properties of high-temperature superconductors. IEEE transactions on microwave theory and techniques, 46(5):469?478, 1998. [56] Oliver Lodge. On lightning, lightning conductors, and lightning protectors. Journal of the Institution of Electrical Engineers, 18(80):386?430, 1889. [57] Simon Ramo, John R Whinnery, and Theodore Van Duzer. Fields and waves in communication electronics. John Wiley & Sons, 1994. [58] Julls Adams Stratton. Electromagnetic theory, mcgrow-hill book company. Inc., New York, and London, pages 205?207, 1941. [59] James C Swihart. Field solution for a thin-film superconducting strip transmission line. Journal of Applied Physics, 32(3):461?469, 1961. [60] Sergey V Yuferev and Nathan Ida. Surface impedance boundary conditions: a comprehensive approach. CRC press, 2018. [61] Sergei A Schelkunoff. The impedance concept and its application to problems of reflection, refraction, shielding and power absorption. Bell System Technical Journal, 17(1):17?48, 1938. [62] Nathan Newman and W Gregory Lyons. High-temperature superconducting microwave devices: fundamental issues in materials, physics, and engineering. Journal of Superconductivity, 6(3):119?160, 1993. [63] JP Turneaure, J Halbritter, and HA Schwettman. The surface impedance of superconductors and normal conductors: The mattis-bardeen theory. Journal of Superconductivity, 4(5):341?355, 1991. [64] Cristoforo Benvenuti, S Calatroni, IE Campisi, P Darriulat, MA Peck, R Russo, and A- M Valente. Study of the surface resistance of superconducting niobium films at 1.5 ghz. Physica C: Superconductivity, 316(3-4):153?188, 1999. [65] InductEx, C J Fourie. HFSS. Stellenbosch Univ., Stellenbosch, South Africa, 2015. [66] MA Leontovich. A new method to solve problems of em wave propagation over the earth surface. USSR Academy of Sciences Trans., Physics Series, 8(1):16?22, 1944. 200 [67] MA Leontovich. Approximate boundary conditions for the electromagnetic field on the surface of a good conductor. Investigations on radiowave propagation, 2:5?12, 1948. [68] Aleksandr N Shchukin. Propagation of radio waves. Svyazizdat, Moscow, 1940. [69] N Klein, H Chaloupka, G Mu?ller, S Orbach, H Piel, B Roas, L Schultz, U Klein, and M Peiniger. The effective microwave surface impedance of high t c thin films. Journal of Applied Physics, 67(11):6940?6945, 1990. [70] Vladimir V Talanov, Lucia V Mercaldo, Steven M Anlage, and John H Claassen. Measurement of the absolute penetration depth and surface resistance of superconductors and normal metals with the variable spacing parallel plate resonator. Review of Scientific Instruments, 71(5):2136?2146, 2000. [71] Anne-Marie Valente-Feliciano. Superconducting rf materials other than bulk niobium: a review. Superconductor Science and Technology, 29(11):113002, 2016. [72] Jin Yong Kim, Moo Sung Hwang, Yoon-Hae Kim, Hyeong Joon Kim, and Young Lee. Origin of low dielectric constant of carbon-incorporated silicon oxide film deposited by plasma enhanced chemical vapor deposition. Journal of Applied Physics, 90(5):2469? 2473, 2001. [73] OF Rocha, CE Viana, LCD Gonc?alves, and NI Morimoto. Electrical characteristics of pecvd silicon oxide deposited with low teos contents at low temperatures. Microelectronics Technology and Devices SBMICRO 2004, pages 295?300, 2004. [74] Babacar Diallo, Konstantina C Topka, Maxime Puyo, Charlotte Lebesgue, Cecile Genevois, Raphael Laloo, Diane Samelor, Helene Lecoq, Mathieu Allix, Hugues Vergnes, et al. Network hydration, ordering and composition interplay of chemical vapor deposited amorphous silica films from tetraethyl orthosilicate. journal of materials research and technology, 13:534?547, 2021. [75] Corey Rae Harrington McRae, Haozhi Wang, Jiansong Gao, Michael R Vissers, Teresa Brecht, Andrew Dunsworth, David P Pappas, and Josh Mutus. Materials loss measurements using superconducting microwave resonators. Review of Scientific Instruments, 91(9):091101, 2020. [76] Vera V Daniel. Dielectric relaxation. Academic press, 1967. [77] RC Zeller and RO Pohl. Thermal conductivity and specific heat of noncrystalline solids. Physical Review B, 4(6):2029, 1971. [78] Peter Debye. Zur theorie der spezifischen wa?rmen. Annalen der Physik, 344(14):789?839, 1912. [79] William A Phillips. Two-level states in glasses. Reports on Progress in Physics, 50(12):1657, 1987. 201 [80] A Anderson. Topics in current physics (amorphous solids: Low temperature properties) vol. 24, 1981. [81] OL Anderson and HE Bo?mmel. Ultrasonic absorption in fused silica at low temperatures and high frequencies. Journal of the American Ceramic Society, 38(4):125?131, 1955. [82] P W Anderson, Bertrand I Halperin, and C M Varma. Anomalous low-temperature thermal properties of glasses and spin glasses. Philosophical Magazine, 25(1):1?9, 1972. [83] William A Phillips. Tunneling states in amorphous solids. Journal of Low Temperature Physics, 7(3):351?360, 1972. [84] M Von Schickfus and S Hunklinger. Saturation of the dielectric absorption of vitreous silica at low temperatures. Physics Letters A, 64(1):144?146, 1977. [85] C Laermans, W Arnold, and S Hunklinger. Influence of an electromagnetic wave on the acoustic absorption of borosilicate glass at low temperatures. Journal of Physics C: Solid State Physics, 10(8):L161, 1977. [86] Sonja Rau, Christian Enss, Siegfried Hunklinger, Peter Neu, and Alois Wu?rger. Acoustic properties of oxide glasses at low temperatures. Physical Review B, 52(10):7179, 1995. [87] M Sargent III. M. 0. scully, and we lamb, jr. Laser Physics, pages 144?155, 1974. [88] CP Slichter. Principles of magnetic resonance springer-verlag, berlin. Heidelberg, New York, 1980. [89] William Andrew Phillips and AC Anderson. Amorphous solids: low-temperature properties, volume 24. Springer, 1981. [90] U Buchenau, Yu M Galperin, VL Gurevich, DA Parshin, MA Ramos, and HR Schober. Interaction of soft modes and sound waves in glasses. Physical Review B, 46(5):2798, 1992. [91] DA Parshin. Soft potential model and universal properties of glasses. Physica Scripta, 1993(T49A):180, 1993. [92] VG Karpov, I Klinger, and FN Ignat?Ev. Theory of the low-temperature anomalies in the thermal properties of amorphous structures. Zh. Eksp. Teor. Fiz, 84:775, 1983. [93] Robert O Pohl, Xiao Liu, and EunJoo Thompson. Low-temperature thermal conductivity and acoustic attenuation in amorphous solids. Reviews of Modern Physics, 74(4):991, 2002. [94] U Strom, M Von Schickfus, and S Hunklinger. Low-temperature anomalies in the microwave dielectric properties of na ?-alumina. Physical Review Letters, 41(13):910, 1978. [95] David M Pozar. Microwave engineering. John wiley & sons, 2011. 202 [96] M Pollak and To H Geballe. Low-frequency conductivity due to hopping processes in silicon. Physical Review, 122(6):1742, 1961. [97] Andrew K Jonscher. The ?universal?dielectric response. nature, 267(5613):673?679, 1977. [98] KL Ngai, AK Jonscher, and CT White. On the origin of the universal dielectric response in condensed matter. Nature, 277(5693):185?189, 1979. [99] Christoph Kaiser. High quality Nb/Al-AlOx/Nb Josephson junctions: technological development and macroscopic quantum experiments, volume 4. KIT Scientific Publishing, 2011. [100] David B Tuckerman, Michael C Hamilton, David J Reilly, Rujun Bai, George A Hernandez, John M Hornibrook, John A Sellers, and Charles D Ellis. Flexible superconducting nb transmission lines on thin film polyimide for quantum computing applications. Superconductor Science and Technology, 29(8):084007, 2016. [101] Jiansong Gao. The physics of superconducting microwave resonators. California Institute of Technology, 2008. [102] Thomas J Licata, Evan G Colgan, James ME Harper, and Stephen E Luce. Interconnect fabrication processes and the development of low-cost wiring for cmos products. IBM Journal of Research and Development, 39(4):419?435, 1995. [103] M David Levenson. Extending optical lithography to the gigabit era. Solid State Technology, 38(2):57?63, 1995. [104] Panayotis C Andricacos, Cyprian Uzoh, John O Dukovic, Jean Horkans, and Hariklia Deligianni. Damascene copper electroplating for chip interconnections. IBM Journal of Research and Development, 42(5):567?574, 1998. [105] Thomas P Moffat, JE Bonevich, WH Huber, A Stanishevsky, DR Kelly, GR Stafford, and D Josell. Superconformal electrodeposition of copper in 500?90 nm features. Journal of The Electrochemical Society, 147(12):4524, 2000. [106] Milan Paunovic and Mordechay Schlesinger. Fundamentals of electrochemical deposition. john wiley & sons, 2006. [107] WD Van Driel. Facing the challenge of designing for cu/low-k reliability. Microelectronics Reliability, 47(12):1969?1974, 2007. [108] Jivaan Kishore Jhothiraman and Rajesh Balachandran. Electroplating: Applications in the semiconductor industry. Advances in Chemical Engineering and Science, 9(2):239?261, 2019. [109] Henry H Radamson, Huilong Zhu, Zhenhua Wu, Xiaobin He, Hongxiao Lin, Jinbiao Liu, Jinjuan Xiang, Zhenzhen Kong, Wenjuan Xiong, Junjie Li, et al. State of the art and future perspectives in advanced cmos technology. Nanomaterials, 10(8):1555, 2020. 203 [110] Daniel T Yohannes, Rick T Hunt, John A Vivalda, Denis Amparo, Alexander Cohen, Igor V Vernik, and Alex F Kirichenko. Planarized, extendible, multilayer fabrication process for superconducting electronics. IEEE Transactions on Applied Superconductivity, 25(3):1?5, 2014. [111] SEEQC. The digital superconducting company, 2019. [112] Patrick A Truitt, Aritoki Suzuki, Daniel Yohannes, Michael Kamkar, Christopher Raum, Benjamin Westbrook, Trevor Sasse, and Adrian Lee. Commercially fabricated low loss superconducting resonators integrated with detectors for frequency domain multiplexing readout of future cosmic microwave background experiments. IEEE Transactions on Applied Superconductivity, 29(5):1?4, 2019. [113] Takumi Ando, Shuichi Nagasawa, Naoki Takeuchi, Naoki Tsuji, Fumihiro China, Mutsuo Hidaka, Yuki Yamanashi, and Nobuyuki Yoshikawa. Three-dimensional adiabatic quantum-flux-parametron fabricated using a double-active-layered niobium process. Superconductor Science and Technology, 30(7):075003, 2017. [114] Mutsuo Hidaka and Shuichi Nagasawa. Fabrication process for superconducting digital circuits. IEICE Transactions on Electronics, 104(9):405?410, 2021. [115] Sergey K Tolpygo, Vladimir Bolkhovsky, Ravi Rastogi, Scott Zarr, Alexandra L Day, Evan Golden, Terence J Weir, Alex Wynn, and Leonard M Johnson. Advanced fabrication processes for superconductor electronics: Current status and new developments. IEEE Transactions on Applied Superconductivity, 29(5):1?13, 2019. [116] Lynn A Abelson and George L Kerber. Superconductor integrated circuit fabrication technology. Proceedings of the IEEE, 92(10):1517?1533, 2004. [117] Carter Kaanta, William Cote, John Cronin, Karey Holland, Pei-Ing Lee, and Terry Wright. Submicron wiring technology with tungsten and planarization. In 1987 International Electron Devices Meeting, pages 209?212. IEEE, 1987. [118] William L Guthrie, William J Patrick, Ernest Levine, Harris C Jones, Ebrahim A Mehter, Thomas F Houghton, George T Chiu, and Michael A Fury. A four-level vlsi bipolar metallization design with chemical-mechanical planarization. IBM journal of research and development, 36(5):845?857, 1992. [119] Mahadevaiyer Krishnan, Jakub W Nalaskowski, and Lee M Cook. Chemical mechanical planarization: slurry chemistry, materials, and mechanisms. Chemical reviews, 110(1):178?204, 2010. [120] Abelaziz Ramzi, Serge A Charlebois, and Philip Krantz. Niobium and aluminum josephson junctions fabricated with a damascene cmp process. Physics Procedia, 36:211? 216, 2012. [121] David A Chan, Bogdan Swedek, Andreas Wiswesser, and Manush Birang. Process control and monitoring with laser interferometry based endpoint detection in chemical mechanical 204 planarization. In IEEE/SEMI 1998 IEEE/SEMI Advanced Semiconductor Manufacturing Conference and Workshop (Cat. No. 98CH36168), pages 377?384. IEEE, 1998. [122] Jeffrey David, B Bonner, T Osterheld, and R Jin. Advanced sti cmp solutions for new device technologies. In Proceedings of SEMI Technology Symposium (STS), volume 99, pages 8?64, 1999. [123] Applied Materials. Leader in materials engineering solutions used to produce virtually every new chip and advanced display in the world.., 1967. [124] Ron Allen, Charles Chen, Tom Trikas, Kurt Lehman, Robert Shinagawa, Vijay Bhaskaran, B Stephenson, and D Watts. In-situ cmp copper endpoint control system. In 2001 IEEE International Symposium on Semiconductor Manufacturing. ISSM 2001. Conference Proceedings (Cat. No. 01CH37203), pages 391?394. IEEE, 2001. [125] Prahalad K Rao, M Brij Bhushan, Satish TS Bukkapatnam, Zhenyu Kong, Sanjay Byalal, Omer F Beyca, Adam Fields, and Ranga Komanduri. Process-machine interaction (pmi) modeling and monitoring of chemical mechanical planarization (cmp) process using wireless vibration sensors. IEEE Transactions on Semiconductor Manufacturing, 27(1):1? 15, 2013. [126] Prahalad Krishna Rao. Sensor-based monitoring and inspection of surface morphology in ultraprecision manufacturing processes. Oklahoma State University, 2013. [127] Donna R Cote, SV Nguyen, William J Cote, Scott L Pennington, Anthony K Stamper, and Dragan V Podlesnik. Low-temperature chemical vapor deposition processes and dielectrics for microelectronic circuit manufacturing at ibm. IBM Journal of Research and Development, 39(4):437?464, 1995. [128] Peter F Cheng, SM Rossnagel, and David N Ruzic. Directional deposition of cu into semiconductor trench structures using ionized magnetron sputtering. Journal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures Processing, Measurement, and Phenomena, 13(2):203?208, 1995. [129] Shin-Pon Ju, Cheng-I Weng, Jee-Gong Chang, and Chi-Chuan Hwang. Molecular dynamics simulation of sputter trench-filling morphology in damascene process. Journal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures Processing, Measurement, and Phenomena, 20(3):946?955, 2002. [130] CC Koch, JO Scarbrough, and DM Kroeger. Effects of interstitial oxygen on the superconductivity of niobium. Physical Review B, 9(3):888, 1974. [131] Matthias Hein. High-temperature-superconductor thin films at microwave frequencies, volume 155. Springer Science & Business Media, 1999. [132] Jerzy Krupka, Krzysztof Derzakowski, Bill Riddle, and James Baker-Jarvis. A dielectric resonator for measurements of complex permittivity of low loss dielectric materials as a function of temperature. Measurement Science and Technology, 9(10):1751, 1998. 205 [133] Harold A Wheeler. Transmission-line properties of parallel strips separated by a dielectric sheet. IEEE Transactions on Microwave Theory and Techniques, 13(2):172?185, 1965. [134] WH Chang. The inductance of a superconducting strip transmission line. Journal of Applied Physics, 50(12):8129?8134, 1979. [135] H-Y Lee and Tatsuo Itoh. Phenomenological loss equivalence method for planar quasi-tem transmission lines with a thin normal conductor or superconductor. IEEE Transactions on Microwave Theory and Techniques, 37(12):1904?1909, 1989. [136] RC Taber. A parallel plate resonator technique for microwave loss measurements on superconductors. Review of scientific instruments, 61(8):2200?2206, 1990. [137] MS Pambianchi, DH Wu, L Ganapathi, and SM Anlage. Dc magnetic field dependence of the surface impedance in superconducting parallel plate transmission line resonators. IEEE transactions on applied superconductivity, 3(1):2774?2777, 1993. [138] David M Sheen, Sami M Ali, Daniel E Oates, Richard S Withers, and J Kong. Current distribution, resistance, and inductance for superconducting strip transmission lines. IEEE Transactions on Applied Superconductivity, 1(2):108?115, 1991. [139] AR Kerr. Surface impedance of superconductors and normal conductors in em simulators. MMA Memo, 21(245):1?17, 1999. [140] Mohammad Hossein Amini and Alireza Mallahzadeh. Modeling of superconducting components in full-wave simulators. Journal of Superconductivity and Novel Magnetism, 34(3):675?681, 2021. [141] Gianluca Piccinini and Andre? Chatel. Simulation and modelling of interconnect networks for cmos quantum bit systems. [142] Release 2021 R2. HFSS. ANSYS, Canonsburg, PA, 2021. [143] S Wuensch, G Benz, E Crocoll, M Fitsilis, M Neuhaus, TA Scherer, and W Jutzi. Normal and superconductor coplanar waveguides with 100 nm line width. IEEE transactions on applied superconductivity, 11(1):115?118, 2001. [144] W Jutzi, S Wuensch, E Crocoll, M Neuhaus, TA Scherer, T Weimann, and J Niemeyer. Microwave and dc properties of niobium coplanar waveguides with 50-nm linewidth on silicon substrates. IEEE transactions on applied superconductivity, 13(2):320?323, 2003. [145] James Wenner, R Barends, RC Bialczak, Yu Chen, J Kelly, Erik Lucero, Matteo Mariantoni, A Megrant, PJJ O?Malley, D Sank, et al. Surface loss simulations of superconducting coplanar waveguide resonators. Applied Physics Letters, 99(11):113513, 2011. [146] Greg Calusine, Alexander Melville, Wayne Woods, Rabindra Das, Corey Stull, Vlad Bolkhovsky, Danielle Braje, David Hover, David K Kim, Xhovalin Miloshi, et al. Analysis and mitigation of interface losses in trenched superconducting coplanar waveguide resonators. Applied Physics Letters, 112(6):062601, 2018. 206 [147] David B Williams and C Barry Carter. The transmission electron microscope. In Transmission electron microscopy, pages 3?17. Springer, 1996. [148] PW Hawkes. Aberration correction past and present. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 367(1903):3637? 3664, 2009. [149] SD Findlay, Naoya Shibata, Hidetaka Sawada, E Okunishi, Yukihito Kondo, T Yamamoto, and Yuichi Ikuhara. Robust atomic resolution imaging of light elements using scanning transmission electron microscopy. Applied Physics Letters, 95(19):191913, 2009. [150] Stephen J Pennycook and Peter D Nellist. Scanning transmission electron microscopy: imaging and analysis. Springer Science & Business Media, 2011. [151] C Barry Carter and David B Williams. Transmission electron microscopy: Diffraction, imaging, and spectrometry. Springer, 2016. [152] AJ d?Alfonso, B Freitag, D Klenov, and LJ Allen. Atomic-resolution chemical mapping using energy-dispersive x-ray spectroscopy. Physical Review B, 81(10):100101, 2010. [153] M-W Chu, SC Liou, C-P Chang, F-S Choa, and CH Chen. Emergent chemical mapping at atomic-column resolution by energy-dispersive x-ray spectroscopy in an aberration- corrected electron microscope. Physical review letters, 104(19):196101, 2010. [154] Leslie J Allen, Adrian J D?Alfonso, Bert Freitag, and Dmitri O Klenov. Chemical mapping at atomic resolution using energy-dispersive x-ray spectroscopy. MRS bulletin, 37(1):47? 52, 2012. [155] LABS. EAG. EAG, www.eag.com, 2022. [156] Lucille A Giannuzzi and Frederick A Stevie. A review of focused ion beam milling techniques for tem specimen preparation. Micron, 30(3):197?204, 1999. [157] Joachim Mayer, Lucille A Giannuzzi, Takeo Kamino, and Joseph Michael. Tem sample preparation and fib-induced damage. MRS bulletin, 32(5):400?407, 2007. [158] Eiji Okunishi, Isamu Ishikawa, Hidetaka Sawada, Fumio Hosokawa, Madoka Hori, and Yukihito Kondo. Visualization of light elements at ultrahigh resolution by stem annular bright field microscopy. Microscopy and Microanalysis, 15(S2):164?165, 2009. [159] M Varela, Andrew R Lupini, K van Benthem, Albina Y Borisevich, Matthew F Chisholm, Naoya Shibata, E Abe, and Stephen J Pennycook. Materials characterization in the aberration-corrected scanning transmission electron microscope. Annu. Rev. Mater. Res., 35:539?569, 2005. [160] KA Mkhoyan, PE Batson, J Cha, WJ Schaff, and J Silcox. Direct determination of local lattice polarity in crystals. Science, 312(5778):1354?1354, 2006. 207 [161] Peter F Cheng, SM Rossnagel, and David N Ruzic. Directional deposition of cu into semiconductor trench structures using ionized magnetron sputtering. Journal of Vacuum Science & Technology B: Microelectronics and Nanometer Structures Processing, Measurement, and Phenomena, 13(2):203?208, 1995. [162] JK Hulm, CK Jones, RA Hein, and JW Gibson. Superconductivity in the tio and nbo systems. Journal of Low Temperature Physics, 7(3):291?307, 1972. [163] E Belohoubek and E Denlinger. Loss considerations for microstrip resonators (short papers). IEEE Transactions on Microwave Theory and Techniques, 23(6):522?526, 1975. [164] AR Sass and WC Stewart. Self and mutual inductances of superconducting structures. Journal of Applied Physics, 39(4):1956?1963, 1968. [165] William M Fairbank. High frequency surface resistivity of tin in the normal and superconducting states. Physical Review, 76(8):1106, 1949. [166] SJ Fiedziuszko and PD Heidmann. Dielectric resonator used as a probe for high t/sub c/superconductor measurements. In IEEE MTT-S International Microwave Symposium Digest, pages 555?558. IEEE, 1989. [167] Victor Belitsky, C Risacher, Miroslav Pantaleev, and Vessen Vassilev. Superconducting microstrip line model studies at millimetre and sub-millimetre waves. International journal of infrared and millimeter waves, 27(6):809?834, 2006. [168] U-Yen Kongpop, Karwan Rostem, and Edward J Wollack. Modeling strategies for superconducting microstrip transmission line structures. IEEE Transactions on Applied Superconductivity, 28(6):1?5, 2018. [169] Ansoft. HFSSv10UserGuide. Ansoft, http://anlage.umd.edu/HFSSv10UserGuide.pdf, 2005. [170] Kenneth K Mei and G-C Liang. Electromagnetics of superconductors. IEEE transactions on microwave theory and techniques, 39(9):1545?1552, 1991. [171] in high-Tc superconductors the c-axis plasmon can be below the gap frequency Although in conventional low-Tc superconductors ?p lays above the superconducting gap frequency. .see t. m. mishonov, phys. rev. b 44, 12033. 1991. [172] RE Glover III and Ms Tinkham. Conductivity of superconducting films for photon energies between 0.3 and 4 0 k t c. Physical Review, 108(2):243, 1957. [173] Jerzy Krupka, Jonathan Breeze, Anthony Centeno, Neil Alford, Thomas Claussen, and Leif Jensen. Measurements of permittivity, dielectric loss tangent, and resistivity of float- zone silicon at microwave frequencies. IEEE Transactions on microwave theory and techniques, 54(11):3995?4001, 2006. 208 [174] Daniel E Oates, Sergey K Tolpygo, and Vladimir Bolkhovsky. Submicron nb microwave transmission lines and components for single-flux-quantum and analog large-scale superconducting integrated circuits. IEEE Transactions on Applied Superconductivity, 27(4):1?5, 2017. [175] E Maxwell, PM Marcus, and John Clarke Slater. Surface impedance of normal and superconductors at 24,000 megacycles per second. Physical Review, 76(9):1332, 1949. [176] APS Khanna and Y Garault. Determination of loaded, unloaded, and external quality factors of a dielectric resonator coupled to a microstrip line. IEEE Transactions on Microwave Theory and Techniques, 31(3):261?264, 1983. [177] M Stoutimore Khalil, MJA Stoutimore, FC Wellstood, and KD Osborn. An analysis method for asymmetric resonator transmission applied to superconducting devices. Journal of Applied Physics, 111(5):054510, 2012. [178] Anthony Megrant, Charles Neill, Rami Barends, Ben Chiaro, Yu Chen, Ludwig Feigl, Julian Kelly, Erik Lucero, Matteo Mariantoni, Peter JJ O?Malley, et al. Planar superconducting resonators with internal quality factors above one million. Applied Physics Letters, 100(11):113510, 2012. [179] Liepe Research Group. SRIMP. Cornell Univ., https://www.classe.cornell.edu/ liepe/webpage/researchsrimp.html, 2012. [180] A Romanenko, CJ Edwardson, PG Coleman, and PJ Simpson. The effect of vacancies on the microwave surface resistance of niobium revealed by positron annihilation spectroscopy. Applied Physics Letters, 102(23):232601, 2013. [181] V Ganni. Helium refrigeration systems for super-conducting accelerators. In AIP Conference Proceedings, volume 1687, page 040001. AIP Publishing LLC, 2015. [182] ICEoxford. Design and manufacture specialist of ultra low temperature (ult) and high magnetic field equipment for cryogenic research community, 2004. [183] DE Oates, Alfredo C Anderson, and PM Mankiewich. Measurement of the surface resistance of yba2cu3o7- x thin films using stripline resonators. Journal of Superconductivity, 3(3):251?259, 1990. [184] Kent D Irwin and Konrad W Lehnert. Microwave squid multiplexer. Applied physics letters, 85(11):2107?2109, 2004. [185] John Arthur Benson Mates. The microwave squid multiplexer. University of Colorado Boulder, 2011. [186] M Von Schickfus and S Hunklinger. The dielectric coupling of low-energy excitations in vitreous silica to electromagnetic waves. Journal of Physics C: Solid State Physics, 9(16):L439, 1976. 209 [187] Milton Abramowitz and Irene A Stegun. Handbook of mathematical functions with formulas, graphs, and mathematical tables, volume 55. US Government printing office, 1964. [188] J Ja?ckle, L Piche?, W Arnold, and S Hunklinger. Elastic effects of structural relaxation in glasses at low temperatures. Journal of Non-Crystalline Solids, 20(3):365?391, 1976. [189] SciPy. minimize(method=?powell?), 2020 [Online]. [190] T Baumann, MV Schickfus, S Hunklinger, and J Ja?ckle. ?glassy? properties of crystalline li3n at low temperatures. Solid State Communications, 35(8):587?590, 1980. [191] Joshua T West, Arthur Kurlej, Alex Wynn, Chad Rogers, Mark A Gouker, and Sergey K Tolpygo. Wafer-scale characterization of a superconductor integrated circuit process, using a cryogenic wafer prober. arXiv preprint arXiv:2112.00705, 2021. [192] Lake Shore Cryotronics. Supporting advanced scientific research, lake shore is a leading global innovator in measurement and control solutions., 1968. 210